Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-29T15:00:25.754Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 May 2015

Jeffrey A. Karson
Affiliation:
Syracuse University, New York
Deborah S. Kelley
Affiliation:
University of Washington
Daniel J. Fornari
Affiliation:
Woods Hole Oceanographic Institution, Massachusetts
Michael R. Perfit
Affiliation:
University of Florida
Timothy M. Shank
Affiliation:
Woods Hole Oceanographic Institution, Massachusetts
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
Discovering the Deep
A Photographic Atlas of the Seafloor and Ocean Crust
, pp. 355 - 400
Publisher: Cambridge University Press
Print publication year: 2015

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

Abbate, E. and Bortolotti, V. (1980). Apennine ophiolites: A peculiar oceanic crust. Ofioliti, 1, 59–96.Google Scholar
Abbot, D. H. and Hoffman, S. E. (1984). Archean plate tectonics revisited: 1. Heat flow, spreading rate, and the age of subducting oceanic lithosphere and their effects on the origin and evolution of continents. Tectonics, 3, 429–448.CrossRefGoogle Scholar
Acton, G. D., Stein, S. and Engeln, J. F. (1991). Block rotation and continental extension in Afar: A comparison to oceanic microplate systems. Tectonophysics, 10, 501–526.Google Scholar
Adamson, A. C. (1985). Basement lithostratigraphy, Deep Sea Drilling Project Hole 504B. In Anderson, R. N., Honnorez, J., Becker, K.et al. (eds.), Initial Reports of the Deep Sea Drilling Project. Washington, DC: US Government Printing Office, pp. 121–127.Google Scholar
Agar, S. M. and Klitgord, K. D. (1995). A mechanism for decoupling within the oceanic lithosphere revealed in the Troodos ophiolite. Nature, 374, 232–238.CrossRefGoogle Scholar
Aghaei, O., Nedimovic', M. R., Carton, H.et al. (2014). Crustal thickness and Moho character of the fast-spreading East Pacific Rise from 9°42′ N to 9°57′ N from poststack-migrated 3-D MCS data. Geochemistry, Geophysics, Geosystems, 15, 634–657, doi:10.1002/2013GC005069.CrossRefGoogle Scholar
Alabaster, T., Pearce, J. A. and Malpas, J. G. (1982). The volcanic stratigraphy and petrogenesis of the Oman ophiolite complex. Contributions to Mineralogy and Petrology, 81, 168–213.CrossRefGoogle Scholar
Alain, K., Callac, N., Guégan, M.et al. (2009). Nautilia abyssai sp. Nov., a thermophilic, chemolithoauthotrophic, sulfur-reducing bacterium isolated from an East Pacific Rise hydrothermal vent. International Journal of Systematic Evolutionary Microbiology, 60, 1182–1186.Google Scholar
Alain, K., Querellou, J., Lesongeur, F.et al. (2002). Caminibacter hydrogeniphilus gen. nov., sp nov., a novel thermophilic, hydrogen-oxidizing bacterium isolated from an East Pacific Rise hydrothermal vent. International Journal of Systematic and Evolutionary Microbiology, 52, 1317–1323.Google ScholarPubMed
Allan, J. F., Batiza, R., Perfit, M. R., Fornari, D. J. and Sack, R. O. (1989). Petrology of lavas from the Lamont Seamount Chain and adjacent East Pacific Rise, 10° N. Journal of Petrology, 30, 1245–1298.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, 1347–1354.CrossRefGoogle Scholar
Alt, J. C. (1995). Subseafloor processes in mid-ocean ridge hydrothermal systems. In Humphris, S., Zierenberg, R., Mullineau, L. and Thomson, R. (eds.), Physical, Chemical, Biological, and Geological Interactions within Seafloor Hydrothermal Systems. Geophysical Monograph 91, Washington, DC: American Geophysical Union, pp. 178–193.Google Scholar
Alt, J. C. (2004). Alteration of the upper oceanic crust: Mineralogy, chemistry and processes. In Davis, E. E. and Elderfield, H. (eds.), Hydrogeology of the Oceanic Lithosphere. Cambridge: Cambridge University Press, pp. 495–533.Google Scholar
Alt, J. C. and Teagle, D. A. H. (2000). Hydrothermal alteration and fluid fluxes in ophiolites and oceanic crust. In Dilek, Y., Moores, E., Elthon, D. and Nicolas, A. (eds.), Ophiolites and Oceanic Crust: New Insights from Field Studies and the Ocean Drilling Program. Geological Society of America Special Papers, 349. Boulder, CO: Geological Society of America, pp. 273–282.CrossRefGoogle Scholar
Alt, J. C., Honnorez, J., Laverne, C. and Emmermann, R. (1986). Hydrothermal alteration of a 1 km section through the upper oceanic crust, Deep Sea Drilling Project Hole 504B: Mineralogy, chemistry, and evolution of seawater-basalt interactions. Journal of Geophysical Research, 91, 10309–10335.CrossRefGoogle Scholar
Alt, J. C., Kinoshita, H., Stokking, L. B. and Michael, P. J. (eds.) 1993. Proceedings of the Ocean Drilling Program, Initial Reports, Leg 148, College Station, TX: Ocean Drilling Program.Google Scholar
Alt, J. C., Laverne, C., Coggon, R. M.et al. (2010). Subsurface structure of a submarine hydrothermal system in oceanic crust formed at the East Pacific Rise, ODP/IODP Site 1256. Geochemistry, Geophysics, Geosystems, 11, Q10010, doi:10.1029/2010GC003144.CrossRefGoogle Scholar
Alt, J., Laverne, C., Vanko, D. A.et al. (1996). Hydrothermal alteration of a section of upper oceanic crust in the eastern equatorial Pacific: A synthesis of results from Site 504 (DSDP Legs 69, 70, and 83, and ODP Legs 111, 137, 140, and 148). In Alt, J. C., Kinoshita, H., Stokking, L. B. and Michael, P. J. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 148. College Station, TX: Ocean Drilling Program, pp. 417–434.CrossRefGoogle Scholar
Anderson, E. M. (1972). The Dynamics of Faulting and Dyke Formation with Applications to Britain. New York: Hafner Publishing Co.Google Scholar
Anderson, R. N., Honnorez, J., Becker, K.et al. (1982). DSDP Hole 504B, the first reference section over 1 km through Layer 2 of the oceanic crust. Nature, 300, 589–594.CrossRefGoogle Scholar
Anderson, R. N., Honnorez, J., Becker, K.et al. (1985). Initial Reports of the Deep Sea Drilling Project, Leg 92. Washington, DC: US Government Printing Office.Google Scholar
Anonymous (1972). Penrose Field Conference on Ophiolites. Geotimes, 17, 24–25.Google Scholar
Arevalo, R. and McDonough, W. (2010). Chemical variations and regional diversity observed in MORB. Chemical Geology, 271, 70–85, doi:10.1016/j.chemgeo.2009.12.013.CrossRefGoogle Scholar
Arnulf, A. F., Harding, A. J., Kent, G. M., Singh, S. C. and Crawford, W. (2014). Constraints on the shallow velocity structure of the Lucky-Strike Volcano, Mid-Atlantic Ridge, from downward continued multi-channel streamer data. Journal of Geophysical Research, doi:10.1002/2013JB010500.CrossRef
Atwater, T. M. and Mudie, J. (1973). Detailed near-bottom geophysical study of the Gorda Rise. Journal of Geophysical Research, 78, 8665–8686.CrossRefGoogle Scholar
Aubouin, J. (1965). Geosynclines. Amsterdam: Elsevier.Google Scholar
Augustin, N., Lackschewitz, K. S., Kuhn, T. and Devey, C. W. (2008). Mineralogical and chemical mass changes in mafic and ultramafic rocks from the Logatchev hydrothermal field (MAR 15° N). Marine Geology, 256, 18–29.CrossRefGoogle Scholar
Aumento, F. and Loubat, H. (1971). The Mid-Atlantic Ridge near 45° N: Serpentinized ultramafic intrusions. Canadian Journal of Earth Sciences, 8, 631–663.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
Auzende, J. M., Cannat, M., Gente, P. et al. (1994). Observation of sections of oceanic crust and mantle cropping out on the southern wall of the Kane Fracture Zone (N. Atlantic). Terra Nova, 6, 143–148.CrossRefGoogle Scholar
Auzende, J. M., Olivet, J. L., Charvet, J. et al. (1978). Sampling and observation of oceanic mantle and crust on Gorringe Bank. Nature, 273, 45–59.CrossRefGoogle Scholar
Baba, K., Chave, A. D., Evans, R. L., Hirth, G. and Mackie, R. L. (2006). Mantle dynamics beneath the East Pacific Rise at 17 S: Insights from the Mantle Electromagnetic and Tomography (MELT) experiment. Journal of Geophysical Research, 111(B2), doi:10.1029/2004JB003598.CrossRefGoogle Scholar
Bach, W., Erzinger, J., Alt, J. C. and Teagle, D. A. H. (1996). Chemistry of the lower sheeted dike complex, Hole 504B (Leg 148): Influence of magmatic differentiation and hydrothermal alteration. In Alt, J. C., Kinoshita, H., Stokking, H. and Michael, P. J. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 148. College Station, TX: Ocean Drilling Program, pp. 39–56.Google Scholar
Bach, W., Garrido, C. J., Paulick, H. and Rosner, M. (2004). Seawater–peridotite interactions: first insights from ODP Leg 209, MAR 15° N. Geochemistry, Geophysics, Geosystems, 5, Q09F26, doi:10.1029/2004GC000744.CrossRefGoogle Scholar
Bachraty, C., Legendre, P. and Desbruyères, D. (2009). Biogeographic relationships among deep-sea hydrothermal vent faunas at global scale. Deep-Sea Research, Part I, 156, 1371–1378.CrossRefGoogle Scholar
Baker, E. T. (2009) Relationships between hydrothermal activity and axial magma chamber distribution, depth, and melt content. Geochemistry, Geophysics, Geosystems, 10(6), doi:10.1029/2009GC002424.CrossRefGoogle Scholar
Baker, E. T. and German, C. R. (2004). On the global distribution of hydrothermal vent fields. In German, C. R., Lin, J. and Parsons, L. M. (eds.), Mid-Ocean Ridges: Hydrothermal Interactions Between the Lithosphere and Oceans. Geophysical Monograph 148, Washington, DC: American Geophysical Union, pp. 245–266.Google Scholar
Baker, E. T. and Lupton, J. E. (1990). Changes in submarine hydrothermal 3He/heat ratios as an indicator of magmatic/tectonic activity. Nature, 346, 556–558.CrossRefGoogle Scholar
Baker, E. T., Cormier, M. -H., Langmuir, C. H. and Zavala, K. (2001). Hydrothermal plumes along segments of contrasting magmatic influence, 15°20′–18°30′N, East Pacific Rise: Influence of axial faulting. Geochemistry, Geophysics, Geosystems, 2(9), doi:10.1029/2000GC000165.CrossRefGoogle Scholar
Baker, E. T., Feely, R. A., Mottl, M. J. and Sansone, F. J. (1994). Hydrothermal plumes along the East Pacific Rise, 8° 40′ to 11° 50′ N: Plume distribution and relationship to the apparent magmatic budget. Earth and Planetary Science Letters, 128, 1–17.CrossRefGoogle Scholar
Baker, E. T., Fox, C. G. and Cowen, J. P. (1999). In situ observations of the onset of hydrothermal discharge during the 1998 submarine eruption of Axial Volcano, Juan de Fuca Ridge. Geophysical Research Letters, 26, 3445–3448.CrossRefGoogle Scholar
Baker, E. T., Haymon, R. M., Resing, J. A. et al. (2008). High-resolution surveys along the hotspot-affected Galápagos Spreading Center: 1. Distribution of hydrothermal activity. Geochemistry, Geophysics, Geosystems, 9, Q09003, doi:10.1029/2008GC002028.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), 1–14, doi: 10.1029/2001JB000651.CrossRefGoogle Scholar
Baker, E. T., Lupton, J. E., Resing, J. A. et al. (2011). Unique event plumes from a 2008 eruption on the Northeast Lau Spreading Center. Geochemistry, Geophysics, Geosystems, 12, Q0AF02, doi:10.1029/2011GC00372.CrossRefGoogle Scholar
Baker, E. T., Massoth, G. J. and Feely, R. A. (1987). Cataclysmic hydrothermal venting on the Juan de Fuca Ridge. Nature, 329, 149–151.CrossRefGoogle Scholar
Baker, E. T., Walker, S. L., Embley, R. W. and de Ronde, C. E. J. (2012). High-resolution hydrothermal mapping of Brothers Caldera, Kermadec Arc. Economic Geology and the Bulletin of the Society of Economic Geologists, 107(8), 1583–1593.CrossRefGoogle Scholar
Ballard, R. D. (1985). The Titanic: Lost and Found – Introduction. Oceanus, 28, 4.Google Scholar
Ballard, R. D. and Moore, J. G. (1977). Photographic Atlas of the Mid-Atlantic Ridge Rift Valley. New York: Springer-Verlag.CrossRefGoogle Scholar
Ballard, R. D. and van Andel, Tj. H. (1977a). Project FAMOUS: Operational techniques and American submersible operations. Geological Society of America Bulletin, 88, 495–506.2.0.CO;2>CrossRefGoogle Scholar
Ballard, R. D. and van Andel, Tj. H. (1977b). Morphology and tectonics of the inner rift valley at lat. 36°50′N on the Mid-Atlantic Ridge. Geological Society of America Bulletin, 88, 507–530.2.0.CO;2>CrossRefGoogle Scholar
Ballard, R. D., Francheteau, J., Juteau, T., Rangan, C. and Normark, W. (1981). EPR at 21°N: Volcanic, tectonic and hydrothermal processes of the central axis. Earth and Planetary Science Letters, 55, 1–10.CrossRefGoogle Scholar
Ballard, R. D., Holcomb, R. T. and van Andel, Tj. H. (1979). The Galapagos Rift at 86°W: 3. Sheet flows, collapse pits and lava lakes of the rift. Journal of Geophysical Research, 84, 4507–5422.CrossRefGoogle Scholar
Ballard, R. D., van Andel, Tj. H. and Holcomb, R. T. (1982). The Galapagos Rift at 86° W. 5. Variations in volcanism, structure, and hydrothermal activity along a 30 km segment of the rift valley. Journal of Geophysical Research, 87, 1149–1161.CrossRefGoogle Scholar
Banda, E., Ranero, C. R., Danobeitia, J. J. and Rivero, A. (1992). Seismic boundaries of the eastern central Atlantic Mesozoic crust from multichannel seismic data. Geological Society of America Bulletin, 104, 1,340–1,349.2.3.CO;2>CrossRefGoogle Scholar
Barager, W. R. A. (1954). Betts Pond Area, Burlington Peninsula, Newfoundland. Newfoundland Department of Mines, Open File Report. St. John's Newfoundland,Canada.Google Scholar
Barany, I. and Karson, J. A. (1989). Basaltic breccias of the Clipperton Fracture Zone: sedimentation and tectonics in a fast-slipping oceanic transform. Geological Society of America Bulletin, 101, 204–220.2.3.CO;2>CrossRefGoogle Scholar
Baranzangi, M. and Dorman, J. (1969). World seismicity maps compiled from ESSA, Coast and Geodetic Survey, epicenter data, 1961–1968. Bulletin of the Seismological Society of America, 59, 369–380.Google Scholar
Barker, A. K., Coogan, L. A. and Gillis, K. M. (2010). Insights into the behaviour of sulphur in mid-ocean ridge axial hydrothermal systems from the composition of the sheeted dyke comlex at Pito Deep. Chemical Geology, 275, 105–115.CrossRefGoogle Scholar
Barnes, I. and O'Neil, J. R. (1969). The relationship between fluids in some fresh alpine-type ultramafics and possible modern serpentinization, Western United States. Bulletin of the Geological Society of America, 80, 1947–1960.CrossRefGoogle Scholar
Baross, J. A. and Deming, J. W. (1995). Growth at high temperatures: Isolation and taxonomy, physiology, and ecology. In Karl, D. M. (ed.), The Microbiology of Deep-Sea Hydrothermal Vents. Boca Raton, FL: CRC Press, pp. 169–217.Google 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, 15, 327–345.Google Scholar
Barreyre, T., Escartín, J., Garcia, R.et al. (2012). Structure, temporal evolution, and heat flux estimates from the Lucky Strike deep-sea hydrothermal field derived from seafloor image mosaics. Geochemistry, Geophysics, Geosystems, 13, 4Q04007, doi:10.1029/2011GC003990.CrossRefGoogle Scholar
Barreyre, T., Soule, S. A. and Sohn, R. A. (2011). Dispersal of volcaniclasts during deep-sea eruptions: Settling velocities and entrainment in buoyant seawater plumes. Journal of Volcanology and Geothermal Research, 205(3), 84–93, doi 10.1016/j.jvolgeores.2011.05.006.CrossRefGoogle Scholar
Barriga, F., Fouquet, Y., Almeida, A.et al. (1998). Discovery of the Saldanha hydrothermal field on the FAMOUS segment of the MAR (36°30′N). Eos, Transactions of the American Geophysical Union, 79, 67.Google Scholar
Bascom, W. (1959). The Mohole. Scientific American, 200(4), 41–49.CrossRefGoogle Scholar
Batiza, R. and Vanko, D. A. (1986). Petrologic evolution of large failed rifts in the eastern Pacific: Petrology of volcanic and plutonic rocks from the Mathematician Ridge area and Guadalupe Trough. Journal of Petrology, 26, 564–602.CrossRefGoogle Scholar
Batiza, R., Niu, Y. L. and Zayac, W. C. (1990). Chemistry of seamounts near the East Pacific Rise: Implications for the geometry of subaxial mantle flow. Geology, 18, 1122–1125.2.3.CO;2>CrossRefGoogle Scholar
Batuev, B. N., Krotov, A. G., Makrov, V. F. et al. (1994). Massive sulfide deposits discovered and sampled at 14°45′N Mid-Atlantic Ridge. Bridge Newsletter, 6, 6–10.Google Scholar
Bazylinski, D. A., Farrington, J. W. and Jannasch, H. W. (1988). Hydrocarbons in surface sediments from a Guaymas Basin hydrothermal vent site. Organic Geochemistry, 12(6), 547–558.CrossRefGoogle Scholar
Beard, J. S. and Hopkinson, L. (2000). A fossil, serpentinization-related hydrothermal vent, Ocean Drilling Program Leg 173, Site 1068 (Iberia Abyssal Plain): Some aspects of mineral and fluid chemistry. Journal of Geophysical Research, 105(B7), 16,527–16,539.CrossRefGoogle Scholar
Becker, K., Sakai, H., Merrill, R. B.et al. (1988). Proceedings of the Ocean Drilling Program, Initial Reports, Leg 111. College Station, TX: Ocean Drilling Program.CrossRefGoogle Scholar
Behn, M. D., Buck, W. R. and Sacks, I. (2006). Topographic controls on dike injection in volcanic rift zones. Earth and Planetary Science Letters, 246(3–4), 188–196, doi:10.1016/j.epsl.2006.04.005.CrossRefGoogle Scholar
Behn, M. D., Sinton, J. M. and Detrick, R. S. (2004). Effect of the Galápagos hotspot on seafloor volcanism along the Galápagos Spreading Center (90.9°–97.6° W). Earth and Planetary Science Letters, 217(3), 331–347.CrossRefGoogle Scholar
Bender, J. F., Hodges, R. N. and Bence, A. E. (1978). Petrogenesis of basalts from the Project FAMOUS area: Experimental study from 0 to 15 kbars. Earth and Planetary Science Letters, 41, 277–302.CrossRefGoogle Scholar
Benthos, I. and Smith, P. F. (Compiler) (1984). Underwater Photography, Scientific and Engineering Applications. New York: Van Nostrand Reinhold.Google Scholar
Bergman, E. A. and Solomon, S. C. (1984). Source mechanisms of earthquakes near mid-ocean ridges from body waveform inversion: Implications for the early evolution of oceanic lithosphere. Journal of Geophysical Research, 89, 11,415–11,441.CrossRefGoogle Scholar
Bergmanis, E. C., Sinton, J. M. and Rubin, K. H. (2007). Eruptive history and magma reservoir dynamics on the East Pacific Rise at 17°30′ N. Geochemistry, Geophysics, Geosystems, 8(12), doi:10.1029/2007GC001742.CrossRefGoogle Scholar
Bezos, A., Escrig, S., Langmuir, C. H., Michael, P. J. and Asimow, P. D. (2009). Origins of chemical diversity of back-arc basin basalts: A segment-scale study of the Eastern Lau Spreading Center. Journal of Geophysical Research, 114(B6), doi:10.1029/2008JB005924.CrossRefGoogle Scholar
Bickle, M. J.et al. (2011). Illuminating Earth's Past, Present, and Future: The International Ocean Discovery Program, Science Plan for 2013–2023. www.iodp.org/Science-Plan-for-2013-2023/.
