Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-28T15:23:30.183Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 June 2014

Roger Searle
Affiliation:
University of Durham
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
Mid-Ocean Ridges , pp. 258 - 308
Publisher: Cambridge University Press
Print publication year: 2013

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

Abrams, L. J., Detrick, R. S. and Fox, P. J. (1988). Morphology and crustal structure of the Kane fracture zone transverse ridge. Journal of Geophysical Research, 93, 3195–3210.CrossRefGoogle Scholar
Alexander, R. T. and Macdonald, K. C. (1996). Sea Beam, SeaMARC II and ALVIN-based studies of faulting on the East Pacific Rise 9°20ʹ N–9°50ʹ N. Marine Geophysical Researches, 18, (5), 557–587.CrossRefGoogle Scholar
Allen, D. E. and Seyfried, W. E. (2004). Serpentinization and heat generation: constraints from Lost City and Rainbow hydrothermal systems. Geochimica et Cosmochimica Acta, 68, (6), 1347–1354.CrossRefGoogle Scholar
Allerton, S. and MacLeod, C. (1999). New wireline seafloor drill augers well. EOS, Transactions of the American Geophysical Union, 80, (33), 367.CrossRefGoogle Scholar
Allerton, S. and Tivey, M. A. (2001). Magnetic polarity structure of the lower oceanic crust. Geophysical Research Letters, 28, (3), 423–426.CrossRefGoogle Scholar
Allerton, S., Murton, B. J., Searle, R. C. and Jones, M. (1995). Extensional faulting and segmentation of the Mid-Atlantic Ridge north of the Kane fracture zone (24°00ʹ N to 24°40ʹ N). Marine Geophysical Researches, 17, 37–61.CrossRefGoogle Scholar
Allerton, S., Escartin, J. and Searle, R. C. (2000). Extremely asymmetric magmatic accretion of oceanic crust at the ends of slow-spreading ridge-segments. Geology, 28, 179–182.2.0.CO;2>CrossRefGoogle Scholar
Alt, J. C. (1995). Subseafloor processes in mid-ocean ridge hydrothermal systems. In Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions, Geophysical Monograph no. 91, ed. Humphris, S. E., Zierenberg, R. A., Mullineaux, L. S. and Thomson, R. E.. Washington, D.C.: American Geophysical Union, pp. 85–114.Google Scholar
Anderson, E. M. (1951). The Dynamics of Faulting and Dyke Formation with Implications to Britain, 2nd edition. Edinburgh & London: Oliver & Boyd.Google Scholar
Anderson, R. N., Honnorez, J., Becker, K. et al. (1982). DSDP hole-504B, the 1st reference section over 1km through layer-2 of the oceanic-crust. Nature, 300, (5893), 589–594.CrossRefGoogle Scholar
Anderson-Fontana, S., Engeln, J. F., Lundgren, P., Larson, R. L. and Stein, S. (1986). Tectonics and evolution of the Juan Fernandez microplate at the Pacific–Nazca–Antarctic triple junction. Journal of Geophysical Research, 91, 2005–2018.CrossRefGoogle Scholar
Anonymous (1993). News from national ridge research programs. InterRidge News, 2, (1), 20–25.Google Scholar
Atwater, T. (1970). Implications of plate tectonics for the Cenozoic tectonic evolution of western North America. Geological Society of America Bulletin, 81, 3513–3536.CrossRefGoogle Scholar
Auzende, J.-M., Bideau, D., Bonatti, E. et al. (1989). Direct observation of a section through slow-spreading oceanic crust. Nature, 337, 726–729.CrossRefGoogle Scholar
Baines, A. G., Cheadle, M. J., John, B. E. and Schwartz, J. J. (2008). The rate of detachment faulting at Atlantis Bank, SW Indian Ridge. Earth and Planetary Science Letters, 273, 105–114.CrossRefGoogle Scholar
Baker, E. T. (1994). A 6-year time-series of hydrothermal plumes over the Cleft segment of the Juan-de-Fuca Ridge. Journal of Geophysical Research-Solid Earth, 99, (B3), 4889–4904.CrossRefGoogle Scholar
Baker, E. T. and German, C. (2004). On the global distribution of hydrothermal vent fields. In Mid-Ocean Ridges: Hydrothermal Interactions Between the Lithosphere and Oceans, Geophysical Monograph 148, ed. German, C., Lin, J. and Parson, L. M.. Washington, D.C.: American Geophysical Union, pp. 1–18.Google Scholar
Baker, E. T., Massoth, G. J. and Feely, R. A. (1987). Cataclysmic hydrothermal venting on the Juan-de-Fuca ridge. Nature, 329, (6135), 149–151.CrossRefGoogle Scholar
Baker, E. T., Lavelle, J. W., Feely, R. A., Massoth, G. J. and Walker, S. L. (1989). Episodic venting of hydrothermal fluids from the Juan de Fuca ridge. Journal of Geophysical Research-Solid Earth and Planets, 94, (B7), 9237–9250.CrossRefGoogle Scholar
Baker, E. T., Feely, R. A., Mottl, M. J. et al. (1994). Hydrothermal plumes along the East Pacific Rise, 8°40′ to 11°50ʹ N – plume distribution and relationship to the apparent magmatic budget. Earth and Planetary Science Letters, 128, (1–2), 1–17.CrossRefGoogle Scholar
Baker, E. T., German, C. R. and Elderfield, H. (1995a). Hydrothermal plumes: global distributions and geological inferences. In Seafloor Hydrothermal Systems: Physical, Chemical, Biological and Geological Interactions, Geophysical Monograph 91, ed. Humphris, S. E., Zierenberg, R. A., Mullineaux, L. S. and Thomson, R. E.. Washington, D.C.: American Geophysical Union, pp. 47–71.Google Scholar
Baker, E. T., Massoth, G. J., Feely, R. A. et al. (1995b). Hydrothermal event plumes from the Coaxial sea-floor eruption site, Juan de Fuca ridge. Geophysical Research Letters, 22, (2), 147–150.CrossRefGoogle Scholar
Baker, E. T., Hey, R. N., Lupton, J. E. et al. (2002). Hydrothermal venting along Earth's fastest spreading center: East Pacific Rise, 27.5°–32.3° S. Journal of Geophysical Research, 107, (B7), EPM2, .CrossRefGoogle Scholar
Ballard, R. D. (1993). The Medea-Jason Remotely Operated Vehicle system. Deep-Sea Research Part I-Oceanographic Research Papers, 40, (8), 1673.CrossRefGoogle Scholar
Ballard, R. D. and Moore, J. G. (1977). Photographic Atlas of the Mid-Atlantic Ridge Rift Valley. Heidelberg: Springer-Verlag, 114 pp.CrossRefGoogle Scholar
Ballard, R. D. and Van Andel, T. H. (1977). Morphology and tectonics of the inner rift valley at lat. 36°50ʹ N on the Mid-Atlantic Ridge. Geological Society of America Bulletin, 88, 507–530.2.0.CO;2>CrossRefGoogle Scholar
Ballard, R. D., Bryan, W. B., Heirtzler, J. R. et al. (1975). Manned submersible observations in FAMOUS area – Mid-Atlantic Ridge. Science, 190, (4210), 103–108.CrossRefGoogle Scholar
Ballard, R. D., Holcomb, R. T. and Van Andel, T. H. (1979). The Galapagos Rift at 86° W: 3. Sheet flows, collapse pits, and lava lakes of the rift valley. Journal of Geophysical Research, 84, (B10), 5407–5422.CrossRefGoogle Scholar
Ballard, R. D., Van Andel, T. H. and Holcomb, R. T. (1982). The Galapagos Rift at 86° W: 5. Variations in volcanism, structure and hydrothermal activity along a 30-kilometer segment of the rift valley. Journal of Geophysical Research, 87, (B2), 1149–1161.CrossRefGoogle Scholar
Ballu, V., Dubois, J., Deplus, C., Diament, M. and Bonvalor, S. (1998). Crustal structure of the Mid-Atlantic Ridge south of Kane Fracture Zone from seafloor and sea surface gravity data. Journal of Geophysical Research, B, 103, (2), 2615–2631.CrossRefGoogle Scholar
Baragar, W., Lambert, M., Baglow, N. and Gibson, I. (1990). The sheeted dyke zone in the Troodos ophiolite. In Proceedings of the Symposium on Ophiolites and Oceanic Lithosphere – TROODOS 87, ed. Malpas, J., Moores, E. M., Panayiotou, A. and Xenophontos, C.. Nicosia, Cyprus: Geological Survey Department, Ministry of Agriculture and Natural Resources, pp. 37–51.Google Scholar
Baran, J. M., Cochran, J. R., Carbotte, S. M. and Nedimovic, M. R. (2005). Variations in upper crustal structure due to variable mantle temperature along the Southeast Indian Ridge. Geochemistry Geophysics Geosystems, 6, Q11002, .CrossRefGoogle Scholar
Barclay, A. H., Toomey, D. R. and Solomon, S. C. (1998). Seismic structure and crustal magmatism at the Mid-Atlantic Ridge, 35° N. Journal of Geophysical Research, 103, (B8), 17827–17844.CrossRefGoogle Scholar
Barclay, A. H., Toomey, D. R. and Solomon, S. C. (2001). Microearthquake characteristics and crustal Vp/Vs structure at the Mid-Atlantic Ridge, 35° N. Journal of Geophysical Research-Solid Earth, 106, (B2), 2017–2034.CrossRefGoogle Scholar
Baross, J. A. and Hoffman, S. E. (1985). Submarine hydrothermal vents and associated gradient environments as sites for the origin and evolution of life. Origins of Life and Evolution of the Biosphere, 15, (4), 327–345.CrossRefGoogle Scholar
Batiza, R., Fox, P. J., Vogt, P. R. et al. (1989). Morphology, abundance, and chemistry of near-ridge seamounts in the vicinity of the Mid-Atlantic Ridge ~26° S. Journal of Geology, 97, 209–220.CrossRefGoogle Scholar
Becker, K., Sakai, H., Adamson, A. C. et al. (1989). Drilling deep into young oceanic-crust, Hole-504B, Costa-Rica Rift. Reviews of Geophysics, 27, (1), 79–102.CrossRefGoogle Scholar
Behn, M. D. and Ito, G. (2008). Magmatic and tectonic extension at mid-ocean ridges: 1. Controls on fault characteristics. Geochemistry Geophysics Geosystems, 9, .CrossRefGoogle Scholar
Bell, D. R. and Rossman, G. R. (1992). Water in Earth's mantle – the role of nominally anhydrous minerals. Science, 255, (5050), 1391–1397.CrossRefGoogle ScholarPubMed
Bell, R. E. and Buck, W. R. (1992). Crustal control of ridge segmentation inferred from observations of the Reykjanes Ridge. Nature, 357, 583–586.CrossRefGoogle Scholar
Beltenev, V., Ivanov, V., Rozhdestvenskaya, I. et al. (2007). New data about hydrothermal fields on the Mid-Atlantic Ridge between 11°–14° N: 32nd cruise of R/V Professor Logatchev. InterRidge News, 18, 14–18.Google Scholar
Bergman, E. A. and Solomon, S. C. (1988). Transform fault earthquakes in the North Atlantic: source mechanisms and depth of faulting. Journal of Geophysical Research, B, 93, (8), 9027–9057.CrossRefGoogle Scholar
Bessonova, E. N., Fishman, V. M., Ryaboyi, V. Z. and Sitnikova, G. A. (1974). The tau method for inversion of travel times – I. Deep seismic sounding data. Geophysical Journal of the Royal Astronomical Society, 36, (2), 377–398.CrossRefGoogle Scholar
Best, M. G. and Christiansen, E. H. (2001). Igneous Petrology. Oxford: Blackwell Science, 458 pp.Google Scholar
Bicknell, J. D., Sempere, J.-C. and Macdonald, K. C. (1988). Tectonics of a fast spreading center: a Deep-Tow and Sea Beam survey on the East Pacific Rise at 19°30ʹ S. Marine Geophysical Researches, 9, (1), 25–45.Google Scholar
Bird, R. T., Naar, D. F., Larson, R. L., Searle, R. C. and Scotese, C. R. (1998). Plate tectonic reconstructions of the Juan Fernandez microplate: transformation from internal shear to rigid rotation. Journal of Geophysical Research, 103, (B4), 7049–7067.CrossRefGoogle Scholar
Bird, R. T., Tebbens, S. F. and Kleinrock, M. C. (1999). Episodic triple-junction migration by rift propagation and microplates. Geology, 27, 911–914.2.3.CO;2>CrossRefGoogle Scholar
Bischoff, J. L. and Rosenbauer, R. J. (1988). Liquid–vapor relations in the critical region of the system NaCl–H2O from 380 °C to 415 °C – a refined determination of the critical-point and 2-phase boundary of seawater. Geochimica et Cosmochimica Acta, 52, (8), 2121–2126.CrossRefGoogle Scholar
Blackinton, J. G., Hussong, D. M. and Kosslos, J. (1983). First results from a combination side scan sonar and seafloor mapping system (SeaMARC II). In OffshoreTechnology Conference, OTC 4478, pp. 307–311.CrossRef
Blackman, D. K. and Forsyth, D. W. (1991). Isostatic compensation of tectonic features of the Mid-Atlantic Ridge: 25–27°30ʹ S. Journal of Geophysical Research, 96, 11741–11758.CrossRefGoogle Scholar
Blackman, D. K., Karson, J. A., Kelley, D. S. et al. (2002). Geology of the Atlantis Massif (Mid-Atlantic Ridge, 30° N): implications for the evolution of an ultramafic oceanic core complex. Marine Geophysical Researches, 23, (5–6), 443–469.CrossRefGoogle Scholar
Blackman, D. K., Ildefonse, B., John, B. E. et al. (2006). Oceanic core complex formation, Atlantis Massif: Expeditions 304 and 305 of the riserless drilling platform from and to Ponta Delgada, Azores (Portugal), Sites U1309–U1311, 17 November 2004–7 January 2005, and from and to Ponta Delgada, Azores (Portugal), Site U1309, 7 January–2 March 2005. Proceedings of the Integrated Ocean Drilling Program, 304/305, .Google Scholar
Blackman, D. K., Karner, G. D. and Searle, R. C. (2008). Three-dimensional structure of oceanic core complexes: effects on gravity signature and ridge flank morphology, Mid-Atlantic Ridge 30° N. Geochemistry Geophysics Geosystems, 9, (6), Q06007, .CrossRefGoogle Scholar
Blackman, D. K., Canales, J. P. and Harding, A. (2009). Geophysical signatures of oceanic core complexes. Geophysical Journal International, 178 (2), 593–613.CrossRefGoogle Scholar
Blackman, D. K., Ildefonse, B., John, B. E. et al. (2011). Drilling constraints on lithospheric accretion and evolution at Atlantis Massif, Mid-Atlantic Ridge 30° N. Journal of Geophysical Research, 116, .CrossRefGoogle Scholar
Blakely, R. and Cox, A. (1972). Identification of short polarity events by transforming marine magnetic profiles to the pole. Journal of Geophysical Research, 77, 4339–4349.CrossRefGoogle Scholar
Bohnenstiehl, D. R. and Carbotte, S. M. (2001). Faulting patterns near 19°30ʹ S on the East Pacific Rise: fault formation and growth at a superfast spreading center. Geochemistry Geophysics Geosystems, 2, art. no. 2001GC000156.CrossRefGoogle Scholar
Bonatti, E. (1976). Serpentine protrusions in the oceanic crust. Earth and Planetary Science Letters, 32, 107–113.CrossRefGoogle Scholar
Bonatti, E. (1978). Vertical tectonism in oceanic fracture zones. Earth and Planetary Science Letters, 37, 369–379.CrossRefGoogle Scholar
Bonatti, E. and Harrison, C. G. A. (1988). Eruption styles of basalt in oceanic spreading ridges and seamounts – effect of magma temperature and viscosity. Journal of Geophysical Research-Solid Earth and Planets, 93, (B4), 2967–2980.CrossRefGoogle Scholar
Bostrom, K., Peterson, M., Joensuu, O. and Fisher, D. (1969). Aluminum-poor ferromanganoan sediments on active oceanic ridges. Journal of Geophysical Research, 74 (12), 3261–3270.CrossRefGoogle Scholar
Bott, M. H. P. (1982). The Interior of the Earth, its Structure, Constitution and Evolution, 2nd edition. London: Edward Arnold.Google Scholar
Boudier, F. and Nicolas, A. (1988). The ophiolites of Oman – preface. Tectonophysics, 151, (1–4), R7–R8.Google Scholar
Bougault, H., Aballea, M., Radford-Knoery, J. et al. (1998). FAMOUS and AMAR segments on the Mid-Atlantic Ridge: ubiquitous hydrothermal Mn, CH4, delta He-3 signals along the rift valley walls and rift offsets. Earth and Planetary Science Letters, 161, (1–4), 1–17.CrossRefGoogle Scholar
Bowen, A. N. and White, R. S. (1986). Deep-tow seismic profiles from the Vema transform and ridge-transform intersection. Journal of the Geological Society of London, 143, 807–817.CrossRefGoogle Scholar
Bown, J. and White, R. (1994). Variation with spreading rate of oceanic crustal thickness and geochemistry. Earth and Planetary Science Letters, 121, (3–4), 435–449.CrossRefGoogle Scholar
Bratt, S. R. and Purdy, G. M. (1984). Structure and variability of oceanic crust on the flanks of the East Pacific Rise between 11° and 13° N. Journal of Geophysical Research, 89, 6111–6125.CrossRefGoogle Scholar
Briais, A., Sloan, H., Parson, L. and Murton, B. (2000). Accretionary processes in the axial valley of the Mid-Atlantic Ridge 27°– 30° N from TOBI side-scan sonar images. Marine Geophysical Researches, 21, 87–119.CrossRefGoogle Scholar
Bryan, W. B., Humphris, S. E., Thompson, G. and Casey, J. F. (1994). Comparative volcanology of small axial eruptive centers in the MARK area. Journal of Geophysical Research, 99, 2973–2984.CrossRefGoogle Scholar
Buck, W. R. (1988). Flexural rotation of normal faults. Tectonics, 7, (5), 959–973.CrossRefGoogle Scholar
Buck, W. R. and Poliakov, A. N. B. (1998). Abyssal hills formed by stretching oceanic lithosphere. Nature, 392, 272–275.CrossRefGoogle Scholar
Buck, W. R. and Su, W. (1989). Focused mantle upwelling below mid-ocean ridges due to feedback between viscosity and melting. Geophysical Research Letters, 16, 641–644.CrossRefGoogle Scholar
Buck, W. R., Lavier, L. L. and Poliakov, A. N. B. (2005). Modes of faulting at mid-ocean ridges. Nature, 434, 719–723.CrossRefGoogle ScholarPubMed
Bullard, E. C. (1952). Heat flow through the floor of the eastern North Pacific ocean. Nature, 170, (4318), 199–200.CrossRefGoogle Scholar
Bullard, E. C. (1963). The flow of heat through the floor of the ocean. In The Sea, volume 3, ed. Hill, M. N.. New York: Interscience Publishers, pp. 218–232.Google Scholar
Bullard, E. C. and Mason, R. G. (1961). The magnetic field astern of a ship. Deep-Sea Research, 8, 20–27.CrossRefGoogle Scholar
Bullard, E. C. and Mason, R. G. (1963). The magnetic field over the oceans. In The Sea, volume 3, ed. Hill, M. N.. New York: Interscience Publishers, pp. 175–217.Google Scholar
Bullard, E. C., Maxwell, A. E. and Revelle, R. (1956). Heat flow through the deep sea floor. In Advances in Geophysics, ed. Landsberg, H. E.. New York: Academic Press, pp. 153–181.Google Scholar
Byerlee, J. D. (1978). Friction of rocks. Pure and Applied Geophysics, 116, 615–626.CrossRefGoogle Scholar
Canales, J. P., Detrick, R. S., Toomey, D. R. and Wilcock, W. S. D. (2003). Segment-scale variations in the crustal structure of 150–300 kyr old fast spreading oceanic crust (East Pacific Rise, 8°15ʹ N–10°5ʹ N) from wide-angle seismic refraction profiles. Geophysical Journal International, 152, (3), 766–794.CrossRefGoogle Scholar
Canales, J. P., Singh, S. C., Detrick, R. S. et al. (2006). Seismic evidence for variations in axial magma chamber properties along the southern Juan de Fuca Ridge. Earth and Planetary Science Letters, 246, (3–4), 353–366.CrossRefGoogle Scholar
Canales, J. P., Tucholke, B. E., Xu, M., Collins, J. A. and DuBois, D. L. (2008). Seismic evidence for large-scale compositional heterogeneity of oceanic core complexes. Geochemistry Geophysics Geosystems, 9, (8), .CrossRefGoogle Scholar
Canales, J. P., Nedimovic, M. R., Kent, G. M., Carbotte, S. M. and Detrick, R. S. (2009). Seismic reflection images of a near-axis melt sill within the lower crust at the Juan de Fuca ridge. Nature, 460, (7251), 89–99.CrossRefGoogle ScholarPubMed
Cande, S. and Kent, D. (1995). Revised calibration of the geomagnetic polarity timescale for the late Cretaceous and Cenozoic. Journal of Geophysical Research, 100, (B4), 6093–6095.CrossRefGoogle Scholar
Cann, J. R. (1970). New model for structure of ocean crust. Nature, 226, (5249), 928–930.CrossRefGoogle ScholarPubMed
Cann, J. R. (1974). Model for oceanic crustal structure developed. Geophysical Journal of the Royal Astronomical Society, 39, (1), 169–187.CrossRefGoogle Scholar
Cann, J. R. and Smith, D. K. (2005). Evolution of volcanism and faulting in a segment of the Mid-Atlantic Ridge at 25° N. Geochemistry Geophysics Geosystems, 6, .CrossRefGoogle Scholar
Cann, J. R. and Strens, M. R. (1982). Black smokers fuelled by freezing magma. Nature, 298, 147–149.CrossRefGoogle Scholar
Cann, J. R., Strens, M. R. and Rice, A. (1985). A simple magma-driven thermal balance model for the formation of volcanogenic massive sulphides. Earth and Planetary Science Letters, 76, 123–134.CrossRefGoogle Scholar
Cann, J. R., Blackman, D. K., Smith, D. K. et al. (1997). Corrugated slip surfaces formed at ridge-transform intersections on the Mid Atlantic Ridge. Nature, 385, 329–332.CrossRefGoogle Scholar
Cann, J. R., Elderfield, H. and Laughton, A. S. (1999). Mid-Ocean Ridges; Dynamics of Processes Associated with the Creation of New Oceanic Crust. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Cann, J. R., Prichard, H. M., Malpas, J. G. and Xenophontos, C. (2001). Oceanic inside corner detachments of the Limassol Forest area, Troodos ophiolite, Cyprus. Journal of the Geological Society, 158, 757–767.CrossRefGoogle Scholar
Cannat, M., Mevel, C., Maia, M. et al. (1995). Thin crust, ultramafic exposures, and rugged faulting patterns at Mid-Atlantic Ridge (22°–24° N). Geology, 23, (1), 49–52.2.3.CO;2>CrossRefGoogle Scholar
Cannat, M., Rommevaux-Jestin, C., Sauter, D., Deplus, C. and Mendel, V. (1999). Formation of the axial relief at the very slow spreading Southwest Indian Ridge (49° to 69° E). Journal of Geophysical Research, 104, (B10), 22825–22843.CrossRefGoogle Scholar
Cannat, M., Cann, J. and Maclennan, J. (2004). Some hard rock constraints on the supply of heat to mid-ocean ridges. In Mid-Ocean Ridges: Hydrothermal Interactions between the Lithosphere and Oceans, Geophysical Monograph, 148, ed. German, , Lin, C. R., J. and Parson, L. M., Washington, D.C.: AGU, pp. 111–149.Google Scholar
Cannat, M., Sauter, D., Mendel, V. et al. (2006). Modes of seafloor generation at a melt-poor ultraslow-spreading ridge. Geology, 34, (7), 605–608; .CrossRefGoogle Scholar
Carbotte, S. M. and Macdonald, K. C. (1990). Causes of variation in fault-facing direction on the ocean floor. Geology, 18, 749–752.2.3.CO;2>CrossRefGoogle Scholar
Carbotte, S. and Macdonald, K. (1992). East Pacific Rise 8°–10°30ʹ N: evolution of ridge segments and discontinuities from SeaMARC II and three-dimensional magnetic studies. Journal of Geophysical Research, 97, (B5), 6959–6982.CrossRefGoogle Scholar
Carbotte, S. M. and Macdonald, K. C. (1994a). The axial topographic high at intermediate and fast spreading ridges. Earth and Planetary Science Letters, 128, (3–4), 85–97.CrossRefGoogle Scholar
Carbotte, S. M. and Macdonald, K. C. (1994b). Comparison of seafloor tectonic fabric at intermediate, fast, and super fast spreading ridges: influence of spreading rate, plate motions, and ridge segmentation on fault patterns. Journal of Geophysical Research, 99, (B7), 13609–13631.CrossRefGoogle Scholar
Carbotte, S. M. and Scheirer, D. S. (2004). Variability of ocean crustal structure created along the global mid-ocean ridge. In Hydrogeology of the Oceanic Lithosphere, ed. Davis, E. E. and Elderfield, H.. Cambridge: Cambridge University Press, pp. 59–107.Google Scholar
Carbotte, S. M., Detrick, R. S., Harding, A. et al. (2006). Rift topography linked to magmatism at the intermediate spreading Juan de Fuca Ridge. Geology, 34, (3), 209–212.CrossRefGoogle Scholar
Carlson, R. L. (1998). Seismic velocities in the uppermost oceanic crust: age dependence and the fate of layer 2A. Journal of Geophysical Research-Solid Earth, 103, (B4), 7069–7077.CrossRefGoogle Scholar
Carlson, R. L. (2011). The effect of hydrothermal alteration on the seismic structure of the upper oceanic crust: evidence from Holes 504B and 1256D. Geochemistry Geophysics Geosystems, 12, .CrossRefGoogle Scholar
Carpine-Lancre, J. (2001). Oceanographic Sovereigns: Prince Albert I of Monaco and King Carlos I of Portugal. In Understanding the Oceans: A Century of Ocean Exploration, ed. Deacon, M., Rice, T. and Summerhayes, C.. London: UCL Press, pp. 56–68.Google Scholar
Casey, J. F. and Karson, J. A. (1981). Magma chamber profiles from the Bay-of-Islands ophiolite complex. Nature, 292, (5821), 295–301.CrossRefGoogle Scholar
Chadwick, W. W. and Embley, R. W. (1994). Lava flows from a mid-1980s submarine eruption on the cleft segment, Juan-de-Fuca Ridge. Journal of Geophysical Research-Solid Earth, 99, (B3), 4761–4776.CrossRefGoogle Scholar