Bischoff, J. L. and Rosenbauer, R. J. (1989). Liquid-vapor relations for the system NaCl–H2O: Summary of the P-T-x surface from 300 °C to 500 °C. American Journal of Science, 289, 217–248.CrossRefGoogle Scholar
Blacic, T. M., Ito, G., Canales, J. P., Detrick, R. S. and Sinton, J. (2004). Constructing the crust along the Galapagos Spreading Center 91.3W-95.5W: Correlation of seismic layer 2A with axial magma lens and topographic characteristics. Journal of Geophysical Research, 109, B10310, 19, doi:10.1029/2004JB003066.CrossRefGoogle Scholar
Blackman, D. K., Cann, J. R., Janssen, B. and Smith, D. K. (1998). Origin of extensional core complexes: Evidence from the Mid-Atlantic Ridge at Atlantis fracture zone. Journal of Geophysical Research, 103, 21,315–21,334.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, 443–469.CrossRefGoogle Scholar
Blackman, D. K., Kelley, D. S., Karson, J. A. and ,Shipboard Scientific Party (2001). Seafloor mapping and sampling of the MAR 30°N Oceanic Core Complex-MARVEL (Mid-Atlantic Ridge Vents in Extending Lithosphere) 2000. InterRidge Newsletter, 10, 33–36.Google Scholar
Blackman, D. K., Kendall, J., Dawson, P. R., Wenk, H., Boyce, D. and Morgan, J. P. (1996). Teleseismic imaging of subaxial flow at mid-ocean ridges: Traveltime effects of anisotropic mineral texture in the mantle. Geophysical Journal International, 127(2), 415–426.CrossRefGoogle Scholar
Blondel, P. (2009). Handbook of Sidescan Sonar. Springer-Praxis.CrossRefGoogle Scholar
Bodinier, J. L. and Godard, M. (2014). Orogenic, Ophiolitic, and Abyssal Peridotites, Treatise on Geochemistry, 2nd edn. Elsevier, pp. 103–151.Google Scholar
Bohnenstiehl, D. R., Dziak, R. R., Tolstoy, M., Fox, C. G. and Fowler, M. (2004). Temporal and spatial history of the 1999–2000 Endeavour Segment seismic series, Juan de Fuca Ridge. Geochemistry, Geophysics, Geosystems, 5, Q09003, doi:10.1029/2004GC000735.CrossRefGoogle Scholar
Bonatti, E. (1976). Serpentinite protrusions in the oceanic crust. Earth and Planetary Science Letters, 32, 107–113.CrossRefGoogle Scholar
Bonatti, E. (1994). The Earth's mantle below the oceans. Scientific American, 270(3), 44–51.Google Scholar
Bonatti, E. and Hamlyn, P. R. (1978). Mantle uplifted blocks in the western Indian Ocean. Science, 201, 249–252.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, 93, 2967–2980.CrossRefGoogle Scholar
Bonatti, E. and Honnorez, J. (1976). Sections of the Earth's crust in the Equatorial Atlantic. Journal of Geophysical Research, 81, 4104–4116.CrossRefGoogle Scholar
Bonatti, E. and Seyler, M. (1987). Crustal underplating and evolution in the Red Sea rift: Uplifted gabbro/gneiss crustal complexes on Zabargad and Brothers Islands. Journal of Geophysical Research, 92, 12,803–12,821.CrossRefGoogle Scholar
Bonatti, E., Ligi, M., Brunelli, D.et al. (2003). Mantle thermal pulses below the Mid-Atlantic Ridge and temporal variation in the formation of oceanic lithosphere. Nature, 423, 499–505.CrossRefGoogle Scholar
Bosch, D., Jamais, M., Boudier, F.et al. (2004). Deep and high-temperature hydrothermal circulation in the Oman Ophiolite: Petrological and isotopic evidence. Journal of Petrology, 45(6), 1181–1208, doi:10.1093/petrology/egh010.CrossRefGoogle Scholar
Boschi, C., Bonatti, E., Ligi, M.et al. (2013). Serpentinization of mantle peridotites along an uplifted lithospheric section, Mid-Atlantic Ridge at 11° N. Lithos, 178, 3–23.CrossRefGoogle Scholar
Boschi, C., Dini, A., Früh-Green, G. L.et al. (2008). Isotopic and element exchange during serpentinization and metasomatism at the Atlantis Massif: Insights from B and Sr isotope data. Geochimica et Cosmochimica Acta, 72, 1801–1823.CrossRefGoogle Scholar
Boschi, C., Früh-Green, G. L., Delacour, A., Karson, J. A. and Kelley, D. S. (2006). Mass transfer and fluid flow during detachment faulting and development of an oceanic core complex. Geochemistry, Geophysics, Geosystems, 7(1), doi:10.1029/2005GC001074.CrossRefGoogle Scholar
Boudier, F. (1978). Microstructural study of three peridotite samples drilled at the western margin of the Mid-Atlantic Ridge. In Melson, W. G., Rabinowitz, P.et al. (eds.), Initial Reports of the Deep Sea Drilling Project, Leg 45. Washington, DC: US Government Printing Office, pp. 603–608.Google Scholar
Boudier, F. and Coleman, R. G. (1981). Cross section through the peridotite in the Samail Ophiolite southeastern Oman Mountains. Journal of Geophysical Research, 86, 2573–2592.CrossRefGoogle Scholar
Boudier, F. and Nicolas, A. (2011). Axial melt lenses at oceanic ridges: A case study in the Oman Ophiolite. Earth and Planetary Science Letters, 304, 313–325.CrossRefGoogle Scholar
Boudier, F., Nicolas, A. and Ildefonse, B. (1996). Magma chambers in the Oman Ophiolite: Fed from the top or bottom?Earth and Planetary Science Letters, 144, 239–250.CrossRefGoogle Scholar
Bougault, H., Charlou, J. -L., Fouquet, Y.et al. (1993). Fast and slow spreading ridges: Structure and hydrothermal activity, ultramafic topographic highs, and CH4 output. Journal of Geophysical Research, 98, 9643–9651.CrossRefGoogle Scholar
Bowen, A. D., Yoerger, D. R., Taylor, C.et al. (2009). The Nereus hybrid underwater robotic vehicle. Underwater Technology, 28, 79–89.CrossRefGoogle Scholar
Brazelton, W. J. and Baross, J. A. (2009). Abundant transposases encoded by the metagenome of a hydrothermal chimney biofilm. The ISME Journal, 3, 1420–1424.CrossRefGoogle ScholarPubMed
Brazelton, W. J., Ludwig, K. A., Sogin, M. L.et al. (2010). Archaea and bacteria with surprising microdiversity show shifts in dominance over 1,000-year time scales in hydrothermal chimneys. Proceedings of the National Academy of Sciences, USA, 107, 1612–1617, doi:1073/pnas.0905369107.CrossRefGoogle ScholarPubMed
Brazelton, W. J., Mehta, M. P., Kelley, D. S. and Baross, J. A. (2011). Physiological differentiation within a single-species biofilm fueled by serpentinization. mBio, 2(4):e00127–11, doi:10.1128/mBio.00127-11.CrossRefGoogle ScholarPubMed
Brazelton, W. J., Schrenk, M. O., Kelley, D. S. and Baross, J. A. (2006). Methane and sulfur metabolizing microbial communities dominate in the Lost City hydrothermal vent ecosystem. Applied Environmental Microbiology, 72, 6257–6270.CrossRefGoogle Scholar
Brocher, T. M., Karson, J. A. and Collins, J. A. (1985). Seismic stratigraphy of the oceanic Moho based on ophiolite models. Geology, 13, 62–65.2.0.CO;2>CrossRefGoogle Scholar
Brongniart, A. (1813). Essai de classification minéralogique des roches mélanges. Journal des Mines, XXXIV, 190–199.Google Scholar
Brown, J. R. and Karson, J. A. (1988). Variations in axial processes on the Mid-Atlantic Ridge: The median valley of the MARK area. Marine Geophysical Researches, 10, 109–138.CrossRefGoogle Scholar
Bryan, W. B. (1983). Systematics of modal phenocryst assemblages in submarine basalts: Petrologic implications. Contributions to Mineralogy and Petrology, 83, 62–74.CrossRefGoogle Scholar
Buck, W. R. (1991). Modes of continental lithospheric extension. Journal of Geophysical Research, 96, 20,161–20,178.CrossRefGoogle Scholar
Buck, W. R., Carbotte, S. M. and Mutter, C. Z. (1997). Controls on extrusion at mid-ocean ridges. Geology, 25, 935–938.2.3.CO;2>CrossRefGoogle Scholar
Buck, W. R., Einarsson, P. and Brandsdóttir, B. (2006). Tectonic stress and magma chamber size as controls on dike propagation: Constraints from the 1975–1984 Krafla rifting episode. Journal of Geophysical Research, 111, B12404, doi:10.1029/2005JB003879.CrossRefGoogle Scholar
Buck, W. R., Lavier, L. L. and Poliakov, A. (2005). Modes of faulting at mid-ocean ridges. Nature, 434, 719–723.CrossRefGoogle ScholarPubMed
Burchardt, S., Tanner, D. C., Troll, V. R., Krumbholz, M. and Gustafsson, I. E. (2011). Three-dimensional geometry of concentric intrusive sheet swarms in the Geitafell and the Dyrfjöll Volcanoes, eastern Iceland. Geochemistry, Geophysics, Geosystems, 12, Q0AB09, doi:10.1029/2011GC003527.CrossRefGoogle Scholar
Burchardt, S., Troll, V. R., Mathieu, L., Emeleus, H. C. and Donaldson, C. H. (2013). Ardnamurchan 3D cone-sheet architecture explained by a single elongate magma chamber. Science Reports, 3, doi:10.1038/srep02891.Google ScholarPubMed
Burke, K. and Kidd, W. S. F. (1978). Were Archean continental gradients much steeper than those of today?Nature, 272, 240–241.CrossRefGoogle Scholar
Bush, V. (1945). Science: The Endless Frontier. Washington, DC: US Government Printing Office.Google Scholar
Butterfield, D. A., Jonasson, I. R., Massoth, G. J.et al. (1997). Seafloor eruptions and evolution of hydrothermal fluid chemistry. Philosophical Transactions of the Royal Society of London, 355, 369–386.CrossRefGoogle Scholar
Butterfield, D. A., Massoth, G. J., McDuff, R. E., Lupton, J. E. and Lilley, M. D. (1990). The geochemistry of hydrothermal fluids from ASHES vent field, Axial Seamount, Juan de Fuca Ridge: Subseafloor boiling and subsequent fluid-rock interaction. Journal of Geophysical Research, 95, 12,895–12,922.CrossRefGoogle Scholar
Butterfield, D. A., McDuff, R. E., Mottl, M. J.et al. (1994). Gradients in the composition of hydrothermal fluids from the Endeavour segment vent field: Phase separation and brine loss. Journal of Geophysical Research, 99, 9561–9583.CrossRefGoogle Scholar
Butterfield, D. A., Roe, K. K., Lilley, M. D.et al. (2004). Mixing, reaction and microbial activity in the sub-seafloor revealed by temporal and spatial variation in diffuse flow vents at Axial Volcano. In Wilcock, W. S. D., DeLong, E. F., Kelley, D. S., Baross, J. A. and Cary, S. C. (eds.), Subseafloor Biosphere at Mid-Ocean Ridges. Geophysical Monograph 144, Washington, DC: American Geophysical Union, pp. 269–290.CrossRefGoogle Scholar
Campbell, I. H. (1978). Some problems with the cumulus theory. Lithos, 11, 311–323.CrossRefGoogle Scholar
Canales, J. P., Carton, H., Carbotte, S. M.et al. (2012a). Network of off-axis melt bodies at the East Pacific Rise. Nature Geoscience Letters, doi: 10.1038/NGEO1377.CrossRef
Canales, J. P., Carton, H., Mutter, J. C.et al. (2012b). Recent advances in multichannel seismic imaging for academic research in deep oceanic environments. Oceanography, 25(1), 113–115, doi.org/10.5670/oceanog.2012.09.CrossRefGoogle Scholar
Canales, J. P., Detrick, R. S., Carbotte, S. M.et al. (2005). Upper crustal structure and axial topography at intermediate spreading ridges: Seismic constraints from the southern Juan de Fuca Ridge. Journal of Geophysical Research, 110, B12104, doi:10.1029/2005JB003630.CrossRefGoogle Scholar
Canales, J. P., Detrick, R. S., Lin, J., Collins, J. A. and Toomey, D. R. (2000). Crustal and upper mantle seismic structure beneath the rift mountains and across a nontransform offset at the Mid-Atlantic Ridge (35°N). Journal of Geophysical Research, 105, 2699–2719.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°15′N–10°5′N) from wide-angle seismic refraction profiles. Geophysical Journal International, 152, 766–794, doi:10:1046/j.1365-246X.2003.10885.x.CrossRefGoogle Scholar
Canales, J. P., Nedimović, M. R., Kent, G. M., Carbotte, S. M. and Detrick, R. S. (2009). Seismic reflection images of near-axis melt sill within the lower crust at the Juan de Fuca Ridge. Nature, 460, 89–93, doi:10.1038/nature08095.CrossRefGoogle ScholarPubMed
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, 353–366.CrossRefGoogle Scholar
Canales, J. P., Sohn, R. A. and deMartin, B. J. (2007). Crustal structure of the Trans-Atlantic Geotraverse (TAG) segment (Mid-Atlantic Ridge, 26°10′N): Implications for the nature of hydrothermal circulation and detachment faulting at slow spreading ridges. Geochemistry, Geophysics, Geosystems, 8(8), doi:10.1029/2007GC001629.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, Q08002, doi:10.1029/2008GC002009.CrossRefGoogle Scholar
Cann, J. R. (1974). A model for oceanic crustal structure developed. Geophysical Journal of the Royal Astronomical Society, 39, 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(9), doi:10.1029/2005GC000954.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., Langseth, M. G., Honnorez, J.et al. (1983). Initial Reports of the Deep Sea Drilling Project, Leg 69. Washington, DC: US Government Printing Office.CrossRefGoogle Scholar
Cannat, M. (1993). Emplacement of mantle rocks in the sea floor at mid-ocean ridges. Journal of Geophysical Research, 98, 4163–4172.CrossRefGoogle Scholar
Cannat, M. (1996). How thick is the magmatic crust at slow spreading oceanic ridges?Journal of Geophysical Research, 101, 2847–2857.CrossRefGoogle Scholar
Cannat, M., Briais, A., Deplus, C.et al. (1999). Mid-Atlantic Ridge–Azores hotspot interactions: Along-axis migration of a hotspot-derived magmatic pulse 14 to 4 myrs ago. Earth and Planetary Science Letters, 173, 257–269, doi:10.1016/S0012-821X(99)00234-4.CrossRefGoogle Scholar
Cannat, M., Juteau, T. and Berger, E. (1990). Petrostructural analysis of the Leg 109 serpentinized peridotites. In Detrick, R. S., Bryan, W. B., Honnorez, J. and Juteau, T. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 109. College Station, TX: Ocean Drilling Program, pp. 47–57.Google Scholar
Cannat, M., Karson, J. A., Miller, D. J.et al. (1995a). Proceedings of the Ocean Drilling Program, Initial Reports, Leg 153. College Station, TX: Ocean Drilling Program.CrossRefGoogle Scholar
Cannat, M., Karson, J. A., Miller, D. J.et al. (1995b). Probing the plutonic foundation of the Mid-Atlantic Ridge. Eos, Transactions of the American Geophysical Union, 76, 129–133.Google Scholar
Cannat, M., Lagabrielle, Y., Coutures, N. D.et al. (1997). Ultramafic and gabbroic exposures at the Mid-Atlantic Ridge: Geological mapping in the 15°N region. Tectonophysics, 279, 193–214.CrossRefGoogle Scholar
Cannat, M., Mével, C., Maia, M.et al. (1995c). Thin crust ultramafic exposures and rugged faulting patterns at the Mid-Atlantic Ridge (22°–24°N). Geology, 23, 49–52.2.3.CO;2>CrossRefGoogle Scholar
Cannat, M., Sauter, D., Mendel, V.et al. (2006). Modes of seafloor generation at a melt-poor ultraslow-spreading ridge. Geology, 34, 605–608, doi:10.1130/G22486A.1.CrossRefGoogle Scholar
Canon-Tapia, E. and Aubourg, C. (2004). Anisotropy of magnetic susceptibility of lava flows and dykes: A historical account. In Martin-Hernandez, F. (ed.), Magnetic Fabric: Methods and Applications. Boulder, CO: Geological Society of America, pp. 205–225.Google Scholar
Carbotte, S. M., Canales, J. P., Nedimović, M. R., Carton, H. and Mutter, J. C. (2012). Recent seismic studies at the East Pacific Rise 8°20′–10°10′N and Endeavour Segment: Insights into mid-ocean ridge hydrothermal and magmatic processes. Oceanography, 25(1), 100–112, doi:org/10.5670/oceanog.2012.08.CrossRefGoogle Scholar
Carbotte, S. M., Marjanovic, M., Carton, H.et al. (2013). Fine-scale segmentation of the crustal magma reservoir beneath the East Pacific Rise. Nature Geoscience, 6, 866–870, doi:10.1038/NGEO1933.CrossRefGoogle Scholar
Carbotte, S. M., Mutter, J. C. and Xu, L. (1997). Contribution of volcanism and tectonism to axial and flank morphology of the southern East Pacific Rise 17°10′–17°40′S from a study of layer 2A geometry. Journal of Geophysical Research, 102, 10,165–10,184.CrossRefGoogle Scholar
Carbotte, S. M., Solomon, A. and Ponce-Correa, G. (2000). Evaluation of morphological indicators of magma supply and segmentation from a seismic reflection study of the East Pacific Rise 15°30′–17°N. Journal of Geophysical Research, 105, 2737–2759.CrossRefGoogle Scholar
Caress, D. W., Clague, D. A., Paduan, J. B.et al. (2012). Repeat bathymetric surveys at 1-metre resolution of lava flows erupted at Axial Seamount in April 2011. Nature Geoscience, 5(7), 483–488, doi:10.1038/NGEO1496.CrossRefGoogle Scholar
Cary, S. C., Fisher, C. R. and Felbeck, H. (1998a). Mussel growth supported by methane as sole carbon and energy source. Science, 240, 78–80.CrossRefGoogle Scholar
Cary, S. C., Shank, T. and Stein, J. (1998b). Worms bask in extreme temperatures. Nature, 391, 545–546.CrossRefGoogle Scholar
Casey, J. F. (1997). Comparison of major- and trace-element geochemistry of abyssal peridotites and mafic plutonic rocks with basalts from the MARK region of the Mid-Atlantic Ridge. In Karson, J. A., Cannat, M., Miller, D. J. and Elthon, D. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 153. College Station, TX: Ocean Drilling Program, pp. 181–242.Google Scholar
Casey, J. F. and Karson, J. A. (1981). Magma chamber profiles from the Bay of Islands ophiolite complex. Nature, 292, 295–301.CrossRefGoogle Scholar
Casey, J. F., Dewey, J. F., Fox, P. J., Karson, J. A. and Rosencrantz, E. (1981). Heterogeneous nature of the oceanic crust and upper mantle: A perspective from the Bay of Islands Ophiolite. In Emiliani, C. (ed.), The Oceanic Lithosphere. New York: Wiley, pp. 305–338.Google Scholar
Casey, J. F., Karson, J. A., Elthon, D. L., Rosencrantz, E. and Titus, M. (1983). Reconstruction of the geometry of accretion during formation of the Bay of Islands Ophiolite Complex. Tectonics, 2, 509–528.CrossRefGoogle Scholar
Cashman, K. V., Mangan, M. T. and Newman, S. (1994). Surface degassing and modifications to vesicle size distributions in active basalt flows. Journal of Volcanology and Geothermal Research, 61, 45–68.CrossRefGoogle Scholar
Cashman, K. V., Thornber, C. and Kauahikaua, J. P. (1999). Cooling and crystallization of lava in open channels, and the transition of pahoehoe lava to ‘A'a. Bulletin Volcanologique, 61, 306–323.CrossRefGoogle Scholar
,CAYTROUGH (1979). Geological and geophysical investigation of the Mid-Cayman Rise spreading center: Initial results and observations. In Talwani, M., Harrison, C. G. and Hayes, D. E. (eds.), Deep Drilling Results in the Atlantic Ocean: Ocean Crust. Maurice Ewing Series, 2, Washington, DC: American Geophysical Union, pp. 66–93.CrossRefGoogle Scholar
Cavanaugh, C. M., Levering, P. R., Maki, J. S., Mitchell, R. and Lidstrom, M. E. (1987). Symbiosis of methylotrophic bacteria and deep-sea mussels. Nature, 325, 346–347.CrossRefGoogle Scholar
Ceuleneer, G. and Cannat, M. (1997). High-temperature ductile deformation of Site 920 peridotites. In Karson, J. A., Cannat, M., Miller, D. J. and Elthon, D. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 153. College Station, TX: Ocean Drilling Program, pp. 23–34.Google Scholar
Ceuleneer, G., Nicolas, A. and Boudier, F. (1988). Mantle flow patterns at an oceanic spreading centre: The Oman peridotites record. Tectonophysics, 151, 1–26.CrossRefGoogle Scholar
Chadwick, J., Perfit, M., Ridley, I.et al. (2005). Magmatic effects of the Cobb hot spot on the Juan de Fuca Ridge. Journal of Geophysical Research Letters, 110, B03101, doi:10.1029/2003JB002767.Google Scholar
Chadwick, W. W. (1997). Is the “axial summit caldera” on the East Pacific Rise really a “graben”?Ridge Events, 8(1), 8–11.Google Scholar
Chadwick, W. W. and Embley, R. W. (1998). Graben formation associated with recent dike intrusions and volcanic eruptions on the mid-ocean ridge. Journal of Geophysical Research, 103, 9807–9826.CrossRefGoogle Scholar
Chadwick, W. W., Cashman, K. V., Embley, R. W.et al. (2008). Direct video and hydrophone observations of submarine explosive eruptions at NW Rota-1 Volcano, Mariana Arc. Journal of Geophysical Research, 113, B08S10, doi:10.1029/2007JB005215.CrossRefGoogle Scholar
Chadwick, W. W., Clague, D. A., Embley, R. W.et al. (2013). The 1998 eruption of Axial Seamount: New insights on submarine lava flow emplacement from high-resolution mapping. Geochemistry, Geophysics, Geosystems, 14(10), 3939–3968, doi:10.1002/ggge.20202.CrossRefGoogle Scholar
Chadwick, W. W., Embley, R. W. and Fox, C. G. (1991). Evidence for volcanic eruption on the southern Juan de Fuca Ridge between 1981 and 1987. Nature, 350, 416–418.CrossRefGoogle Scholar
Chadwick, W. W., Embley, R. W. and Fox, C. G. (2003). Quantitative constraints on the growth of submarine lava pillars from a monitoring instrument that was caught in a lava flow. Journal of Geophysical Research, 108(B11), 2534, doi:2510.1029/2003JB002422.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 II, 45, 2547–2569.CrossRefGoogle Scholar
Chadwick, W. W., Gregg, T. K. P. and Embley, R. W. (1999). Submarine lineated sheet flows: A unique lava morphology formed on subsiding lava ponds. Bulletin of Volcanology, 61, 194–206.CrossRefGoogle Scholar
Chadwick, W. W., Nooner, S. L., Butterfield, D. A. and Lilley, M. D. (2012). Seafloor deformation and forecasts of the April 2011 eruption at Axial Seamount. Nature Geoscience, 5, 474–477, doi: 10.1038/NGEO1464.CrossRefGoogle Scholar
Chadwick, W. W., Scheirer, D. S., Embley, R. W. and Johnson, H. P. (2001). High-resolution bathymetric surveys using scanning sonars: Lava flow morphology, hydrothermal vent and geologic structure at recent eruption sites on the Juan de Fuca Ridge. Journal of Geophysical Research, 106, 16,075–16,100.CrossRefGoogle Scholar
Charlou, J. L., Dmitried, L., Bougault, H. and Needham, H. D. (1988). Hydrothermal CH4 between 12°N and 15°N over the Mid-Atlantic Ridge. Deep-Sea Research Part A, 35, 121–131.CrossRefGoogle Scholar
Charlou, J. L., Donval, J. P., Douville, E.et al. (2000). Compared geochemical signature and evolution of Menez Gwen (37°50′N) and Lucky Strike (37°17′N) hydrothermal fields, south of the Azores Triple Junction on the Mid-Atlantic Ridge. Chemical Geology, 171, 49–75.CrossRefGoogle Scholar
Charlou, J. L., Donval, J. P., Fouquet, Y., Jean-Baptiste, P. and Holm, N. (2002). Geochemistry of high H2 and CH4 vent fluids issuing from ultramafic rocks at the Rainbow hydrothermal field (36°14′N, MAR). Chemical Geology, 191, 345–359.CrossRefGoogle Scholar
Charlou, J. L., Donval, J. P., Jean-Baptiste, P. and Dapoigny, A. (1996) Gases and helium isotopes in high temperature solutions sampled before and after ODP Leg 158 drilling a TAG hydrothermal field (26°N, MAR). Geophysical Research Letters, 23, 3491–3494.CrossRefGoogle Scholar
Charlou, J. L., Donval, J. P., Konn, C., Ondreas, H. and Fouquet, Y. (2010). High production and fluxes of H2 and CH4 and evidence of abiotic hydrocarbon synthesis by serpentinization in ultramafic-hosted hydrothermal systems on the Mid-Atlantic Ridge. In Rona, P., Devey, C. and Murton, B. (eds.), Diversity of Hydrothermal Systems on Slow Spreading Ocean Ridges. Geophysical Monograph 188, Washington, DC: American Geophysical Union, pp. 265–296.CrossRefGoogle Scholar
Cheadle, M. J. and Grimes, C. (2010). Structural geology: To fault or not to fault. Nature Geoscience, 3, 454–456, doi:10.1038/ngeo910.CrossRefGoogle Scholar
Chevaldonné, P., Desbruyères, D. and Childress, J. J. (1992). Some like it hot… and some even hotter. Nature, 359, 593–594.CrossRefGoogle Scholar
Childress, J. J. and Fisher, C. R. (1992). The biology of hydrothermal vent animals: Physiology, biochemistry, and autotrophic symbiosis. Oceanography and Marine Biology, 30, 337–441.Google Scholar
Christensen, N. I. (1978). Ophiolites, seismic velocities and oceanic crustal structure. Tectonophysics, 47, 131–157.CrossRefGoogle Scholar
Christensen, N. I. (1984). The magnitude, symmetry and origin of upper mantle anisotropy based on fabric analyses of ultramafic tectonites. Geophysical Journal of the Royal Astronomical Society, 76, 89.CrossRefGoogle Scholar
Christensen, N. I. and Mooney, W. D. (1995). Seismic velocity structure and composition of the continental crust: A global view. Journal of Geophysical Research, 100, 9761–9788.CrossRefGoogle Scholar
Christensen, N. I. and Salisbury, M. H. (1975). Structure and composition of the lower oceanic crust. Reviews of Geophysics and Space Physics, 13, 57–86.CrossRefGoogle Scholar
Christensen, N. I. and Salisbury, M. H. (1982). Lateral heterogeneity in the seismic structure of the oceanic crust inferred from velocity studies in the Bay of Islands ophiolites, Newfoundland. Geophysical Journal of the Royal Astronomical Society, 68, 675–688.CrossRefGoogle Scholar
Christeson, G. L., Karson, J. A. and McIntosh, K. D. (2010). Mapping of seismic layer 2A/2B boundary above the sheeted dike unit at intermediate-spreading crust exposed near the Blanco Transform. Geochemistry, Geophysics, Geosystems, 11, Q03015, doi: 10.1029/2009GC002864.CrossRefGoogle Scholar
Christeson, G. L., Kent, G. M., Purdy, G. M. and Detrick, R. S. (1996). Extrusive thickness variability at the East Pacific Rise 9°–10°N: Constraints from seismic techniques. Journal of Geophysical Research, 101, 2859–2783.CrossRefGoogle Scholar
Christeson, G. L., McIntosh, K. D. and Karson, J. A. (2007). Inconsistent correlation of seismic layer 2a and lava layer thickness in oceanic crust. Nature, 445, 418–421, doi:10.1038/nature05517.CrossRefGoogle ScholarPubMed
Christeson, G. L., Morgan, J. V. and Warner, M. R. (2012). Shallow oceanic crust: Full waveform tomographic images of the seismic layer 2A/2B boundary. Journal of Geophysical Research, 117, B05101, doi:10.1029/2011JB008972.CrossRefGoogle Scholar
Christeson, G. L., Purdy, G. M. and Fryer, G. J. (1992). Structure of young upper crust at the East Pacific Rise near 9°30′N. Geophysical Research Letters, 19, 1045–1048.CrossRefGoogle Scholar
Christie, D. M. and Sinton, J. M. (1981). Evolution of abyssal lavas along propagating segments of the Galápagos spreading center. Earth and Planetary Science Letters, 56, 321–335.CrossRefGoogle Scholar
Christie, D. M. and Sinton, J. M. (1986). Mantle processes in the generation of MORB from the 95°W Galápagos propagating rift area. Contributions to Mineralogy and Petrology, 94, 274–288.CrossRefGoogle Scholar
Church, W. R. and Stevens, R. K. (1971). Early Paleozoic complexes of the Newfoundland Appalachians as mantle–oceanic crust sequences. Journal of Geophysical Research, 76, 1460–1466.CrossRefGoogle Scholar
Chutas, L. A. (2007). Structures in Upper Oceanic Crust: Perspectives from Pito Deep and Iceland. Masters thesis, Duke University.Google Scholar
Ciccarelli, F. D., Doerks, T., von Mering, C.et al. (2006). Toward automatic reconstruction of a highly resolved tree of life. Science, 311, 1283–1286.CrossRefGoogle ScholarPubMed
Cipriani, A., Bonatti, E., Brunelli, D. and Ligi, M. (2009). 26 million years of mantle upwelling below a segment of the Mid Atlantic Ridge: The Vema Lithospheric Section revisited. Earth and Planetary Science Letters, 285, 87–95.CrossRefGoogle Scholar
Clague, D. A. and Paduan, J. B. (2009). Submarine basaltic volcanism. In Cousens, B. and Piercey, S. J. (eds.), Submarine Volcanism and Mineralization: Modern Through Ancient. Geological Association of Canada, pp. 41–60, doi:10.1029/2011GC003791.Google Scholar
Clague, D. A., Caress, D. W., Thomas, H.et al. (2008). Abundance and distribution of hydrothermal chimneys and mounds on the Endeavour Ridge determined by 1-m resolution AUV multibeam mapping surveys. Eos, Transactions of the American Geophysical Union, 89(53), V41B-2079.Google Scholar
Clague, D. A., Davis, A. S. and Dixon, J. E. (2003). Submarine strombolian eruptions along the Gorda mid-ocean ridge. In White, J. D. L., Smellie, J. L. and Clague, D. A. (eds.), Explosive Subaequeous Volcanism. Geophysical Monograph 140, Washington, DC: American Geophysical Union, pp. 1–128.Google Scholar
Clague, D. A., Dreyer, B. M., Paduan, J. B.et al. (2013). Geologic history of the summit of Axial Seamount, Juan de Fuca Ridge. Geochemistry, Geophysics, Geosystems, 14, 4403–4443, doi:10.1002/ggge.20240.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, 171–188.CrossRefGoogle Scholar
Clague, D. A., Paduan, J. B., Caress, D. W., Thomas, H. and Chadwick, W. W. (2011). Volcanic morphology of West Mata Volcano, NE Lau Basin, based on high-resolution bathymetry and depth changes. Geochemistry, Geophysics, Geosystems, 12(11), doi:10.1029/2011GC003791.CrossRefGoogle Scholar
Clark, A. C. (1991) In Mackay, A. L. (ed.), A Dictionary of Scientific Quotations. Boca Raton, FL: CRC Press.Google Scholar
Clark, M. R., Rowden, A. A., Schlacher, T.et al. (2010). The ecology of seamounts: Structure, function and human impacts. Annual Review of Marine Science, 2, 253–278.CrossRefGoogle ScholarPubMed
Claypool, G. E. and Kvenvolden, K. A. (1983). Methane and other hydrocarbon gases in marine sediments. Annual Review of Earth and Planetary Sciences, 11, 299–327.CrossRefGoogle Scholar
Cochran, J. R., Goff, J. A., Malinverno, A.et al. (1993). Morphology of a ‘superfast’ mid-ocean ridge crest and flanks: The East Pacific Rise, 7–9° S. Marine Geophysical Researches, 15(1), 65–75.CrossRefGoogle Scholar
Coffin, M. F., Duncan, R. A., Eldholm, O.et al. (2006). Large igneous provinces and scientific ocean drilling: Status quo and a look ahead. Oceanography, 19(4), 150–160, doi.org/10.5670/oceanog.2006.13.CrossRefGoogle Scholar
Colaco, A., Blandin, J., Cannat, M.et al. (2011). MoMAR-D: A technological challenge to monitor the dynamics of the Lucky Strike vent ecosystem. ICES Journal of Marine Science, 68(2), 416–424, doi:10.1093/icesjms/fsq075.CrossRefGoogle Scholar
Colaco, A., Dehairs, F. and Desbruyères, D. (2002). Nutritional relations of deep-sea hydrothermal fields at the Mid-Atlantic Ridge: A stable isotope approach. Deep-Sea Research I, 49, 395–412.CrossRefGoogle Scholar
Coleman, R. G. (1981). Tectonic setting for ophiolite obduction in Oman. Journal of Geophysical Research, 86, 2497–2508.CrossRefGoogle Scholar
Collins, J. A., Brocher, T. M. and Karson, J. A. (1986). Two-dimensional seismic reflection modeling of the oceanic crust/mantle transition in the Bay of Islands ophiolite complex. Journal of Geophysical Research, 91, 12,520–12,538.CrossRefGoogle Scholar
Collins, J. A., Tucholke, B. E. and Canales, J. P. (2001). Structure of Mid-Atlantic Ridge megamullions from seismic refraction experiments and multichannel seismic reflection profiling. Eos, Transactions of the American Geophysical Union, 82, F1100.Google Scholar
Collins, P. C., Croot, P., Carlsson, C.et al. (2013a). A primer for the environmental impact assessment of mining at seafloor massive sulfide deposits. Marine Policy, 42, 198–209.CrossRefGoogle Scholar
Collins, P. C., Kennedy, R., Copley, J.et al. (2013b). VentBase: Developing a consensus among stakeholders in the deep-sea regarding environmental impact assessment for deep-sea mining. Marine Policy, 42, 334–336.CrossRefGoogle Scholar
Colman, A., Sinton, J. M., White, S. M.et al. (2012). Effects of variable magma supply on mid-ocean ridge eruptions: Constraints from mapped lava flows along the Galápagos Spreading Center. Geochemistry, Geophysics, Geosystems, 13, Q08014, doi:10.1029/2012GC004163.CrossRefGoogle Scholar
Connelly, D. P., Germa, C. R., Asada, M.et al. (2007). Hydrothermal activity on the ultra-slow spreading southern Knipovich Ridge. Geochemistry, Geophysics, Geosystems, 8, Q08013, doi:10.1029/2007GC001652.CrossRefGoogle Scholar