Chadwick, W. W., Embley, R. W. and Fox, C. G. (1995). Seabeam depth changes associated with recent lava flows, Coaxial segment, Juan-de-Fuca Ridge – evidence for multiple eruptions between 1981–1993. Geophysical Research Letters, 22, (2), 167–170.CrossRefGoogle Scholar
Chadwick, W. W., Embley, R. W. and Shank, T. M. (1998). The 1996 Gorda Ridge eruption: geologic mapping, sidescan sonar, and SeaBeam comparison results. Deep-Sea Research Part II-Topical Studies in Oceanography, 45, (12), 2547–2569.CrossRefGoogle Scholar
Chadwick, W. W., Embley, R. W., Milburn, H. B., Meinig, C. and Stapp, M. (1999). Evidence for deformation associated with the 1998 eruption of Axial Volcano, Juan de Fuca Ridge, from acoustic extensometer measurements. Geophysical Research Letters, 26, (23), 3441–3444.CrossRefGoogle Scholar
Chadwick, W. W., Nooner, S. L., Zumberge, M. A., Embley, R. W. and Fox, C. G. (2006). Vertical deformation monitoring at Axial Seamount since its 1998 eruption using deep-sea pressure sensors. Journal of Volcanology and Geothermal Research, 150, (1–3), 313–327.CrossRefGoogle Scholar
Chapman, C. H. and Drummond, R. (1982). Body-wave seismograms in inhomogeneous media using Maslov asymptotic theory. Bulletin of the Seismological Society of America, 72, S277–S317.Google Scholar
Chave, A. D. and Cox, C. S. (1982). Controlled electromagnetic sources for measuring electrical-conductivity beneath the oceans .1. Forward problem and model study. Journal of Geophysical Research, 87, (B7), 5327–5338.CrossRefGoogle Scholar
Chayes, D. N. (1983). Evolution of Sea MARC I. In IEEE Proceedings of the 3rd Working Symposium on Oceanographic Data Systems. New York: IEEE Computer Society Press, pp. 103–108.Google Scholar
Chen, Y. J. (1992). Oceanic crustal thickness versus spreading rate. Geophysical Research Letters, 19, (8), 753–756.CrossRefGoogle Scholar
Chen, Y. (2004). Modelling the thermal structure of the oceanic crust. In Mid-Ocean Ridges: Hydrothermal Interactions between the Lithosphere and Oceans, Geophysical Monograph 148, ed. German, C., Lin, J. and Parson, L. M.. Washington, D.C.: American Geophysical Union, pp. 95–110.Google Scholar
Chen, Y. S. J. and Lin, J. (2004). High sensitivity of ocean ridge thermal structure to changes in magma supply: the Galapagos Spreading Center. Earth and Planetary Science Letters, 221, (1–4), 263–273.CrossRefGoogle Scholar
Chen, Y. and Morgan, W. J. (1990a). A nonlinear rheology model for mid-ocean ridge axis topography. Journal of Geophysical Research, 95, 17583–17604.CrossRefGoogle Scholar
Chen, Y. and Morgan, W. J. (1990b). Rift valley/no rift valley transition at mid-ocean ridges. Journal of Geophysical Research, 95, 17571–17581.CrossRefGoogle Scholar
Chen, Y. J. and Phipps Morgan, J. (1996). The effects of spreading rate, the magma budget, and the geometry of magma emplacement on the axial heat flux at mid-ocean ridges. Journal of Geophysical Research, 101, (B5), 11475–11482.CrossRefGoogle Scholar
Chesterman, W. D., Clynick, P. R. and Stride, A. H. (1958). An acoustic aid to sea bed survey. Acustica, 8, 285–290.Google Scholar
Chin, C. S., Coale, K. H., Elrod, V. A. et al. (1994). In-situ observations of dissolved iron and manganese in hydrothermal vent plumes, Juan-de-Fuca Ridge. Journal of Geophysical Research-Solid Earth, 99, (B3), 4969–4984.CrossRefGoogle Scholar
Christeson, G. L., Purdy, G. M. and Fryer, G. J. (1994). Seismic constraints on shallow crustal emplacement processes at the fast spreading East Pacific Rise. Journal of Geophysical Research, 99, (B9), 17957–17973.CrossRefGoogle Scholar
Christeson, G. L., McIntosh, K. D. and Karson, J. A. (2010). Inconsistent correlation of seismic layer 2a and lava layer thickness in oceanic crust. Nature, 445, 418–421.CrossRefGoogle Scholar
Clague, D. A., Moore, J. G. and Reynolds, J. R. (2000a). Formation of submarine flat-topped volcanic cones in Hawai'i. Bulletin of Volcanology, 62, (3), 214–233.CrossRefGoogle Scholar
Clague, D. A., Reynolds, J. R. and Davis, A. S. (2000b). Near-ridge seamount chains in the northeastern Pacific Ocean. Journal of Geophysical Research-Solid Earth, 105, (B7), 16541–16561.CrossRefGoogle Scholar
Clague, D. A., Paduan, J. B. and Davis, A. S. (2009). Widespread strombolian eruptions of mid-ocean ridge basalt. Journal of Volcanology and Geothermal Research, 180, (2–4), 171–188.CrossRefGoogle Scholar
Cochran, J. R. (1979). An analysis of isostasy in the worlds oceans: 2. Mid-ocean ridge crests. Journal of Geophysical Research, 84, 4713–4729.CrossRefGoogle Scholar
Cochran, J. R. (1986). Variations in subsidence rates along intermediate and fast spreading midocean ridges. Geophysical Journal of the Royal Astronomical Society, 87, (2), 421–454.CrossRefGoogle Scholar
Cochran, J. R. (2008). Seamount volcanism along the Gakkel Ridge, Arctic Ocean. Geophysical Journal International, 174, 1153–1173.CrossRefGoogle Scholar
Cochran, J. R., Kurras, G. J., Edwards, M. H. and Coakley, B. J. (2003). The Gakkel Ridge: bathymetry, gravity anomalies, and crustal accretion at extremely slow spreading rates. Journal of Geophysical Research, 108, (B2), .CrossRefGoogle Scholar
Coffin, M. F. and 30 others (2001). Earth, Oceans and Life: Scientific Investigation of the Earth System Using Multiple Drilling Platforms and New Technologies. Integrated Ocean Drilling Program Initial Science Plan, 2003–2013. Washington, D.C., 20036–2102: IODP International Working Group Support Office.
Cogné, J. P., Francheteau, J., Coutillot, V. et al. (1995). Large rotation of the Easter microplate as evidenced by oriented paleomagnetic samples from the ocean floor. Earth and Planetary Science Letters, 136, 213–222.CrossRefGoogle Scholar
Collette, B. J. (1974). Thermal contraction joints in a spreading seafloor as origin of fracture zones. Nature, 251, 299–300.CrossRefGoogle Scholar
Collier, J. S. and Singh, S. C. (1998). Poisson's ratio structure of young oceanic crust. Journal of Geophysical Research, B, 103, 20981–20996.CrossRefGoogle Scholar
Collier, J. and Sinha, M. (1990). Seismic images of a magma chamber beneath the Lau Basin back-arc spreading center. Nature, 346, (6285), 646–648.CrossRefGoogle Scholar
Collins, J. A., Blackman, D. K., Harris, A. and Carlson, R. L. (2009). Seismic and drilling constraints on velocity structure and reflectivity near IODP Hole U1309D on the central dome of Atlantis Massif, Mid-Atlantic Ridge 30° N. Geochemistry Geophysics Geosystems, 10, .CrossRefGoogle Scholar
Connelly, D., Copley, J., Murton, B. et al. (2012). Hydrothermal vent fields and chemosynthetic biota on the world's deepest seafloor spreading centre. Nature Communications, 3, .CrossRefGoogle ScholarPubMed
Cooper, K. M., Goldstein, S. J., Sims, K. W. W. and Murrell, M. T. (2003). Uranium-series chronology of Gorda Ridge volcanism: new evidence from the 1996 eruption. Earth and Planetary Science Letters, 206, (3–4), 459–475.CrossRefGoogle Scholar
Corliss, J. B. (1990). Hot-springs and the origin of life. Nature, 347, (6294), 624.CrossRefGoogle Scholar
Corliss, J. B., Dymond, J., Gordon, L. I. et al. (1979). Submarine thermal springs on the Galapagos Rift. Science, 203, 1073–1083.CrossRefGoogle ScholarPubMed
Cormier, M.-H. (1997). The ultra-fast East Pacific Rise: instability of the plate boundary and implications for accretionary processes. Philosophical Transactions of the Royal Society of London, Series A, 355, 341–367.CrossRefGoogle Scholar
Cormier, M.-H. and Macdonald, K. C. (1994). East Pacific Rise at 18°–19° S: asymmetric spreading and ridge reorientation by ultrafast migration of axial discontinuities. Journal of Geophysical Research, 99, (B1), 543–564.CrossRefGoogle Scholar
Cormier, M.-H., Scheirer, D. S. and Macdonald, K. C. (1996). Evolution of the East Pacific Rise at 16°–19° S since 5 Ma: bisection of overlapping spreading centers by new, rapidly propagating ridge segments. Marine Geophysical Researches, 18, 53–84.CrossRefGoogle Scholar
Cormier, M. H., Ryan, W. B. F., Shah, A. K. et al. (2003). Waxing and waning volcanism along the East Pacific Rise on a millennium time scale. Geology, 31, (7), 633–636.2.0.CO;2>CrossRefGoogle Scholar
Courtillot, V. (1982). Propagating rifts and continental breakup. Tectonics, 1, 239–250.CrossRefGoogle Scholar
Cowie, P. A., Vanneste, C. and Sornette, D. (1993a). Statistical physics model for the spatio-temporal evolution of faults. Journal of Geophysical Research, 98, 21809–21821.CrossRefGoogle Scholar
Cowie, P. A., Scholz, C. H., Edwards, M. and Malinverno, A. (1993b). Fault strain and seismic coupling on mid-ocean ridges. Journal of Geophysical Research, 98, 17911–17920.CrossRefGoogle Scholar
Cowie, P. A., Malinverno, A., Ryan, W. B. F. and Edwards, M. H. (1994). Quantitative fault studies on the East Pacific Rise: a comparison of sonar imaging techniques. Journal of Geophysical Research, 99, (B8), 15205–15218.CrossRefGoogle Scholar
Cox, A. and Hart, R. B. (1986). Plate Tectonics – How It Works. Blackwell Scientific Publications, Inc., 392 pp.Google Scholar
Cox, A., Dalrymple, G. B. and Doell, R. R. (1963). Geomagnetic polarity epochs and pleistocene geochronometry. Nature, 198, (488), 1049–1051.CrossRefGoogle Scholar
Crane, K. (1976). The intersection of the Siqueiros transform fault and the East Pacific Rise. Marine Geology, 21, 25–46.CrossRefGoogle Scholar
Crane, K. (1987). Structural evolution of the East Pacific Rise axis from 13°10ʹ N to 10°35ʹ N – interpretations from SeaMARC-I data. Tectonophysics, 136, (1–2), 65–124.CrossRefGoogle Scholar
Crawford, W. C. and Webb, S. C. (2002). Variations in the distribution of magma in the lower crust and at the Moho beneath the East Pacific Rise at 9°–10° N. Earth and Planetary Science Letters, 203, (1), 117–130.CrossRefGoogle Scholar
Crawford, W. C., Webb, S. C. and Hildebrand, J. A. (1991). Sea-floor compliance observed by long-period pressure and displacement measurements. Journal of Geophysical Research-Solid Earth, 96, (B10), 16151–16160.CrossRefGoogle Scholar
Crawford, W. C., Webb, S. C. and Hildebrand, J. A. (1998). Estimating shear velocities in the oceanic crust from compliance measurements by two-dimensional finite difference modeling. Journal of Geophysical Research, B, 103, (5), 9895–9916.CrossRefGoogle Scholar
Crawford, W. C., Webb, S. C. and Hildebrand, J. A. (1999). Constraints on melt in the lower crust and Moho at the East Pacific Rise, 9°48ʹ N, using seafloor compliance measurements. Journal of Geophysical Research-Solid Earth, 104, (B2), 2923–2939.CrossRefGoogle Scholar
Crawford, W. C., Singh, S. C., Seher, T. et al. (2010). Crustal structure, magma chamber, and faulting beneath the Lucky Strike hydrothermal vent field. In Diversity of Hydrothermal Systems on Slow Spreading Ocean Ridges, Geophysical Monograph 188, ed. Rona, P. A., Devey, C. W., Dyment, J. and Murton, B. J.. Washington, D.C.: American Geophysical Union, pp. 113–152.CrossRefGoogle Scholar
Crowder, L. K. and Macdonald, K. C. (2000). New constraints on the width of the zone of active faulting on the East Pacific Rise 8°30ʹ N–10°00ʹ N from Sea Beam Bathymetry and SeaMARC II Side-scan Sonar. Marine Geophysical Researches 21, (6), 513–527.CrossRefGoogle Scholar
CYAMEX (1981). First manned submersible dives on the East Pacific Rise at 21° N (Project RITA): general results. Marine Geophysical Researches, 4, 345–379.CrossRefGoogle Scholar
Danchik, R. J. (1984). The navy navigation satellite system (Transit). Johns Hopkins APL Technical Digest, 5, (4), 323–329.Google Scholar
Davis, E. E. (1982). Evidence for extensive basalt flows on the sea-floor. Geological Society of America Bulletin, 93, (10), 1023–1029.2.0.CO;2>CrossRefGoogle Scholar
Davis, E. E. and Lister, C. R. B. (1974). Fundamentals of ridge crest topography. Earth and Planetary Science Letters, 21, 405–413.CrossRefGoogle Scholar
Delaney, J. R. (1989). RIDGE Initial Science Plan February 1989. Seattle: School of Oceanography, University of Washington, 90 pp.
deMartin, B. J., Sohn, R. A., Canales, J. P. and Humphris, S. E. (2007). Kinematics and geometry of active detachment faulting beneath the Trans-Atlantic Geotraverse (TAG) hydrothermal field on the Mid-Atlantic Ridge. Geology, 35, (8), 711–714; .CrossRefGoogle Scholar
DeMets, C., Gordon, R. G., Argus, D. F. and Stein, S. (1990). Current plate motions. Geophysical Journal International, 101, 425–478.CrossRefGoogle Scholar
DeMets, C., Gordon, R. G., Argus, D. F. and Stein, S. (1994). Effect of recent revisions to the geomagnetic reversal timescale on estimates of current plate motions. Geophysical Research Letters, 21, (20), 2191–2194.CrossRefGoogle Scholar
Deschamps, A., Tivey, M.Embley, R. W. and Chadwick, W. W. (2007). Quantitative study of the deformation at Southern Explorer Ridge using high-resolution bathymetric data. Earth and Planetary Science Letters, 259, (1–2), 1–17.CrossRefGoogle Scholar
Detrick, R. S., Cormier, M. H., Prince, R. A., Forsyth, D. W. and Ambos, E. L. (1982). Seismic constraints on the crustal structure of the Vema fracture zone. Journal of Geophysical Research, 87, 10599–10612.CrossRefGoogle Scholar
Detrick, R. S., Buhl, P., Vera, E. et al. (1987). Multi-channel seismic imaging of a crustal magma chamber along the East Pacific Rise. Nature, 326, 35–41.CrossRefGoogle Scholar
Detrick, R., White, R. and Purdy, G. (1993a). Crustal structure of north-Atlantic fracture-zones. Reviews of Geophysics, 31, (4), 439–458.CrossRefGoogle Scholar
Detrick, R. S., Harding, A. J., Kent, G. M. et al. (1993b). Seismic structure of the southern East Pacific Rise. Science, 259, (5094), 499–503.CrossRefGoogle ScholarPubMed
Detrick, R., Collins, J. and Swift, S. (1994). In-situ evidence for the nature of the seismic layer 2/3 boundary in oceanic crust. Nature, 370, 288–290.CrossRefGoogle Scholar
Detrick, R. S., Needham, H. D. and Renard, V. (1995). Gravity anomalies and crustal thickness variations along the Mid-Atlantic Ridge between 33° N and 40° N. Journal of Geophysical Research, 100, (B3), 3767–3787.CrossRefGoogle Scholar
Detrick, R. S., Sinton, J. M., Ito, G. et al. (2002). Correlated geophysical, geochemical, and volcanological manifestations of plume-ridge interaction along the Galapagos Spreading Center. Geochemistry Geophysics Geosystems, 3, .CrossRefGoogle Scholar
Dick, H. J. B., Thompson, W. B. and Bryan, W. B. (1981). Low angle faulting and steady-state emplacement of plutonic rocks at ridge-transform intersections. EOS, Transactions of the American Geophysical Union, 62, 406.Google Scholar
Dick, H. J. B., Fisher, R. L. and Bryan, W. B. (1984). Mineralogic variability of the uppermost mantle along mid-ocean ridges. Earth and Planetary Science Letters, 69, (1), 88–106.CrossRefGoogle Scholar
Dick, H. J. B., Meyer, P. S., Bloomer, S., Stakes, D. and Mawer, C. (1991). Lithostratigraphic evolution of an in situ section of oceanic layer 3. Proceedings of the Ocean Drilling Program, Scientific Results, 118, 439–538.Google Scholar
Dick, H. J. B., Natland, J. H., Alt, J. C. et al. (2000). A long in situ section of the lower ocean crust: results of ODP Leg 176 drilling at the Southwest Indian Ridge. Earth and Planetary Science Letters, 179, (1), 31–51.CrossRefGoogle Scholar
Dick, H. J. B., Lin, J. and Schouten, H. (2003). An ultraslow-spreading class of ocean ridge. Nature, 426, (6965), 405–412.CrossRefGoogle ScholarPubMed
Dick, H. J. B., Tivey, M. A. and Tucholke, B. E. (2008). Plutonic foundation of a slow-spreading ridge segment: oceanic core complex at Kane Megamullion, 23°30ʹ N, 45°20ʹ W. Geochemistry Geophysics Geosystems, 9, (9), Q05014.CrossRefGoogle Scholar
Dietz, R. S. (1961). Continental and ocean basin evolution by spreading of the sea floor. Nature, 190, 854–7.CrossRefGoogle Scholar
Dilek, Y., Moores, E., Elthon, D. and Nicolas, A. (2000). Ophiolites and Oceanic Crust: New Insights from Field Studies and the Ocean Drilling Program. Boulder: Geological Society of America, 552 pp.Google Scholar
Donovan, D. T. and Stride, A. H. (1961). An acoustic survey of the sea floor south of Dorset and its geological interpretation. Philosophical Transactions of the Royal Society of London, B244, 299–330.CrossRefGoogle Scholar
Donval, J. P. et al. (1997). High H2 and CH4 content in hydrothermal fluids from Rainbow site newly sampled at 36°14ʹ N on the AMAR segment, Mid-Atlantic Ridge (diving FLORES cruise, July 1997). Comparison with other MAR sites. EOS, Transactions of the American Geophysical Union, 78, 832.Google Scholar
Douville, E., Charlou, J. L., Oelkers, E. H. et al. (2002). The rainbow vent fluids (36°14ʹ N, MAR): the influence of ultramafic rocks and phase separation on trace metal content in Mid-Atlantic Ridge hydrothermal fluids. Chemical Geology, 184, (1–2), 37–48.CrossRefGoogle Scholar
Dunlop, D. J. and Prévot, M. (1982). Magnetic properties and opaque mineralogy of drilled submarine intrusive rocks. Geophysical Journal of the Royal Astronomical Society, 69, (3), 763–802.CrossRefGoogle Scholar
Dunn, R. A., Toomey, D. R. and Solomon, S. C. (2000). Three-dimensional seismic structure and physical properties of the crust and shallow mantle beneath the East Pacific Rise at 9°30ʹ N. Journal of Geophysical Research-Solid Earth, 105, (B10), 23537–23555.CrossRefGoogle Scholar
Dunn, R. A., Lekic, V., Detrick, R. S. and Toomey, D. R. (2005). Three-dimensional seismic structure of the Mid-Atlantic Ridge (35° N): evidence for focused melt supply and lower crustal dike injection. Journal of Geophysical Research-Solid Earth, 110, (B9), .CrossRefGoogle Scholar
Duven, D. J. and Artis, D. A. (1985). Global Positioning System surface navigation accuracy study. Marine Geodesy, 9, (2), 145–173.CrossRefGoogle Scholar
Dyment, J., Arkani-Hamed, J. and Ghods, A. (1997). Contribution of serpentinized ultramafics to marine magnetic anomalies at slow and intermediate spreading centres: insights from the shape of the anomalies. Geophysical Journal International, 129, (3), 691–701.CrossRefGoogle Scholar
Dziak, R. P. and Fox, C. G. (1999). Long-term seismicity and ground deformation at Axial Volcano, Juan de Fuca Ridge. Geophysical Research Letters, 26, (24), 3641–3644.CrossRefGoogle Scholar
Dziak, R. P., Hammond, S. R. and Fox, C. G. (2011). A 20-year hydroacoustic time series of seismic and volcanic events in the northeast Pacific Ocean. Oceanography, 24, (3), 280–293.CrossRefGoogle Scholar
Eakins, B. W. and Lonsdale, P. F. (2003). Structural patterns and tectonic history of the Bauer microplate, Eastern Tropical Pacific. Marine Geophysical Researches, 24, (3–4), 171–205.CrossRefGoogle Scholar
Edwards, R. N., Law, L. K. and Delaurier, J. M. (1981). On measuring the electrical-conductivity of the oceanic-crust by a modified magnetometric resistivity method. Journal of Geophysical Research, 86, (B12), 1609–1615.CrossRefGoogle Scholar
Eittreim, S. and Ewing, J. (1975). Vema Fracture Zone transform fault. Geology, 3, (10), 555–558.2.0.CO;2>CrossRefGoogle Scholar
Elthon, D. (1991). Geochemical evidence for formation of the Bay-of-Islands ophiolite above a subduction zone. Nature, 354, (6349), 140–143.CrossRefGoogle Scholar
Embley, R. W. and Chadwick, W. W. (1994). Volcanic and hydrothermal processes associated with a recent phase of sea-floor spreading at the northern Cleft segment – Juan-de-Fuca Ridge. Journal of Geophysical Research-Solid Earth, 99, (B3), 4741–4760.CrossRefGoogle Scholar
Embley, R. W. and Wilson, D. S. (1992). Morphology of the Blanco Transform Fault Zone – NE Pacific: implications for its tectonic evolution. Marine Geophysical Researches, 14, 25–45.CrossRefGoogle Scholar
Embley, R. W., Chadwick, W. W., Jonasson, I. R., Butterfield, D. A. and Baker, E. T. (1995). Initial results of the rapid response to the 1993 Coaxial event – relationships between hydrothermal and volcanic processes. Geophysical Research Letters, 22, (2), 143–146.CrossRefGoogle Scholar
Embley, R. W., Chadwick, W. W., Clague, D. and Stakes, D. (1999). 1998 Eruption of Axial Volcano: multibeam anomalies and sea-floor observations. Geophysical Research Letters, 26, (23), 3425–3428.CrossRefGoogle Scholar
Engeln, J. F., Wiens, D. A. and Stein, S. (1986). Mechanisms and depths of Atlantic transform earthquakes. Journal of Geophysical Research-Solid Earth and Planets, 91, (B1), 548–577.CrossRefGoogle Scholar
Escartin, J. and Canales, J. P. (2011). Detachments in Oceanic Lithosphere: Deformation, Magmatism, Fluid Flow, and Ecosystems, AGU Chapman Conference on Oceanic Detachments; Agros, Cyprus, 8–15 May 2010. EOS, Transactions of the American Geophysical Union, 92, (4), 31–32.Google Scholar
Escartin, J. and Lin, J. (1995). Ridge offsets, normal faulting, and gravity anomalies of slow spreading ridges. Journal of Geophysical Research, 100, (B4), 6163–6177.CrossRefGoogle Scholar
Escartin, J. and Lin, J. (1998). Tectonic modification of axial structure: evidence from spectral analyses of gravity and bathymetry of the Mid-Atlantic Ridge flanks (25.5°–17.5° N). Earth and Planetary Science Letters, 154, (1–4), 279–293.CrossRefGoogle Scholar
Escartin, J., Hirth, G. and Evans, B. (1997). Effects of serpentinization on the lithospheric strength and the style of normal faulting at slow-spreading ridges. Earth and Planetary Science Letters, 151, 181–189.CrossRefGoogle Scholar
Escartin, J., Cowie, P. A., Searle, R. C. et al. (1999). Quantifying tectonic strain and magmatic accretion at a slow-spreading ridge segment, Mid-Atlantic Ridge, 29° N. Journal of Geophysical Research, B, 104, 10421–10437.CrossRefGoogle Scholar
Escartin, J., Soule, S. A., Fornari, D. J. et al. (2007). Interplay between faults and lava flows in construction of the upper oceanic crust: the East Pacific Rise crest 9°25ʹ–9°58ʹ N. Geochemistry Geophysics Geosystems, 8, .CrossRefGoogle Scholar
Escartin, J., Smith, D. K., Cann, J. et al. (2008). Central role of detachment faults in accretion of slow-spreading oceanic lithosphere. Nature, 455, 790–795.CrossRefGoogle ScholarPubMed
Evans, R. L., Constable, S. C., Sinha, M. C., Cox, C. S. and Unsworth, M. J. (1991). Upper crustal resistivity structure of the East Pacific Rise near 13° N. Geophysical Research Letters, 18, (10), 1917–1920.CrossRefGoogle Scholar
Ewing, J. and Ewing, M. (1959). Seismic refraction measurements in the Atlantic Ocean basins, in the Mediterranean, on the Mid-Atlantic Ridge, and in the Norwegian Sea. Bulletin of the Geological Society of America, 70, 291–318.CrossRefGoogle Scholar
Ewing, J. I. and Tirey, G. B. (1961). Seismic profiler. Journal of Geophysical Research, 66, (9), 2917.CrossRefGoogle Scholar
Ewing, J. I. and Zaunere, R. (1964). Seismic profiling with a pneumatic sound source. Journal of Geophysical Research, 69, 4913–4915.CrossRefGoogle Scholar
Ewing, M., Crary, A. P. and Rutherford, H. M. (1937). Geophysical investigations in the emerged and submerged Atlantic coastal plain, part I: methods and results. Bulletin of the Geological Society of America, 48, 753–802.CrossRefGoogle Scholar
Ewing, M., Worzel, J. L. and Vine, A. C. (1967). Early development of ocean-bottom photography at Woods Hole Oceanographic Institution and Lamont Geological Observatory. In Deep-Sea Photography, ed. J. B. Hersey. Baltimore: The Johns Hopkins Press, pp. 13–41.