Constantin, M. (1999). Gabbroic intrusions and magmatic metasomatism in harzburgites from the Garrett transform fault: Implications for the nature of the mantle–crust transition at fast-spreading ridges. Contributions to Mineralogy and Petrology, 136, 111–130.CrossRefGoogle Scholar
Constantin, M., Hékinian, R., Bideau, D. and Hébert, R. (1996). Construction of the oceanic lithosphere by magmatic intrusions: Petrologic evidence from plutonic rocks formed along the fast-spreading East Pacific Rise. Geology, 24(8), 731–734.2.3.CO;2>CrossRefGoogle Scholar
Coogan, L. A. (2007). The lower oceanic crust. In Turekian, K. and Holland, H. D. (eds.), Treatise on Geochemistry. Oxford: Pergamon, pp. 1–45, doi:10.1016/B978-008043751-4/00230-3.Google Scholar
Coogan, L. A. (2014). The lower oceanic crust. In Holland, H. and Turekian, K. (eds.), Reference Module in Earth Systems and Environmental Sciences, from Treatise on Geochemistry (Second Edition), Vol. 4: The Crust. Oxford: Elsevier, pp. 497–541, doi.org/10.1016/B978-0-08-095975-7.00316-8.Google Scholar
Coogan, L. A., Gillis, K. M., MacLeod, C. J., Thompson, G. M. and Hékinian, R. (2002). Petrology and geochemistry of the lower oceanic crust formed at the East Pacific Rise and exposed at Hess Deep: A synthesis and new results. Geochemistry, Geophysics, Geosystems, 3(11), 8604, doi:10.1029/2001/GC000230.CrossRefGoogle Scholar
Coogan, L. A., Howard, K. A., Gillis, K. M.et al. (2006). Chemical and thermal constraints on focused fluid flow in the lower oceanic crust. American Journal of Science, 306, 389–427, doi:10.2475/06.2006.01.CrossRefGoogle Scholar
Coogan, L. A., Kempton, P. D., Saunders, A. D. and Norry, M. J. (2000). Melt aggregation within the crust beneath the Mid-Atlantic Ridge: Evidence from plagioclase and clinopyroxene major and trace element compositions. Earth and Planetary Science Letters, 176, 245–257.CrossRefGoogle Scholar
Coogan, L. A., Mitchell, N. C. and O'Hara, M. J. (2003). Roof assimilation at fast-spreading ridges: An investigation combining geophysical, geochemical and field evidence. Journal of Geophysical Research, 108(B1), 2002, doi:10.1029/2001JB001171.CrossRefGoogle Scholar
Coogan, L. A., Wilson, R. N., Gillis, K. M. and MacLeod, C. J. (2001). Near-solidous evolution of oceanic gabbros: Insights from amphibole geochemistry. Geochimica et Cosmochimica Acta, 65, 4339–4357.CrossRefGoogle Scholar
Cordier, C., Caroff, M., Juteau, T.et al. (2007). Bulk-rock geochemistry and plagioclase zoning in lavas exposed along the northern flank of the Western Blanco Depression (Northeast Pacific): Insight into open-system magma chamber processes. Lithos, 99, 289–311.CrossRefGoogle Scholar
Corliss, J. B., Dymond, J., Gordon, L. I.et al. (1979). Submarine hydrothermal springs on the Galápagos Rift. Science, 203, 1073–1083.CrossRefGoogle ScholarPubMed
Cousteau, J.-Y. (1953). The Silent World. New York: Harper & Brothers.Google Scholar
Cowen, J. P., Betram, M. A., Baker, E. T.et al. (1998). Geomicrobial transformations of manganese in Gorda Ridge event plumes. Deep-Sea Research II, 45, 2713–2738.CrossRefGoogle Scholar
Cowen, J. P., Bertram, M. A., Wakeham, S.et al. (2001). Biogeochemical cycling in hydrothermal plumes. In Halbach, P. E., Tunnicliffe, V. and Hein, J. R. (eds.), Energy and Mass Transfer in Marine Hydrothermal Systems. Berlin: Dahlem University Press, pp. 303–316.Google Scholar
Cowen, J. P., Fornari, D. J., Shank, T. M.et al. (2007). Volcanic eruptions at East Pacific Rise crest near 9°50′N. Eos, Transactions of the American Geophysical Union, 88(7), 81–82.CrossRefGoogle Scholar
Cox, A. (1969). Geomagnetic reversals. Science, 163, 000–000.CrossRefGoogle ScholarPubMed
Cravo, A., Foster, P., Almeida, C., Bebianno, M. J. and Company, R. (2008). Metal concentrations in the shell of Bathymodiolus azoricus from contrasting hydrothermal vent fields on the Mid-Atlantic Ridge. Marine Environmental Research, 65, 338–348.CrossRefGoogle ScholarPubMed
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° 48′ N, using seafloor compliance measurements. Journal of Geophysical Research, 104(B2), 2923–2939.CrossRefGoogle Scholar
Crone, T. J., Wilcock, W. S., Barclay, A. H. and Parsons, J. D. (2006). The sound generated by mid-ocean ridge black smoker hydrothermal vents. PLoS ONE, 1(1), doi:10.1371/journal.one.0000133.CrossRefGoogle ScholarPubMed
Croon, M. B., Cande, S. C. and Stock, J. M. (2010). Abyssal hill deflections at Pacific–Antarctic ridge–transform intersections. Geochemistry, Geophysics, Geosystems, 11(11), Q11004, doi:10.1029/2010GC003236.CrossRefGoogle Scholar
Curewitz, D. and Karson, J. A. (1998). Geological consequences of dike intrusion at mid-ocean ridge spreading centers. In Buck, W. R., Delaney, P. T., Karson, J. A. and Lagabrielle, Y. (eds.), Faulting and Magmatism at Mid-Ocean Ridges. Washington, DC: American Geophysical Union, pp. 117–136.Google Scholar
Curtis, A. C., Wheat, C. G., Fryer, P. and Moyer, C. L. (2013). Mariana forearc serpentinite mud volcanoes harbor novel communities of extremophilic Archaea. Geomicrobiology Journal, 30, 430–441.CrossRefGoogle Scholar
Cuvelier, D., Sarrazin, J., Colaco, A.et al. (2009). Distribution and spatial variation of hydrothermal faunal assemblages at Lucky Strike (Mid-Atlantic Ridge) revealed by high resolution video image analysis. Deep-Sea Research, 1(156), 2026–2040.CrossRefGoogle Scholar
,CYAMEX Scientific Team (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
Daly, R. A. (1926). Our Mobile Earth. New York: Charles Scribner's Sons.Google Scholar
Davies, A., Roberts, J. M. and Hall-Spencer, J. M. (2007). Preserving deep-sea natural heritage: emerging issues in offshore conservation and management. Biological Conservation, 138, 299–312.CrossRefGoogle Scholar
Davies, H. L. (1971). Peridotite-Gabbro-Basalt Complex in Eastern Papua: An Overthrust Plate of Oceanic Mantle and Crust. Bureau of Mineral Resources, 128, Geology and Geophysics, Australia.Google Scholar
Davis, A. S. and Clague, D. A. (1987). Geochemistry, mineralogy, and petrogenesis of basalt from the Gorda Ridge. Journal of Geophysical Research, 92, 10,467–10,483.CrossRefGoogle Scholar
Davis, A. S. and Clague, D. A. (1990). Gabbroic xenoliths from the northern Gorda Ridge: Implications for magma chamber processes under slow-spreading centers. Journal of Geophysical Research, 95, 10,885–10,906.CrossRefGoogle Scholar
Davis, E. E. and Elderfield, H. (2004). Hydrogeology of the Oceanic Crust. Cambridge: Cambridge University Press.Google Scholar
Davis, E. E. and Fisher, A. T. (1994). On the nature and consequences of hydrothermal circulation in the Middle Valley sedimented rift: Inferences from geophysical and geochemical observations, Leg 139. In Mottl, M. J., Davis, E. E., Fisher, A. T. and Slack, J. F. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 139. College Station, TX: Ocean Drilling Program, pp. 695–717.Google Scholar
Davis, E. E., Wang, K., Thompson, R. E., Becker, K. and Cassidy, J. F. (2001). An episode of seafloor spreading and associated plate deformation inferred from crustal fluid pressure transients. Journal of Geophysical Research, 106, 21,953–21,963.CrossRefGoogle Scholar
Deardorff, N. D., Cashman, K. V. and Chadwick, W. W. (2011). Observations of eruptive plumes and pyroclastic deposits from submarine explosive eruptions at NW Rota-1, Mariana Arc. Journal of Volcanology and Geothermal Research, 202(1–2), 47–59, doi:10.1016/j.jvolgeores.2011.01.003.CrossRefGoogle Scholar
Dechaine, E. G., Bates, A. E., Shank, T. M. and Cavanaugh, C. M. (2006). Off-axis symbiosis found: Characterization and biogeography of bacterial symbionts of Bathymodiolus mussels from Lost City hydrothermal vents. Environmental Microbiology, 8(11), 1902–1912.CrossRefGoogle ScholarPubMed
Delacour, A., Früh-Green, G. L., Bernasconi, S. M., Schaeffer, P. H. and Kelley, D. S. (2008a). Carbon geochemistry of serpentinites at the Lost City Hydrothermal System (30°N, MAR). Geochimica et Cosmochimica Acta, 72, 3681–3702.CrossRefGoogle Scholar
Delacour, A., Früh-Green, G. L., Frank, M., Gutjahr, M., and Kelley, D. S. (2008b). Sr- and Nd-isotope geochemistry of the Atlantis Massif (30°N, MAR): Implications for fluid fluxes and lithospheric heterogeneity. Chemical Geology, 254, 19–35.CrossRefGoogle Scholar
Delaney, J. R. and Barga, R. S. (2009). A 2020 vision for ocean science. In Hey, T., Tansley, S. and Tolle, K. (eds.), The Fourth Paradigm Data-Intensive Scientific Discovery. Microsoft Research, pp. 27–38.Google Scholar
Delaney, J. R. and Kelley, D. S. (2014). Understanding the planetary life support system: Next generation science in the ocean basins. In Favali, P., de Santis, A. and Beranzoli, L. (eds.), Seafloor Observatories: A New Vision of the Earth from the Abyss. Springer Praxis.Google Scholar
Delaney, J. R., Kelley, D. S., Lilley, M. D. (1998). The quantum event of oceanic crustal accretion: Impacts of diking at mid-ocean ridges. Science, 281, 222–230.CrossRefGoogle ScholarPubMed
Delaney, J. R., Kelley, D. S., Lilley, M. D. (1997). The Endeavour hydrothermal system I: Cellular circulation above an active cracking front yields large sulfide structures, “fresh” vent water, and hyperthermophile archaea. RIDGE Events, 8, 11–19.Google Scholar
Delaney, J. R., Kelley, D. S. and Mathez, E. A. (2001). “Edifice Rex” sulfide recovery project: Analyses of submarine hydrothermal, microbial habitat. Eos, Transactions of the American Geophysical Union, 82(6), 67, 72–73.CrossRefGoogle Scholar
Delaney, J. R., McDuff, R. E. and Lupton, J. E. (1984). Hydrothermal fluid temperatures of 400 °C on the Endeavour Segment, northern Juan de Fuca Ridge. Eos, Transactions of the American Geophysical Union, 65(45), 973.Google Scholar
Delaney, J. R., Mogk, D. W. and Mottl, M. J. (1987a). Quartz-cemented breccias from the Mid-Atlantic Ridge: Samples of a high-salinity hydrothermal upflow zone. Journal of Geophysical Research, 92, 9175–9192.CrossRefGoogle Scholar
Delaney, J. R., Robigou, V., McDuff, R. E. and Tivey, M. K. (1992). Geology of a vigorous hydrothermal system on the Endeavour Segment, Juan de Fuca Ridge. Geophysical Research Letters, 97, 19,663–19,682.CrossRefGoogle Scholar
Delaney, J. R., Spiess, F. N., Solomon, S. C.et al. (1987b). Scientific rationale for establishing long-term ocean bottom observatory/laboratory systems. In Teleki, P.et al. (eds.), Marine Minerals: Resource Assessment Strategies. Springer, pp. 389–411.CrossRefGoogle Scholar
DeLong, S. E., Dewey, J. F. and Fox, P. J. (1977). Displacement history of oceanic fracture zones. Geology, 5, 199–202.2.0.CO;2>CrossRefGoogle Scholar
DeLong, S. E., Dewey, J. F. and Fox, P. J. (1979). Topographic and geologic evolution of fracture zones. Journal of the Geological Society of London, 136, 303–310.CrossRefGoogle Scholar
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, 711–714, doi:10.1130/G23718A.1.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
Desbruyères, D., Alayse-Danet, A.-M., Ohta, S. and the Scientific Parties of Biolau and Starmer Cruises (1994). Deep-sea hydrothermal communities in Southwestern Pacific back-arc basins (the North Fiji and Lau Basins): Composition, micro-distribution and food web. Marine Geology, 116, 227–242.CrossRefGoogle Scholar
Desbruyères, D., Almeida, A., Biscoito, M.et al. (2000). A review of the distribution of hydrothermal vent communities along the northern Mid-Atlantic Ridge: Dispersal vs. environmental controls. Hydrobiologia, 440, 201–216.CrossRefGoogle Scholar
Desbruyères, D., Biscoito, M.Caprais, J. C.et al. (2001). Variations in deep-sea hydrothermal vent communities on the Mid-Atlantic Ridge near the Azores plateau. Deep-Sea Research Part I, 48, 1325–1346.CrossRefGoogle Scholar
Desbruyères, D., Chevaldonné, P, Alayse, A. M.et al. (1998). Biology and ecology of the ‘Pompeii worm’ (Alvinella pompejana Desbruyères and Laubier), a normal dweller of an extreme deep-sea environment: A synthesis of current knowledge and recent developments. Deep-Sea Research, 45, 383–422.CrossRefGoogle Scholar
Desbruyères, D., Segonzac, M. and Bright, M. (2006). Handbook of Deep-Sea Hydrothermal Vent Fauna: Denisia, Vol. 18, Linz-Dornach, Austria: Oberösterreichischen Landesmuseen.Google Scholar
Detrick, R. S., Buhl, P., Vera, E.et al. (1987). Multichannel seismic imaging of a crustal magma chamber along the East Pacific Rise. Nature, 326, 35–41.CrossRefGoogle Scholar
Detrick, R. S., Collins, J., Stephen, R. 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., Harding, A. J.Kent, G. M.et al. (1993a). Seismic structure of the southern East Pacific Rise. Science, 259, 499–503.CrossRefGoogle ScholarPubMed
Detrick, R. S., White, R. S. and Purdy, G. M. (1993b). Crustal structure of North Atlantic fracture zones. Reviews of Geophysics, 31, 439–458.CrossRefGoogle Scholar
Devey, C. W., Lackschewitz, K. S. and Baker, E. (2005). Hydrothermal and volcanic activity found on the Southern Mid-Atlantic Ridge. Eos, Transactions of the American Geophysical Union, 86, 209–216.CrossRefGoogle Scholar
Dewey, J. F. and Bird, J. M. (1971). Origin and emplacement of the ophiolite suite: Appalachian ophiolites in Newfoundland. Journal of Geophysical Research, 76, 3179–3206.CrossRefGoogle Scholar
Dewey, J. F. and Kidd, W. S. F. (1977). Geometry of plate accretion. Geological Society of America Bulletin, 88, 960–968.2.0.CO;2>CrossRefGoogle Scholar
Dick, H. J. B. (1989). Abyssal peridotites, very slow spreading ridges and ocean ridge magmatism. In Saunders, A. D. and Norry, M. J. (eds.), Magmatism in the Ocean Basins. Special Publication 42, London: Geological Society of London, pp. 77–105.Google Scholar
Dick, H. J. B. and Bullen, T. (1984). Chromian spinel as a petrogenetic indicator in abyssal and alpine-type peridotites and spacially associated lavas. Contributions to Mineralogy and Petrology, 86, 54–76.CrossRefGoogle Scholar
Dick, H. J. B. and Natland, J. H. (1996). Late-stage melt evolution and transport in the shallow mantle beneath the East Pacific Rise. In Mével, C., Gillis, K. M., Allan, J. F. and Meyer, P. S. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 147. College Station, TX: Ocean Drilling Program, pp. 439–538.Google Scholar
Dick, H. J. B., Lin, J. and Schouten, H. (2003). An ultraslow-spreading class of ocean ridge. Nature, 426, 405–412.CrossRefGoogle ScholarPubMed
Dick, H. J. B., Lissenberg, C. J. and Warren, J. M. (2010). Mantle melting, melt transport, and delivery beneath a slow-spreading ridge: The paleo-MAR from 23°15′N to 23°45′N. Journal of Petrology, 51(1–2), 425–467, doi:10.1093/petrology/egp088.CrossRefGoogle Scholar
Dick, H. J. B., Meyer, P., Bloomer, S.et al. (1991). Lithostratigraphic evolution of an in situ section of oceanic layer 3. In von Herzen, R. P. and Robinson, P. T. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 118. College Station, TX: Ocean Drilling Program, pp. 439–515.Google Scholar
Dick, H. J. B., Natland, J. H., Alt, J. C. et al. (2000). A long in situ section of the lower oceanic crust: Results of ODP Leg 176 drilling at the Southwest Indian Ridge. Earth and Planetary Science Letters, 179, 31–51.CrossRefGoogle Scholar
Dick, H. J. B., Schouten, H., Meyer, P. S. et al. (1991). Tectonic evolution of the Atlantis II fracture zone. In Proceedings of the Ocean Drilling Program, Scientific Results, Leg 118. pp. 359–398.Google Scholar
Dick, H. J. B., Tivey, M. A. and Tucholke, B. E. (2008). Plutonic foundation of a slow-spreading ridge segment: The oceanic core complex at Kane megamullion, 23°30′N, 42°20′W. Geochemistry, Geophysics, Geosystems, 9, Q05014, doi:10.1029/2007GC001645.CrossRefGoogle Scholar
Dietz, R. S. (1961). Continental and ocean basin evolution by spreading of the sea floor. Nature, 190, 854–857.CrossRefGoogle Scholar
Dilek, Y., Harper, G. D., Pezard, P. A. and Tartarotti, P. (1996). Structure of the sheeted dike complex in Hole 504B (Leg 148). In Alt, J. C., Kinoshita, H., Stokking, L. B. and Michael, P. J. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 148. College Station, TX: Ocean Drilling Program, pp. 229–243.Google Scholar
Dilek, Y., Moores, E. M., Delaloye, M. and Karson, J. A. (1991). Amagmatic extension and tectonic denudation in the Kizildag ophiolite, southern Turkey: Implications for the evolution of Neotethyan oceanic crust. In Peters, T., Nicolas, A. and Coleman, R. G. (eds.), Ophiolite Genesis and Evolution of Oceanic Lithosphere. Dordrecht: Kluwer, pp. 485–500.CrossRefGoogle Scholar
Dixon, J. E., Clague, D. A. and Eissen, J. P. (1986). Gabbroic xenoliths and host ferrobasalt from the southern Juan de Fuca Ridge. Journal of Geophysical Research, 91(B3), 3795–3820.CrossRefGoogle Scholar
Dixon, J. E., Stolper, E. M. and Holloway, J. R. (1995). An experimental study of water and carbon dioxide solubilities in mid-ocean ridge basaltic liquids. Part I: Calibration and solubility models. Journal of Petrology, 36(6), 1607–1631.Google Scholar
Douville, J. L., Charlou, E. H., Oelkers, P. et al. (2002). The rainbow vent fluids (36°14′N, MAR): The influence of ultramafic rocks and phase separation on trace metal content in Mid-Atlantic Ridge hydrothermal fluids. Chemical Geology, 184, 37–48.CrossRefGoogle Scholar
Drake, C. L. and Girdler, R. W. (1964). A geophysical study of the Red Sea. Geophysical Journal of the Royal Astronomical Society, 8(5), 473–495.CrossRefGoogle Scholar
Dunkelman, T. J., Karson, J. A. and Rosendahl, B. R. (1988). Structural style of the Turkana Rift, Kenya. Geology, 16, 258–261.2.3.CO;2>CrossRefGoogle Scholar
Dunn, R. A. and Toomey, D. R. (1997). Seismological evidence for three-dimensional melt migration beneath the East Pacific Rise. Nature, 388, 259–262.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°30′N. Journal of Geophysical Research, 105(B10), 23,537–22,555.CrossRefGoogle Scholar
Dziak, R. P. and Fox, C. G. (1990). Long-term seismicity and ground deformation at Axial Volcano, Juan de Fuca Ridge. Geophysical Research Letters, 26, 3641–3644.CrossRefGoogle Scholar
Dziak, R. P. and Fox, C. G. (1999). The January 1998 earthquake swarm at Axial Volcano, Juan de Fuca Ridge: Hydroacoustic evidence of seafloor volcanic activity. Geophysical Research Letters, 26, doi:10.1029/1999GL002332.CrossRefGoogle Scholar
Dziak, R. P., Baker, E. T., Shaw, A. M. et al. (2012a). Flux measurements of explosive degassing using a yearlong hydroacoustic record at an erupting submarine volcano. Geochemistry, Geophysics, Geosystems, 13, Q0AF07, doi:10.1029/2012GC004211.CrossRefGoogle Scholar
Dziak, R. P., Bohnenstiehl, D. R. and Smith, D. K. (2012b). Hydroacoustic monitoring of oceanic spreading centers: Past, present, and future. Oceanography, 25(1), 116–127.CrossRefGoogle Scholar
Dziak, R. P., Bohnenstiehl, D. R., Cowen, J. P. et al. (2007). Rapid dike emplacement leads to eruptions and hydrothermal plume release during seafloor spreading events. Geology, 35, 579–582.CrossRefGoogle Scholar
Dziak, R. P., Bohnenstiehl, D. R., Matsumoto, H. et al. (2009). January 2006 seafloor-spreading event at 9°50′N, East Pacific Rise: Ridge dike intrusion and transform fault interactions from regional hydroacoustic data. Geochemistry, Geophysics, Geosystems, 10, Q06T06, doi:10.1029/2009GC002388.CrossRefGoogle Scholar
Dziak, R. P., Fox, C. G. and Schreiner, A. E. (1995). The June-July 1993 seismo-acoustic event at CoAxial segment, Juan de Fuca Ridge: Evidence for a lateral dike injection. Geophysical Research Letters, 22, 135–138.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, 280–293, doi:10.5670/oceanog.2011.79.CrossRefGoogle Scholar
Dziak, R. P., Haxel, J. H., Bohnenstiehl, D. R. et al. (2012c). Seismic precursors and magma ascent before the April 2011 eruption at Axial Seamount. Nature Geosciences, 5, 478–482.CrossRefGoogle Scholar
Eason, D. E. and Sinton, J. M. (2009). Lava shields and fissure eruptions of the Western Volcanic Zone, Iceland: Evidence for magma chambers and crustal interaction. Journal of Volcanology and Geothermal Research, 186, 331–348.CrossRefGoogle Scholar
Edmond, J. M., Measures, B., Mangum, B. et al. (1979a). On the formation of metal-rich deposits at ridge crests. Earth and Planetary Science Letters, 46, 19–30.CrossRefGoogle Scholar
Edmond, J. M., Measures, C., McDuff, R. E. et al. (1979b). Ridge crest hydrothermal activity and the balances of the major and minor elements in the ocean: The Galápagos data. Earth and Planetary Science Letters, 46, 1–18.CrossRefGoogle Scholar
Edmonds, H. N., Michael, P. J., Baker, E. T. et al. (2003). Discovery of abundant hydrothermal venting on the ultraslow-spreading Gakkel ridge in the Arctic Ocean. Nature, 421, 252–256.CrossRefGoogle ScholarPubMed
Edwards, M. H., Fornari, D. J., Malinverno, A., Ryan, W. B. F. and Madsen, J. (1991). The regional tectonic fabric of the East Pacific Rise from 12°50′N to 15°10′N. Journal of Geophysical Research, 96, 7995–8017.CrossRefGoogle Scholar
Ehlmann, B. L., Mustard, J. F. and Murchie, S. L. (2010). Geologic setting of serpentine deposits on Mars. Geophysical Research Letters, 37, L06201, doi:10.1029/2010GL042596.CrossRefGoogle Scholar
Einarsson, P. and Ericksson, J. (1982). Earthquake fractures in the districts land and Rangarvellir in the South Iceland Seismic Zone. Jokull, 32, 113–119.Google Scholar
Einarsson, P. and Brandsdóttir, B. (1980). Seismological evidence for lateral magma intrusion during the July 1978 deflation of the Krafla volcano in NE-Iceland. Journal of Geophysics, 47, 160–165.Google Scholar
Einsele, G. (1982). Mechanism of sill intrusion into soft sediment and expulsion of pore water. In Curray, J. R. and Moore, D. G. (eds.), Initial Reports of the Deep Sea Drilling Project, Leg 64, Part 2. Washington, DC: US Government Printing Office, pp. 1169–1176.Google Scholar
Einsele, G. (1985). Basaltic sill-sediment complexes in young spreading centers: genesis and significance. Geology, 13, 249–252.2.0.CO;2>CrossRefGoogle Scholar
Einsele, G., Gieskes, J., Curray, J.et al. (1980). Intrusion of basaltic sills into highly porous sediments, and resulting hydrothermal activity. Nature, 283, 441–445.CrossRefGoogle Scholar
Elthon, D. (1987). Petrology of gabbroic rocks from the Mid-Cayman Rise spreading center. Journal of Geophysical Research, 92, 658–682.CrossRefGoogle Scholar
Elthon, D. and Stern, C. (1978). Metamorphic petrology of the Sarmiento Ophiolite Complex, Chile. Geology, 6, 464–468.2.0.CO;2>CrossRefGoogle Scholar
Elthon, D., Ross, D. K. and Meen, J. K. (1995). Compositional variations of basaltic glasses from the Mid-Cayman Rise spreading center. Journal of Geophysical Research, 100(B7), 12,497–12,512, doi:10.1029/94JB02777.CrossRefGoogle Scholar
Elthon, D., Stewart, M. and Ross, D. K. (1992). Compositional trends of minerals in oceanic cumulates. Journal of Geophysical Research, 97, 15,189–15,200.CrossRefGoogle Scholar
Embley, R. W. and Lupton, J. E. (2004). Diking, event plumes, and the subsurface biosphere at mid-ocean ridges. In Wilcock, W. D., Kelley, D. S., Baross, J. A., DeLong, E. and Cary, C. (eds.), The Subseafloor Biosphere at Mid-Ocean Ridges. Geophysical Monograph 144, Washington, DC: American Geophysical Union, pp. 75–97.CrossRefGoogle Scholar
Embley, R. W., Chadwick, W. W., Baker, E. T.et al. (2006). Long-term eruptive activity at a submarine arc volcano. Nature, 441, 494–497.CrossRefGoogle Scholar
Embley, R. W., Chadwick, W. W., Clague, D. A. and Stakes, D. (1999). 1998 eruption of Axial Volcano: Multibeam anomalies and sea-floor observations. Geophysical Research Letters, 26, 3425–3428.CrossRefGoogle Scholar
Embley, R. W., Chadwick, W. W., Jonasson, I. R., Butterfield, D. A. and Baker, E. T. (1995). Initial results of the response to the 1993 CoAxial event: Relationships between hydrothermal and volcanic processes. Geophysical Research Letters, 22, 143–146, doi:.org/10.1029/94GL02281.CrossRefGoogle Scholar
Embley, R. W., Chadwick, W. W., Perfit, M. R. and Baker, E. T. (1991). Geology of the northern Cleft segment, Juan de Fuca Ridge: Recent eruptions, seafloor spreading, and the formation of megaplumes. Geology, 19, 771–775.2.3.CO;2>CrossRefGoogle Scholar
Embley, R. W., Chadwick, W. W., Perfit, M. R., Smith, M. C. and Delaney, J. F. (2000). Recent eruptions on the CoAxial segment of the Juan de Fuca Ridge: Implications for mid-ocean ridge accretion processes. Journal of Geophysical Research, 105, 16,501–16,525.CrossRefGoogle Scholar
Embley, R. W., Jonasson, I. R., Perfit, M. R.et al. (1988). Submersible investigation of an extinct hydrothermal system on the Galápagos Ridge: Sulfide mounds, stockwork zone, and differentiated lavas. Canadian Mineralogist, 26, 517–539.Google Scholar
Embley, R., Chadwick, W. W., Baker, E. T.et al. (2006). Long-term eruptive activity at a submarine arc volcano. Nature, 441, 494–497.CrossRefGoogle Scholar
Engel, A. E. J., Engel, C. G. and Havens, R. G. (1965). Chemical characteristics of oceanic basalts and the upper mantle. Geological Society of America Bulletin, 76(7), 719–734.CrossRefGoogle Scholar
Engel, C. G. and Fisher, R. L. (1975). Granitic to ultramafic rock complexes of the Indian Ocean ridge system western Indian Ocean. Geological Society of America Bulletin, 86, 1553–1578.2.0.CO;2>CrossRefGoogle Scholar
Engels, J. L., Edwards, M. H., Fornari, D. J., Perfit, M. R. and Cann, J. R. (2003). A new model for submarine volcanic collapse formation. Geochemistry, Geophysics, Geosystems, 4(9), doi:10.1029/2002GC000483.CrossRefGoogle Scholar
Escartín, J. and Canales, J. P. (2010). Chapman Conference on Detachments in Oceanic Lithosphere: Deformation, Magmatism, Fluid Flow and Ecosystems (Conference report). Eos, Transactions of the American Geophysical Union, 92, 31, doi:10.1029/2011EO040003.CrossRefGoogle Scholar
Escartín, J., Cannat, M., Pouliquen, G., Rabain, A. and Lin, J. (2001a). Crustal thickness of V-shaped ridges south of the Azores: Interaction of the Mid-Atlantic Ridge (36°–39° N) and the Azores hot spot. Journal of Geophysical Research, 106, 21,719–21,735, doi:10.1029/2001JB000224.CrossRefGoogle Scholar
Escartín, J., García, R., Delaunoy, J.et al. (2008a). Globally aligned photomosaic of the Lucky Strike hydrothermal vent field (Mid-Atlantic Ridge, 37°18.5′N): Release of georeferenced data, mosaic construction, and viewing software. Geochemistry, Geophysics, Geosystems, 9(12), doi:10.1029/2008GC002204.CrossRefGoogle Scholar
Escartín, 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, 181–190.CrossRefGoogle Scholar
Escartín, J., Hirth, G. and Evans, B. (2001b). Strength of slightly serpentinized peridotites: Implications for the tectonics of oceanic lithosphere. Geology, 29, 1023–1026.2.0.CO;2>CrossRefGoogle Scholar
Escartín, J., Mevel, C., MacLeod, C. J. and McCaig, A. M. (2003). Constraints on deformation conditions and the origin of oceanic detachments: The Mid-Atlantic Ridge core complex at 15°45′N. Geochemistry, Geophysics, Geosystems, 4(8), doi:1029/2002GC000472.CrossRefGoogle Scholar
Escartín, J., Smith, D. K., Cann, J. R.et al. (2008b). Central role of detachment faults in accretion of slow-spreading oceanic lithosphere. Nature, 455, 790–794, doi:10.1038/nature07333.CrossRefGoogle ScholarPubMed
Escartín, 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° 25′–9° 58′ N. Geochemistry, Geophysics, Geosystems, 8(6), doi:10.1029/2006GC001399.CrossRefGoogle Scholar
Ewing, M. A. and Engel, L. (1962). Seismic shooting at sea. Scientific American, 5, 116–126.CrossRefGoogle Scholar
Ewing, M. A., Vine, A. C. and Worzel, J. L. (1946). Photography of the ocean bottom. Journal of the Optical Society of America, 36, 307.CrossRefGoogle Scholar
Ferrini, V. L., Fornari, D. J., Shank, T. M.et al. (2007). Sub-meter bathymetric mapping of volcanic and hydrothermal features on the East Pacific Rise Crest at 9° 50′ N. Geochemistry, Geophysics, Geosystems, 8, Q01006, doi:10.1029/2006GC001333.CrossRefGoogle Scholar
Ferrini, V. L., Tivey, M. K., Carbotte, S. M., Martinez, F. and Roman, C. N. (2008). Variable morphologic expression of volcanic, tectonic, and hydrothermal processes at six hydrothermal vent fields in the Lau Back-arc Basin. Geochemistry, Geophysics, Geosystems, 9, Q07022, doi:10.1029/2008GC002047.CrossRefGoogle Scholar
Fisher, A. T. (1998). Permeability within basaltic oceanic crust. Reviews in Geophysics, 36, 143–182.CrossRefGoogle Scholar
Fisher, A. T. and Wheat, G. (2010). Seamounts as conduits for massive fluid, heat, and solute fluxes on ridge flanks. Oceanography Magazine, 23(1), 74–87.CrossRefGoogle Scholar
Fisher, A. T., Davis, E. E. and Becker, K. (2008). Borehole-to-borehole hydrologic response across 2.4 km in the upper oceanic crust: Implications for crustal-scale properties. Journal of Geophysical Research, 113, doi:10.1029/2007JB005447.CrossRefGoogle Scholar
Fisher, A. T., Davis, E. E., Hutnak, M.et al. (2003a). Hydrothermal recharge and discharge across 50 km guided by seamounts on a young ridge flank. Nature, 421, 618–621.CrossRefGoogle ScholarPubMed
Fisher, A. T., Stein, C. A., Harris, R. N.et al. (2003b). Abrupt thermal transition reveals hydrothermal boundary and role of seamounts within the Cocos Plate. Geophysical Research Letters, 30(11), 1550, doi:1510.1029/2002GL016766.CrossRefGoogle Scholar
Fisher, C. R., Childress, J. J., Arp, A. J.et al. (1988). Microhabitat variation in the hydrothermal vent mussel, Bathymodiolus thermophiles, at the Rose Garden vent on the Galápagos Rift. Deep-Sea Research Part A. Oceanographic Research Papers, 35, 1769–1791, doi.org/10.1016/0198-0149(88) 90049-0.CrossRefGoogle Scholar
Fontaine, F. and Wilcock, W. S. D. (2006). Dynamics and storage of brine in mid-ocean ridge hydrothermal systems. Journal of Geophysical Research, 111, 1–16, B06102, doi:10.1029/2005JB003866.CrossRefGoogle Scholar
Fornari, D. J. (2003). A new deep-sea towed digital camera and multi-rock coring system. Eos, Transactions of the American Geophysical Union, 84, 69 and 73.CrossRefGoogle Scholar
Fornari, D. J. (2004). Realizing the dreams of da Vinci and Verne. Oceanus, 42, 1–5.Google Scholar
Fornari, D. J., Beaulieu, S. E., Holden, J. F., Mullineaux, L. S. and Tolstoy, M. (2012b). Introduction to the Special Issue: From RIDGE to Ridge 2000. Oceanography, 25(1), 12–17.CrossRefGoogle Scholar
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. (1998a). 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, 103, 9827–9855.CrossRefGoogle Scholar
Fornari, D. J., Humphris, S. E. and Perfit, M. R. (1997). Deep submergence science takes a new approach. Eos, Transactions of the American Geophysical Union, 78, 402 and 408.CrossRefGoogle Scholar
Fornari, D. J., Humphris, S. E.et al. (1996). LUSTRE '96 (LUcky STRike Exploration) R/V Knorr Cruise 145–19, Multidisciplinary investigations of hydrothermal vents on Lucky Strike Seamount and the tectonic and volcanic structure of the Mid-Atlantic Ridge rift valley between 37°10′-25′N: Near-bottom studies using the DSL-120kHz Sonar, ARGO-II and ROV Jason: Final Cruise Report, WHOI Technical Report. Woods Hole Oceanographic Institution.