Fleming, J. A. (1937). Magnetic survey of the oceans. In International Aspects of Oceanography, ed. Vaughan, T. W.. Washington, D.C.: National Academy of Sciences.Google Scholar
Flewellen, C., Millard, N. and Rouse, I. (1993). TOBI, a vehicle for deep ocean survey. Electronics and Communication Engineering Journal, (April 1993), 85–93.CrossRef
Fornari, D. J., Gallo, D. G., Edwards, M. H. et al. (1989). Structure and topography of the Siqueiros Transform Fault System: evidence for the development of intra-transform spreading centers. Marine Geophysical Researches, 11, 263–299.CrossRefGoogle Scholar
Fornari, D. J., Haymon, R. M., Perfit, M. R., Gregg, T. K. P. and Edwards, M. H. (1998). Axial summit trough of the East Pacific Rise 9°–10° N: geological characteristics and evolution of the axial zone on fast spreading mid-ocean ridges. Journal of Geophysical Research, B, 103, (5), 9827–9855.CrossRefGoogle Scholar
Fornari, D., Tivey, M., Schouten, H. et al. (2004). Submarine lava flow emplacement at the East Pacific Rise 9°50ʹ N: implications for uppermost ocean crust stratigraphy and hydrothermal fluid circulation. In Mid-Ocean Ridges: Hydrothermal Interactions Between the Lithosphere and Oceans: Geophysical Monograph 148, ed. German, C. R., Lin, J. and Parson, L. M.. Washington, D.C.: American Geophysical Union, pp. 187–217.Google Scholar
Forsyth, D. W. (1975). Early structural evolution and anisotropy of oceanic upper mantle. Geophysical Journal of the Royal Astronomical Society, 43, (1), 103–162.CrossRefGoogle Scholar
Forsyth, D. and Uyeda, S. (1975). Relative importance of driving forces of plate motion. Geophysical Journal of the Royal Astronomical Society, 43, (1), 163–200.CrossRefGoogle Scholar
Fouquet, Y., Ondreas, H., Charlou, J. L. et al. (1995). Atlantic lava lakes and hot vents. Nature, 377, 201.CrossRefGoogle Scholar
Fouquet, Y., Cambon, P., Etoubleau, J. et al. (2010). Geodiversity of hydrothermal processes along the Mid-Atlantic Ridge and ultramafic-hosted mineralization: a new type of oceanic Cu-Zn-Co-Au volcanogenic massive sulfide deposit. In Diversity of Hydrothermal Systems on Slow Spreading Ocean Ridges, Geophysical Monograph 188, ed. Rona, P. A., Devey, C. W., Dyment, J. and Murton, B. J.. Washington, D.C.: American Geophysical Union, pp. 321–367.CrossRefGoogle Scholar
Fowler, C. M. R. (2005). The Solid Earth: An Introduction to Global Geophysics, 2nd edition. Cambridge: Cambridge University Press, 685 pp.Google Scholar
Fox, C. G., Murphy, K. M. and Embley, R. W. (1988). Automated display and statistical-analysis of interpreted deep-sea bottom photographs. Marine Geology, 78, (3–4), 199–216.CrossRefGoogle Scholar
Fox, C. G., Matsumoto, H. and Lau, T. K. A. (2001). Monitoring Pacific Ocean seismicity from an autonomous hydrophone array. Journal of Geophysical Research-Solid Earth, 106, (B3), 4183–4206.CrossRefGoogle Scholar
Fox, P. J. and Gallo, D. G. (1984). A tectonic model for ridge–transform–ridge plate boundaries: implications for the structure of oceanic lithosphere. Tectonophysics, 104, (3–4), 205–242.CrossRefGoogle Scholar
Fox, P. J. and Stroup, J. B. (1981). The plutonic foundation of the oceanic crust. In The Sea, volume 7, ed. Emiliani, C.. New York: Wiley, pp. 119–218.Google Scholar
Fox, P. J., Schreiber, E and Peterson, J. J. (1973). Geology of oceanic crust – compressional wave velocities of oceanic rocks. Journal of Geophysical Research, 78, (23), 5155–5172.CrossRefGoogle Scholar
Fox, P. J., Grindlay, N. R. and Macdonald, K. C. (1991). The Mid-Atlantic Ridge (31° S–34°30ʹ S): temporal and spatial variations of accretionary processes. Marine Geophysical Researches, 13, 1–20.CrossRefGoogle Scholar
Francheteau, J. and Ballard, R. D. (1983). The East Pacific Rise near 21° N, 13° N, and 20° S: inferences for along-strike variability of axial processes of the mid-ocean ridge. Earth and Planetary Science Letters, 64, 93–116.CrossRefGoogle Scholar
Francheteau, J., Choukroune, P., Hekinian, R., Le Pichon, X. and Needham, H. D. (1976). Oceanic fracture zones do not provide deep sections in the crust. Canadian Journal of Earth Sciences, 13, 1223–1235.CrossRefGoogle Scholar
Francheteau, J., Juteau, T. and Rangan, C. (1979). Basaltic pillars in collapsed lava-pools on the deep ocean-floor. Nature, 281, (5728), 209–211.CrossRefGoogle Scholar
Francis, T. J. G. (1985). Resistivity measurements of an ocean floor sulphide mineral deposit from the submersible Cyana. Marine Geophysical Researches, 7, 419–438.CrossRefGoogle Scholar
Freed, A. M., Lin, J. and Melosh, H. J. (1995). Long-term survival of the axial valley morphology at abandoned slow-spreading centers. Geology, 23, (11), 971–974.2.3.CO;2>CrossRefGoogle Scholar
Freudenthal, T. and Wefer, G. (2007). Scientific drilling with the sea floor drill rig MeBo. Scientific Drilling, 5, 63–66.CrossRefGoogle Scholar
Früh-Green, G. L., Kelley, D. S., Bernasconi, S. M. et al. (2003). 30,000 years of hydrothermal activity at the Lost City vent field. Science, 301, 495–498.CrossRefGoogle ScholarPubMed
Fuchs, K. (1968). The reflection of spherical waves from transition zones with arbitrary depth-dependent elastic moduli and density. Journal of the Physics of the Earth, 16 (Special Issue), 27.CrossRefGoogle Scholar
Fuchs, K. and Müller, G. (1971). Computation of synthetic seismograms with the reflectivity method and comparison with observations. Geophysical Journal of the Royal Astronomical Society, 23, 417–433.CrossRefGoogle Scholar
Fujiwara, T., Lin, J., Matsumoto, T. et al. (2003). Crustal evolution of the Mid-Atlantic Ridge near Fifteen-Twenty Fracture Zone in the last 5 Ma. Geochemistry, Geophysics, Geosystems, 4, article 1024, .CrossRefGoogle Scholar
Fundis, A. T., Soule, S. A., Fornari, D. J. and Perfit, M. R. (2010). Paving the seafloor: volcanic emplacement processes during the 2005–2006 eruptions at the fast spreading East Pacific Rise, 9°50ʹ N. Geochemistry Geophysics Geosystems, 11, .CrossRefGoogle Scholar
Gac, S., Dyment, J., Tisseau, C. and Goslin, J. (2003). Axial magnetic anomalies over slow-spreading ridge segments: insights from numerical 3-D thermal and physical modelling. Geophysical Journal International, 154, (3), 618–632.CrossRefGoogle Scholar
Gallo, D. G., Kidd, W. S. F., Fox, P. J. et al. (1984). Tectonics at the intersection of the East Pacific Rise with the Tamayo Transform Fault. Marine Geophysical Researches, 6, (2), 159–185.Google Scholar
Gallo, D. G., Fox, P. J. and Macdonald, K. C. (1986). A Sea Beam investigation of the Clipperton Transform Fault: the morphotectonic expression of a fast-slipping transform boundary. Journal of Geophysical Research, 91, 3455–3467.CrossRefGoogle Scholar
Galperin, E. I. and Kosminskaya, I. P. (1958). Characteristics of the methods of deep seismic sounding on the sea. Bulletin of the Academy of Sciences USSR, Geophysical series, (7), 475–483.
Gaherty, J. B., Kato, M. and Jordan, T. H. (1999). Seismological structure of the upper mantle: a regional comparison of seismic layering. Physics of the Earth and Planetary Interiors, 110, (1–2), 21–41.CrossRefGoogle Scholar
Garcés, M. and Gee, J. S. (2007). Paleomagnetic evidence of large footwall rotations associated with low-angle faults at the Mid-Atlantic Ridge. Geology, 35, (3), 279–282; .CrossRefGoogle Scholar
Gass, I. G. (1968). Is Troodos Massif of Cyprus a fragment of Mesozoic ocean floor? Nature, 220, (5162), 39.CrossRefGoogle Scholar
Gass, I. G. (1990). Ophiolites and oceanic lithosphere. In Proceedings of the Symposium on Ophiolites and Oceanic Lithosphere – TROODOS 87, ed. Malpas, J., Moores, E. M., Panayiotou, A. and Xenophontos, C.. Nicosia, Cyprus: Geological Survey Department, pp. 1–10.Google Scholar
Gass, I. G. and Masson Smith, D. C. (1963). The geology and gravity anomalies of the Troodos Massif, Cyprus. Philosophical Transactions of the Royal Society of London, Series A, 255, 417–467.CrossRefGoogle Scholar
Gass, I. G., Lippard, S. J. and Shelton, A. W. (1984). Ophiolites and Oceanic Lithosphere, Special Publication 13. London: Geological Society, 413 pp.Google Scholar
Gee, J. S. and Kent, D. V. (2007). Source of oceanic magnetic anomalies and the geomagnetic polarity timescale. In Treatise on Geophysics, Volume 5, ed. Schubert, G.. Amsterdam: Elsevier, pp. 455–507.CrossRefGoogle Scholar
Gee, J. and Meurer, W. P. (2002). Slow cooling of middle and lower oceanic crust inferred from multicomponent magnetizations of gabbroic rocks from the Mid-Atlantic Ridge south of the Kane fracture zone. Journal of Geophysical Research, 107, (B7), EPM3, doi10.1029/2001JB000062.CrossRefGoogle Scholar
Gente, P., Pockalny, R. A., Durand, C. et al. (1995). Characteristics and evolution of the segmentation of the Mid-Atlantic Ridge between 20° N and 24° N during the last 10 million years. Earth and Planetary Science Letters, 129, (1–4), 55–71.CrossRefGoogle Scholar
Gee, J., Schneider, D. A. and Kent., D. V. (1996). Marine magnetic anomalies as recorders of geomagnetic intensity variations. Earth and Planetary Science Letters, 144, 327–335.CrossRefGoogle Scholar
Gerard, R., Ewing, M. and Langseth, M. G. (1962). Thermal gradient measurements in water and bottom sediment of western Atlantic. Journal of Geophysical Research, 67, (2), 785–803.CrossRefGoogle Scholar
German, C. R. and Lin, J. (2004). The thermal structure of the oceanic crust, ridge-spreading and hydrothermal circulation: how well do we understand their inter-connections?Mid-Ocean Ridges: Hydrothermal Interactions between the Lithosphere and Oceans, 148, 1–18.Google Scholar
German, C. R. and Parson, L. M. (1998). Hydrothermal activity along the Mid-Atlantic Ridge: an interplay between magmatic and tectonic processes. Earth and Planetary Science Letters, 160, 327–341.CrossRefGoogle Scholar
German, C. R., Briem, J., Chin, C. et al. (1994). Hydrothermal activity on the Reykjanes Ridge: the Steinahóll vent field at 63°06ʹ N. Earth and Planetary Science Letters, 121, (3–4), 647–654.CrossRefGoogle Scholar
German, C. R., Klinkhammer, G. P. and Rudnicki, M. D. (1996a). The Rainbow hydrothermal plume, 36.15 N, MAR. Geophysical Research Letters, 23, (21), 2979–2982.CrossRefGoogle Scholar
German, C. R., Parson, L. M., Bougault, H. et al. (1996b). Hydrothermal exploration near the Azores Triple-Junction: tectonic control of venting at slow-spreading ridges?Earth and Planetary Science Letters, 138, (1–4), 93–104.CrossRefGoogle Scholar
German, C. R., Baker, E. T., Mével, C. A., Tamaki, K. and FUJI Scientific Team (1998). Hydrothermal activity along the South West Indian Ridge. Nature, 395, 490–492.CrossRefGoogle Scholar
German, C., Tyler, P. and Griffiths, G. (2003). The maiden voyage of UK ROV “Isis”. Ocean Challenge, 12, (3), 16–18.Google Scholar
German, C., Lin, J. and Parson, L. M. (2004). Mid-Ocean Ridges: Hydrothermal Interactions Between the Lithosphere and Oceans. Washington, D.C.: American Geophysical Union, 311 pp.CrossRefGoogle Scholar
German, C. R., Bennett, S. A., Connelly, D. P. et al. (2008). Hydrothermal activity on the southern Mid-Atlantic Ridge: tectonically- and volcanically-controlled venting at 4–5° S. Earth and Planetary Science Letters, 273, (3–4), 332–344.CrossRefGoogle Scholar
German, C., Thurnherr, A., Knoery, J. et al. (2010). Heat, volume and chemical fluxes from submarine venting: a synthesis of results from the Rainbow hydrothermal field, 36° N MAR. Deep-Sea Research Part I-Oceanographic Research Papers, 57 (4), 518–527.CrossRefGoogle Scholar
Gill, R. (2010). Igneous Rocks and Processes. Chichester: Wiley-Blackwell, 428 pp.Google Scholar
Ginster, U., Mottl, M. J. and Von Herzen, R. P. (1994). Heat-flux from black smokers on the Endeavor and Cleft segments, Juan de Fuca Ridge. Journal of Geophysical Research-Solid Earth, 99, (B3), 4937–4950.CrossRefGoogle Scholar
Girardeau, J. and Nicolas, A. (1981). The structures of two ophiolite massifs, Bay-of-Islands, Newfoundland: a model for the oceanic-crust and upper mantle. Tectonophysics, 77, (1–2), 1–34.CrossRefGoogle Scholar
Glen, W. (1982). The Road to Jaramillo: Critical Years of the Revolution in Earth Science. Stanford: Stanford University Press, 480 pp.Google Scholar
Glenn, M. F. (1970). Introducing an operational multi-beam array sonar. International Hydrographic Review, 47, 35–39.Google Scholar
Goetze, C. (1978). The mechanism of solid state creep. Philosophical Transactions of the Royal Society of London, Series A, 288, 99–119.CrossRefGoogle Scholar
Goldstein, S. J., Perfit, M. R., Batiza, R., Fornari, D. J. and Murrell, M. T. (1994). Off-axis volcanism at the East Pacific Rise detected by uranium-series dating of basalts. Nature, 367, (6459), 157–159.CrossRefGoogle Scholar
Goss, A. R., Perfit, M. R., Ridley, W. I. et al. (2010). Geochemistry of lavas from the 2005–2006 eruption at the East Pacific Rise, 9°46ʹ N–9°56ʹ N: implications for ridge crest plumbing and decadal changes in magma chamber compositions. Geochemistry Geophysics Geosystems, 11, .Google Scholar
Gregg, T. K. P. and Fink, J. H. (1995). Quantification of submarine lava-flow morphology through analog experiments. Geology, 23, 73–76.2.3.CO;2>CrossRefGoogle Scholar
Gregg, T. K. P. and Smith, D. K. (2003). Volcanic investigations of the Puna Ridge, Hawai'i: relations of lava flow morphologies and underlying slopes. Journal of Volcanology and Geothermal Research, 126, (1–2), 63–77.CrossRefGoogle Scholar
Gregg, T. K. P., Fornari, D. J., Perfit, M. R., Haymon, R. M. and Fink, J. H. (1996). Rapid emplacement of a mid-ocean ridge lava flow on the East Pacific Rise at 9°46′–51ʹ N. Earth and Planetary Science Letters, 144, (3–4), E1–E7.CrossRefGoogle Scholar
Griffiths, R. W. and Fink, J. H. (1992). Solidification and morphology of submarine lavas: a dependence on extrusion rate. Journal of Geophysical Research, 97, 19729–19737.CrossRefGoogle Scholar
Grimes, C. B., John, B. E., Cheadle, M. J. and Wooden, J. L. (2008). Protracted construction of gabbroic crust at a slow spreading ridge: constraints from 206Pb/238U zircon ages from Atlantis Massif and IODP Hole U1309D (30° N, MAR). Geochemistry Geophysics Geosystems, 9, doi10.1029/2008GC002063.CrossRef
Grindlay, N. R., Fox, P. J. and Macdonald, K. C. (1991). Second-order ridge axis discontinuities in the South Atlantic: morphology, structure, and evolution. Marine Geophysical Researches, 13, 21–49.CrossRefGoogle Scholar
Grindlay, N. R., Fox, P. J. and Vogt, P. R. (1992). Morphology and tectonics of the Mid-Atlantic Ridge (25°–27°30ʹ S) from Sea Beam and magnetic data. Journal of Geophysical Research, 97, (B5), 6983–7010.CrossRefGoogle Scholar
Grindlay, N. R., Madsen, J. A., Rommevaux-Jestin, C. and Sclater, J. (1998). A different pattern of ridge segmentation and mantle Bouguer gravity anomalies along the ultra-slow spreading Southwest Indian Ridge (15°30ʹ E–25° E). Earth and Planetary Science Letters, 161, 243–253.CrossRefGoogle Scholar
Guspi, F. (1987). Frequency-domain reduction of potential field measurements to a horizontal plane. Geoexploration, 24, 87–98.CrossRefGoogle Scholar
Gutenberg, B. and Richter, C. F. (1954). Seismicity of the Earth and Associated Phenomena. Princeton, NJ: Princeton University Press.Google Scholar
Hall, J. M., Fisher, B. E., Walls, C. C. et al. (1987). Vertical distribution and alteration of dikes in a profile through the Troodos ophiolite. Nature, 326, (6115), 780–782.CrossRefGoogle Scholar
Hannington, M. D., Jonasson, I. R., Herzig, P. M. and Petersen, S. (1995). Physical and chemical processes of seafloor mineralization at mid-ocean ridges. In Seafloor Hydrothermal Systems: Physical, Chemical, Biological and Geological Interactions, Geophysical Monograph 91, ed. Humphris, S. E., Zierenberg, R. A., Mullineaux, L. S. and Thomson, R. E.. Washington, D.C.: American Geophysical Union, pp. 115–157.Google Scholar
Harding, A. J., Kent, G. M. and Orcutt, J. A. (1993). A multichannel seismic investigation of upper crustal structure at 9° N on the East Pacific Rise: implications for crustal accretion. Journal of Geophysical Research-Solid Earth, 98, (B8), 13925–13944.CrossRefGoogle Scholar
Harrison, C. G. A. (1976). Magnetization of oceanic-crust. Geophysical Journal of the Royal Astronomical Society, 47, (2), 257–283.CrossRefGoogle Scholar
Hashimoto, J., Ohta, S., Gamo, T. et al. (2001). First hydrothermal vent communities from the Indian Ocean discovered. Zoological Science, 18, (5), 717–721.CrossRefGoogle Scholar
Hasterok, D., Chapman, D. S. and Davis, E. E. (2011). Oceanic heat flow: implications for global heat loss. Earth and Planetary Science Letters, 311, (3–4), 386–395.CrossRefGoogle Scholar
Haxby, W. F. and Weissel, J. K. (1986). Evidence for small-scale mantle convection from Seasat altimeter data. Journal of Geophysical Research, 91, 3507–3520.CrossRefGoogle Scholar
Hayes, D. E. (1988). Age–depth relationships and depth anomalies in the southeast Indian Ocean and south Atlantic Ocean. Journal of Geophysical Research, 93, (B4), 2937–2954.CrossRefGoogle Scholar
Haymon, R. M. (1983). Growth history of hydrothermal black smoker chimneys. Nature, 301, (5902), 695–698.CrossRefGoogle Scholar
Haymon, R. M. (1996). The response of ridge crest hydrothermal systems to segmented, episodic magma supply. In Tectonic, Hydrothermal and Biological Segmentation at Mid-Ocean Ridges, ed. MacLeod, C. J., Tyler, P. A. and Walker, C. L.. London: Geological Society, pp. 157–168.Google Scholar
Haymon, R. and White, S. (2004). Fine-scale segmentation of volcanic/hydrothermal systems along fast-spreading ridge crests. Earth and Planetary Science Letters, 226, (3–4), 367–382.CrossRefGoogle Scholar
Haymon, R. M., Fornari, D. J., Edwards, M. H. et al. (1991). Hydrothermal vent distribution along the East Pacific Rise crest (9°09′–54ʹ N) and its relationship to magmnatic and tectonic processes on fast-spreading mid-ocean ridges. Earth and Planetary Science Letters, 104, 513–534.CrossRefGoogle Scholar
Haymon, R. M., Fornari, D. J., Von Damm, K. L. et al. (1993). Volcanic eruption of the mid-ocean ridge along the East Pacific Rise crest at 9°45–52ʹ N: direct submersible observations of seafloor phenomena associated with an eruption. Earth and Planetary Science Letters, 119, 85–101.CrossRefGoogle Scholar
Head, J. W., Wilson, L. and Smith, D. K. (1996). Mid-ocean ridge eruptive vent morphology and structure: evidence for dike widths, eruption rates, and evolution of eruptions and axial volcanic ridges. Journal of Geophysical Research, 101, (B12), 28265–28280.CrossRefGoogle Scholar
Heezen, B. C. (1960). The rift in the ocean floor. Scientific American, 203, (4), 98–110.CrossRefGoogle Scholar
Heezen, B. C. (1962). The deep-sea floor. In Continental Drift, ed. Runcorn, S. K.. New York: Academic Press, pp. 235–268.CrossRefGoogle Scholar
Heezen, B. C. (1969). World rift system – an introduction to the symposium. Tectonophysics, 8, (4–6), 269–279.CrossRefGoogle Scholar
Heezen, B. C. and Menard, H. W. (1963). Chapter 12: Topography of the deep-sea floor. In The Sea, ed. M. N. Hill. New York: John Wiley, pp. 233–280.