Fornari, D. J., Malahoff, A. and Heezen, B. C. (1979). Visual observations of the volcanic micromorphology of Tortuga, Lorraine and Tutu seamounts, and petrology and chemistry of ridge and seamount features in and around the Panama Basin. Marine Geology, 31(1), 1–30.CrossRefGoogle Scholar
Fornari, D. J., Perfit, M. R., Allan, J. F.et al. (1988). Geochemical and structural studies of the Lamont seamounts: Seamounts as indicators of mantle processes. Earth and Planetary Science Letters, 89, 63–83.CrossRefGoogle Scholar
Fornari, D. J., Perfit, M. R., Malahoff, A. and Embley, R. (1983). Geochemical studies of abyssal lavas recovered by DSRV ALVIN from eastern Galápagos Rift, Inca Transform and Ecuador Rift: 1. Major element variations in natural glasses and special distribution of lavas. Journal of Geophysical Research, 88, 10,519–10,529.CrossRefGoogle Scholar
Fornari, D. J., Ryan, W. B. F. and Fox, P. J. (1984). The evolution of craters and calderas on young seamounts: Insights from Sea MARC I and Sea Beam sonar surveys of a small seamount group near the axis of the East Pacific Rise at ∼ 10oN. Journal of Geophysical Research, 89, 11,069–11,083.CrossRefGoogle Scholar
Fornari, D. J., Shank, T., Von Damm, K. L.et al. (1998b). Time-series temperature measurements at high-temperature hydrothermal vents, East Pacific Rise 9° 49′–51′N: Monitoring a crustal cracking event. Earth and Planetary Science Letters, 160, 419–431.CrossRefGoogle Scholar
Fornari, D. J., Tivey, M. A., Schouten, H.et al. (2004). Submarine lava flow emplacement at the East Pacific Rise 9°50′N: Implications for uppermost oceanic crust stratigraphy and hydrothermal fluid circulation. In German, C. R., Lin, J. and Parson, L. M. (eds.), Mid-Ocean Ridges: Hydrothermal Interactions Between the Lithosphere and Oceans. Geophysical Monograph 148, Washington, DC: American Geophysical Union, pp. 187–218.Google Scholar
Fornari, D. J., Von Damm, K. L., Bryce, J. G.et al. (2012a). The East Pacific Rise between 9°N and 10°N: Twenty-five years of integrated, multidisciplinary oceanic spreading center studies. Oceanography, 25(1), 18–43, doi.org/10.5670/oceanog.2012.02.CrossRefGoogle Scholar
Forsyth, D. (1992). Geophysical constraints on mantle flow and melt generation beneath mid-ocean ridges. In Phipps Morgan, J., Blackman, D. K. and Sinton, J. M. (eds.), Mantle Flow and Melt Generation at Mid-Ocean Ridges. Washington, D C: American Geophysical Union, pp. 1–66.Google Scholar
Fouquet, Y., Charlou, J.-L., Costa, I.et al. (1995). Atlantic lava lakes and hot vents. Nature, 377, 201.CrossRefGoogle Scholar
Fournier, R. O. (1987). Conceptual models of brine evolution in magmatic-hydrothermal systems. In Decker, R. W., Wright, T. L. and Stauffer, P. H., Volcanism in Hawaii. US Geological Survey Professional Paper, 1350, pp. 1487–1506.Google Scholar
Foustoukos, D. I. and Seyfried, W. E. (2004). Hydrocarbons in hydrothermal vent fluids: The role of chromium-bearing catalysts. Science, 304, 1002–1005.CrossRefGoogle ScholarPubMed
Foustoukos, D. I. and Seyfried, W. E. (2007a). Fluid phase separation processes in submarine hydrothermal systems. In Liebscher, A. and Heinrich, C. A. (eds.), Fluid-Fluid Interactions. Mineralogical Society of America, pp. 213–233.Google Scholar
Foustoukos, D. I. and Seyfried, W. E. (2007b). Quartz solubility in the two-phase and critical region of the NaCl-KCl-H2O system: Implications for submarine hydrothermal vent systems at 9°50′N East Pacific Rise. Geochimica et Cosmochimica Acta, 71, 186–201CrossRefGoogle Scholar
Fowler, C. M. R. (1976). Crustal structure of the Mid-Atlantic Ridge crest at 37°N. Geophysical Journal of the Royal Astronomical Society, 47, 459–491.CrossRefGoogle Scholar
Fox, C. G. (1990). Consequences of phase separation on the distribution of hydrothermal fluids at ASHES vent field, Axial Volcano, Juan de Fuca Ridge. Journal of Geophysical Research, 95, 12,923–12,926.CrossRefGoogle Scholar
Fox, C. G. and Dziak, R. P. (1998). Hydroacoustic detection of volcanic activity on the Gorda Ridge, February–March 1996. Deep-Sea Research Part II: Topical Studies in Oceanography, 45(12), 2513–2530.CrossRefGoogle Scholar
Fox, C. G., Chadwick, W. W. and Embley, R. W. (1992). Detection of changes in ridge-crest morphology using repeated multibeam surveys. Journal of Geophysical Research, 97, 11,149–11,162.CrossRefGoogle Scholar
Fox, C. G., Chadwick, W. W. and Embley, R. W. (2001). Direct observation of a submarine volcanic eruption from a sea-floor instrument caught in a lava flow. Nature, 412, 727–729.CrossRefGoogle Scholar
Fox, C. G., Dziak, R. P., Matsumoto, H. and Schreiner, A. E. (1994). Potential for monitoring low-level seismicity on the Juan de Fuca Ridge using military hydrophone arrays. Marine Technology Society Journal, 27(4), 22–30.Google Scholar
Fox, C. G., Radford, W. E., Dziak, R. P.et al. (1995). Acoustic detection of a seafloor spreading episode on the Juan de Fuca Ridge using military hydrophone arrays. Geophysical Research Letters, 22(2), 131–134.CrossRefGoogle Scholar
Fox, P. J. and Gallo, D. G. (1986). The geology of the North Atlantic transform plate boundaries and their aseismic extensions. In Vogt, P. R. and Tucholke, B. E. (eds.), The Geology of North America: The Western North Atlantic Region. DNAG Series, Boulder, CO: Geological Society of America, pp. 157–172.Google Scholar
Fox, P. J. and Gallo, D. G. (1989). Transforms of the eastern central Pacific. In Winterer, E. L., Hussong, D. M. and Decker, R. W. (eds.), The Eastern Pacific and Hawaii. Boulder, CO: Geological Society of America, pp. 111–124.Google Scholar
Fox, P. J. and Stroup, J. B. (1981). The plutonic foundation of the oceanic crust. In Emiliani, C. (ed.), The Oceanic Lithosphere. New York: Wiley, pp. 119–218.Google Scholar
Fox, P. J., Detrick, R. S. and Purdy, G. M. (1980). Evidence for crustal thinning near fracture zones: Implications for ophiolites. In Panayiotou, A. (ed.), Proceedings of the Ophiolite Symposium. Nicosia, Cyprus: Geological Survey Department, pp. 161–168.Google Scholar
Fox, P. J., Schreiber, E. and Peterson, J. (1972). Compressional wave velocities in basalt and altered basalt recovered during Leg 14. In Hayes, D. E., Pimm, A. C.et al. (eds.), Initial Reports of the Deep Sea Drilling Project, Leg 14. Washington, DC: US Government Printing Office, pp. 773–775.Google Scholar
Fox, P. J., Schreiber, E. and Peterson, J. J. (1973). The geology of the oceanic crust: Compressional wave velocities of oceanic rocks. Journal of Geophysical Research, 78, 5155–5172.CrossRefGoogle Scholar
Fox, P. J., Schreiber, E., Rowlett, H. and Mccamy, K. (1976). The geology of the Oceanographer Fracture Zone: A model for fracture zones. Journal of Geophysical Research, 81, 4117–4128.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., Armijo, R., Cheminee, J. L.et al. (1990). 1 Ma East Pacific Rise oceanic crust and uppermost mantle exposed by rifting in Hess Deep (equatorial Pacific Ocean). Earth and Planetary Science Letters, 101, 281–295.CrossRefGoogle Scholar
Francheteau, J., Armijo, R., Cheminee, J. L.et al. (1992). Dyke complex of the East Pacific Rise exposed in the walls of Hess Deep and the structure of the upper oceanic crust. Earth and Planetary Science Letters, 111, 109–121.CrossRefGoogle Scholar
Francheteau, J., Armijo, R., Cogné, J. P.et al. (1994). Submersible observations of the Easter microplate and its boundary. Eos, Transactions of the American Geophysical Union, 75, 582.Google Scholar
Francheteau, J., Juteau, T. and Rangan, C. (1979b). Basaltic pillars in collapsed lava-pools on the deep ocean-floor. Nature, 281, 209–211.CrossRefGoogle Scholar
Francheteau, J., Needham, H. D., Choukroune, P.et al. (1979a). Massive deep sea sulfide ore deposits discovered on the East Pacific Rise. Nature, 277, 523–528.CrossRefGoogle Scholar
Francheteau, J., Patriat, P., Segoufin, J.et al. (1988). Pito and Orongo fracture zones: The northern and southern boundaries of the Easter microplate (southeast Pacific). Earth and Planetary Science Letters, 89, 363–374.Google Scholar
Francheteau, J., Yelles-Chaouche, A. and Craig, H. (1987). The Juan Fernandez microplate north of the Pacific-Nazca-Antarctic plate junction at 35°S. Earth and Planetary Science Letters, 86, 253–286.CrossRefGoogle Scholar
Francis, T. J. G. (1981). Serpentinization faults and their role in the tectonics of slow spreading ridges. Journal of Geophysical Research, 86, 11,616–11,622.CrossRefGoogle Scholar
Frey, F. A., Walker, N., Stakes, D., Hart, S. R. and Nielsen, R. (1993). Geochemical characteristics of basaltic glasses from the AMAR and FAMOUS axial valleys, Mid-Atlantic Ridge (36°–37° N): Petrogenetic implications. Earth and Planetary Science Letters, 115(1), 117–136.CrossRefGoogle Scholar
Fricke, H., Giere, O., Stetter, K.et al. (1989). Hydrothermal vent communities at the shallow subpolar Mid-Atlantic Ridge. Marine Biology, 102, 425–429.CrossRefGoogle Scholar
Früh-Green, G. L., Connolly, J. A. D., Kelley, D. S., Plas, A. and Grobety, B. (2004). Serpentinization of oceanic peridotites: Implications for geochemical cycles and biological activity. In Wilcock, W. D., Kelley, D. S., Baross, J. A., DeLong, E. and Cary, C. (eds.), The Subseafloor Biosphere at Mid-Ocean Ridges. Geophysical Monograph 144, Washington, DC: American Geophysical Union, pp. 119–135.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, 302, 495–498.CrossRefGoogle Scholar
Fryer, P., Wheat, C. G. and Mottl, M. J. (1999). Mariana blueschist mud volcanism: Implications for conditions within the subduction zone. Geology, 27, 103–106.2.3.CO;2>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°50'′N. Geochemistry, Geophysics, Geosystems, 11(8), Q08024, doi:10.1029/2010GC003058.CrossRefGoogle Scholar
Fustec, A., Desbruyères, D. and Laubier, L. (1988). Estimation de la biomase des peuplements associes aux sources hydrothermales profoundes de la dorsale du Pacifique oriental a 13°N. Biologie et ecologie des sources hydrothermales. Oceanologica Acta, 8, 15–22.Google Scholar
Gablina, I. F., Semkova, T. A., Stepanova, T. V. and Gor'kova, N. V. (2006). Diagenetic alteration of copper sulfides in modern ore-bearing sediments of the Logatchev-1 hydrothermal field (Mid-Atlantic Ridge 14°45′N). Lithology and Mineral Resources, 41, 27–44.CrossRefGoogle Scholar
Gale, A., Dalton, C. A., Langmuir, C. H., Su, Y. and Schilling, J.-G. (2013). The mean composition of ocean ridge basalts. Geochemistry, Geophysics, Geosystems, 14, 489–518, doi:10.1029/2012GC004334.CrossRefGoogle Scholar
Gallant, R. M. and Von Damm, K. L. (2006). Geochemical controls on hydrothermal fluids from the Kairei and Edmond vent fields, 23°–25°S, Central Indian Ridge. Geochemistry, Geophysics, Geosystems, 7(6). Q06018; doi:10.1029/2005GC001067.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
Gamo, T., Chiba, H., Yamanaka, T.et al. (2001). Chemical characteristics of newly discovered black smoker fluids and associated hydrothermal plumes at the Rodriguez Triple Junction, Central Indian Ridge. Earth and Planetary Science Letters, 193, 371–379.CrossRefGoogle Scholar
Gamo, T., Nakayama, E., Shitashima, K.et al. (1996). Hydrothermal plumes at the Rodriguez Triple Junction, Indian Ridge. Earth and Planetary Science Letters, 142, 261–270.CrossRefGoogle Scholar
Gao, Y., Hoefs, J., Hellebrand, E., Von Der Handt, A. and Snow, J. E. (2007). Trace element zoning in pyroxenes from ODP Hole 735B gabbros: Diffusive exchange or synkinematic crystal fractionation?Contributions to Mineralogy and Petrology, 152, 429–442, doi:10.1007/s00410-006-0158-4.CrossRefGoogle Scholar
Garces, M. and Gee, J. S. (2007). Paleomagnetic evidence of large footwall rotations associated with low-angle faults at the Mid-Atlantic Ridge. Geology, 35, 279–282.CrossRefGoogle Scholar
Gardner, J. V., Armstrong, A. A., Calder, B. R. and Beaudoin, J. (2014). So, how deep is the Mariana Trench?Marine Geodesy, 42, doi:10.1080/01490419.2013.837849.Google Scholar
Garry, W. B., Gregg, T. K. P., Soule, S. A. and Fornari, D. J. (2006). Formation of submarine lava channel textures: Insights from laboratory simulations. Journal of Geophysical Research, 111, B03104, doi:10.1029/2005JB003796.CrossRefGoogle Scholar
Gass, I. G. (1968). Is the Troodos massif of Cyprus a fragment of Mesozoic ocean floor?Nature, 220, 39–42.CrossRefGoogle Scholar
Gast, P. W. (1968). Trace element fractionation and the origin of tholeiitic and alkaline magma types. Geochimica et Cosmochimica Acta, 32, (10), 1057–1086.CrossRefGoogle Scholar
Gebruk, A. V., Chevaldonne, P., Shank, T., Lutz, R. A. and Vrijenhoek, R. C. (2000). Deep-sea hydrothermal vent communities of the Logatchev area (14°45′N, Mid-Atlantic Ridge): Diverse biotopes and high biomass. Journal of the Marine Biological Association, 80, 383–393.CrossRefGoogle Scholar
Gebruk, A. V., Pimenov, N. V. and Savvichev, A. S. (1993). Feeding specialization of bresiliid shrimps in the TAG site hydrothermal community. Marine Ecology Progress Series, 98, 247–253.CrossRefGoogle Scholar
Gente, P., Dyment, J., Maia, M. and Goslin, M. (2003). Interaction between the Mid-Atlantic Ridge and the Azores hot spot during the last 85 Myr: Emplacement and rifting of the hot spot-derived plateaus. Geochemistry, Geophysics, Geosystems, 4(10), 8514 doi:10.1029/2003GC000527.CrossRefGoogle Scholar
German, C. R., Bennett, S. A., Connelly, D. P.et al. (2008a). Hydrothermal activity on the southern Mid-Atlantic Ridge: Tectonically- and volcanically-controlled venting at 4–5°S. Earth and Planetary Science Letters, 273, 332–344.CrossRefGoogle Scholar
German, C. R., Thurnherr, A. M., 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, 57, 518–527.CrossRefGoogle Scholar
German, C. R., Yoerger, D. R., Jakuba, M.et al. (2008b). Hydrothermal exploration with the Autonomous Benthic Explorer. Deep-Sea Research Part 1, 55, 203–219.CrossRefGoogle Scholar
Gill, J. B. (1976). Composition and age of Lau Basin and Ridge volcanic rocks: Implications for evolution of an interarc basin and remnant arc. Geological Society of America Bulletin, 87(10), 1384–1395.2.0.CO;2>CrossRefGoogle Scholar
Gillis, K. M. (1995). Controls on hydrothermal alteration in fast-spreading oceanic crust. Earth and Planetary Science Letters, 134, 473–489.CrossRefGoogle Scholar
Gillis, K. M. (2002). The root zone of an ancient hydrothermal system exposed in the Troodos ophiolite, Cyprus. Journal of Geology, 110, 57–74.CrossRefGoogle Scholar
Gillis, K. M. (2008). The roof of an axial magma chamber: A hornfelsic heat exchanger. Geology, 36(4), 299–302, doi:10.1130/G24590A.1.CrossRefGoogle Scholar
Gillis, K. M. and Coogan, L. A. (2002). Anatectic migmatites from the roof of an ocean ridge magma spreading centre. Journal of Petrology, 43, 2075–2095.CrossRefGoogle Scholar
Gillis, K. M., and Roberts, M. D. (1999). Cracking at the magma-hydrothermal transition: Evidence from the Troodos Ophiolite, Cyprus. Earth and Planetary Science Letters, 169, 227–244.CrossRefGoogle Scholar
Gillis, K. M. and Robinson, P. T. (1988). Distribution of alteration zones in the upper oceanic crust. Geology, 16, 262–266.2.3.CO;2>CrossRefGoogle Scholar
Gillis, K. M. and Robinson, P. T. (1990). Patterns and processes of alteration in the lavas and dikes of the Troodos Ophiolite, Cyprus. Journal of Geophysical Research, 95, 21,523–21,548.CrossRefGoogle Scholar
Gillis, K. M., Coogan, L. A. and Pedersen, R. B. (2005). Strontium isotope constraints on fluid flow in the upper oceanic crust at the East Pacific Rise. Earth and Planetary Science Letters, 232, 82–94.CrossRefGoogle Scholar
Gillis, K. M., Mével, C., Allan, J.et al. (1993a). Proceedings of the Ocean Drilling Program, Initial Reports, Leg 147. College Station, TX: Ocean Drilling Program.Google Scholar
Gillis, K. M., Muehlenbachs, K., Stewart, M., Gleeson, T. and Karson, J. A. (2001). Fluid flow patterns in fast-spreading East Pacific Rise crust exposed at Hess Deep. Journal of Geophysical Research, 106, 26,311–26,329.CrossRefGoogle Scholar
Gillis, K. M., Snow, J. E., Klaus, A.et al. (2013). Primitive layered gabbros from fast-spreading lower oceanic crust. Nature, doi:10.1038/nature12778.CrossRef
Gillis, K. M., Thompson, G. and Kelley, D. S. (1993b). A view of the lower crustal component of hydrothermal systems at the Mid-Atlantic Ridge. Journal of Geophysical Research, 98, 19,597–19,619.CrossRefGoogle Scholar
Girardeau, J. and Nicolas, A. (1981). Structures in two of the Bay of Islands (Newfoundland) ophiolite massifs: A model for oceanic crust and upper mantle. Tectonophysics, 77, 1–34.CrossRefGoogle Scholar
Girardeau, J., Cornen, G., Beslier, M.-O.et al. (1998). Extensional tectonics in the Gorringe Bank rocks, Eastern Atlantic Ocean: Evidence of an oceanic ultra-slow mantellic accreting centre. Terra Nova, 10, 330–336.CrossRefGoogle Scholar
Girguis, P. R. and Holden, J. F. (2012). On the potential for bioenergy and biofuels from hydrothermal vent microbes. Oceanography, 25(1), 213–217, doi:org/10.5670/oceanog.2012.20.CrossRefGoogle Scholar
Glennie, K. W., Boeuf, M. G. A., Hughes Clark, M. W.et al. (1973). Late Cretaceous nappes in Oman mountains and their geologic evolution. American Association of Petroleum Geologists Bulletin, 57, 5–27.Google Scholar
Glickson, D. A., Kelley, D. S. and Delaney, J. R. (2007). Geology and hydrothermal evolution of the Mothra Hydrothermal Field, Endeavour Segment, Juan de Fuca Ridge. Geochemistry, Geophysics, Geosystems, 8(6), Q06010, doi:10.1029/2007GC001588.CrossRefGoogle Scholar
Godard, M., Awaji, S., Hansen, H.-E.et al. (2009). Geochemistry of a long in-situ section of intrusive slow-spread lithosphere: Results from IODP Site U1309 (Atlantis Massif, 30°N Mid-Atlantic Ridge). Earth and Planetary Science Letters, 279, 110–122.CrossRefGoogle Scholar
Godard, M., Dautria, J.-M. and Perrin, M. (2003). Geochemical variability of the Oman ophiolite lavas: Relationship with spatial distribution and paleomagnetic directions. Geochemistry, Geophysics, Geosystems, 4(6), doi:2002GC000452.CrossRefGoogle Scholar
Godard, M., Jousselin, D. and Bodinier, J.-L. (2000). Relationships between geochemistry and structure beneath a paleo-spreading centre: A study of the mantle section in the Oman Ophiolite. Earth and Planetary Science Letters, 180, 133–148.CrossRefGoogle Scholar
Goff, J. A., Fornari, D. J., Cochran, J. R., Keeley, C. and Malinverno, A. (1993). Wilkes transform system and “nannoplate”. Geology, 21(7), 623–626.2.3.CO;2>CrossRefGoogle Scholar
Goffredi, S. K., Warén, A., Orphan, V. J., Van Dover, C. L. and Vrijenhoek, R. C. (2004). Novel forms of structural integration between microbes and a hydrothermal vent gastropod from the Indian Ocean. Applied Environmental Microbiology, 70(5), 3082–3090, doi:10.1128/AEM.70.5.3082-3090.2004.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°46′N–9°56′N: Implications for ridge crest plumbing and decadal changes in magma chamber compositions. Geochemistry, Geophysics, Geosystems, 11, Q05T09, doi:10.1029/2009GC00297.CrossRefGoogle Scholar