Heezen, B. C. and Rawson, M. (1977). Visual observations of sea-floor subduction line in Middle-America Trench. Science, 196, (4288), 423–426.CrossRefGoogle ScholarPubMed
Heezen, B. C. and Tharp, M. (1954). Physiographic diagram of the western North Atlantic. Geological Society of America Bulletin, 65, (12), 1260–1261.Google Scholar
Heezen, B. C. and Tharp, M. (1957). Physiographic Diagram of the North Atlantic. New York: Geological Society of America.Google Scholar
Heezen, B. C. and Tharp, M. (1961). Physiographic Diagram of the South Atlantic, the Caribbean, the Scotia Sea, and the Eastern Margin of the South Pacific Ocean. New York: Geological Society of America.Google Scholar
Heezen, B. C. and Tharp, M. (1964). Physiographic Diagram of the Indian Ocean. New York: Geological Society of America.Google Scholar
Heezen, B. C. and Tharp, M. (1971). Physiographic Diagram of the Western Pacific Ocean. Boulder, Colorado: Geological Society of America.Google Scholar
Heezen, B. C. and Tharp, M. (1977). World Ocean Floor Panorama, Marie Tharp Maps, NY 10976.
Heezen, B. C., Tharp, M. and Gerard, R. D. (1964a). Vema fracture zone in equatorial Atlantic. Journal of Geophysical Research, 69, (4), 733–739.CrossRefGoogle Scholar
Heezen, B. C., Bunce, E. T., Hersey, J. B. and Tharp, M. (1964b). Chain and Romanche fracture zones. Deep-Sea Research, 11, 11–33.Google Scholar
Heirtzler, J. R., Le Pichon, X. and Baron, J. G. (1966). Magnetic anomalies over the Reykjanes Ridge. Deep-Sea Research, 13, 427–443.Google Scholar
Hekinian, R., Thompson, G. and Bideau, D. (1989). Axial and off-axial heterogeneity of basaltic rocks from the East Pacific Rise at 12°35ʹ N–12°51ʹ N and 11°26ʹ N–11°30ʹ N. Journal of Geophysical Research-Solid Earth and Planets, 94, (B12), 17437–17463.CrossRefGoogle Scholar
Helmberger, D. V. and Moms, G. B. (1969). A travel time and amplitude interpretation of a marine refraction profile: primary waves. Journal of Geophysical Research, 74, 483–494.CrossRefGoogle Scholar
Helo, C., Longpre, M.-A., Shimizu, N., Clague, D. A. and Stix, J. (2011). Explosive eruptions at mid-ocean ridges driven by CO2-rich magmas. Nature Geoscience, 4, (4), 260–263.CrossRefGoogle Scholar
Hersey, J. B. (1963). 4. Continuous reflection profiling. In The Sea, volume 3, ed. Hill, M. N.. New York: Interscience Publishers, pp. 47–72.Google Scholar
Hess, H. H. (1962). History of ocean basins. In Petrologic Studies: A Volume in Honour of A. F. Buddington, ed. Engel, A. E. J., James, H. L. and Leonard, B. F.. New York: Geological Society of America, pp. 599–620.Google Scholar
Hessler, R. R. and Kaharl, V. A. (1995). Tectonic and volcanic controls on hydrothermal processes at the mid-ocean ridge: an overview based on near-bottom and submersible studies. In Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions, Geophysical Monograph 91, ed. Humphris, S. E., Zierenberg, R. A., Mullineaux, L. S. and Thomson, R. E.. Washington, D.C.: American Geophysical Union, pp. 72–84.Google Scholar
Hey, R. N. (1977). A new class of pseudofaults and their bearing on plate tectonics: a propagating rift model. Earth and Planetary Science Letters, 37, 321–325.CrossRefGoogle Scholar
Hey, R. N. and Wilson, D. S. (1982). Propagating rift explanation for the tectonic evolution of the NE Pacific – the pseudomovie. Earth and Planetary Science Letters, 58, 147–188.CrossRefGoogle Scholar
Hey, R. N., Duennebier, F. K. and Morgan, W. J. (1980). Propagating rifts on mid-ocean ridges. Journal of Geophysical Research, 85, 2647–2658.CrossRefGoogle Scholar
Hey, R. N., Naar, D. F., Kleinrock, M. C. et al. (1985). Microplate tectonics along a superfast seafloor spreading system near Easter Island. Nature, 317, 320–325.CrossRefGoogle Scholar
Hey, R. N., Kleinrock, M. C., Miller, S. P., Atwater, T. M. and Searle, R. C. (1986). Sea Beam/Deep-tow investigation of an active oceanic propagating rift system, Galapagos 95.5° W. Journal of Geophysical Research, 91, 3369–3393.CrossRefGoogle Scholar
Hill, M. N. (1952). Seismic refraction shooting in an area of the Eastern Atlantic. Philosophical Transactions of the Royal Society of London, 244, 561–596.CrossRefGoogle Scholar
Hill, M. (1959). A ship-borne nuclear-spin magnetometer. Deep-Sea Research, 5, 309–311.CrossRefGoogle Scholar
Hill, M. N. (1957). Recent geophysical exploration of the ocean floor. Physics and Chemistry of the Earth, 2, 129–163.CrossRefGoogle Scholar
Hill, M. N. (1963). 3. Single-ship seismic refraction shooting. In The Sea, volume 3, The Earth Beneath the Sea, ed. Hill, M. N.. New York: John Wiley & Sons, pp. 39–46.Google Scholar
Hooft, E. E. E., Schouten, H. and Detrick, R. S. (1996). Constraining crustal emplacement processes from the variation in seismic layer 2A thickness at the East Pacific Rise. Earth and Planetary Science Letters, 142, (3–4), 289–309.CrossRefGoogle Scholar
Hooft, E. E. E., Detrick, R. S., Toomey, D. R., Collins, J. A. and Lin, J. (2000). Crustal thickness and structure along three contrasting spreading segments of the Mid-Atlantic Ridge, 33.5°–35° N. Journal of Geophysical Research, B, 105, 8205–8226.CrossRefGoogle Scholar
Horton, E. E. (1961). Preliminary drilling phase of Mohole project I. Summary of drilling operations (La Jolla and Guadalupe sites). Bulletin of the American Association of Petroleum Geologists, 45, (11), 1789–1792.Google Scholar
Hosford, A., Lin, J. and Detrick, R. S. (2001). Crustal evolution over the last 2 m.y. at the Mid-Atlantic Ridge OH-1 segment, 35° N. Journal of Geophysical Research, 106, (B7), 13269–13285.CrossRefGoogle Scholar
Houtz, R. E. and Ewing, J. I. (1976). Upper crustal structure as a function of plate age. Journal of Geophysical Research, 81, 2490–2498.CrossRefGoogle Scholar
Huang, P. Y. and Solomon, S. C. (1988). Centroid depths of mid-ocean ridge earthquakes: dependence on spreading rate. Journal of Geophysical Research, 93, 13445–13477.CrossRefGoogle Scholar
Huang, P. Y., Solomon, S. C., Bergman, E. A. and Nabelek, J. L. (1986). Focal depths and mechanisms of Mid-Atlantic Ridge earthquakes from body wave-form inversion. Journal of Geophysical Research-Solid Earth and Planets, 91, (B1), 579–598.CrossRefGoogle Scholar
Huggett, Q. (1990). Long-range underwater photography in the deep ocean. Marine Geophysical Researches, 12, (1–2), 69–81.CrossRefGoogle Scholar
Humphris, S. E. and Cann, J. R. (2000). Constraints on the energy and chemical balances of the modern TAG and ancient Cyprus seafloor sulfide deposits. Journal of Geophysical Research-Solid Earth, 105, (B12), 28477–28488.CrossRefGoogle Scholar
Humphris, S. E., Bryan, W. B., Thompson, G. and Autio, L. K. (1990). Morphology, geochemistry, and evolution of Serocki Volcano. In Proceedings of the Ocean Drilling Program, Scientific Results, ed. Detrick, R., Honnorez, J., Bryan, W. B., Juteau, T. et al. College Station, TX, Ocean Drilling Program, pp. 67–84.Google Scholar
Humphris, S. E., Zierenberg, R. A., Mullineaux, L. S. and Thomson, R. E. (1995). Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions, Geophysical Monograph no. 91. Washington, D.C.: American Geophysical Union.CrossRefGoogle Scholar
Humphris, S. E., Herzig, P. M., Miller, D. J. and others (1996). 1. Introduction and principal results. In Proceedings of the Ocean Drilling Program, Initial Reports, 158, ed. Humphris, S. E., Herzig, P. M. and Miller, D. J.. College Station, Texas, pp. 5–14.CrossRefGoogle Scholar
Hussenoeder, S., Tivey, M. A. and Schouten, H. (1995). Direct inversion of potential field data from an irregular observation surface with application to the Mid-Atlantic Ridge. Geophysical Research Letters, 22, (23), 3131–3134.CrossRefGoogle Scholar
Hussenoeder, S. A., Tivey, M. A., Schouten, H. and Searle, R. C. (1996). Near-bottom magnetic survey of the Mid-Atlantic Ridge axis, 24°–24°40ʹ N: implications for crustal accretion at slow spreading ridges. Journal of Geophysical Research, 101, (B10), 22051–22069.CrossRefGoogle Scholar
IAGA Working Group V-MOD (2010). International Geomagnetic Reference Field: the eleventh generation. Geophysical Journal International, 183, (3), 1216–1230.CrossRefGoogle Scholar
Ildefonse, B., Blackman, D., John, B. E. et al. (2007). Oceanic core complexes and crustal accretion at slow-spreading ridges. Geology, 35, (7), 623–626; .CrossRefGoogle Scholar
Irving, E., Robertson, W. A. and Aumento, F. (1970). The Mid-Atlantic Ridge near 45° N, VI: remanent intensity, susceptibility and iron content of dredge sample. Canadian Journal of Earth Sciences, 7, 226–238.CrossRefGoogle Scholar
Isacks, B. L., Oliver, J. and Sykes, L. R. (1968). Seismology and the new global tectonics. Journal of Geophysical Research, 73, 5855–5900.CrossRefGoogle Scholar
Isezaki, N. (1986). A new shipboard three-component magnetometer. Geophysics, 51, (10), 1992–1998.CrossRefGoogle Scholar
Ito, G. and Behn, M. D. (2008). Magmatic and tectonic extension at mid-ocean ridges: 2. Origin of axial morphology. Geochemistry Geophysics Geosystems, 9, .CrossRefGoogle Scholar
Ito, G. and Martel, S. J. (2002). Focusing of magma in the upper mantle through dike interaction. Journal of Geophysical Research-Solid Earth, 107, (B10), art. no. 2223, pp. ECV 6-1–ECV 6-17.CrossRefGoogle Scholar
Jackson, H. R., Reid, I. and Falconer, R. K. H. (1982). Crustal structure near the Arctic mid-ocean ridge. Journal of Geophysical Research, 87, (B3), 1773–1783.CrossRefGoogle Scholar
Jakosky, B. M. and Shock, E. L. (1998). The biological potential of Mars, the early Earth, and Europa. Journal of Geophysical Research-Planets, 103, (E8), 19359–19364.CrossRefGoogle ScholarPubMed
Jannasch, H. W. (1995). Microbial interactions with hydrothermal fluids. In Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions, Geophysical Monograph no. 91, ed. Humphris, S. E., Zierenberg, R. A., Mullineaux, S. and Thomson, R. E.. Washington D.C.: American Geophysical Union, pp. 273–296.Google Scholar
Johnson, H. P. and Atwater, T. (1977). Magnetic study of basalts from the Mid-Atlantic Ridge, lat 37° N. Geological Society of America Bulletin, 88, 637–647.2.0.CO;2>CrossRefGoogle Scholar
Johnson, H. P. and Merrill, R. (1973). Low temperature oxidation of a titanomagnetite and the implications for paleomagnetism. Journal of Geophysical Research, 78, 4938–4949.CrossRefGoogle Scholar
Johnson, H. P. and Tivey, M. A. (1995). Magnetic-properties of zero-age oceanic-crust – a new submarine lava flow on the Juan-De-Fuca Ridge. Geophysical Research Letters, 22, (2), 175–178.CrossRefGoogle Scholar
Johnson, H. P., Becker, K. and Von Herzen, R. (1993). Near-axis heat-flow measurements on the northern Juan-de-Fuca ridge – implications for fluid circulation in oceanic-crust. Geophysical Research Letters, 20, (17), 1875–1878.CrossRefGoogle Scholar
Jones, E. J. W. (1967). Seismic reflection profiling at sea with a pneumatic sound source. PhD thesis, Cambridge University.Google Scholar
Jones, E. J. W. (1999). Marine Geophysics. Chichester: John Wiley & Sons Ltd., 466 pp.Google Scholar
Kappel, E. S. and Ryan, W. B. F. (1986). Volcanic episodicity and a non-steady state rift valley along northeast Pacific spreading centers: evidence from SeaMARC I. Journal of Geophysical Research, 91, 13925–13940.CrossRefGoogle Scholar
Karson, J. A. (1990). Seafloor spreading on the Mid-Atlantic Ridge: implications for the structure of ophiolites and oceanic lithosphere produced in slow-spreading environments. In Proceedings of the Symposium on Ophiolites and Oceanic Lithosphere – TROODOS 87, ed. Malpas, J., Moores, E. M., Panayiotou, A. and Xenophontos, C.. Nicosia, Cyprus: Geological Survey Department, Ministry of Agriculture and Natural Resources, pp. 547–555.Google Scholar
Karson, J. A. (1998). Internal structure of oceanic lithosphere: a perspective from tectonic windows. In Faulting and Magmatism at Mid-Ocean Ridges: Geophysical Monograph 106, ed. Buck, W. R., Delaney, P. T., Karson, J. A. and Lagabrielle, Y.. Washington, D.C.: American Geophysical Union, pp. 177–218.Google Scholar
Karson, J. A. (1999). Geological investigation of a lineated massif at the Kane Transform Fault: implications for oceanic core complexes. Philosophical Transactions of the Royal Society of London Series A-Mathematical Physical and Engineering Sciences, 357, (1753), 713–736.CrossRefGoogle Scholar
Karson, J. A. and Brown, J. R. (1988). Geologic setting of the Snake Pit hydrothermal site: an active vent field on the Mid-Atlantic Ridge. Marine Geophysical Researches, 10, (1/2), 91–107.CrossRefGoogle Scholar
Karson, J. A. and Dick, H. J. B. (1983). Tectonics of ridge-transform intersections at the Kane fracture-zone. Marine Geophysical Researches, 6, (1), 51–98.CrossRefGoogle Scholar
Karson, J. A., Thompson, G., Humphris, S. E. et al. (1987). Along-axis variations in seafloor spreading in the MARK area. Nature, 328, 681–685.CrossRefGoogle Scholar
Karson, J. A., Tivey, M. A. and Delaney, J. R. (2002a). Internal structure of uppermost oceanic crust along the Western Blanco Transform Scarp: implications for subaxial accretion and deformation at the Juan de Fuca Ridge. Journal of Geophysical Research-Solid Earth, 107, (B9), .CrossRefGoogle Scholar
Karson, J. A., Klein, E. M., Hurst, S. D. et al. (2002b). Structure of uppermost fast-spread oceanic crust exposed at the Hess Deep Rift: implications for subaxial processes at the East Pacific Rise. Geochemistry Geophysics Geosystems, 3, doi10.1029/2001GC000155.CrossRefGoogle Scholar
Karson, J. A., Frueh-Green, G. L., Kelley, D. S. et al. (2006). Detachment shear zone of Atlantic Massif core complex, Mid-Atlantic Ridge, 30° N. Geochemistry Geophysics Geosystems, 7, .CrossRefGoogle Scholar
Kashefi, K. and Lovley, D. (2003). Extending the upper temperature limit for life. Science, 301, 934.CrossRefGoogle ScholarPubMed
Kastens, K. A., Macdonald, K. C., Becker, K. and Crane, K. (1979). The Tamayo transform fault in the mouth of the Gulf of California. Marine Geophysical Researches, 4, 129–151.CrossRefGoogle Scholar
Katsumata, K., Sato, T., Kasahara, J. et al. (2001). Microearthquake seismicity and focal mechanisms at the Rodriguez Triple Junction in the Indian Ocean using ocean bottom seismometers. Journal of Geophysical Research-Solid Earth, 106, (B12), 30689–30699.CrossRefGoogle Scholar
Kearey, P. and Brooks, M. (1991). An Introduction to Geophysical Exploration, 2nd edition, Oxford: Blackwell Scientific Publications, 254 pp.Google Scholar
Kearey, P., Klepeis, K. A. and Vine, F. J. (2009). Global Tectonics, 3rd edition. Wiley-Blackwell 482 pp.Google Scholar
Keen, C. and Tramontini, C. (1970). A seismic refraction survey on the Mid-Atlantic Ridge. Geophysical Journal of the Royal Astronomical Society, 20, (5), 473–491.CrossRefGoogle Scholar
Kelemen, P. B. and Aharonov, E. (1998). Periodic formation of magma fractures and generation of layered gabbros in the lower crust beneath oceanic spreading ridges. In Faulting and Magmatism at Mid-Ocean Ridges – Geophysical Monograph 106, ed. Buck, W. R., Delaney, P. T., Karson, J. A. and Lagabrielle, Y.. Washington, D.C.: American Geophysical Union, pp. 267–289.Google Scholar
Kelemen, P. B., Shimizu, N. and Salters, V. J. M. (1995). Extraction of mid-ocean-ridge basalt from the upwelling mantle by focused flow of melt in dunite channels. Nature, 375, (6534), 747–753.CrossRefGoogle Scholar
Kelemen, P. B., Braun, M. and Hirth, G. (2000). Spatial distribution of melt conduits in the mantle beneath oceanic spreading centers: observations from the Ingalls and Oman ophiolites. Geochemistry Geophysics Geosystems, 1, .CrossRefGoogle Scholar
Kelemen, P. B., Matter, J., Streit, E. E. et al. (2011). Rates and mechanisms of mineral carbonation in peridotite: natural processes and recipes for enhanced, in situ CO2 capture and storage. In Annual Review of Earth and Planetary Sciences, Vol 39, ed. Jeanloz, R. and Freeman, K. H., pp. 545–576.CrossRef
Kelley, D. S., Karson, J. A., Blackman, D. K. et al. (2001). An off-axis hydrothermal vent field near the Mid-Atlantic Ridge at 30° N. Nature, 412, (6843), 145–149.CrossRefGoogle ScholarPubMed
Kelley, D. S., Karson, J. A., Fruh-Green, G. L. et al. (2005). A serpentinite-hosted ecosystem: the Lost City hydrothermal field. Science, 307, (5714), 1428–1434.CrossRefGoogle ScholarPubMed
Kennett, B. L. N. and Orcutt, J. A. (1976). Comparison of travel time inversions for marine refraction profiles. Journal of Geophysical Research, 81, (23), 4061–4070.CrossRefGoogle Scholar
Kennett, B. L. N., Bunch, A. W. H., Orcutt, J. A. and Raitt, R. W. (1977). Variations in crustal structure on East Pacific Rise crest – travel time inversion approach. Earth and Planetary Science Letters, 34, (3), 439–444.CrossRefGoogle Scholar
Kennish, M. J. and Lutz, R. A. (1998). Morphology and distribution of lava flows on mid-ocean ridges: a review. Earth-Science Reviews, 43, (3–4), 63–90.CrossRefGoogle Scholar
Kent, G. M., Harding, A. J. and Orcutt, J. A. (1990). Evidence for a smaller magma chamber beneath the East Pacific Rise at 9°30ʹ N. Nature, 344, 650–653.CrossRefGoogle Scholar
Kent, G. M., Harding, A. J. and Orcutt, J. A. (1993). Distribution of magma beneath the East Pacific Rise between the Clipperton Transform and the 9°17ʹ N deval from forward modeling of common depth point data. Journal of Geophysical Research, 98, (B8), 13971–13995.Google Scholar
Kent, G. M., Harding, A. J., Orcutt, J. A. et al. (1994). Uniform accretion of oceanic-crust south of the Garrett Transform at 14°15ʹ S on the East Pacific Rise. Journal of Geophysical Research-Solid Earth, 99, (B5), 9097–9116.CrossRefGoogle Scholar
Kirk, R. E., Whitmarsh, R. B. and Langford, J. J. (1982). A 3-component ocean bottom seismograph for controlled source and earthquake seismology. Marine Geophysical Researches, 5, (3), 327–341.CrossRefGoogle Scholar
Kitahara, A., Seama, N. and Isezaki, N. (1994). A subduction angle of the Pacific Plate beneath the Japan Trench inferred from three component geomagnetic anomalies. Journal of Geomagnetism and Geolectricity, 46, 455–462.CrossRefGoogle Scholar
Kleinrock, M. C. and Hey, R. N. (1989a). Detailed tectonics near the tip of the Galapagos 95.5° W propagator: how the lithosphere tears and a spreading center develops. Journal of Geophysical Research, 94, 13801–13838.CrossRefGoogle Scholar
Kleinrock, M. C. and Hey, R. N. (1989b). Migrating transform zone and lithosphere transfer at the Galapagos 95.5° W propagator. Journal of Geophysical Research, 94, 13859–13878.CrossRefGoogle Scholar
Kleinrock, M. C. and Humphris, S. E. (1996). Structural control on sea-floor hydrothermal activity at the TAG active mound. Nature, 382, (6587), 149–153.CrossRefGoogle Scholar
Koelsch, D. E., Witzell, W. E. S., Broda, J. E., Wooding, F. B. and Purdy, G. M. (1986). A deep towed explosive source for seismic experiments on the ocean floor. Marine Geophysical Researches, 8, 345–361.CrossRefGoogle Scholar
Kohnen, W. (2009). Human exploration of the deep seas: fifty years and the inspiration continues. Marine Technology Society Journal, 43, (5), 42–62.CrossRefGoogle Scholar
Kong, L. S. L., Solomon, S. C. and Purdy, G. M. (1986). Microearthquakes near the TAG hydrothermal field, Mid-Atlantic Ridge, 26° N. EOS, Transactions of the American Geophysical Union, 67, 1021.Google Scholar
Kong, L. S., Detrick, R. S., Fox, P. J., Mayer, L. A. and Ryan, W. B. F. (1988/89). The morphology and tectonics of the MARK area from Sea Beam and Sea MARC I observations (Mid-Atlantic Ridge 23° N). Marine Geophysical Researches, 10, 59–90.CrossRefGoogle Scholar
Kong, L. S. L., Solomon, S. C. and Purdy, G. M. (1992). Microearthquake characteristics of a midocean ridge along-axis high. Journal of Geophysical Research, 97, (B2), 1659–1685.CrossRefGoogle Scholar
Koschinsky, A. et al. (2008). Hydrothermal venting at pressure–temperature conditions above the critical point of seawater, 5° S on the Mid-Atlantic Ridge. Geology, 36, 615–618.CrossRefGoogle Scholar
Krasnov, S. G., Cherkashev, G. A., Sepanova, T. V. et al. (1995). Detailed geological studies of hydrothermal fields in the North Atlantic. In Hydrothermal Vents and Processes, Special Publication 87, ed. Parson, L. M., Walker, C. L. and Dixon, D. R.. London: Geological Society, pp. 43–64.Google Scholar
Kuo, B. Y. and Forsyth, D. W. (1988). Gravity anomalies of the ridge-transform system in the South Atlantic between 31 and 34.5° S: upwelling centers and variations in crustal thickness. Marine Geophysical Researches, 10, (3–4), 205–232.CrossRefGoogle Scholar
Kuo, B. Y., Forsyth, D. W. and Parmentier, E. M. (1986). Flexure and thickening of the lithosphere at the East Pacific Rise. Geophysical Research Letters, 13, 681–684.CrossRefGoogle Scholar
Lalou, C., Reyss, J. L., Brichet, E., Rona, P. A. and Thompson, G. (1995). Hydrothermal activity on a 105 year scale at a slow-spreading ridge, TAG hydrothermal field, Mid-Atlantic Ridge 26° N. Journal of Geophysical Research-Solid Earth, 100, (B9), 17855–17862.CrossRefGoogle Scholar
Langmuir, C. H., Klein, E. M. and Plank, T. (1992). Petrological systematics of mid-ocean ridge basalts: constraints on melt generation beneath ocean ridges. In Mantle Flow and Melt Generation at Mid-Ocean Ridges, Geophysical Monograph 71, ed. Morgan, J. P., Blackman, D. K. and Sinton, J. M.. Washington, D.C.: American Geophysical Union, pp. 183–280.Google Scholar
Langmuir, C., Humphris, S., Fornari, D. et al. (1997). Hydrothermal vents near a mantle hot spot: the Lucky Strike vent field at 37° N on the Mid-Atlantic Ridge. Earth and Planetary Science Letters, 148, (1–2), 69–91.CrossRefGoogle Scholar
Langston, C. A. (1981). Source inversion of seismic waveforms – the Koyna, India, earthquakes of 13 September 1967. Bulletin of the Seismological Society of America, 71, (1), 1–24.Google Scholar
Larson, R. L. and Spiess, F. N. (1969). East Pacific Rise crest – a near-bottom geophysical profile. Science, 163, (3862), 68–71.CrossRefGoogle Scholar
Larson, R. L., Searle, R. C., Kleinrock, M. C. et al. (1992). Roller-bearing tectonic evolution of the Juan Fernandez microplate. Nature, 356, 571–576.CrossRefGoogle Scholar
Laughton, A. S. (1981). The first decade of GLORIA. Journal of Geophysical Research, 86, 11511–11534.CrossRefGoogle Scholar
Laughton, A. S., Searle, R. C. and Roberts, D. G. (1979). The Reykjanes Ridge crest and the transition between its rifted and non-rifted regions. Tectonophysics, 55, 173–177.CrossRefGoogle Scholar
Lavier, L. L., Buck, W. R. and Poliakov, A. N. B. (1999). Self-consistent rolling-hinge model for the evolution of large-offset low-angle normal faults. Geology, 27, (12), 1127–1130.2.3.CO;2>CrossRefGoogle Scholar
Lavier, L. L., Buck, W. R. and Poliakov, A. N. B. (2000). Factors controlling normal fault offset in an ideal brittle layer. Journal of Geophysical Research-Solid Earth, 105, (B10), 23431–23442.CrossRefGoogle Scholar
Lawson, K., Searle, R. C., Pearce, J. A., Browning, P. and Kempton, P. (1996). Detailed volcanic geology of the MARNOK area, Mid-Atlantic Ridge north of Kane transform. In Tectonic, Magmatic, Hydrothermal and Biological Segmentation of Mid-Ocean Ridges, Geological Society of London, Special Publication, 118, ed. MacLeod, C. J., Tyler, P. A. and Walker, C. L.. London: Geological Society, pp. 61–102.CrossRefGoogle Scholar
Le Pichon, X. (1968). Sea-floor spreading and continental drift. Journal of Geophysical Research, 73, 3661–3697.CrossRefGoogle Scholar
Le Pichon, X. and Hayes, D. E. (1971). Marginal offsets, fracture zones and the early opening of the South Atlantic. Journal of Geophysical Research, 76, 6283–6293.CrossRefGoogle Scholar
Lee, S.-M. and Searle, R. C. (2000). Crustal magnetization of the Reykjanes Ridge and implications for its along-axis variability and the formation of axial volcanic ridges. Journal of Geophysical Research, 105, 5907–5930.CrossRefGoogle Scholar
Leeds, A. R., Kausel, E. and Knopoff, L. (1974). Variations of upper mantle structure under the Pacific Ocean. Science, 186, 141–143.CrossRefGoogle ScholarPubMed
Lilwall, R. C., Francis, T. J. G. and Porter, I. T. (1977). Ocean bottom seismograph observations on the Mid-Atlantic Ridge near 45° N. Geophysical Journal of the Royal Astronomical Society, 51, 357–370.CrossRefGoogle Scholar
Lilwall, R. C., Francis, T. J. G. and Porter, I. T. (1978). Ocean bottom seismograph observations on the Mid-Atlantic Ridge near 45° N – further results. Geophysical Journal of the Royal Astronomical Society, 55, 255–262.CrossRefGoogle Scholar
Lilwall, R. C., Francis, T. J. G. and Porter, I. T. (1981). A microearthquake survey at the junction of the East Pacific Rise and the Wilkes (9° S) fracture zone. Geophysical Journal of the Royal Astronomical Society, 66, 407–416.CrossRefGoogle Scholar
Lin, J. and Phipps Morgan, J. (1992). The spreading rate dependence of three-dimensional mid-ocean ridge gravity structure. Geophysical Research Letters, 19, (1), 13–16.CrossRefGoogle Scholar
Lin, J., Purdy, G. M., Schouten, H., Sempéré, J.-C. and Zervas, C. (1990). Evidence from gravity data for focused magmatic accretion along the Mid-Atlantic Ridge. Nature, 344, 627–632.CrossRefGoogle Scholar
Lippard, S. J., Shelton, A. W. and Gass, I. G. (1986). The Ophiolite of Northern Oman, Memoir 11. London: Geological Society.Google Scholar
Lister, C. R. B. (1970). Measurement of in-situ sediment conductivity by means of a Bullard-type probe. Geophysical Journal of the Royal Astronomical Society, 19, (5), 521–532.CrossRefGoogle Scholar
Lister, C. R. B. (1972). Thermal balance of a mid-ocean ridge. Geophysical Journal of the Royal Astronomical Society, 26, (5), 515.CrossRefGoogle Scholar
Lister, C. R. B. (1974). Penetration of water into hot rock. Geophysical Journal of the Royal Astronomical Society, 39, (3), 465–509.CrossRefGoogle Scholar
Lonsdale, P. (1977). Structural geomorphology of a fast-spreading rise crest: the East Pacific Rise near 3°25ʹ S. Marine Geophysical Researches, 3, 251–293.CrossRefGoogle Scholar
Lonsdale, P. (1978). Near-bottom reconnaissance of a fast-slipping transform fault zone at the Pacific–Nazca plate boundary. Journal of Geology, 86, 451–472.CrossRefGoogle Scholar
Lonsdale, P. (1988). Structural pattern of the Galapagos microplate and evolution of the Galapagos triple junctions. Journal of Geophysical Research, 93, 13551–13574.CrossRefGoogle Scholar
Lonsdale, P. (1994). Geomorphology and structural segmentation of the crest of the southern (Pacific–Antarctic) East Pacific Rise. Journal of Geophysical Research-Solid Earth, 99, (B3), 4683–4702.CrossRefGoogle Scholar
Lonsdale, P. (1995). Segmentation and disruption of the East Pacific Rise in the mouth of the Gulf of California. Marine Geophysical Researches, 17, (4), 323–359.CrossRefGoogle Scholar
Lonsdale, P. (2005). Creation of the Cocos and Nazca plates by fission of the Farallon plate. Tectonophysics, 404, (3–4), 237–264.CrossRefGoogle Scholar
Louden, K. E., White, R. S., Potts, C. G. and Forsyth, D. W. (1986). Structure and seismotectonics of the Vema Fracture Zone, Atlantic Ocean. Journal of the Geological Society of London, 143, 795–805.CrossRefGoogle Scholar
Lowell, R. P., Rona, P. A. and Von Herzen, R. P. (1995). Sea-floor hydrothermal systems. Journal of Geophysical Research-Solid Earth, 100, (B1), 327–352.CrossRefGoogle Scholar
Lowrie, W. (1974). Oceanic basalt magnetic-properties and Vine and Matthews hypothesis. Journal of Geophysics-Zeitschrift Fur Geophysik, 40, (4), 513–536.Google Scholar
Lowrie, W. (1997). Fundamentals of Geophysics. Cambridge: Cambridge University Press, 354 pp.Google Scholar
Ludwig, K. A., Kelley, D. S., Butterfield, D. A., Nelson, B. K. and Früh-Green, G. L. (2006). Formation and evolution of carbonate chimneys at the Lost City Hydrothermal Field. Geochimica et Cosmochimica Acta, 70, 3625–3645.CrossRefGoogle Scholar
Ludwig, K. A., Shen, C.-C., Kelley, D. S., Cheng, H. and Edwards, R. L. (2011). U–Th systematic and 230Th ages of carbonate chimneys at the Lost City Hydrothermal Field. Geochimica et Cosmochimica Acta, 75, 1869–1888.CrossRefGoogle Scholar
Lupton, J. E. (1995). Hydrothermal plumes: near and far field. In Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions, Geophysical Monograph no. 91, ed. Humphris, S. E., Zierenberg, R. A., Mullineaux, S. and Thomson, R. E.. Washington, D.C.: American Geophysical Union, pp. 317–346.Google Scholar
Lupton, J. E. and Craig, H. (1981). A major helium-3 source at 15° S on the East Pacific Rise. Science, 214, 13–18.CrossRefGoogle Scholar
Lupton, J., Delaney, J., Johnson, H. and Tivey, M. (1985). Entrainment and vertical transport of deep-ocean water by buoyant hydrothermal plumes. Nature, 316, 621–623.CrossRefGoogle Scholar
Lutz, R. A. and Kennish, M. J. (1993). Ecology of deep-sea hydrothermal vent communities: a review. Reviews of Geophysics, 31, (3), 211–242.CrossRefGoogle Scholar
Macdonald, K. C. (1977). Near-bottom magnetic anomalies, asymmetric spreading, oblique spreading, and tectonics of the Mid-Atlantic Ridge near lat 37° N. Geological Society of America Bulletin, 88, 541–555.2.0.CO;2>CrossRefGoogle Scholar
Macdonald, K. C. (1982). Mid-ocean ridges: fine sale tectonic, volcanic and hydrothermal processes within the plate boundary zone. Annual Reviews of Earth and Planetary Science, 10, 155–189.CrossRefGoogle Scholar
Macdonald, K. C. (1998). Linkages between faulting, volcanism, hydrothermal activity and segmentation on fast-spreading centers. In Faulting and Magmatism at Mid-Ocean Ridges – Geophysical Monograph 106, ed. Buck, W. R., Delaney, P. T., Karson, J. A. and Lagabrielle, Y.. Washington, D.C.: American Geophysical Union, pp. 27–59.Google Scholar
Macdonald, K. C. and Fox, P. J. (1983). Overlapping spreading centres: new accretion geometry on the East Pacific Rise. Nature, 302, 55–57.CrossRefGoogle Scholar
Macdonald, K. C. and Fox, P. J. (1988). The axial summit graben and cross-sectional shape of the East Pacific Rise as indicators of axial magma chambers and recent volcanic eruptions. Earth and Planetary Science Letters, 88, 119–131.CrossRefGoogle Scholar
Macdonald, K. C., Kastens, K. A., Spiess, F. N. and Miller, S. P. (1979). Deep tow studies of the Tamayo Transform Fault. Marine Geophysical Researches, 4, 37–70.CrossRefGoogle Scholar
Macdonald, K. C., Becker, K., Spiess, F. N. and Ballard, R. D. (1980a). Hydrothermal heat flux of the “black smoker” vents on the East Pacific Rise. Earth and Planetary Science Letters, 48, 1–7.CrossRefGoogle Scholar
Macdonald, K. C., Miller, S. P., Huestis, S. P. and Spiess, F. N. (1980b). Three-dimensional modelling of a magnetic reversal boundary from inversion of deep-tow measurements. Journal of Geophysical Research, 85, 3670–3680.CrossRefGoogle Scholar
Macdonald, K. C., Sempere, J.-C. and Fox, P. J. (1984). East Pacific Rise from Siqueiros to Orozco fracture zones: along-strike continuity of the axial neovolcanic zone and structure and evolution of overlapping spreading centers. Journal of Geophysical Research, 89, 6049–6069.CrossRefGoogle Scholar
Macdonald, K. C., Sempere, J.-C. and Fox, P. J. (1986a). Reply: the debate concerning overlapping spreading centers and mid-ocean ridge processes. Journal of Geophysical Research, 91, 10501–10510.CrossRefGoogle Scholar
Macdonald, K. C., Castillo, D., Miller, S. et al. (1986b). Deep-tow studies of the Vema Fracture Zone: 1 – The tectonics of a major slow-slipping transform fault and its intersection with the Mid-Atlantic Ridge. Journal of Geophysical Research, 91, 3334–3354.CrossRefGoogle Scholar
Macdonald, K. C., Sempere, J.-C., Fox, P. J. and Tyce, R. (1987). Tectonic evolution of ridge-axis discontinuities by the meeting, linking, or self-decapitation of neighbouring ridge segments. Geology, 15, 993–997.2.0.CO;2>CrossRefGoogle Scholar
Macdonald, K. C., Haymon, R. M., Miller, S. P., Sempere, J.-C. and Fox, P. J. (1988a). Deep-tow and Sea Beam studies of duelling propagating ridges on the East Pacific Rise near 20°40ʹ S. Journal of Geophysical Research, 93, 2875–2898.CrossRefGoogle Scholar
Macdonald, K. C., Fox, P. J., Perram, L. J. et al. (1988b). A new view of the mid-ocean ridge from the behaviour of ridge-axis discontinuities. Nature, 335, 217–225.CrossRefGoogle Scholar
Macdonald, K. C., Haymon, R. and Shor, A. (1989). A 220 km2 recently erupted lava field on the East Pacific Rise near lat. 8° S. Geology, 17, 212–216.2.3.CO;2>CrossRefGoogle Scholar
Macdonald, K. C., Scheirer, D. S. and Carbotte, S. M. (1991). Mid-ocean ridges: discontinuities, segments and giant cracks. Science, 253, 986–994.CrossRefGoogle ScholarPubMed
Macdonald, K. C., Fox, P. J., Miller, S. et al. (1992). The East Pacific Rise and its flanks 8–18° N – history of segmentation, propagation and spreading direction based on Seamarc-II and Sea Beam studies. Marine Geophysical Researches, 14, (4), 299.CrossRefGoogle Scholar
Macdonald, K. C., Fox, P. J., Alexander, R. T., Pockalny, R. and Gente, P. (1996). Volcanic growth faults and the origin of Pacific abyssal hills. Nature, 380, (6570), 125–129.CrossRefGoogle Scholar
MacGregor, L. M., Constable, S. and Sinha, M. C. (1998). The RAMESSES experiment–III. Controlled-source electromagnetic sounding of the Reykjanes Ridge at 57°45ʹ N. Geophysical Journal International, 135, (3), 773–789.CrossRefGoogle Scholar
Macleod, C. and Rothery, D. (1992). Ridge axial segmentation in the Oman ophiolite: evidence from along-strike variations in the sheeted dyke complex. In Ophiolites and their Modern Oceanic Analogues, ed. Parson, L. M., Murton, B. and Browning, P.. London: Geological Society, pp. 39–63.Google Scholar
MacLeod, C. J. and Yaouancq, G. (2000). A fossil melt lens in the Oman ophiolite: implications for magma chamber processes at fast spreading ridges. Earth and Planetary Science Letters, 176, (3–4), 357–373.CrossRefGoogle Scholar
MacLeod, C. J., Dick, H. J. B., Allerton, S. et al. (1998). Geological mapping of slow-spread lower ocean crust: a deep-towed video and wireline rock drilling survey of Atlantis Bank (ODP Site 735, Southwest Indian Ridge). InterRidge News, 7, 39–43.Google Scholar
MacLeod, C. J., Escartin, J., Banerji, D. et al. (2002). Direct geological evidence for oceanic detachment faulting: the Mid-Atlantic Ridge, 15°45ʹ N. Geology, 30, 879–882.2.0.CO;2>CrossRefGoogle Scholar
MacLeod, C. J., Searle, R. C., Murton, B. J. et al. (2009). Life cycle and internal structure of oceanic core complexes. Earth and Planetary Science Letters, 287, (3–4), 333–344.CrossRefGoogle Scholar
Madsen, J. A., Forsyth, D. W. and Detrick, R. S. (1984). A new isostatic model for the East Pacific Rise Crest. Journal of Geophysical Research, 89, 9997–10015.CrossRefGoogle Scholar
Madsen, J. A., Fox, P. J. and Macdonald, K. C. (1986). Morphotectonic fabric of the Orozco Transform Fault: results from a Sea Beam investigation. Journal of Geophysical Research, 91, 3439–3454.CrossRefGoogle Scholar
Magde, L. S. and Smith, D. K. (1995). Seamount volcanism at the Reykjanes Ridge: relationship to the Iceland hot spot. Journal of Geophysical Research, 100, (B5), 8449–8468.CrossRefGoogle Scholar
Magde, L. S. and Sparks, D. W. (1997). Three-dimensional mantle upwelling, melt generation, and melt migration beneath segment slow spreading ridges. Journal of Geophysical Research, B, 102, (9), 20571–20583.CrossRefGoogle Scholar
Magde, L. S., Detrick, R. S., Kent, G. M. et al. (1995). Crustal and upper-mantle contribution to the axial gravity-anomaly at the southern East Pacific Rise. Journal of Geophysical Research-Solid Earth, 100, (B3), 3747–3766.CrossRefGoogle Scholar
Magde, L. S., Sparks, D. W. and Detrick, R. S. (1997). The relationship between buoyant mantle flow, melt migration, and gravity bull's eyes at the Mid-Atlantic Ridge between 33° N and 35° N. Earth and Planetary Science Letters, 148, (1–2), 59–67.CrossRefGoogle Scholar
Magde, L. S., Barclay, A. H., Toomey, D. R., Detrick, R. S. and Collins, J. A. (2000). Crustal magma plumbing within a segment of the Mid-Atlantic Ridge, 35° N. Earth and Planetary Science Letters, 175, 55–67.CrossRefGoogle Scholar
Maia, M. and Gente, P. (1998). Three-dimensional gravity and bathymetric analysis of the Mid-Atlantic Ridge between 20° N and 24° N: flow geometry and temporal evolution of segmentation. Journal of Geophysical Research, 103B, (1), 951–974.CrossRefGoogle Scholar
Mallows, C. and Searle, R. C. (2012). A geophysical study of oceanic core complexes and surrounding terrain, Mid-Atlantic Ridge at 13°–14° N. Geochemistry Geophysics Geosystems, 13, .CrossRefGoogle Scholar
Malpas, J. (1978). Magma generation in upper mantle, field evidence from ophiolite suites, and application to generation of oceanic lithosphere. Philosophical Transactions of the Royal Society of London Series A-Mathematical Physical and Engineering Sciences, 288, (1355), 527–545.CrossRefGoogle Scholar
Malpas, J. (1990). Crustal accretionary processes in the Troodos ophiolite, Cyprus: evidence from field mapping and deep crustal drilling. In Proceedings of the Symposium on Ophiolites and Oceanic Lithosphere – TROODOS 87, ed. Malpas, J., Moores, E. M., Panayiotou, A. and Xenophontas, C.. Nicosia, Cyprus: Geological Survey Department, pp. 65–74.Google Scholar
Malpas, J., Moores, E. M., Panayiotou, A. and Xenophontos, C. (1990). Proceedings of the Symposium on Ophiolites and Oceanic Lithosphere – TROODOS 87. Nicosia, Cyprus: Geological Survey Department, Ministry of Agriculture and Natural Resources, 733 pp.Google Scholar
Mammerickx, J., Naar, D. F. and Tyce, R. L. (1988). The Mathematician paleoplate. Journal of Geophysical Research, 93, 3025–3040.CrossRefGoogle Scholar
Marks, K. M., Vogt, P. R. and Hall, S. A. (1990). Residual depth anomalies and the origin of the Australian–Antarctic Discordance zone. Journal of Geophysical Research, 95, (B11), 17325–17337.CrossRefGoogle Scholar
Marmer, H. A. (1933). Recent major oceanographic expeditions: a review of the work of the Meteor, Carnegie, Dana, and Snellius expeditions. Geographical Review, 23, (2), 299–305.CrossRefGoogle Scholar
Martin, W., Baross, J., Kelley, D. and Russell, M. J. (2008). Hydrothermal vents and the origins of life. Nature Reviews Microbiology, 6, 805–814.CrossRefGoogle Scholar
Mason, R. (1985). Ophiolites. Geology Today, 1, 136–140.CrossRefGoogle Scholar
Mason, R. G. (1958). A magnetic survey off the west coast of the United States between latitudes 32° and 36° N, longitudes 121° and 128° W. Geophysical Journal of the Royal Astronomical Society, 1, 320–329.Google Scholar
Mason, R. G. and Raff, A. D. (1961). Magnetic survey off the west coast of North America, 32° N latitude to 42° N latitude. Geological Society of America Bulletin, 72, (8), 1259–1265.CrossRefGoogle Scholar
Mastin, L. G. and Pollard, D. D. (1988). Surface deformation and shallow dike intrusion processes at Inyo Craters, Long Valley, California. Journal of Geophysical Research-Solid Earth and Planets, 93, (B11), 13221–13235.CrossRefGoogle Scholar
Maury, M. F. (1860). The Physical Geography of the Sea. London: T. Nelson and Sons.Google Scholar
Maxwell, A. E., von Herzen, R. P., Hsu, K. J. et al. (1970). Deep sea drilling in the South Atlantic. Science, 168, 1047–1059.CrossRefGoogle ScholarPubMed
McAllister, E. and Cann, J. (1996). Initiation and evolution of boundary wall faults along the Mid-Atlantic Ridge, 25–29° N. In Tectonic, Magmatic, Hydrothermal and Biological Segmentation of Mid-Ocean Ridges, ed. MacLeod, C. J., Tyler, P. A. and Walker, C. L.. London: Geological Society Special Publication 118, pp. 29–48.Google Scholar
McCaig, A. M., Cliff, R. A., Escartin, J., Fallick, A. E. and MacLeod, C. J. (2007). Oceanic detachment faults focus very large volumes of black smoker fluids. Geology, 35, (10), 935–938; .CrossRefGoogle Scholar
McKenzie, D. P. (1967). Some remarks on heat flow and gravity anomalies. Journal of Geophysical Research, 72, (24), 6261–6273.CrossRefGoogle Scholar
McKenzie, D. (1986). The geometry of propagating rifts. Earth and Planetary Science Letters, 77, 176–186.CrossRefGoogle Scholar
McKenzie, D. P. and Bickle, M. J. (1988). The volume and composition of melt generated by extension of the lithosphere. Journal of Petrology, 29, 625–679.CrossRefGoogle Scholar
McKenzie, D. P. and Bowin, C. O. (1976). The relationship between bathymetry and gravity in the Atlantic Ocean. Journal of Geophysical Research, 81, 1903–1915.CrossRefGoogle Scholar
McKenzie, D. P. and Morgan, W. J. (1969). Evolution of triple junctions. Nature, 224, (5215), 125–133.CrossRefGoogle Scholar
McKenzie, D. P. and Parker, R. L. (1967). The North Pacific: an example of tectonics on a sphere. Nature, 216, 1276–1280.CrossRefGoogle Scholar
Melchert, B., Devey, C. W., German, C. R. et al. (2008). First evidence for high-temperature off-axis venting of deep crustal/mantle heat: the Nibelungen hydrothermal vent field, southern Mid-Atlantic Ridge. Earth and Planetary Science Letters, 275, 61–69.CrossRefGoogle Scholar
Menard, H. W. (1960). East Pacific Rise. Science, 132, (3441), 1737–1746.CrossRefGoogle ScholarPubMed
Menard, H. W. (1967). Extension of Northeastern–Pacific fracture zones. Science, 155, 72–74.CrossRefGoogle ScholarPubMed
Menard, H. W. and Atwater, T. (1968). Changes in direction of seafloor spreading. Nature, 219, 463–467.CrossRefGoogle Scholar
Menard, H. W. and Atwater, T. (1969). Origin of fracture-zone topography. Nature, 222, 1037–1040.CrossRefGoogle Scholar
Mendel, V. and Sauter, D. (1997). Seamount volcanism at the super slow-spreading southwest Indian ridge between 57° E and 70° E. Geology, 25, (2), 99–102.2.3.CO;2>CrossRefGoogle Scholar
Mendel, V., Sauter, D., Parson, L., Vanney, J.-R. and Munschy, M. (1997). Segmentation and morphotectonic variations along a super slow-spreading center: the Southwest Indian Ridge (57° E–70° E). Marine Geophysical Researches, 19, (6), 505–533.CrossRefGoogle Scholar
Meurer, W. P. and Gee, J. (2002). Evidence for the protracted construction of slow-spread oceanic crust by small magmatic injections. Earth and Planetary Science Letters, 201, 45–55.CrossRefGoogle Scholar
Michael, P. J., Langmuir, C. H., Dick, H. J. B. et al. (2003). Magmatic and amagmatic seafloor generation at the ultraslow-spreading Gakkel ridge, Arctic Ocean. Nature, 423, (6943), 956–961.CrossRefGoogle ScholarPubMed
Millard, N. W., Griffiths, G., Finegan, G. et al. (1998). Versatile autonomous submersibles – the realising and testing of a practical vehicle. Underwater Technology, 23, (1), 7–17.CrossRefGoogle Scholar
Miller, E. and Ewing, M. (1956). Geomagnetic measurements in the Gulf of Mexico and in the vicinity of Caryn Peak. Geophysics, 21, 406–432.CrossRefGoogle Scholar
Minshull, T. A., Muller, M. R. and White, R. S. (2006). Crustal structure of the Southwest Indian Ridge at 66° E: seismic constraints. Geophysical Journal International, 166, (1), 135–147.CrossRefGoogle Scholar
Minster, J. B., Jordan, T. H., Molnar, P. and Haines, E. (1974). Numerical modelling of instantaneous plate tectonics. Geophysical Journal of the Royal Astronomical Society, 36, 541–576.CrossRefGoogle Scholar
Mitchell, N. C. and Searle, R. C. (1999). Fault scarp statistics at the Galapagos Spreading Centre from Deep Tow data. Marine Geophysical Researches, 20, (3), 183–193.CrossRefGoogle Scholar
Moore, D. E. and Lockner, D. A. (2007). Comparative deformation behavior of minerals in serpentinized ultramafic rock: application to the slab–mantle interface in subduction zones. International Geology Review, 49, (5), 401–415.CrossRefGoogle Scholar
Moore, D. E. and Rymer, M. J. (2007). Talc-bearing serpentinite and the creeping section of the San Andreas fault. Nature, 448, .CrossRefGoogle ScholarPubMed
Moores, E. M. (1982). Origin and emplacement of ophiolites. Reviews of Geophysics, 20, (4), 735–760.CrossRefGoogle Scholar
Morgan, W. J. (1968). Rises, trenches, great faults and crustal blocks. Journal of Geophysical Research, 73, 1959–1982.CrossRefGoogle Scholar
Moritz, H. (1980). Geodetic Reference System 1980: International Union of Geodesy and Geophysics Resolution no. 7, pp. 128–133.CrossRef
Morley, L. W. and Larochelle, A. (1964). Paleomagnetism as a means of dating geological events. Royal Society of Canada, Special Publication, 9, 40–51.Google Scholar
Morris, A., Gee, J. S., Pressling, N. et al. (2009). Footwall rotation in an oceanic core complex quantified using reoriented Integrated Ocean Drilling Program core samples. Earth and Planetary Science Letters, 28, 217–228.CrossRefGoogle Scholar
Morton, J. L. and Sleep, N. H. (1985). Seismic reflections from a Lau Basin magma chamber. In Geology and Offshore Resources of Pacific Island Arcs – Tonga Region, ed. Scholl, D. W. and Vallier, T. L.. Houston: Circum-Pacific Council for Energy and Mineral Resources, pp. 441–453.Google Scholar
Muller, M. R., Robinson, C. J., Minshull, T. A., White, R. S. and Bickle, M. J. (1997). Thin crust beneath Ocean Drilling Program borehole 735B at the Southwest Indian Ridge? Earth and Planetary Science Letters, 148, 93–107.CrossRefGoogle Scholar
Muller, M. R., Minshull, T. A. and White, R. S. (1999). Segmentation and melt supply at the Southwest Indian Ridge. Geology, 27, (10), 867–870.2.3.CO;2>CrossRefGoogle Scholar
Muller, M. R., Minshull, T. A. and White, R. S. (2000). Crustal structure of the Southwest Indian Ridge at the Atlantis II Fracture Zone. Journal of Geophysical Research, B, 105, 25809–25828.CrossRefGoogle Scholar
Müller, R. D., Sdrolias, M., Gaina, C. and Roest, W. R. (2008). Age, spreading rates, and spreading asymmetry of the world's ocean crust. Geochemistry Geophysics Geosystems, 9, .CrossRefGoogle Scholar
Mullineaux, L. S., Adams, D. K., Mills, S. W. and Beaulieu, S. E. (2010). Larvae from afar colonize deep-sea hydrothermal vents after a catastrophic eruption. Proceedings of the National Academy of Sciences, 107 (17), 7829–7834.CrossRefGoogle ScholarPubMed
Murray, J. and Hjort, J. (1912). The Depths of the Ocean. London: Macmillan & Co.Google Scholar
Murton, B. J., Schroth, N., LeBas, T. et al. (2013). Formation of volcanic crust at slow spreading mid-ocean ridges by steady state processes. Nature Geoscience, under review.