Goud, M. R. and Karson, J. A. (1985). Tectonics of short-offset, slow-slipping transform zones in the FAMOUS area, Mid-Atlantic ridge. Marine Geophysical Researches, 7, 489–514.CrossRefGoogle Scholar
Gràcia, E., Charlou, J. C., Radford-Knoery, J. and Parson, L. M. (2000). Non-transform offsets along the Mid-Atlantic Ridge south of the Azores (38°N–34°N): Ultramafic exposures and hosting of hydrothermal vents. Earth and Planetary Science Letters, 177, 89–103.CrossRefGoogle Scholar
Grandin, R., Socquet, A., Jacques, E.et al. (2010). Sequence of rifting in Afar, Manda-Hararo Rift, Ethiopia, 2005–2009: Time-space evolution and interactions between dikes from interferometric synthetic aperture radar and static stress change modeling. Journal of Geophysical Research, 115(B10), doi:10.1029/2009JB000815.CrossRefGoogle Scholar
Grassle, J. F. (1985). Hydrothermal vent animals: Distribution and biology. Science, 229, 713–717.CrossRefGoogle ScholarPubMed
Grassle, J. F. and Petrecca, R. (1994). Macrofaunal response to artificial enrichments and depressions in a deep-sea habitat. Journal of Marine Research, 52, 345–369.Google Scholar
Grassle, J. F., Berg, C. J., Childress, J. J.et al. (1979). Galápagos '79: Initial findings of a deep-sea biological quest. Oceanus, 22(2), 2–10.Google Scholar
Greenbaum, D. (1972). Magmatic processes at ocean ridges and evidence from the Troodos massif, Cyprus. Nature Physical Science, 238, 18–21.CrossRefGoogle Scholar
Greenberg, D. S. (1964). Mohole: The project that went awry. Science, 143, 115–119.CrossRefGoogle Scholar
Gregg, T. K. P. and Chadwick, W. W. (1996). Submarine lava-flow inflation: A model for the formation of lava pillars. Geology, 24, 981–984.2.3.CO;2>CrossRefGoogle 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 Fornari, D. J. (1998). Long submarine lava flows: Observations and results from numerical modeling. Journal of Geophysical Research, 103, 27,517–27,531.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: The East Pacific Rise at 9°46′–51′N. Earth and Planetary Science Letters, 144, E1–E7.CrossRefGoogle Scholar
Gregg, T. K. P., Fornari, D. J., Perfit, M. R., Ridley, W. I. and Kurz, M. D. (2000). Using submarine lava pillars to record mid-ocean ridge eruption dynamics. Earth and Planetary Science Letters, 178, 195–214.CrossRefGoogle Scholar
Griffiths, R. W. and Fink, J. H. (1992a). Solidification and morphology of submarine lavas: A dependence on extrusion rate. Journal of Geophysical Research, 97, 19,729–19,737.CrossRefGoogle Scholar
Griffiths, R. W. and Fink, J. H. (1992b). The morphology of lava flows in planetary environments: Predictions from analog experiments. Journal of Geophysical Research, 97, 19,739–19,748.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(8), Q08012, doi:10.1029/2008GC002063.CrossRefGoogle Scholar
Grove, T. L., Kinzler, R. J. and Bryan, W. B. (1992). Fractionation of mid-ocean ridge basalts. In Phipps Morgan, J., Blackman, D. K. and Sinton, J. (eds.), Mantle Flow and Melt Generation at Mid-Ocean Ridges. Washington, DC: American Geophysical Union, pp. 281–310, doi:org/10.1029/GM071p0281.Google Scholar
Gudmundsson, A. (1983). Form and dimensions of dykes in eastern Iceland. Tectonophysics, 95, 295–307.CrossRefGoogle Scholar
Gudmundsson, A. (1990). Emplacement of dikes, sills and crustal magma chambers at divergent plate boundaries. Tectonophysics, 176, 257–275.CrossRefGoogle Scholar
Haase, K. M., Koschinsky, A. and Petersen, S. (2009). Diking, young volcanism and diffuse hydrothermal activity on the southern Mid-Atlantic Ridge: the Lilliput field at 9°33′S. Marine Geology, 266, 52–64.CrossRefGoogle Scholar
Haase, K. M., Petersen, S., Koschinsky, A.et al. (2007). Young volcanism and related hydrothermal activity at 5°S on the slow-spreading southern Mid-Atlantic Ridge. Geochemistry, Geophysics, Geosystems, 8(11), Q11002, doi:10.1029/2006GC001509.CrossRefGoogle Scholar
Hall, M. (1876). Note upon a portion of basalt from Mid-Atlantic. Mineralogical Magazine, 1(1), 1–3, doi:10.1180/minmag.1876.001.1.03.Google Scholar
Hammond, S. R. (1990). Relationships between lava types, seafloor morphology, and the occurrence of hydrothermal venting in the ASHES vent field of Axial Volcano. Journal of Geophysical Research, 95, 12,875–12,893.CrossRefGoogle Scholar
Han, S., Carbotte, S. M., Carton, H.et al. (2014). Architecture of on- and off-axis magma bodies at EPR 9°37′–40′ N and implications for oceanic crustal accretion. Earth and Planetary Science Letters, 390, 31–44.CrossRefGoogle Scholar
Hanna, H. D. (2004). Geochemical Variations in Basaltic Glasses from an Incipient Rift and Upper Level Gabbros from Hess Deep, Eastern Equatorial Pacific. M.S. thesis, Duke University.Google Scholar
Hannington, M. D. and Scott, S. D. (1988). Mineralogy and geochemistry of a hydrothermal silica-sulfide-sulfate spire in the caldera of Axial Seamount, Juan de Fuca Ridge. Canadian Mineralogist, 26, 603–625.Google Scholar
Hannington, M. D., de Rhonde, C. E. J. and Petersen, S. (2005). Sea-floor tectonics and submarine hydrothermal systems. In Hedquist, J. W., Thompson, J. F. H., Goldfarb, R. J. and Richards, J. R. (eds.), Economic Geology, One Hundredth Anniversary Volume, 1905–2005. Littleton, CO: Society of Economic Geologists, pp. 111–142.Google Scholar
Hannington, M. D., Galley, A. G.Herzig, P. M. and Petersen, S. (1998). Comparison of the TAG mound and stockwork complex with Cyprus-type massive sulfide deposits. In Herzig, P. M., Humphris, S. E., Miller, D. J., and Zierenberg, R. A. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results Leg 158. College Station, TX: Ocean Drilling Program, pp. 389–415.Google 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 Humphris, S., Zierenberg, R., Mullineau, L. and Thomson, R. (eds.), Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions. Geophysical Monograph 91, Washington, DC: American Geophysical Union, pp. 115–157.Google Scholar
Hannington, M., Herzig, P.Stoffers, P.et al. (2001). First observations of high-temperature submarine hydrothermal vents and massive anhydrite deposits off the north coast of Iceland. Marine Geology, 177, 199–220.CrossRefGoogle 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, 98, 13,925–13,944.CrossRefGoogle Scholar
Harding, A. J., Orcutt, J., Kappus, M.et al. (1989). The structure of young oceanic crust at 13°N on the East Pacific Rise from expanding spread profiles. Journal of Geophysical Research, 94, 12,163–12,196.CrossRefGoogle Scholar
Harper, G. D. (1982). Evidence for large-scale rotations at spreading centers from the Josephine Ophiolite. Tectonophysics, 82, 25–44.CrossRefGoogle Scholar
Harper, G. D. (1988). Episodic magma chambers and amagmatic extension in the Josephine ophiolite. Geology, 16, 831–834.2.3.CO;2>CrossRefGoogle Scholar
Harris, A., Bailey, J., Calvari, S. and Dehn, J. (2005). Heat loss measured at a lava channel and its implications for down-channel cooling and rheology. In Manga, M. and Ventura, G. (eds.), Kinematics and Dynamics of Lava Flows. Boulder, CO: Geological Society of America, Special Paper 396, pp. 125–146.CrossRefGoogle Scholar
Hart, S. R., Staudigel, H., Koppers, A. A. P.et al. (2000). Vailulu'u undersea volcano: The new Samoa. Geochemistry, Geophysics, Geosystems, 1(12), 1056, doi:10.1029/2000GC000108.CrossRefGoogle Scholar
Hartley, M. E. and Thordarsson, T. (2013). The 1874–1876 volcano-tectonic episode at Askja, North Iceland: Lateral flow revisited. Geochemistry, Geophysics, Geosystems, 14, 2286–2309, doi:10.1002/ggge.20151.CrossRefGoogle Scholar
Hashimoto, J., Ohta, S., Gamo, T.et al. (2001). Hydrothermal vents and associated biological communities in the Indian Ocean. InterRidge News, 10, 21–22.Google Scholar
Hawkins, J. W. (2003). Geology of supra-subduction zones: Implications for the origin of ophiolites. In Dilek, Y. and Newcomb, S. (eds.), Ophiolite Concept and the Evolution of Geological Thought. Boulder, CO: Geological Society of America, pp. 227–268.CrossRefGoogle Scholar
Hayman, N. W. and Karson, J. A. (2007). Faults and damage zones in fast-spread crust exposed on the north wall of the Hess Deep Rift: Conduits and seals in seafloor hydrothermal systems. Geochemistry, Geophysics, Geosystems, 8(10), Q10002, doi:10.1029/2007GC001623.CrossRefGoogle Scholar
Hayman, N. W., Grindlay, N. R., Perfit, M. R.et al. (2011). Oceanic core complex development at the ultraslow spreading Mid-Cayman Spreading Center. Geochemistry, Geophysics, Geosystems, 12(3), Q0AG02, doi:10.1029/2010GC003240.CrossRefGoogle Scholar
Haymon, R. M. (1983). Growth history of black smoker hydrothermal chimneys. Nature, 301, 695–698.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′– 54′N) and its relationship to magmatic and tectonic processes on fast-spreading mid-ocean ridges. Earth and Planetary Science Letters, 104, 513–534CrossRefGoogle 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°42–52′N: Direct submersible observations of seafloor phenomena associated with an eruption event in April, 1991. Earth and Planetary Science Letters, 119, 85–101.CrossRefGoogle Scholar
Haymon, R. M., Koski, R. A. and Abrams, M. J. (1989). Hydrothermal discharge zones beneath massive sulfide deposits mapped in the Oman ophiolite. Geology, 17, 531–535.2.3.CO;2>CrossRefGoogle Scholar
Haymon, R. M., Koski, R. A. and Sinclair, C. (1984). Fossils of hydrothermal vent worms from cretaceous sulfide ores of the Samail Ophiolite, Oman. Science, 223(4643), 1407–1409, doi: 10.1126/science.223.4643.1407.CrossRefGoogle ScholarPubMed
Haymon, R. M., White, S. M., Baker, E. T.et al. (2008). High-resolution surveys along the hot spot-affected Galápagos Spreading Center: 3. Black smoker discoveries and the implications for geological controls on hydrothermal activity. Geochemistry, Geophysics, Geosystems, 9 (12), 1–30.CrossRefGoogle Scholar
Head, J. W. and Wilson, L. (2003). Deep submarine pyroclastic eruptions: Theory and predicted landforms and deposits. Journal of Volcanology and Geothermal Research, 121(3), 155–193.CrossRefGoogle Scholar
Head, J. W., Hiesinger, H., Ivanov, M. A.et al. (1999). Possible ancient oceans on Mars: Evidence from Mars orbiter laser altimeter data. Science, 286, 2134–2137.CrossRefGoogle ScholarPubMed
Hébert, L. B. and Montési, L. G. J. (2010). Generation of permeability barriers during melt extraction at mid-ocean ridges. Geochemistry, Geophysics, Geosystems, 11, Q12008, doi:10.1029/2010GC003270.CrossRefGoogle Scholar
Hébert, R., Bideau, D. and Hékinian, R. (1983). Ultramafic and mafic rocks from the Garrett Transform Fault near 13°30'S on the East Pacific Rise: Igneous petrology. Earth and Planetary Science Letters, 65, 107–125.CrossRefGoogle Scholar
Hedrick, D. B., Pledger, R. J., White, D. C. and Baross, J. A. (1992). In situ microbial ecology of hydrothermal vent sediments. FEMS Microbial Ecology, 101, 1–10.CrossRefGoogle Scholar
Heezen, B. C. (1960). The rift in the ocean floor. Scientific American, 203, 98–110.CrossRefGoogle Scholar
Heezen, B. C. and Hollister, C. D. (1971). The Face of the Deep. New York: Oxford University Press.Google Scholar
Heezen, B. C. and Tharp, M. (1977). World Ocean Floor. Palisades, NY: Office of Naval Research, US Navy.Google Scholar
Heezen, B. C., Tharp, M. and Ewing, M. (1959). The Floors of the Oceans. 1. The North Atlantic. Special Paper 65, Boulder, CO: Geological Society of America, pp. 1–126.Google Scholar
Heft, K., Gillis, K. M., Pollock, M. A., Karson, J. A. and Klein, E. M. (2008). Constraints on the nature of axial hydrothermal systems from the sheeted dike complex exposed at Pito Deep. Geochemistry, Geophysics, Geosystems, 9(5), doi:10.1029/2007GC001926.Google Scholar
Heirtzler, J. R. and LePichon, X. (1974). FAMOUS: A plate tectonics study of the genesis of the lithosphere. Geology, June, 273–274.2.0.CO;2>CrossRef
Hékinian, R., Auzende, J. M., Francheteau, J.et al. (1985). Offset spreading centers near 12°53′N on the East Pacific Rise: Submersible observations and composition of the volcanics. Marine Geophysical Researches, 7, 359–377.CrossRefGoogle Scholar
Hékinian, R., Bideau, D., Cannat, M., Francheteau, J. and Hébert, R. (1992). Volcanic activity and crust-mantle exposure in the ultrafast Garrett Transform Fault near 13°18′S in the Pacific. Earth and Planetary Science Letters, 108, 259–275.CrossRefGoogle Scholar
Hékinian, R., Bideau, D., Francheteau, J.et al. (1993). Petrology of the East Pacific Rise crust and upper mantle exposed in Hess Deep (Eastern Equatorial Pacific). Journal of Geophysical Research, 98, 8069–8094.CrossRefGoogle Scholar
Hekinian, R. M., Fevrier, F., Avedik, P.et al. (1983a). East Pacific Rise near 13°N: Geology of new hydrothermal fields. Science, 219, 1321–1324.CrossRefGoogle ScholarPubMed
Hékinian, R., Francheteau, J., Renard, V.et al. (1983b). Intense hydrothermal activity at the axis of the East Pacific Rise near 13°N: Submersible witnesses the growth of sulfide chimney. Marine Geophysical Research, 6(1), 1–14, doi:10.1007/BF00300395.CrossRefGoogle Scholar
Helgason, J. and Zentilli, M. (1985). Field characteristics of laterally emplaced dikes: Anatomy of exhumed Miocene dike swarm in Reydarfjordur, eastern Iceland. Tectonophysics, 115, 247–274.CrossRefGoogle Scholar
Hellebrand, E., Snow, J. E., Mostefaoui, S. and Hoppe, P. (2005). Trace element distribution between orthopyroxene and clinopyroxene in peridotites from the Gakkel Ridge: A SIMS and NanoSIMS study. Contributions to Mineralogy and Petrology, 150(5), 486–504.CrossRefGoogle Scholar
Hellebrand, E., Snow, J. E. and Mühe, R. (2002). Mantle melting beneath Gakkel Ridge (Arctic Ocean): Abyssal peridotite spinel compositions. Chemical Geology, 182, (2), 227–235.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, 260–263, doi:10.1038/NGEO1104.CrossRefGoogle Scholar
Henstock, T. J., Woods, A. W. and White, R. S. (1993). The accretion of oceanic crust by episodic sill intrusion. Journal of Geophysical Research, 98, 4143–4161.CrossRefGoogle Scholar
Hersey, J. B. (ed.) (1967). Deep-Sea Photography. Baltimore: The Johns Hopkins University Press.Google Scholar
Herzberg, C. (2004). Partial crystallization of mid-ocean ridge basalts in the crust and mantle. Journal of Petrology, 45(12), 2389–2405.CrossRefGoogle Scholar
Herzberg, C., Asimow, P. D., Arndt, N.et al. (2007). Temperatures in ambient mantle and plumes: Constraints from basalts, picrites, and komatiites. Geochemistry, Geophysics, Geosystems, 8, Q02006, doi:10.1029/2006GC001390.CrossRefGoogle Scholar
Herzig, P. M. and Plüger, W. L. (1988). Exploration for hydrothermal activity near the Rodriguez Triple Junction, Indian Ocean. Canadian Mineralogist, 26, 721–736.Google Scholar
Hess, H. H. (1954). Geological hypotheses and the Earth's crust under the Oceans. Proceedings of the Royal Society, A, 222, 341–348.CrossRefGoogle Scholar
Hess, H. H. (1955). Serpentinites, orogeny and epeirogeny. In Crust of the Earth: A Symposium. Boulder, CO: Geological Society of America, Special Paper 62, pp. 391–408.CrossRefGoogle Scholar
Hess, H. H. (1962). History of ocean basins. In Engel, A. E. J., Hall, H. L. and Leonard, B. F. (eds.), Petrologic Studies: A Volume in Honor of A. F. Buddington. Boulder, CO: Geological Society of America, pp. 599–620.Google Scholar
Hess, H. H. (1965). Mid-oceanic ridges and tectonics of the sea floor. In Whittard, W. F. and Bradshaw, R. (eds.), Submarine Geology and Geophysics, Proceedings of the Seventeenth Symposium of the Colston Research Society. London: Butterworths, pp. 317–334.Google Scholar
Hessler, R. R., Smithey, W. M., Boudrias, M. A.et al. (1988). Temporal changes in megafauna at the Rose Garden hydrothermal vent, Galápagos Rift, eastern tropical Pacific. Deep-Sea Research, 35, 1681–1709.CrossRefGoogle Scholar
Hessler, R. R., Smithey, W. M. and Keller, M. A. (1985). Spatial and temporal variation of giant clams, tube worms, and mussels at deep-sea hydrothermal vents. Bulletin of the Biological Society of Washington, 6, 411–428.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. (2004). Propagating rifts and microplates at mid-ocean ridges. In Selley, R. C. (ed.), Encylopedia of Geology. London: Academic Press, pp. 396–405.Google Scholar
Hey, R. N., Baker, E., Bohnenstiehl, D.et al. (2004). Tectonic/volcanic segmentation and controls on hydrothermal venting along Earth's fastest seafloor spreading system, EPR 27°–32°S. Geochemistry, Geophysics, Geosystems, 5,(12, Q12007, doi:10.1029/2004GC000764.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, Galápagos 95.5°W. Journal of Geophysical Research, 91, 3369–3393.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., Sinton, J. M. and Duennebier, F. K. (1989). Propagating rifts and spreading centers. In Winterer, E. L., Hussong, D. M. and Decker, R. W. (eds.), The Geology of North America: The Eastern Pacific Ocean and Hawaii. Boulder, CO: Geological Society of America, pp. 161–176.Google Scholar
Hoagland, P., Beaulieu, S., Tivey, M. A.et al. (2010). Deep-sea mining of seafloor massive sulfides. Marine Policy, 34, 728–732.CrossRefGoogle Scholar
Hodges, F. N. and Papike, J. J. (1976). DSDP site 334: Magmatic cumulates from oceanic layer 3. Journal of Geophysical Research, 81, 4135–4151.CrossRefGoogle Scholar
Holden, J. F., Summit, M., Baross, J. A. (1998). Thermophilic and hyperthermophilic microorganisms in 3–30 °C fluids following a deep-sea volcanic eruption. FEMS Microbiology Ecology, 25, 33–41.Google Scholar
Holdsworth, R. E., Van Diggelen, E. W. E., Spiers, C. J.et al. (2011). Fault rocks from the SAFOD core samples: Implications for weakening at shallow depths along the San Andreas Fault, California. Journal of Structural Geology, 33(2), 132–144, doi:10.1016/j.jsg.2010.11.010 .CrossRefGoogle Scholar
Holm, N. G. and Baltscheffsky, H. (2011). Links between hydrothermal environments, pyrophosphate, Na+ and early evolution. Origins of Life and Evolution of the Biosphere, 41, 483–493.CrossRefGoogle ScholarPubMed
Holm, N. G. and Charlou, J. L. (2001). Initial indications of abiotic formation of hydrocarbons in the Rainbow ultramafic hydrothermal system, Mid-Atlantic Ridge. Earth and Planetary Science Letters, 191, 1–8.CrossRefGoogle Scholar
Hon, K., Kauahikaua, J. P., Denlinger, R. and Mackay, K. (1994). Emplacement and inflation of pahoehoe sheet flows: Observations and measurements of active lava flows on Kilauea Volcano, Hawaii. Geological Society of America Bulletin, 106(3), 351–370.2.3.CO;2>CrossRefGoogle Scholar
Hooft, E. E. and Detrick, R. S. (1993). The role of density in the accumulation of basaltic melts at mid-ocean ridges. Geophysical Research Letters, 20, 423–426.CrossRefGoogle Scholar