Mutter, C. Z. and Mutter, J. C. (1993). Variations in thickness of layer-3 dominate oceanic crustal structure. Earth and Planetary Science Letters, 117, (1–2), 295–317.CrossRefGoogle Scholar
Naar, D. F. and Hey, R. N. (1989). Speed limit for oceanic transform faults. Geology, 17, 420–422.2.3.CO;2>CrossRefGoogle Scholar
Naar, D. F. and Hey, R. N. (1991). Tectonic evolution of the Easter microplate. Journal of Geophysical Research, 96, 7961–7973.CrossRefGoogle Scholar
Navin, D. A., Peirce, C. and Sinha, M. C. (1998). The RAMESSES experiment – II. Evidence for accumulated melt beneath a slow spreading ridge from wide-angle refraction and multichannel reflection seismic profiles. Geophysical Journal International, 135, (3), 746–772.CrossRefGoogle Scholar
Nazarova, K. (1994). Serpentinized peridotites as a possible source for oceanic magnetic anomalies. Marine Geophysical Researches, 16, 455–462.CrossRefGoogle Scholar
Neumann, G. A. and Forsyth, D. W. (1993). The paradox of the axial profile: isostatic compensation along the axis of the Mid-Atlantic Ridge?Journal of Geophysical Research, 98, (B10), 17891–17910.CrossRefGoogle Scholar
Neves, M. C., Searle, R. C. and Bott, M. H. P. (2003). Easter microplate dynamics. Journal of Geophysical Research, 108, (B4), ETG14, .CrossRefGoogle Scholar
Nicolas, A. (1989). Structure of Ophiolites and Dynamics of Oceanic Lithosphere. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Nicolas, A., Boudier, E., Ildefonse, B. and Ball, E. (2000). Accretion of Oman and United Arab Emirates ophiolite – discussion of a new structural map. Marine Geophysical Research, 21, (3–4), 147–179.CrossRefGoogle Scholar
Niu, Y. and O’Hara, M. J. (2007). Global correlations of ocean ridge basalt chemistry with axial depth: a new perspective. Journal of Petrology, 49, (4), 633–664.CrossRefGoogle Scholar
Nooner, S. L. and Chadwick, W. W. (2009). Volcanic inflation measured in the caldera of Axial Seamount: implications for magma supply and future eruptions. Geochemistry Geophysics Geosystems, 10, .CrossRefGoogle Scholar
Normark, W. R., Morton, J. L. and Ross, S. L. (1987). Submersible observations along the southern Juan-de-Fuca Ridge – 1984 Alvin program. Journal of Geophysical Research-Solid Earth and Planets, 92, (B11), 11283–11290.CrossRefGoogle Scholar
Officer, C. B., Ewing, J. I., Hennion, J. F., Harkrider, D. G. and Miller, D. E. (1959). Geophysical investigations in the eastern Caribbean: summary of 1955 and 1956 cruises. In Physics and Chemistry of the Earth, volume 3. London: Pergamon, pp. 17–26.Google Scholar
Ohara, Y., Fujioka, K., Ishii, T. and Yurimoto, H. (2003). Peridotites and gabbros from the Parece Vela backarc basin: unique tectonic window in an extinct backarc spreading ridge. Geochemistry Geophysics Geosystems, 4, 8611, .CrossRefGoogle Scholar
Okino, K., Matsuda, K., Christie, D. M., Nogi, Y. and Koizumi, K.-I. (2004). Development of oceanic detachment and asymmetric spreading at the Australian–Antarctic Discordance. Geochemistry Geophysics Geosystems, 5, (12), Q12012, .CrossRefGoogle Scholar
Oldenburg, D. W. (1975). A physical model for the creation of the lithosphere. Geophysical Journal of the Royal Astronomical Society, 43, (2), 425–451.CrossRefGoogle Scholar
Oldenburg, D. W. and Brune, J. N. (1972). Ridge transform fault spreading pattern in freezing wax. Science, 178, 301–304.CrossRefGoogle ScholarPubMed
Oldenburg, D. W. and Brune, J. N. (1975). An explanation for the orthogonality of ocean ridges and transform faults. Journal of Geophysical Research, 80, 2575–2585.CrossRefGoogle Scholar
Olive, J.-A., Behn, M. D. and Tucholke, B. E. (2010). The structure of oceanic core complexes controlled by the depth distribution of magma emplacement. Nature Geoscience, 3, 491–495.CrossRefGoogle Scholar
O’Neill, H. S. C. and Jenner, F. E. (2012). The global pattern of trace-element distributions in ocean floor basalts. Nature, 491, (7426), 698–705.CrossRefGoogle ScholarPubMed
Oufi, O., Cannat, M. and Horen, H. (2002). Magnetic properties of variably serpentinized abyssal peridotites. Journal of Geophysical Research-Solid Earth, 107, (B5), EPM 3–1–EPM 3–19, .CrossRefGoogle Scholar
Paduan, J. B., Caress, D. W., Clague, D. A., Paull, C. K. and Thomas, H. (2009). High-resolution mapping of mass wasting, tectonic, and volcanic hazards using the MBARI Mapping AUV. In Extended Abstracts of the International Conference on Seafloor Mapping for Geohazard Assessment, May 11–13, Forio d’Ischia, ed. Chiocci, F. L., Ridente, D., Casalbore, D. and Bosman, A., Forio d’Ischia, Italy: Società Geologica Italiana.Google Scholar
Pariso, J. E. and Johnson, H. P. (1991). Alteration processes at Deep-Sea Drilling Project Ocean Drilling Program Hole 504B at the Costa-Rica Rift – implications for magnetization of oceanic-crust. Journal of Geophysical Research-Solid Earth and Planets, 96, (B7), 11703–11722.CrossRefGoogle Scholar
Pariso, J. E. and Johnson, H. P. (1993a). Do lower crustal rocks record reversals of the earth's magnetic-field? Magnetic petrology of oceanic gabbros from Ocean Drilling Program hole-735b. Journal of Geophysical Research, 98, (B9), 16013–16032.CrossRefGoogle Scholar
Pariso, J. E. and Johnson, H. P. (1993b). Do layer-3 rocks make a significant contribution to marine magnetic-anomalies – in-situ magnetization of gabbros at Ocean Drilling Program Hole-735b. Journal of Geophysical Research-Solid Earth, 98, (B9), 16033–16052.CrossRefGoogle Scholar
Parker, R. L. (1972). The rapid calculation of potential anomalies. Geophysical Journal of the Royal Astronomical Society, 31, 447–455.CrossRefGoogle Scholar
Parker, R. L. and Huestis, S. P. (1974). The inversion of magnetic anomalies in the presence of topography. Journal of Geophysical Research, 79, 1587–1593.CrossRefGoogle Scholar
Parker, R. L. and Klitgord, K. D. (1972). Magnetic upward continuation from an uneven track. Geophysics, 37, (4), 662–668.CrossRefGoogle Scholar
Parker, R. L. and Oldenburg, D. W. (1973). Thermal model of ocean ridges. Nature Physical Science, 242, 137–139.CrossRefGoogle Scholar
Parson, L. M., Murton, B. J. and Browning, P. (1992). Ophiolites and their Modern Oceanic Analogues, Special Publication, 60. London: Geological Society, 330 pp.Google Scholar
Parson, L. M., Murton, B. J., Searle, R. C. et al. (1993). En echelon volcanic ridges at the Reykjanes Ridge: a life cycle of volcanism and tectonics. Earth and Planetary Science Letters, 117, 73–87.CrossRefGoogle Scholar
Parsons, B. and Sclater, J. G. (1977). An analysis of the variation of ocean floor bathymetry and heat flow with age. Journal of Geophysical Research, 82, 803–827.CrossRefGoogle Scholar
Patriat, P. and Courtillot, V. (1984). On the stability of triple junctions and its relation to episodicity in spreading. Tectonics, 3, 317–332.CrossRefGoogle Scholar
Patterson, R. B. (1972). Increased Ranges for Conventional Underwater Cameras. Proc. S.P.I.E. (Proceedings of the International Society for Optics and Photonics), 24, 153–161.Google Scholar
Pearce, J. A., Lippard, S. J. and Roberts, S. (1984). Characteristics and tectonic significance of supra-subduction zone ophiolites. In Marginal Basin Geology, Special Publication, 16, ed. Kokelaar, B. P. and Howells, M. F.. London: Geological Society, pp. 77−94.Google Scholar
Pedersen, R. B., Thorseth, I. H., Nygård, T. E., Lilley, M. D. and Kelley, D. S. (2010a). Hydrothermal activity at the Arctic mid-ocean ridges. In Diversity of Hydrothermal Systems on Slow Spreading Ocean Ridges, Geophysical Monograph 188, ed. Rona, P. A., Devey, C. W., Dyment, J. and Murton, B. J.. Washington, D.C.: American Geophysical Union, pp. 67–89.CrossRefGoogle Scholar
Pedersen, R. B., Rapp, H. T., Thorseth, I. H. et al. (2010b). Discovery of a black smoker vent field and vent fauna at the Arctic Mid-Ocean Ridge. Nature Communications, 1, 126, .CrossRefGoogle ScholarPubMed
Pelli, D. G. and Chamberlain, S. C. (1989). The visibility of 350 °C black-body radiation by the shrimp Rimicaris exoculata and man. Nature, 337, (6206), 460–461.CrossRefGoogle ScholarPubMed
Penrose, (1972). Penrose field conference on ophiolites. Geotimes, 17, 24–25.Google Scholar
Perfit, M. R. and Chadwick, W. W. (1998). Magmatism at mid-ocean ridges: constraints from volcanological and geochemical investigations. In Faulting and Magmatism at Mid-Ocean Ridges, Geophysical Monograph 106, ed. Buck, W. R., Delaney, P. T., Karson, J. A. and Lagabrielle, Y.. Washington, D.C.: American Geophysical Union, pp. 59–115.Google Scholar
Perfit, M. R., Fornari, D. J., Smith, M. C. et al. (1994). Small-scale spatial and temporal variations in midocean ridge crest magmatic processes. Geology, 22, (4), 375–379.2.3.CO;2>CrossRefGoogle Scholar
Perk, N. W., Coogan, L. A., Karson, J. A., Klein, E. M. and Hanna, H. D. (2007). Petrology and geochemistry of primitive lower oceanic crust from Pito Deep: implications for the accretion of the lower crust at the Southern East Pacific Rise. Contributions to Mineralogy and Petrology, 154, (5), 575–590.CrossRefGoogle Scholar
Phillips, J. D., Driscoll, A. H., Peal, K. R., Marquet, W. M. and Owen, D. M. (1979). A new undersea geological survey tool: ANGUS. Deep Sea Research Part A, 26, (2), 211–225.CrossRefGoogle Scholar
Phipps Morgan, J. and Chen, Y. J. (1993). Dependence of ridge-axis morphology on magma supply and spreading rate. Nature, 364, (19 August), 706–708.CrossRefGoogle Scholar
Phipps Morgan, J. and Forsyth, D. W. (1988). 3-D flow and temperature perturbations due to a transform offset: effects on oceanic crustal and upper mantle structure. Journal of Geophysical Research, 93, 2955–2966.CrossRefGoogle Scholar
Phipps Morgan, J. and Parmentier, E. M. (1985). Causes and rate-limiting mechanism of ridge propagations: a fracture mechanics model. Journal of Geophysical Research, 90, 8603–8612.CrossRefGoogle Scholar
Pockalny, R. A., Detrick, R. S. and Fox, P. J. (1988). Morphology and tectonics of the Kane Transform from Sea Beam bathymetry data. Journal of Geophysical Research, 93, 3179–3193.CrossRefGoogle Scholar
Pockalny, R. A., Smith, A. and Gente, P. (1995). Spatial and temporal variability of crustal magnetization of a slowly spreading ridge: Mid-Atlantic Ridge (20°–24° N). Marine Geophysical Researches, 17, (3), 301–320.CrossRefGoogle Scholar
Pockalny, R. A., Gente, P. and Buck, R. (1996). Oceanic transverse ridges: a flexural response to fracture-zone-normal extension. Geology, 24, (1), 71–74.2.3.CO;2>CrossRefGoogle Scholar
Poliakov, A. N. B. and Buck, W. R. (1998). Mechanics of stretching elastic-plastic-viscous layers: applications to slow-spreading mid-ocean ridges. In Faulting and Magmatism at Mid-Ocean Ridges, ed. Buck, W. R., Delaney, P. T., Karson, J. A. and Lagabrielle, Y.. Washington, D.C.: American Geophysical Union, pp. 305–323.Google Scholar
Pollack, H. N., Hurter, S. J. and Johnson, J. R. (1993). Heat-flow from the Earth's interior – analysis of the global data set. Reviews of Geophysics, 31, (3), 267–280.CrossRefGoogle Scholar
Pollard, D. D. and Aydin, A. (1984). Propagation and linkage of oceanic ridge segments. Journal of Geophysical Research, 89, 10017–10028.CrossRefGoogle Scholar
Pollard, D. D., Delaney, P. T., Duffield, W. A., Endo, E. T. and Okamura, A. T. (1983). Surface deformation in volcanic rift zones. Tectonophysics, 94, (1–4), 541–584.CrossRefGoogle Scholar
Powell, C. (1971). Decca, LORAN and Omega – amphibious aids to navigation. Radio and Electronic Engineer, 41, (12), S184.Google Scholar
Prévot, M. and Grommé, S. (1975). Intensity of magnetization of sub-aerial and submarine basalts and its possible change with time. Geophysical Journal of the Royal Astronomical Society, 40, 207–224.CrossRefGoogle Scholar
Prince, R. A. and Forsyth, D. W. (1988). Horizontal extent of anomalously thin crust near the Vema fracture-zone from the 3-dimensional analysis of gravity-anomalies. Journal of Geophysical Research-Solid Earth and Planets, 93, (B7), 8051–8063.CrossRefGoogle Scholar
Pruis, M. J. and Johnson, H. P. (2004). Tapping into the sub-seafloor: examining diffuse flow and temperature from an active seamount on the Juan de Fuca Ridge. Earth and Planetary Science Letters, 217, (3–4), 379–388.CrossRefGoogle Scholar
Purdy, G. M. and Detrick, R. S. (1986). Crustal structure of the Mid-Atlantic Ridge at 23° N from seismic refraction studies. Journal of Geophysical Research-Solid Earth and Planets, 91, (B3), 3739–3762.CrossRefGoogle Scholar
Purdy, G. M., Sempere, J.-C., Schouten, H., DuBois, D. L. and Goldsmith, R. (1990). Bathymetry of the Mid-Atlantic Ridge, 24°–31° N: a map series. Marine Geophysical Researches, 12, 247–252.CrossRefGoogle Scholar
Quick, J. E. and Delinger, R. P. (1993). Ductile deformation and the origin of layered gabbro in ophiolites. Journal of Geophysical Research, 98, (B8), 14015–14027.CrossRefGoogle Scholar
Raff, A. D. and Mason, R. G. (1961). Magnetic survey off the west coast of North-America, 40° N latitude to 52° N latitude. Geological Society of America Bulletin, 72, (8), 1267–1270.CrossRefGoogle Scholar
Raitt, R. W. (1956). Seismic-refraction studies of the Pacific ocean basin: part I: crustal thickness of the central equatorial Pacific. Bulletin of the Geological Society of America, 67, 1623–1640.CrossRefGoogle Scholar
Raitt, R. W. (1963). Chapter 6. The crustal rocks. In The Sea, volume 3, ed. Hill, M. N.. New York: John Wiley, pp. 85–102.Google Scholar
Ranero, C. R. and Reston, T. J. (1999). Detachment faulting at ocean core complexes. Geology, 27, (11), 983–986.2.3.CO;2>CrossRefGoogle Scholar
Reid, J. L. (1982). Evidence of an effect of heat-flux from the East Pacific Rise upon the characteristics of the mid-depth waters. Geophysical Research Letters, 9, (4), 381–384.CrossRefGoogle Scholar
Reinke-Kunze, C. (1994). Welt der Forschungsschiffe. Hamburg: DSV-Verlag GmbH, 192 pp.Google Scholar
Renard, V. and Allenou, J. P. (1979). Sea Beam, multi-beam echo-sounding in Jean Charcot – description, evaluation and 1st results. International Hydrographic Review, 56, (1), 35–67.Google Scholar
Reston, T. J. and Ranero, C. R. (2011). The 3-D geometry of detachment faulting at mid-ocean ridges. Geochemistry Geophysics Geosystems, 12, .CrossRefGoogle Scholar
Reynolds, J. R., Langmuir, C. H., Bender, J. F., Kastens, K. A. and Ryan, W. B. F. (1992). Spatial and temporal variability in the geochemistry of basalts from the East Pacific Rise. Nature, 359, (6395), 493–499.CrossRefGoogle Scholar
Riedel, W. R., Ladd, H. S., Tracey, J. I. and Bramlette, M. N. (1961). Preliminary drilling phase of Mohole project. Bulletin of the American Association of Petroleum Geologists, 45, 1793–1798.Google Scholar
Ritsema, J., Deuss, A., van Heijst, H. J. and Woodhouse, J. H. (2011). S40RTS: a degree-40 shear-velocity model for the mantle from new Rayleigh wave dispersion, teleseismic traveltime and normal-mode splitting function measurements. Geophysical Journal International, 184, (3), 1223–1236.CrossRefGoogle Scholar
Robertson, A. and Xenophontos, C. (1993). Development of concepts concerning the Troodos ophiolite and adjacent units in Cyprus. In Magmatic Processes and Plate Tectonics, ed. Prichard, H. M., Alabaster, T., Harris, N. B. W. and Neary, C. R.. London: Geological Society, pp. 85–119.Google Scholar
Robinson, P. T., Melson, W. G., Ohearn, T. and Schmincke, H. U. (1983). Volcanic glass compositions of the Troodos ophiolite, Cyprus. Geology, 11, (7), 400–404.2.0.CO;2>CrossRefGoogle Scholar
Rogers, A. D., Tyler, P. A., Connelly, D. P. et al. (2012). The discovery of new deep-sea hydrothermal vent communities in the Southern Ocean and implications for biogeography. Plos Biology, 10, (1), .CrossRefGoogle ScholarPubMed
Rona, P. A. (2010). Emerging diversity of hydrothermal systems on slow spreading ocean ridges. In Diversity of Hydrothermal Systems on Slow Spreading Ocean Ridges, Geophysical Monograph 188, ed. Rona, P. A., Devey, C. W., Dyment, J. and Murton, B. J.. Washington, D.C.: American Geophysical Union, pp. 5–10.CrossRefGoogle Scholar
Rona, P. A., Boström, K., Laubier, L. and Smith, K. L. (1983). Hydrothermal Processes at Seafloor Spreading Centers. New York: Plenum, 796 pp.CrossRefGoogle Scholar
Rona, P. A., Klinkhammer, G., Nelsen, T. A., Trefry, J. H. and Elderfield, H. (1986). Black smoker, massive sulfides and vent biota at the Mid-Atlantic Ridge. Nature, 321, 33–37.CrossRefGoogle Scholar
Rona, P. A., Jackson, D. R., Wen, T. et al. (1997). Acoustic mapping of diffuse flow at a seafloor hydrothermal site: Monolith Vent, Juan de Fuca Ridge. Geophysical Research Letters, 24, (19), 2351–2354.CrossRefGoogle Scholar
Rona, P. A., Devey, C. W., Dyment, J. and Murton, B. J. (editors) (2010). Diversity of Hydrothermal Systems on Slow Spreading Ocean Ridges, Geophysical Monograph 188. Washington, D.C.: American Geophysical Union, 431 pp.CrossRefGoogle Scholar
Rubin, A. M. (1995). Propagation of magma-filled cracks. Annual Review of Earth and Planetary Sciences, 23, 287–336.CrossRefGoogle Scholar
Rusby, J. S. M., Dobson, R., Edge, R. H., Pierce, F. E. and Somers, M. L. (1969). Records obtained from the trials of a long range side-scan sonar (GLORIA Project). Nature, 223, (5212), 125–126.CrossRefGoogle Scholar
Rusby, R. I. and Searle, R. C. (1993). Intraplate thrusting near the Easter microplate. Geology, 21, 311–314.2.3.CO;2>CrossRefGoogle Scholar
Rusby, R. I. and Searle, R. C. (1995). A history of the Easter microplate, 5.25 Ma to present. Journal of Geophysical Research, 100, (B7), 12617–12640.CrossRefGoogle Scholar
Russell, M. J., Hall, A. J. and Martin, W. (2010). Serpentinization as a source of energy at the origin of life. Geobiology, 8, (5), 355–371.CrossRefGoogle ScholarPubMed
Ryan, W. B. F., Carbotte, S. M., Coplan, J. O. et al. (2009). Global multi-resolution topography synthesis. Geochemistry Geophysics Geosystems, 10, .CrossRefGoogle Scholar
Rychert, C. A. and Shearer, P. M. (2009). A global view of the lithosphere–asthenosphere boundary. Science 324, 495–498.CrossRefGoogle ScholarPubMed
Rychert, C. A. and Shearer, P. M. (2011). Imaging the lithosphere–asthenosphere boundary beneath the Pacific using SS waveform modeling. Journal of Geophysical Research, 116, .CrossRefGoogle Scholar
Salisbury, M. H. and Christensen, N. I. (1978). Seismic velocity structure of a traverse through Bay of Islands ophiolite complex, Newfoundland, an exposure of oceanic-crust and upper mantle. Journal of Geophysical Research, 83, (B2), 805–817.CrossRefGoogle Scholar
Sandwell, D. T. (1991). Geophysical applications of satellite altimetry. Reviews of Geophysics, 29, part 1, supplement S, 132–137.CrossRefGoogle Scholar
Sandwell, D. T. and Smith, W. H. F. (2009). Global marine gravity from retracked Geosat and ERS-1 altimetry: Ridge segmentation versus spreading rate. Journal of Geophysical Research, 114, B0141, .CrossRefGoogle Scholar
Sarrazin, J., Rodier, P., Tivey, M. K. et al. (2009). A dual sensor device to estimate fluid flow velocity at diffuse hydrothermal vents. Deep-Sea Research Part I-Oceanographic Research Papers, 56, (11), 2065–2074.CrossRefGoogle Scholar
Sauter, D., Parson, L., Mendel, V. et al. (2002). TOBI sidescan sonar imagery of the very slow-spreading Southwest Indian Ridge: evidence for along-axis magma distribution. Earth and Planetary Science Letters, 199, (1–2), 81–95.CrossRefGoogle Scholar
Sauter, D., Mendel, V., Rommevau-Jestin, C. et al. (2004). Focused magmatism versus amagmatic spreading along the ultra-slow spreading Southwest Indian Ridge: evidence from TOBI side scan imagery. Geochemistry Geophysics Geosystems, 5, (10), .CrossRefGoogle Scholar
Sauter, D., Cannat, M., Rouméjon, S. et al. (2013). 11 Myr-continuous exhumation of mantle-derived rocks at the Southwest Indian Ridge. Nature Geoscience, .Google Scholar
Scheirer, D. S. and Macdonald, K. C. (1995). Near-axis seamounts on the flanks of the East Pacific Rise, 8° N to 17° N. Journal of Geophysical Research, 100, (B2), 2239–2259.CrossRefGoogle Scholar
Schmerr, N. (2012). The Gutenberg discontinuity: melt at the lithosphere–asthenosphere boundary. Science, 335, (6075), 1480–1483.CrossRefGoogle ScholarPubMed
Schouten, H. and Denham, C. (1979). Modelling the oceanic magnetic source layer. In Deep Drilling Results in the Atlantic Ocean, ed. Talwani, M.. Washington, D.C.: American Geophysical Union, pp. 151–159.CrossRefGoogle Scholar
Schouten, H. and McCamy, K. (1972). Filtering marine magnetic anomalies. Journal of Geophysical Research, 77, 7089–7099.CrossRefGoogle Scholar
Schouten, H. and White, R. S. (1980). Zero-offset fracture zones. Geology, 8, 175–179.2.0.CO;2>CrossRefGoogle Scholar
Schouten, H., Klitgord, K. D. and Whitehead, J. A. (1985). Segmentation of mid-ocean ridges. Nature, 317, 225–229.CrossRefGoogle Scholar
Schouten, H., Klitgord, K. D. and Gallo, D. G. (1993). Edge-driven microplate kinematics. Journal of Geophysical Research, 98, 6689–6701.CrossRefGoogle Scholar
Schouten, H., Tivey, M. A., Fornari, D. J. and Cochran, J. R. (1999). Central anomaly magnetization high: constraints on the volcanic construction and architecture of seismic layer 2A at a fast-spreading mid-ocean ridge, the East Pacific Rise at 9°30′–50ʹ N. Earth and Planetary Science Letters, 169, 37–50.CrossRefGoogle Scholar
Schouten, H., Smith, D. K., Cann, J. R. and Escartín, J. (2010). Tectonic versus magmatic extension in the presence of core complexes at slow-spreading ridges from a visualization of faulted seafloor topography. Geology, 38, (7), 615–618.CrossRefGoogle Scholar
Schroeder, T., John, B. and Frost, B. R. (2002). Geologic implications of seawater circulation through peridotite exposed at slow-spreading mid-ocean ridges. Geology, 30, (4), 367–370.2.0.CO;2>CrossRefGoogle Scholar
Schultz, A. and Elderfield, H. (1997). Controls on the physics and chemistry of seafloor hydrothermal circulation. Philosophical Transactions of the Royal Society of London, Series A, 355, (1723), 387–425.CrossRefGoogle Scholar
Schultz, A., Delaney, J. R. and McDuff, R. E. (1992). On the partitioning of heat-flux between diffuse and point-source sea-floor venting. Journal of Geophysical Research-Solid Earth, 97, (B9), 12299–12314.CrossRefGoogle Scholar
Schultz, A., Dickson, P. and Elderfield, H. (1996). Temporal variations in diffuse hydrothermal flow at TAG. Geophysical Research Letters, 23, (23), 3471–3474.CrossRefGoogle Scholar
Schwarz, A. (2011). Papers from Origins 2011, Origins of Life and Evolution of Biospheres, Special Issue: 41, 495–632.
Sclater, J. G., Anderson, R. N. and Bell, M. L. (1971). Elevation of ridges and evolution of central eastern Pacific. Journal of Geophysical Research, 76, (32), 7888–7915.CrossRefGoogle Scholar
Sclater, J. G., Parsons, B. and Jaupart, C. (1981). Oceans and continents: similarities and differences in the mechanisms of heat loss. Journal of Geophysical Research, 86, 11535–11552.CrossRefGoogle Scholar
Scott, R. B., Rona, P. A., McGregor, B. A. and Scott, M. R. (1974). TAG hydrothermal field. Nature, 251, (5473), 301–302.CrossRefGoogle Scholar
Seama, N., Nogi, Y. and Isezaki, N. (1993). A new method for precise determination of the position and strike of magnetic boundaries using vector data of the geomagnetic anomaly field. Geophysical Journal International, 113, 155–164.CrossRefGoogle Scholar
Searle, R. C. (1980). Tectonic pattern of the Azores spreading centre and triple junction. Earth and Planetary Science Letters, 51, 415–434.CrossRefGoogle Scholar
Searle, R. C. (1981). The active part of Charlie–Gibbs Fracture Zone: a study using sonar and other geophysical techniques. Journal of Geophysical Research, 86, 243–262.CrossRefGoogle Scholar
Searle, R. C. (1983). Multiple, closely spaced transform faults in fast-slipping fracture zones. Geology, 11, 607–610.2.0.CO;2>CrossRefGoogle Scholar
Searle, R. C. (1984). GLORIA survey of the East Pacific Rise near 3.5° S: tectonic and volcanic characteristics of a fast-spreading mid-ocean rise. Tectonophysics, 101, 319–344.CrossRefGoogle Scholar
Searle, R. C. (1986). GLORIA investigations of oceanic fracture zones: comparative studies of the transform fault zone. Journal of the Geological Society of London, 143, 743–756.CrossRefGoogle Scholar
Searle, R. C. (2012). Are axial volcanic ridges where all the (volcanic) action is? Paper OS11E-06, American Geophysical Union Fall Meeting, San Francisco.
Searle, R. C. and Escartín, J. (2004). The rheology and morphology of oceanic lithosphere and mid-ocean ridges. In Mid-Ocean Ridges: Hydrothermal Interactions between the Lithosphere and Oceans, Geophysical Monograph 148, ed. German, C., Lin, J. and Parson, L. M.. Washington, D.C.: American Geophysical Union, pp. 63–94.Google Scholar
Searle, R. C. and Hey, R. N. (1983). GLORIA observations of the propagating rift at 95.5° W on the Cocos–Nazca spreading center. Journal of Geophysical Research, 88, 6433–6447.CrossRefGoogle Scholar
Searle, R. C. and Laughton, A. S. (1977). Sonar studies of the Mid-Atlantic Ridge crest near Kurchatov Fracture Zone. Journal of Geophysical Research, 82, 5313–5328.CrossRefGoogle Scholar
Searle, R. C., Rusby, R. I., Engeln, J. et al. (1989). Comprehensive sonar imaging of the Easter microplate. Nature, 341, 701–705.CrossRefGoogle Scholar
Searle, R. C., Le Bas, T. P., Mitchell, N. C. et al. (1990). GLORIA image processing: the state of the art. Marine Geophysical Researches, 12, 21–39.CrossRefGoogle Scholar
Searle, R. C., Bird, R. T., Rusby, R. I. and Naar, D. F. (1993). The development of two oceanic microplates: Easter and Juan Fernandez microplates, East Pacific Rise. Journal of the Geological Society of London, 150, 965–976.CrossRefGoogle Scholar
Searle, R. C., Keeton, J. A., Lee, S. M. et al. (1998a). The Reykjanes Ridge: structure and tectonics of a hot-spot influenced, slow-spreading ridge, from multibeam bathymetric, gravity and magnetic investigations. Earth and Planetary Science Letters, 160, 463–478.CrossRefGoogle Scholar
Searle, R. C., Cowie, P. A., Mitchell, N. C. et al. (1998b). Fault structure and detailed evolution of a slow spreading ridge segment: the Mid-Atlantic Ridge at 29° N. Earth and Planetary Science Letters, 154, (1–4), 167–183.CrossRefGoogle Scholar
Searle, R. C., Cannat, M., Fujioka, K. et al. (2003). The FUJI Dome: A large detachment fault near 64° E on the very slow-spreading southwest Indian Ridge. Geochemistry Geophysics Geosystems, 4, (8), 9105, .CrossRefGoogle Scholar
Searle, R. C., Francheteau, J. and Armijo, R. (2006). Compressional deformation north of the Easter microplate: a manned submersible and seafloor gravity investigation. Geophysical Journal International, 164, 359–369, .CrossRefGoogle Scholar
Searle, R. C., Murton, B. J., Achenbach, K. et al. (2010). Structure and development of an axial volcanic ridge: Mid-Atlantic Ridge, 45° N. Earth and Planetary Science Letters, 299, 228–241.CrossRefGoogle Scholar
Sempéré, J.-C. and Macdonald, K. C. (1986). Overlapping spreading centers: implications from crack growth by the displacement discontinuity method. Tectonics, 5, 151–163.CrossRefGoogle Scholar
Sempéré, J.-C., Lin, J., Brown, H. S., Schouten, H. and Purdy, G. M. (1993). Segmentation and morphotectonic variations along a slow spreading center: the Mid-Atlantic Ridge (24°00ʹ N–30°40ʹ N). Marine Geophysical Researches, 15, (3), 153–200.CrossRefGoogle Scholar
Sempéré, J.-C., Cochran, J. R. and SEIR Scientific Team (1997). The Southeast Indian Ridge between 88° E and 118° E: variations in crustal accretion at constant spreading rate. Journal of Geophysical Research, 102, (B7), 15489–15505.CrossRefGoogle Scholar
Severinghaus, J. P. and Macdonald, K. C. (1988). High inside corners at ridge-transform intersections. Marine Geophysical Researches, 9, 353–367.CrossRefGoogle Scholar
Shah, A. K. and Buck, W. R. (2001). Causes for axial high topography at mid-ocean ridges and the role of crustal thermal structure. Journal of Geophysical Research-Solid Earth, 106, (B12), 30865–30879.CrossRefGoogle Scholar
Shah, A. K. and Buck, W. R. (2003). Plate bending stresses at axial highs, and implications for faulting behavior. Earth and Planetary Science Letters, 211, (3–4), 343–356.CrossRefGoogle Scholar
Shaw, P. (1992). Ridge segmentation, faulting and crustal thickness in the Atlantic Ocean. Nature, 358, 490–493.CrossRefGoogle Scholar
Shaw, P. R. and Lin, J. (1993). Causes and consequences of variations in faulting style at the Mid-Atlantic Ridge. Journal of Geophysical Research, 98, (B12), 21839–21851.CrossRefGoogle Scholar
Shemenda, A. I. and Grocholsky, A. L. (1994). Physical modelling of slow seafloor spreading. Journal of Geophysical Research, 99, (B5), 9137–9153.CrossRefGoogle Scholar
Shepard, F. P. (1948). Submarine Geology. New York: Harper and Brothers, 348 pp.Google Scholar
Shepard, F. P. (1959). The Earth Beneath the Sea. London: Oxford University Press, 275 pp.Google Scholar
Sichler, B. and Hekinian, R. (2002). Three-dimensional inversion of marine magnetic anomalies on the equatorial Atlantic Ridge (St. Paul Fracture Zone): delayed magnetization in a magmatically starved spreading center?Journal of Geophysical Research-Solid Earth, 107, (B12), .CrossRefGoogle Scholar
Sims, K. W. W., Blichert-Toft, J., Fornari, D. J. et al. (2003). Aberrant youth: chemical and isotopic constraints on the origin of off-axis lavas from the East Pacific Rise, 9°–10° N. Geochemistry Geophysics Geosystems, 4, (10), 8621, .CrossRefGoogle Scholar
Singh, S. C., Crawford, W. C., Carton, H. et al. (2006a). Discovery of a magma chamber and faults beneath a Mid-Atlantic Ridge hydrothermal field. Nature, 442, (7106), 1029–1032.CrossRefGoogle ScholarPubMed
Singh, S. C., Harding, A. J., Kent, G. M. et al. (2006b). Seismic reflection images of the Moho underlying melt sills at the East Pacific Rise. Nature, 442, (7100), 287–290.CrossRefGoogle ScholarPubMed
Sinha, M. and Evans, R. L. (2004). Geophysical constraints on the thermal regime of the oceanic crust. In Mid-Ocean Ridges: Hydrothermal Interactions between Lithosphere and Oceans, Geophysical Monograph 148, ed. German, C., Lin, J., Tribuzio, R. et al. Washington, D.C.: American Geophysical Union, pp. 19–62.Google Scholar
Sinha, M. C., Patel, P. D., Unsworth, M. J., Owen, T. R. E. and MacCormack, M. R. G. (1990). An active source electromagnetic sounding system for marine use. Marine Geophysical Researches, 12, (1–2), 59–68.CrossRefGoogle Scholar
Sinha, M. C., Constable, S. C., Peirce, C. et al. (1998). Magmatic processes at slow-spreading ridges: implications of the RAMESSES experiment at 57°45ʹ N on the Mid-Atlantic Ridge. Geophysical Journal International, 135, (3), 731–745.CrossRefGoogle Scholar
Sinton, J. M. and Detrick, R. S. (1992). Mid-ocean ridge magma chambers. Journal of Geophysical Research, 97, 197–216.CrossRefGoogle Scholar
Sinton, J. M., Smaglik, S. M., Mahoney, J. J. and Macdonald, K. C. (1991). Magmatic processes at superfast spreading midocean ridges – glass compositional variations along the East Pacific Rise 13°–23° S. Journal of Geophysical Research-Solid Earth and Planets, 96, (B4), 6133–6155.CrossRefGoogle Scholar
Sinton, J. M., Bergmanis, E., Rubin, K. et al. (2002). Volcanic eruptions on mid-ocean ridges: new evidence from the superfast spreading East Pacific Rise, 17°–19° S. Journal of Geophysical Research, 107, (B6), ECV3 1–20, .CrossRefGoogle Scholar
Sleep, N. H. and Biehler, S. (1970). Topography and tectonics at the intersection of fracture zones with central rifts. Journal of Geophysical Research, 75, 2748–2752.CrossRefGoogle Scholar
Small, C. and Sandwell, D. T. (1989). An abrupt change in ridge-axis gravity with spreading rate. Journal of Geophysical Research, 94, (B12), 17388–17392.CrossRefGoogle Scholar
Smallwood, J. R. and White, R. S. (1998). Crustal accretion at the Reykjanes-Ridge, 61–62° N. Journal of Geophysical Research, 103, 5185–5201.CrossRefGoogle Scholar
Smith, D. K. and Cann, J. R. (1990). Hundreds of small volcanoes on the median valley floor of the Mid-Atlantic Ridge at 24–30° N. Nature, 348, 152–155.CrossRefGoogle Scholar
Smith, D. K. and Cann, J. R. (1992). The role of seamount volcanism in crustal construction at the Mid-Atlantic Ridge (24°–30° N). Journal of Geophysical Research, 97, 1645–1658.CrossRefGoogle Scholar
Smith, D. K. and Cann, J. R. (1999). Constructing the upper crust of the Mid-Atlantic Ridge: a reinterpretation based on Puna Ridge, Kilauea Volcano. Journal of Geophysical Research, B, 104, 25379–25399.CrossRefGoogle Scholar
Smith, D. K., Humphris, S. E. and Bryan, W. B. (1995a). A comparison of volcanic edifices at the Reykjanes Ridge and the Mid-Atlantic Ridge at 24°–30° N. Journal of Geophysical Research, 100, (B11), 22485–22498.CrossRefGoogle Scholar
Smith, D. K., Cann, J. R., Dougherty, M. E. et al. (1995b). Mid-Atlantic Ridge volcanism from deep-towed side-scan sonar images, 25°–29° N. Journal of Volcanology and Geothermal Research, 67, (4), 233–262.CrossRefGoogle Scholar
Smith, D. K., Tolstoy, M., Fox, C. G. et al. (2002). Hydroacoustic monitoring of seismicity at the slow-spreading Mid-Atlantic Ridge. Geophysical Research Letters, 29, (11), 13-1–13-4.CrossRefGoogle Scholar
Smith, D. K., Escartin, J., Cannat, M. et al. (2003). Spatial and temporal distribution of seismicity along the northern Mid-Atlantic Ridge (15°–35° N). Journal of Geophysical Research, 108, (B3), 2167, .CrossRefGoogle Scholar
Smith, D. K., Cann, J. R. and Escartin, J. (2006). Widespread active detachment faulting and core complex formation near 13° N on the Mid-Atlantic Ridge. Nature, 442, 440–443, .CrossRefGoogle ScholarPubMed
Smith, D. K., Escartin, J., Schouten, H. and Cann, J. R. (2008). Fault rotation and core complex formation: significant processes in seafloor formation at slow-spreading midocean ridges (Mid-Atlantic Ridge, 13°–15° N). Geochemistry Geophysics Geosystems, 9, (3), .CrossRefGoogle Scholar
Smith, M. C., Perfit, M. R. and Jonasson, I. R. (1994). Petrology and geochemistry of basalts from the southern Juan-de-Fuca Ridge – controls on the spatial and temporal evolution of midocean ridge basalt. Journal of Geophysical Research-Solid Earth, 99, (B3), 4787–4812.CrossRefGoogle Scholar
Sohn, R. A. and Sims, K. W. W. (2005). Bending as a mechanism for triggering off-axis volcanism on the East Pacific Rise. Geology, 33, 93–96.CrossRefGoogle Scholar
Sohn, R. A., Webb, S. C., Hildebrand, J. A. and Cornuelle, B. D. (1997). Three-dimensional tomographic velocity structure of upper crust, co-axial segment, Juan de Fuca Ridge: implications for on-axis evolution and hydrothermal circulation. Journal of Geophysical Research, 102, (B8), 17679–17695.CrossRefGoogle Scholar
Sohn, R. A., Hildebrand, J. A. and Webb, S. C. (1998). Postrifting seismicity and a model for the 1993 diking event on the CoAxial segment, Juan de Fuca Ridge. Journal of Geophysical Research, B, 103, (5), 9867–9877.CrossRefGoogle Scholar
Sohn, R. A., Hildebrand, J. A. and Webb, S. C. (1999). A microearthquake survey of the high-temperature vent fields on the volcanically active East Pacific Rise (9°50ʹ N). Journal of Geophysical Research-Solid Earth, 104, (B11), 25367–25377.CrossRefGoogle Scholar
Sohn, R. and 21 others (2008). Explosive volcanism on the ultraslow-spreading Gakkel ridge, Arctic Ocean. Nature, 453, (7199), 1236–1238.CrossRefGoogle ScholarPubMed
Soule, S. A., Fornari, D. J., Perfit, M. R. et al. (2005). Channelized lava flows at the East Pacific Rise crest 9°–10° N: the importance of off-axis lava transport in developing the architecture of young oceanic crust. Geochemistry Geophysics Geosystems, 6, .CrossRefGoogle Scholar
Soule, S. A., Fornari, D. J., Perfit, M. R. and Rubin, K. H. (2007). New insights into mid-ocean ridge volcanic processes from the 2005–2006 eruption of the East Pacific Rise, 9°46ʹ N–9°56ʹ N. Geology, 35, (12), 1079–1082.CrossRefGoogle Scholar
Soule, S. A., Escartin, J. and Fornari, D. J. (2009). A record of eruption and intrusion at a fast spreading ridge axis: axial summit trough of the East Pacific Rise at 9–10° N. Geochemistry Geophysics Geosystems, 10, .CrossRefGoogle Scholar
Sparks, D. W. and Parmentier, E. M. (1991). Melt extraction from the mantle beneath spreading centers. Earth and Planetary Science Letters, 105, 368–377.CrossRefGoogle Scholar
Sparks, D. W., Parmentier, E. M. and Morgan, J. P. (1993). Three-dimensional mantle convection beneath a segmented spreading center: implications for along-axis variations in crustal thickness and gravity. Journal of Geophysical Research, 98, 21977–21995.CrossRefGoogle Scholar
Spencer, J. E. (1999). Geologic continuous casting below continental and deep-sea detachment faults and at the striated extrusion of Sacsayhuaman, Peru. Geology, 27, (4), 327–330.2.3.CO;2>CrossRefGoogle Scholar
Spiess, F. N. and Maxwell, A. E. (1964). Search for Thresher. Science, 145, (3630), 349–355.CrossRefGoogle ScholarPubMed
Spiess, F. N. and Mudie, J. D. (1970). Small scale topographic and magnetic features. In The Sea, volume 4 (Part I), ed. Maxwell, A. E., Bullard, E. and Worzel, J. L.. New York: Wiley-Interscience, pp. 205–250.Google Scholar
Spiess, F. N., Macdonald, K. C., Atwater, T. et al. (1980). East Pacific Rise – hot springs and geophysical experiments. Science, 207, (4438), 1421–1433.CrossRefGoogle ScholarPubMed
Spudich, P. and Orcutt, J. (1980). A new look at the seismic velocity structure of the oceanic-crust. Reviews of Geophysics, 18, (3), 627–645.CrossRefGoogle Scholar
Stakes, D. S., Holloway, G. L., Tucker, P. et al. (1997). Diamond rotary coring from an ROV or submersible for hardrock sample recovery and instrument deployment: the MBARI Multiple-Barrel Rock Coring System. Marine Technology Society Journal, 31, (3), 11–20.Google Scholar
Stakes, D. S., Perfit, M. R., Tivey, M. A. et al. (2006). The Cleft revealed: geologic, magnetic, and morphologic evidence for construction of upper oceanic crust along the southern Juan de Fuca Ridge. Geochemistry Geophysics Geosystems, 7, .CrossRefGoogle Scholar
Standish, J. J. and Sims, K. W. W. (2006). Lava emplacement and crustal architecture within an ultraslow-spreading rift valley. Geochimica et Cosmochimica Acta, 70, (18), A611.CrossRefGoogle Scholar
Standish, J. J. and Sims, K. W. W. (2010). Young off-axis volcanism along the ultraslow-spreading Southwest Indian Ridge. Nature Geoscience, 3, (4), 286–292.CrossRefGoogle Scholar
Stein, C. A. and Stein, S. (1992). A model for the global variation in oceanic depth and heat flow with lithospheric age. Nature, 359, 123–129.CrossRefGoogle Scholar
Stein, C. A. and Stein, S. (1994). Constraints on hydrothermal heat flux through the oceanic lithosphere from global heat flow. Journal of Geophysical Research, 99, 3081–3096.CrossRefGoogle Scholar
Stein, C. A. and Stein, S. (1996). Thermo-mechanical evolution of oceanic lithosphere: implications for the subduction process and deep earthquakes (overview). In Subduction: Top to Bottom. Geophysical Monograph 96, ed. G. E. Bebout, et al. Washington: American Geophysical Union, pp. 1–17.CrossRef
Stommel, H. (1982). Is the South Pacific He-3 plume dynamically active? Earth and Planetary Science Letters, 61, (1), 63–67.CrossRefGoogle Scholar
Strens, M. R. and Cann, J. R. (1986). A fracture-loop thermal balance model of black smoker circulation. Tectonophysics, 122, (3–4), 307–324.CrossRefGoogle Scholar
Stride, A. (2010). 14. Side-scan sonar – a tool for seafloor geology. In Of Seas and Ships and Scientists, ed. Laughton, A., Gould, J., Tucker, T. and Roe, H.. Cambridge: Lutterworth Press, pp. 193–207.Google Scholar
Sturm, M. E., Goldstein, S. J., Klein, E. M., Karson, J. A. and Murrell, M. T. (2000). Uranium-series age constraints on lavas from the median valley of the Mid-Atlantic Ridge, MARK area. Earth and Planetary Science Letters, 181, (1–2), 61–70.CrossRefGoogle Scholar
Swift, S. A. and Stephen, R. A. (1992). How much gabbro is in ocean seismic layer-3? Geophysical Research Letters, 19, (18), 1871–1874.CrossRefGoogle Scholar
Swift, S. A., Lizarralde, D., Stephen, R. A. and Hoskins, H. (1998). Velocity structure in upper ocean crust at Hole 504B from vertical seismic profiles. Journal of Geophysical Research-Solid Earth, 103, (B7), 15361–15376.Google Scholar
Sykes, L. R. (1967). Mechanism of earthquakes and nature of faulting on the mid-oceanic ridges. Journal of Geophysical Research, 72, 5–27.CrossRefGoogle Scholar
Talwani, M. (1965). Computation with the help of a digital computer of magnetic anomalies caused by bodies of arbitrary shape. Geophysics, 30, 797–817.CrossRefGoogle Scholar
Talwani, M. and Heirtzler, J. R. (1964). Computation of magnetic anomalies caused by two-dimensional structures of arbitrary shape. Computers in the Mineral Industries, part 1, Stanford University Publications in the Geological Sciences, 9, 464–480.Google Scholar
Talwani, M., Heezen, B. C. and Worzel, J. L. (1961). Gravity anomalies, physiography, and crustal structure of Mid-Atlantic Ridge. Journal of Geophysical Research, 66, (8), 2565.Google Scholar
Talwani, M., Le Pichon, X. and Ewing, M. (1965). Crustal structure of the mid-ocean ridges 2. Computed model from gravity and seismic refraction data. Journal of Geophysical Research, 70, 341–352.CrossRefGoogle Scholar
Talwani, M., Windisch, C. C. and Langseth, M. G. (1971). Reykjanes Ridge crest: a detailed geophysical study. Journal of Geophysical Research, 76, 473–577.CrossRefGoogle Scholar
Tan, Y. and Helmberger, D. V. (2007). Trans-Pacific upper mantle shear velocity structure. Journal of Geophysical Research-Solid Earth, 112, (B8), .CrossRefGoogle Scholar
Tani, K., Dunkley, D. J. and Ohara, Y. (2011). Termination of backarc spreading: zircon dating of a giant oceanic core complex. Geology, 39, (1), 47–50.CrossRefGoogle Scholar
Tao, C. et al. (2009). New hydrothermal fields found along the SWIR during the Legs 5–7 of the Chinese DY115–20 Expedition. In American Geophysical Union, Fall Meeting, paper OS21A-1150.
Tao, C., Lin, J., Guo, S. et al. (2012). First active hydrothermal vents on an ultraslow-spreading center: Southwest Indian Ridge. Geology, 40, (1), 47–50.CrossRefGoogle Scholar
Tapponnier, P. and Francheteau, J. (1978). Necking of the lithosphere and the mechanics of slowly accreting plate boundaries. Journal of Geophysical Research, 83, 3955–3970.CrossRefGoogle Scholar
Tauxe, L. (1998). Paleomagnetic Principles and Practice. Dordrecht: Kluwer Academic Publishers, 299 pp.Google Scholar
Taylor, B., Crook, K. and Sinton, J. (1994). Extensional transform zones and oblique spreading centers. Journal of Geophysical Research, B, 99, (10), 19707–19718.CrossRefGoogle Scholar
Teagle, D. A. H., Ildefonse, B., Blum, P. and IODP Expedition 335 Scientists (2012). IODP Expedition 335: Deep Sampling in ODP Hole 1256D. Scientific Drilling, (13), 28–34.CrossRef
Telford, W. M., Geldart, L. P. and Sheriff, R. E. (1990). Applied Geophysics. 2nd edition. Cambridge: Cambridge University Press, 792 pp.CrossRefGoogle Scholar
Tharp, M. (1982). Mapping the ocean floor – 1947 to 1977. In The Ocean Floor, ed. Scrutton, R. A. and Talwani, M.. Chichester: John Wiley & Sons Ltd., pp. 19–31.Google Scholar
Thatcher, W. and Hill, D. P. (1995). A simple model for the fault-generated morphology of slow-spreading mid-oceanic ridges. Journal of Geophysical Research, 100, (B1), 561–570.CrossRefGoogle Scholar
Thomson, C. W. (1877). The Atlantic, A Preliminary Account of the General Results of the Exploring Voyage of H. M. S. Challenger. London: Macmillan.Google Scholar
Thurber, C. H. (1983). Earthquake locations and three-dimensional crustal structure in the Coyote Lake area, central California. Journal of Geophysical Research, 88, 8226–8236.CrossRefGoogle Scholar
Tivey, M. A. (1996). Vertical magnetic structure of ocean crust determined from near-bottom magnetic field measurements. Journal of Geophysical Research-Solid Earth, 101, (B9), 20275–20296.CrossRefGoogle Scholar
Tivey, M. A. and Johnson, H. P. (1987). The central anomaly magnetic high: implications for ocean crust construction and evolution. Journal of Geophysical Research, 92, 12685–12694.CrossRefGoogle Scholar
Tivey, M. A. and Johnson, H. P. (1993). Variations in oceanic crustal structure and implications for the fine-scale magnetic anomaly signal. Geophysical Research Letters, 20, (17), 1879–1882.CrossRefGoogle Scholar
Tivey, M. K., Olson, L. O., Miller, V. W. and Light, R. D. (1990). Temperature measurements during initiation and growth of a black smoker chimney. Nature 346, 51–54.CrossRefGoogle Scholar
Tivey, M. K., Humphris, S. E., Thompson, G., Hannington, M. D. and Rona, P. A. (1995). Deducing patterns of fluid-flow and mixing within the TAG active hydrothermal mound using mineralogical and geochemical data. Journal of Geophysical Research-Solid Earth, 100, (B7), 12527–12555.CrossRefGoogle Scholar
Tivey, M. A., Schouten, H. and Kleinrock, M. C. (2003). A near-bottom magnetic survey of the Mid-Atlantic Ridge axis at 26° N: implications for the tectonic evolution of the TAG segment. Journal of Geophysical Research-Solid Earth, 108, (B5), .CrossRefGoogle Scholar
Tizard, T. H. (1876). ‘Report on Temperatures’ and ‘General Summary of Atlantic Ocean temperatures’. In HMS Challenger, no. 7. Report on ocean soundings and temperatures. London: Admiralty.Google Scholar
Tolstoy, M., Harding, A. J. and Orcutt, J. A. (1993). Crustal thickness on the Mid-Atlantic Ridge – Bull's-eye gravity-anomalies and focused accretion. Science, 262, (5134), 726–729.CrossRefGoogle ScholarPubMed
Tolstoy, M., Harding, A. J., Orcutt, J. A. and the TERA Group (1997). Deepening of the axial magma chamber on the southern East Pacific Rise toward the Garrett Fracture Zone. Journal of Geophysical Research, 102, (B2), 3097–3108.CrossRefGoogle Scholar
Tolstoy, M., Cowen, J. P., Baker, E. T. et al. (2006). A sea-floor spreading event captured by seismometers. Science 314, 1920–1922.CrossRefGoogle ScholarPubMed
Tolstoy, M., Waldhauser, F., Bohnenstiehl, D. R., Weekly, R. T. and Kim, W.-Y. (2008). Seismic identification of along-axis hydrothermal flow on the East Pacific Rise. Nature, 451, (7175), .CrossRefGoogle ScholarPubMed
Toomey, D. R., Solomon, S. C., Purdy, G. M. and Murray, M. H. (1985). Microearthquakes beneath the median valley of the Mid-Atlantic Ridge near 23°N: hypocenters and focal mechanisms. Journal of Geophysical Research, 90, 5443–5458.CrossRefGoogle Scholar
Toomey, D. R., Solomon, S. C. and Purdy, G. M. (1988). Microearthquakes beneath the median valley of the Mid-Atlantic Ridge near 23° N: tomography and tectonics. Journal of Geophysical Research, 93, (B8), 9093–9112.CrossRefGoogle Scholar
Toomey, D. R., Purdy, G. M., Solomon, S. C. and Wilcock, W. S. D. (1990). The three-dimensional seismic velocity structure of the East Pacific Rise near latitude 9°30ʹ N. Nature, 347, 639–645.CrossRefGoogle Scholar
Toomey, D. R., Purdy, G. M., Barclay, A. H., Wolfe, C. J. and Solomon, S. C. (1993). FARA microearthquake experiments IV: implications of Mid-Atlantic Ridge seismicity for models of young oceanic lithosphere. EOS, Transactions of the American Geophysical Union, 74, (43), 601.Google Scholar
Toomey, D. R., Solomon, S. C. and Purdy, G. M. (1994). Tomographic imaging of the shallow crustal structure of the East Pacific Rise at 9°30ʹ N. Journal of Geophysical Research, 99, (B12), 24135–24157.CrossRefGoogle Scholar
Toomey, D. R., Wilcock, W. S. D., Conder, J. A. et al. (2002). Asymmetric mantle dynamics in the MELT region of the East Pacific Rise. Earth and Planetary Science Letters, 200, (3–4), 287–295.CrossRefGoogle Scholar
Tréhu, A. M. (1975). Depth versus (age)1/2: a perspective on mid-ocean rises. Earth and Planetary Science Letters, 27, 287–304.CrossRefGoogle Scholar
Triantafyllou, M. and Hoover, F. (1990). Cable dynamics for tethered underwater vehicles. In MIT SeaGrant College Program Report 90–4. Cambridge, Massachussetts: Massachussetts Institute of Technology.Google Scholar
Tucholke, B. E. and Lin, J. (1994). A geological model for the structure of ridge segments in slow spreading ocean crust. Journal of Geophysical Research, 99, 11937–11958.CrossRefGoogle Scholar
Tucholke, B. E. and Schouten, H. (1988/89). Kane Fracture Zone. Marine Geophysical Researches, 10, 1–39.CrossRefGoogle Scholar
Tucholke, B., Lin, J., Kleinrock, M. et al. (1997). Segmentation and crustal structure of the western Mid-Atlantic Ridge flank, 25°25ʹ–27°10ʹ N and 0–29 m.y. Journal of Geophysical Research, 102, (B5), 10203–10223.CrossRefGoogle Scholar
Tucholke, B., Lin, J. and Kleinrock, M. (1998). Megamullions and mullion structure defining oceanic metamorphic core complexes on the Mid-Atlantic Ridge. Journal of Geophysical Research, B, 103, (5), 9857–9866.CrossRefGoogle Scholar
Tucholke, B. E., Fujioka, K., Ishihara, T., Hirth, G. and Kinoshita, M. (2001). Submersible study of an oceanic megamullion in the central North Atlantic. Journal of Geophysical Research, B, 106, 16145–16161.CrossRefGoogle Scholar
Tucholke, B. E., Behn, M. D., Buck, W. R. and Lin, J. (2008). Role of melt supply in oceanic detachment faulting and formation of megamullions. Geology, 36, 455–458.CrossRefGoogle Scholar
Tucker, T. and McCartney, B. (2010). 16: Engineering and applied physics. In Of Seas and Ships and Scientists, ed. Laughton, A., Gould, J., Tucker, T. and Roe, H.. Cambridge: Lutterworth Press, pp. 233–267.Google Scholar
Tunnicliffe, V. (1991). The biology of hydrothermal vents – ecology and evolution. Oceanography and Marine Biology, 29, 319–407.Google Scholar
Tunnicliffe, V. and Fowler, C. M. R. (1996). Influence of sea-floor spreading on the global hydrothermal vent fauna. Nature, 379, (6565), 531–533.CrossRefGoogle Scholar
Tunnicliffe, V., McArthur, A. G. and McHugh, D. (1998). A biogeographical perspective of the deep-sea hydrothermal vent fauna. Advances in Marine Biology, 34, 353–442.CrossRefGoogle Scholar
Turcotte, D. L. (1974). Are transform faults thermal contraction cracks?Journal of Geophysical Research, 79, 2573–2577.CrossRefGoogle Scholar
Turcotte, D. L. and Schubert, G. (1982). Geodynamics: Applications of Continuum Physics to Geological Problems. New York: John Wiley & Sons, 450 pp.Google Scholar
Turner, S., Beier, C., Niu, Y. and Cook, C. (2011). U–Th–Ra disequilibria and the extent of off-axis volcanism across the East Pacific Rise at 9°30ʹ N, 10°30ʹ N, and 11°20ʹ N. Geochemistry Geophysics Geosystems, 12, .CrossRefGoogle Scholar
Unsworth, M. (1994). Exploration of midocean ridges with a frequency-domain electromagnetic system. Geophysical Journal International, 116, (2), 447–467.CrossRefGoogle Scholar
Van Andel, T. H. and Ballard, R. D. (1979). The Galapagos Rift at 86° W: 2. Volcanism, structure and evolution of the rift valley. Journal of Geophysical Research, 84, (B10), 5390–5406.CrossRefGoogle Scholar
Van Avendonk, H. J. A., Harding, A. J., Orcutt, J. A. and McClain, J. S. (2001). Contrast in crustal structure across the Clipperton transform fault from travel time tomography. Journal of Geophysical Research-Solid Earth, 106, (B6), 10961–10981.CrossRefGoogle Scholar
Van Dover, C. L., German, C. R., Speer, K. G., Parson, L. M. and Vrijenhoek, R. C. (2002). Evolution and biogeography of deep-sea vent and seep invertebrates. Science, 295, (5558), 1253–1257.CrossRefGoogle ScholarPubMed
Vening Meinesz, F. A. (1929). Theory and Practice of Pendulum Observations at Sea. Delft: Waltman.Google Scholar
Vera, E. E. and Diebold, J. B. (1994). Seismic imaging of oceanic layer 2a between 9°30ʹ N and 10° N on the East Pacific Rise from 2-ship wide-aperture profiles. Journal of Geophysical Research-Solid Earth, 99, (B2), 3031–3041.CrossRefGoogle Scholar
Vereshchaka, A. L. (1996). A new genus and species of caridean shrimp (crustacea: decapoda: alvinocarididae) from North Atlantic hydrothermal vents. Journal of the Marine Biological Association, U.K., 76, 951–961.CrossRefGoogle Scholar
Vine, F. J. (1966). Spreading of the ocean floor: new evidence. Science, 154, 1405–1415.CrossRefGoogle ScholarPubMed
Vine, F. J. and Matthews, D. H. (1963). Magnetic anomalies over ocean ridges. Nature, 199, 947–949.CrossRefGoogle Scholar
Vine, F. J. and Smith, G. C. (1990). Structure and physical properties of the Troodos crustal section at ICRDG drillholes CY1, 1A and 4. In Proceedings of the Symposium on Ophiolites and Oceanic Lithosphere – TROODOS 87, ed. Malpas, J., Moores, E. M., Panayiotou, A. and Xenophontas, C.. Nicosia, Cyprus: Geological Survey Department, pp. 113–130.Google Scholar
Vogt, P. R. and deBoer, J. (1976). Morphology, magnetic anomalies, and basalt magnetization at the ends of the Galapagos high-amplitude zone. Earth and Planetary Science Letters, 33, 145–163.CrossRefGoogle Scholar
Von Damm, K. L. (1995). Controls on the chemistry and temporal variability of seafloor hydrothermal fluids. In Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions, Geophysical Monograph, 91, ed. Humphris, S. E., Zierenberg, R. A., Mullineaux, S. and Thomson, R. E.. Washington D.C.: American Geophysical Union, pp. 222–247.Google Scholar
Von Herzen, R. P., Cordery, M. J., Detrick, R. S. and Fang, C. (1989). Heat-flow and the thermal origin of hot spot swells – the Hawaiian swell revisited. Journal of Geophysical Research-Solid Earth and Planets, 94, (B10), 13783–13799.CrossRefGoogle Scholar
Von Herzen, R. P., Kirklin, J. and Becker, K. (1996). Geoelectrical measurements at the TAG hydrothermal mound. Geophysical Research Letters, 23, 3451–3454.CrossRefGoogle Scholar
Wang, X. and Cochran, J. R. (1993). Gravity anomalies, isostasy, and mantle flow at the East Pacific Rise crest. Journal of Geophysical Research, 98, 19505–19532.CrossRefGoogle Scholar
Watts, A. B. (1978). An analysis of isostasy in the worlds oceans: 1. Hawaiian–Emperor seamount chain. Journal of Geophysical Research, 83, 5989–6004.CrossRefGoogle Scholar
Watts, A. B., Bodine, J. H. and Steckler, M. S. (1980). Observations of flexure and the state of stress in the oceanic lithosphere. Journal of Geophysical Research, 85, 6369–6376.CrossRefGoogle Scholar
Wegener, A. (1912). Die Entstehung der Kontinente. Geologische Rundschau, 3, (4), 276–292.CrossRefGoogle Scholar
Wegener, A. (1966). The Origin of the Continents and Oceans (translated from 4th German edition by J. Biram). London: Methuen.Google Scholar
White, R. S. (1984). Atlantic oceanic crust: seismic structure of a slow-spreading ridge. In Ophiolites and Oceanic Lithosphere, Special Publication of the Geological Society, London, 13, ed. Gass, I. G., Lippard, S. J. and Shelton, A. W.. Oxford: Blackwell Scientific Publications, pp. 101–111.Google Scholar
White, R. S., Detrick, R. S., Sinha, M. C. and Cormier, M. H. (1984). Anomalous seismic crustal structure of oceanic fracture zones. Geophysical Journal of the Royal Astronomical Society, 79, 779–798.CrossRefGoogle Scholar
White, R., Mckenzie, D. and O’Nions, R. (1992). Oceanic crustal thickness from seismic measurements and rare-earth element inversions. Journal of Geophysical Research, 97, (B13), 19683–19715.CrossRefGoogle Scholar
White, S. M., Haymon, R. M., Fornari, D. J., Perfit, M. R. and Macdonald, K. C. (2002). Correlation between tectonic and volcanic segmentation of fast-spreading ridges: evidence from volcanic structures and lava flow morphology on the East Pacific Rise at 9°–10° N. Journal of Geophysical Research, B, 107, (8), EPM7, 1–20, .CrossRefGoogle Scholar
White, S. M., McClinton, J. T., Sinton, J. M. et al. (2010). Resolving volcanic eruptions: new fine-scale mapping by AUV Sentry of Galápagos Spreading Center 92° W and 95° W. In American Geophysical Union Fall Meeting. San Francisco: American Geophysical Union, pp. V52A-07.Google Scholar
Whitehead, J. A., Dick, H. J. B. and Schouten, H. (1984). A mechanism for magmatic accretion under spreading centres. Nature, 312, 146–148.CrossRefGoogle Scholar
Whitmarsh, R. B. (1970). An ocean bottom pop-up seismic recorder. Marine Geophysical Researches, 1, 91–98.CrossRefGoogle Scholar
Whitmarsh, R. B. and Laughton, A. S. (1975). The fault pattern of a slow-spreading ridge near a fracture zone. Nature, 258, 509–510.CrossRefGoogle Scholar
Wiens, D. A. and Stein, S. (1983). Age dependence of oceanic intraplate seismicity and implications for lithospheric evolution. Journal of Geophysical Research, 88, 6455–6468.CrossRefGoogle Scholar
Wilcock, W. S. D., Purdy, G. M. and Solomon, S. C. (1990). Microearthquake evidence for extension across the Kane transform fault. Journal of Geophysical Research, B, 95, (10), 15439–15462.CrossRefGoogle Scholar
Wilcock, W. S. D., Purdy, G. M., Solomon, S. C., Dubois, D. L. and Toomey, D. R. (1992). Microearthquakes on and near the East Pacific Rise, 9° N–10° N. Geophysical Research Letters, 19, (21), 2131–2134.CrossRefGoogle Scholar
Wilcock, W. S. D., Archer, S. D. and Purdy, G. M. (2002). Microearthquakes on the Endeavour segment of the Juan de Fuca Ridge. Journal of Geophysical Research-Solid Earth, 107, (B12), .CrossRefGoogle Scholar
Williams, A. B. and Rona, P. A. (1986). Two new caridean shrimps (bresiliidae) from a hydrothermal field on the Mid-Atlantic Ridge. Journal of Crustacean Biology, 6, (3), 446–462.CrossRefGoogle Scholar
Williams, C. M., Tivey, M. A., Schouten, H. and Fornari, D. J. (2008). Central Anomaly Magnetization High documentation of crustal accretion along the East Pacific Rise (9°55ʹ–9°25ʹ N). Geochemistry Geophysics Geosystems, 9, .CrossRefGoogle Scholar
Wilson, D. S., Teagle, D. A. H., Alt, J. C. et al. (2006). Drilling to gabbro in intact ocean crust. Science, 312, (5776), 1016–1020.CrossRefGoogle ScholarPubMed
Wilson, J. T. (1965). A new class of faults and their bearing on continental drift. Nature, 207, 343–347.CrossRefGoogle Scholar
Wing, C. G. (1969). MIT vibrating string surface-ship gravimeter. Journal of Geophysical Research, 74, (25), 5882–5894.CrossRefGoogle Scholar
Woese, C. R., Kandler, O. and Wheelis, M. L. (1990). Evolution towards a natural system of organisms: proposal for the domains Archaea, Bacteria, and Eucarya. Proceedings of the National Academy of Science USA, 87, 4576–4579.CrossRefGoogle ScholarPubMed
Wolery, T. J. and Sleep, N. H. (1976). Hydrothermal circulation and geochemical flux at mid-ocean ridges. Journal of Geology, 84, 249–275.CrossRefGoogle Scholar
Wolfe, C. J., Purdy, G. M., Toomey, D. R. and Solomon, S. C. (1995). Microearthquake characteristics and crustal velocity structure at 29° N on the Mid-Atlantic Ridge: the architecture of a slow-spreading segment. Journal of Geophysical Research, 100, (B12), 24449–24472.CrossRefGoogle Scholar
Woods Hole Oceanographic Institution (2011). Woods Hole Oceanographic Institution: A Brief History. Woods Hole, Massachusetts: Woods Hole Oceanographic Institution, 2 pp.
Worm, H.-U. (2001). Magnetic stability of oceanic gabbros from ODP Hole 735B. Earth and Planetary Science Letters, 193, 287–302.CrossRefGoogle Scholar
Worzel, J. L. (1959). Continuous gravity measurements on a surface ship with the Graf sea gravimeter. Journal of Geophysical Research, 64, (9), 1299–1315.CrossRefGoogle Scholar
Worzel, J. L. and Harrison, J. (1963). Gravity at sea. In The Sea, volume 3, ed. Hill, M. N.. New York: Interscience Publishers, pp. 134–174.Google Scholar
Wright, D. J. (1998). Formation and development of fissures at the East Pacific Rise: implications for faulting and magmatism at mid-ocean ridges. In Faulting and Magmatism at Mid-Ocean Ridges – Geophysical Monograph 106, ed. Buck, W. R., Delaney, P. T., Karson, J. A. and Lagabrielle, Y.. Washington, D.C.: American Geophysical Union, pp. 137–151.Google Scholar
Wylie, J. J., Helfrich, K. R., Dade, B., Lister, J. R. and Salzig, J. F. (1999). Flow localization in fissure eruptions. Bulletin of Volcanology, 60, 432–440.CrossRefGoogle Scholar
Xu, M., Canales, J. P., Tucholke, B. E. and DuBois, D. L. (2009). Heterogeneous seismic velocity structure of the upper lithosphere at Kane oceanic core complex, Mid-Atlantic Ridge. Geochemistry, Geophysics, Geosystems, 10, (10), .CrossRefGoogle Scholar
Yamazaki, T., Seama, N., Okino, K., Kitada, K. and Naka, J. (2003). Spreading process of the northern Mariana Trough: rifting–spreading transition at 22° N. Geochemistry Geophysics Geosystems, 4, .CrossRefGoogle Scholar
Yeo, I. (2012). Detailed studies of mid-ocean ridge volcanism – Mid-Atlantic Ridge 45° N and elsewhere. Ph.D. thesis, Durham University, Durham,
Yeo, I. A. and Searle, R. C. (2013). High resolution ROV mapping of a slow-spreading ridge: Mid-Atlantic Ridge 45° N. Geochemistry Geophysics Geosystems, 14, .CrossRef
Yeo, I., Searle, R. C., Achenbach, K., LeBas, T. and JC24 Shipboard Scientific Party (2012). Eruptive hummocks: building blocks of the upper ocean crust. Geology, 40, (1), 91–94.CrossRefGoogle Scholar
Yoerger, D. R., Bradley, A. M., Singh, H. et al. (1998). Multisensor mapping of the deep seafloor with the Autonomous Benthic Explorer. In Proceedings of the 2000 International Symposium on Underwater Technology. New York: IEEE, pp. 248–253.Google Scholar
Zelt, C. A. and Smith, R. B. (1992). Seismic travel-time inversion for 2-D crustal velocity structure. Geophysical Journal International, 108, 16–34.CrossRefGoogle 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.

  • References
  • Roger Searle, University of Durham
  • Book: Mid-Ocean Ridges
  • Online publication: 05 June 2014
  • Chapter DOI: https://doi.org/10.1017/CBO9781139084260.013
Available formats
×

Save book to Dropbox

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

  • References
  • Roger Searle, University of Durham
  • Book: Mid-Ocean Ridges
  • Online publication: 05 June 2014
  • Chapter DOI: https://doi.org/10.1017/CBO9781139084260.013
Available formats
×

Save book to Google Drive

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

  • References
  • Roger Searle, University of Durham
  • Book: Mid-Ocean Ridges
  • Online publication: 05 June 2014
  • Chapter DOI: https://doi.org/10.1017/CBO9781139084260.013
Available formats
×