Hooft, E. E., Detrick, R. S. and Kent, G. M. (1997). Seismic structure and indicators of magma budget along the southern East Pacific Rise. Journal of Geophysical Research, 102, 27,319–27,340.CrossRefGoogle Scholar
Hooft, 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, 105, 8205–8226.CrossRefGoogle Scholar
Hooft, E. E., Patel, H., Wilcock, W., Becker, K.et al. (2010). A seismic swarm and regional hydrothermal and hydrologic perturbations: The northern Endeavour segment, February 2005. Geochemistry, Geophysics, Geosystems, 11, Q12015, doi:10.1029/2010GC003264.CrossRefGoogle Scholar
Hooft, 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, 289–310.CrossRefGoogle Scholar
Horita, J. and Berndt, M. E. (1999). Abiogenic ethane formation and isotopic fractionation under hydrothermal conditions. Science, 285, 1055–1057.CrossRefGoogle Scholar
Horst, A. J., Varga, R. G., Gee, J. S. and Karson, J. A. (2006). Magnetic remanence and anisotropy of magnetic susceptibility of dikes from super-fast spread crust exposed at Pito Deep Rift. Transactions of the American Geophysical Union, 87, T51C–1546.Google Scholar
Horst, A. J., Varga, R. J., Gee, J. S. and Karson, J. A. (2011). Paleomagnetic constraints on constructional deformation of superfast-spread oceanic crust exposed at Pito Deep Rift. Journal of Geophysical Research, 116(12), doi:10.1029/2011JB008268.CrossRefGoogle Scholar
Horst, A. J., Varga, R. J., Gee, J. S. and Karson, J. A. (2014). Diverse magma flow directions during construction of sheeted dike complexes at fast- to superfast-spreading centers. Earth and Planetary Science Letters, 408, 119–131.CrossRefGoogle Scholar
Hosford, A., Tivey, M. A., Matsumoto, T.et al. (2003). Crustal magnetization and accretion at the Southwest Indian Ridge near the Atlantis II fracture zone, 0–25 Ma. Journal of Geophysical Research, 108, doi:1029/2001JB000604.CrossRefGoogle Scholar
Houtz, R. E. (1976). Seismic properties of layer 2A in the Pacific. Journal of Geophysical Research, 81, 6321–6340.CrossRefGoogle Scholar
Huber, J. A., Butterfield, D. A. and Baross, J. A. (2003). Bacterial diversity in a subseafloor habitat following a deep-sea volcanic eruption. FEMS Microbiology Ecology, 42, 393–409.CrossRefGoogle Scholar
Hull, E. (1892). Volcanoes: Past and Present. New York: Charles Scribner's Sons and London: Walter Scott.Google 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, 105, 28,477–28,488.CrossRefGoogle Scholar
Humphris, S. E. and Kleinrock, M. C. (1996). Detailed morphology of the TAG active hydrothermal mound: Insights into its formation and growth. Geophysical Research Letters, 23, 3443–3446.CrossRefGoogle Scholar
Humphris, S. E. and Tivey, M. K. (2000). A synthesis of geological and geochemical investigations of the TAG hydrothermal field: Insights into fluid-flow and mixing processes in a hydrothermal system. In Dilek, Y., Moores, E., Elthon, D. and Nicolas, A. (eds.), Ophiolites and Oceanic Crust: New Insights from Field Studies and the Ocean Drilling Program. Special Paper, 349, Boulder, CO: Geological Society of America, pp. 213–235.CrossRefGoogle Scholar
Humphris, S. E., DeMenocal, P. B., Edwards, K. J., Fisher, A. T. and Saffer, D. (2011). The need for scientific ocean drilling. Eos, Transactions of the American Geophysical Union, 92(10), 83–84.CrossRefGoogle Scholar
Humphris, S. E., Fornari, D. J., Scheirer, D. S., German, C. R. and Parson, L. (2002). Geotectonic setting of hydrothermal activity on the summit of Lucky Strike Seamount (37°17′N, Mid-Atlantic Ridge). Geochemistry, Geophysics, Geosystems, 3(8), doi: 10.1029/2001GC000284.CrossRefGoogle Scholar
Humphris, S. E., Herzig, P. M.et al. (1996). Proceedings of the Ocean Drilling Program, Initial Reports Leg, 158. College Station, TX: Ocean Drilling Program.CrossRefGoogle Scholar
Hurst, S. D., Moores, E. M. and Varga, R. J. (1994). Structure and geophysical expression of the Solea Graben, Troodos Ophiolite Cyprus. Tectonics, 13, 139–156.CrossRefGoogle Scholar
Hussenoeder, S. A., Detrick, R. S., Kent, G. M., Schouten, H. and Harding, A. J. (2002a). Fine-scale seismic structure of young upper crust at 17°20′S on the fast spreading East Pacific Rise. Journal of Geophysical Research, 107(B8), doi:10.1029/2001JB001688.CrossRefGoogle Scholar
Hussenoeder, S. A., Kent, G. M. and Detrick, R. S. (2002b). Upper crustal seismic structure of the slow spreading Mid-Atlantic Ridge, 35 degrees N: Constraints on volcanic emplacement processes. Journal of Geophysical Research, 107(B8), 2156, doi:10.1029/2001JB001691.CrossRefGoogle Scholar
Hynes, A. (1975). Comment on “The Troodos ophiolite was probably formed in an island arc” by A. Miyashiro. Earth and Planetary Science Letters, 25, 213–216.CrossRefGoogle Scholar
Ildefonse, B., Blackman, D. K., John, B. E.et al. (2007). Oceanic core complexes and crustal accretion at slow-spreading ridges. Geology, 35(7), 623–626, doi:10.1130/G23531A.1.CrossRefGoogle Scholar
Irvine, T. N. (1982). Terminology for layered intrusions. Journal of Petrology, 23, 127–162.CrossRefGoogle Scholar
Irvine, T. N., Anderson, J. C. ø. and Brooks, C. K. (1998). Included blocks (and blocks with blocks) in the Skaergaard intrusion: Geologic relations and the origins of rhythmic modally graded layers. Geological Society of America Bulletin, 110(11), 1398–1447.2.3.CO;2>CrossRefGoogle Scholar
Isaaks, B., Oliver, J. and Sykes, L. R. (1968). Seismology and the new global tectonics. Journal of Geophysical Research, 73, 5855–5899.CrossRefGoogle Scholar
Jambon, A. (1994). Earth degassing and large-scale geochemical cycling of volatile elements. In Caroll, M. R. and Holloway, J. R. (eds.), Volatiles and Magmas. Reviews in Mineralogy, 30, Washington DC: Mineralogy Society of America, pp. 479–517.Google Scholar
Jamieson, J. W., Hannington, M. D., Clague, D.et al. (2013). Sulfide geochronology along the Endeavour Segment of the Juan de Fuca Ridge. Geochemistry, Geophysics, Geosystems, 14, 2084–2099, doi:10.1002/ggge.20133.CrossRefGoogle Scholar
Jannasch, H. W. (1995). Microbial interaction with hydrothermal fluids. In Humphris, S. E., Zierenberg, R. A., Mullineaux, L. S. and Thomson, R. E. (eds.), Seafloor Hydrothermal Systems: Physical, Chemical, Biological and Geological Interactions. Geophysical Monograph 91, Washington, DC: American Geophysical Union Press, pp. 273–296.Google Scholar
Jaroslow, G. E., Hirth, G. and Dick, H. J. B. (1996). Abyssal peridotite mylonites: Implications for grain-size sensitive flow and strain localization in the oceanic lithosphere. Tectonophysics, 256, 17–37.CrossRefGoogle Scholar
Jean-Baptiste, P., Mantisi, F., Pauwells, H., Grimaud, D. and Patriat, P. (1992). Hydrothermal 3He and manganese plumes at 19°29′S on the Central Indian Ridge. Geophysical Research Letters, 19, 1787–1790.CrossRefGoogle Scholar
Jenner, F. E. and O'Neill, H. S. C. (2012). Analysis of 60 elements in 616 ocean floor basaltic glasses. Geochemistry, Geophysics, Geosystems, 13, Q02005, doi:10.1029/2011GC004009.CrossRefGoogle Scholar
Johnson, H. P., Hutnak, M., Dziak, R. P.et al. (2000). Earthquake-induced in a hydrothermal system on the Juan de Fuca Ridge. Nature, 407, 174–177.CrossRefGoogle Scholar
Johnson, H. P., Tivey, M. A., Bjorklund, T. A. and Salmi, M. S. (2010). Hydrothermal circulation within the Endeavour Segment, Juan de Fuca Ridge. Geochemistry, Geophysics, Geosystems, 11(5), Q05002, doi.org/10.1029/2009GC002957.CrossRefGoogle Scholar
Johnson, K. S., Childress, J. J. and Beehler, C. L. (1988a). Short-term temperature variability in the Rose Garden hydrothermal vent field: An unstable deep-sea environment. Deep-Sea Research Part 1, 35, 1711–1721.CrossRefGoogle Scholar
Johnson, K. S., Childress, J. J., Hessler, R. R.et al. (1988b). Chemical and biological interactions in the Rose Garden [eastern Pacific Ocean] hydrothermal vent field, Galápagos spreading center. Deep-Sea Research Part 1. 35, 1723–1744.CrossRefGoogle Scholar
Johnson, K. T. M. (1998). Experimental determination of partition coefficients for rare earth and high-field-strength elements between clinopyroxene, garnet, and basaltic melt at high pressures. Contributions to Mineralogy and Petrology, 133, 60–68.CrossRefGoogle Scholar
Johnson, K. T. M. and Dick, H. J. B. (1992). Open system melting and temporal and spatial variation of peridotite and basalt at the Atlantis II fracture zone. Journal of Geophysical Research, 97, 9219–9241.CrossRefGoogle Scholar
Johnson, K. T. M., Dick, H. J. B. and Shimizu, N. (1990). Melting in the oceanic upper mantle: An ion microprobe study of diopsides in abyssal peridotites. Journal of Geophysical Research, 95, 2661–2678.CrossRefGoogle Scholar
Jolly, R. J. H. and Sanderson, D. J. (1995). Variations in the form and distribution of dykes in the Mull swarm, Scotland. Journal of Structural Geology, 17, 1543–1557.CrossRefGoogle Scholar
Jousselin, D. and Nicolas, A. (2000a). Structural mapping of an off-axis mantle diapir in the Mansah region (Sumail massif in the Oman ophiolite). Marine Geophysical Researches, 21, 243–257.CrossRefGoogle Scholar
Jousselin, D. and Nicolas, A. (2000b). The Moho transition zone in the Oman ophiolite: Relation with wehrlites in the crust and dunites in the mantle. Marine Geophysical Researches, 21, 229–241.CrossRefGoogle Scholar
Jousselin, D., Dunn, R. and Toomey, D. R. (2003). Modeling the seismic signature of structural data from the Oman Ophiolite: Can a mantle diapir be detected beneath the East Pacific Rise?Geochemistry, Geophysics, Geosystems, 4(7), 8610, doi:10.1029/2002GC000418.CrossRefGoogle Scholar
Jousselin, D., Morales, L. F. G., Nicolle, M. and Stephant, A. (2012). Gabbro layering induced by simple shear in the Oman Ophiolite Moho transition zone. Earth and Planetary Science Letters, 331–332, 55–66.CrossRefGoogle Scholar
Jousselin, D., Nicolas, A. and Boudier, F. (1998). Detailed mapping of a mantle diapir below a paleo-spreading center in the Oman ophiolite. Journal of Geophysical Research, 103, 18,153–18,170.CrossRefGoogle Scholar
Juniper, S. K. and Sarrazin, J. (1995). Interaction of vent biota and hydrothermal deposits: Present evidence and future experimentation. In Humphris, S., Zierenberg, R., Mullineau, L. and Thomson, R., Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions. Geophysical Monograph 91, Washington, DC: American Geophysical Union, pp. 178–193.Google Scholar
Juniper, S. K. and Tunnicliffe, V. (1997). Crustal accretion and the hot vent ecosystem. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 355, 459–474.CrossRefGoogle Scholar
Juniper, S. K., Martineu, P., Sarrazin, J. and Gelinas, Y. (1995). Microbial-mineral floc associated with nascent hydrothermal activity on CoAxial Segment, Juan de Fuca Ridge. Geophysical Research Letters, 22, 179–182.CrossRefGoogle Scholar
Juteau, T., Bideau, O., Dauteuil, G.et al. (1995). A submersible study in the western Blanco Fracture Zone, N.E. Pacific: Lithostratigraphy, magnetic structure and magmatic and tectonic evolution during the last 1.6 Ma. Marine Geophysical Researches, 17, 399–430.CrossRefGoogle Scholar
Juteau, T., Cannat, M. and Lagabrielle, Y. (1990). Serpentinized peridotites in the upper oceanic crust away from transform zones: A comparison of the results of previous DSDP and ODP legs. In Detrick, R. S., Honnorez, J., Bryan, W. B. and Juteau, T. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 81. College Station, TX: Ocean Drilling Program, pp. 303–308.Google Scholar
Kadar, E., Costa, V. and Santos, R. S. (2006). Distribution of micro-essential (Fe, Cu, Zn) and toxic (Hg) metals in tissues of two nutritionally distinct hydrothermal shrimps. Science of the Total Environment, 358, 143–150.CrossRefGoogle ScholarPubMed
Kappel, E. and Ryan, W. B. F. (1986). Volcanic episodicity and a non-steady state rift valley along the Northeast Pacific spreading centers: Evidence from Sea MARC I. Journal of Geophysical Research, 91, 13,925–13,940.CrossRefGoogle Scholar
Kappel, E. S. and Normark, W. R. (1987). Morphometric variability within the axial zone of the southern Juan de Fuca Ridge: Interpretation from Sea MARC II and Sea MARC I, and deep-sea photography. Journal of Geophysical Research, 92, 11,291–11,302.CrossRefGoogle Scholar
Karl, D. M. (1995). The Microbiology of Deep-Sea Hydrothermal Vents. Boca Raton, FL; New York: CRC Press, pp. 1–299.Google Scholar
Karson, J. A. (1977). The Geology of the Northern Lewis Hills, Western Newfoundland. Ph.D. dissertation, State University of New York at Albany.Google Scholar
Karson, J. A. (1987). Factors controlling the orientations of dykes in ophiolites and oceanic crust. In Halls, H. C. and Fahrig, W. F. (eds.), Mafic Dyke Swarms. Special Paper 34 Ottawa: Geological Association of Canada, pp. 229–241.Google 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 Malpas, J., Moores, E. M., Panayioutou, A. and Xenophontos, C. (eds.), Ophiolites and Oceanic Crustal Analogues: Proceedings of the Symposium “Troodos 1987”. Nicosia, Cyprus: Geological Survey Department, pp. 125–130.Google Scholar
Karson, J. A. (1998). Internal structure of oceanic lithosphere: A perspective from tectonic windows. In Buck, W. R., Delaney, P. T., Karson, J. A. and Lagabrielle, Y. (eds.), Faulting and Magmatism at Mid-Ocean Ridges. Geophysical Monograph 106, Washington, DC: American Geophysical Union, pp. 177–218.Google Scholar
Karson, J. A. (1999). Geological investigation of a lineated massif at the Kane transform: Implications for oceanic core complexes. Philosophical Transactions of the Royal Society, A, 357, 713–740.CrossRefGoogle Scholar
Karson, J. A. (2002). Geologic structure of uppermost oceanic crust created at fast- to intermediate-rate spreading centers. Annual Review of Earth and Planetary Sciences, 30, 347–384.CrossRefGoogle Scholar
Karson, J. A. and Dewey, J. F. (1978). Coastal Complex, western Newfoundland: An Early Ordovician oceanic fracture zone. Geological Society of America Bulletin, 89, 1037–1049.2.0.CO;2>CrossRefGoogle Scholar
Karson, J. A. and Dick, H. J. B. (1983). Tectonics of ridge-transform intersections at the Kane Fracture Zone, 24°N on the Mid-Atlantic Ridge. Marine Geophysical Researches, 6, 51–98.CrossRefGoogle Scholar
Karson, J. A. and Lawrence, R. M. (1997). Tectonic setting of serpentinite exposures on the western median valley wall of the MARK Area in the vicinity of Site 920. In Karson, J. A., Cannat, M., Miller, D. J. and Elthon, D. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 153. College Station, TX: Ocean Drilling Program, pp. 5–22.CrossRefGoogle Scholar
Karson, J. A. and Rona, P. A. (1990). Block-tilting, transfer faults and structural control of magmatic and hydrothermal processes in the TAG Area, Mid-Atlantic Ridge 26°N. Geological Society of America Bulletin, 102, 1635–1645.2.3.CO;2>CrossRefGoogle Scholar
Karson, J. A., Collins, J. A. and Casey, J. F. (1984). Geologic and seismic velocity structure of the crust/mantle transition in the Bay of Islands Ophiolite Complex. Journal of Geophysical Research, 89, 6129–6138.CrossRefGoogle Scholar
Karson, J. A., Elthon, D. L. and DeLong, S. E. (1983). Ultramafic intrusions in the Lewis Hills Massif, Bay of Islands ophiolite complex, Newfoundland: Implications for igneous processes at oceanic fracture zones. Geological Society of America Bulletin, 94, 15–29.2.0.CO;2>CrossRefGoogle Scholar
Karson, J. A., Francheteau, J., Gee, J. S.et al. (2005). Nested-scale investigation of tectonic windows into super-fast spread crust exposed at the Pito Deep Rift, Easter Microplate, SE Pacific. InterRidge News, 14, 5–8.Google Scholar
Karson, J. A., Früh-Green, G. L., Kelley, D. S.et al. (2006a). Detachment shear zone of the Atlantis Massif core complex, Mid-Atlantic Ridge, 30° N. Geochemistry, Geophysics, Geosystems, 7(6), Q06016, doi:10.1029/2005GC001109.CrossRefGoogle Scholar
Karson, J. A., Hurst, S. D. and Lonsdale, P. (1992). Tectonic rotations of dikes in fast-spread oceanic crust exposed near Hess Deep. Geology, 20, 685–688.2.3.CO;2>CrossRefGoogle Scholar
Karson, J. A., Klein, E. M., Hurst, S. D.et al. (2002a). 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, 1002, doi: 10.1029/2001GC000155.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. (2002b). 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, 107(B9), 2181, doi:10.1029/2000JB000051.CrossRefGoogle Scholar
Karson, J. A., Williams, E. A., Früh-Green, G. L.et al. (2006b). Detachment shear zone on the Atlantis Massif Core Complex, Mid-Atlantic Ridge 30°N. Geochemistry, Geophysics, Geosystems, 7, (6), doi:10.1029/2005GC001109.CrossRefGoogle Scholar
Kashefi, K. and Lovley, D. R. (2003). Extending the upper temperature limit for life. Science, 301, 934.CrossRefGoogle ScholarPubMed
Katz, R. F. and Weatherley, S. M. (2012). Consequences of mantle heterogeneity for melt extraction at mid-ocean ridges. Earth and Planetary Science Letters, 335–336, 226–237.CrossRefGoogle Scholar
Kay, M. (1951). North American Geosynclines. Memoir 48, Boulder, CO: Geological Society of America.CrossRefGoogle Scholar
Kelemen, P. B., Braun, M. and Hirth, G. (2000). Spatial distribution of melt conduits in the mantle beneath oceanic spreading ridges: Observations from the Ingalls and Oman ophiolites. Geochemistry, Geophysics, Geosystems, 1(1), 1005, doi:10.1029/1999GC000012.CrossRefGoogle Scholar
Kelemen, P. B., Hirth, G., Shimizu, N., Spiegelman, M. and Dick, H. J. B. (1997a). A review of melt migration processes in the adiabatically upwelling mantle beneath oceanic spreading ridges. Philosophical Transactions of the Royal Society of London A, 355, 283–318.CrossRefGoogle Scholar
Kelemen, P. B., Kikawa, E., Miller, D. J. and Shipboard Scientific Party (2007). Leg 209 summary: Processes in a 20-km-thick conductive boundary layer beneath the Mid-Atlantic Ridge, 14°–16°N. In Kelemen, P. B., Kikawa, E. and Miller, D. J. (eds.), Proceedings of the Ocean Drilling Program, Leg 209. College Station, TX: Ocean Drilling Program, pp. 1–33, doi:10.2973/odp.proc.sr.209.001.2007.Google Scholar
Kelemen, P. B., Koga, K. and Shimizu, N. (1997b). Geochemistry of gabbro sills in the crust/mantle transition zone of the Oman ophiolite: Implications for the origin of the oceanic lower crust. Earth and Planetary Science Letters, 146, 475–488.CrossRefGoogle Scholar
Kelemen, P. B., Quick, J. E. and Dick, H. J. B. (1992). Formation of harzburgite by pervasive melt/rock reaction in the upper mantle. Nature, 358, 635–641.CrossRefGoogle Scholar
Kelley, D. S. (1996). Methane-rich fluids in the oceanic crust. Journal of Geophysical Research, 101, 2943–2962.CrossRefGoogle Scholar
Kelley, D. S. and Delaney, J. R. (1987). Two-phase separation and fracturing in mid-ocean ridge gabbros at temperatures greater than 700 °C. Earth and Planetary Science Letters, 83, 53–66.CrossRefGoogle Scholar
Kelley, D. S. and Früh-Green, G. (1999). Abiogenic methane in deep-seated mid-ocean ridge environments: Insights from stable isotope analyses. Journal of Geophysical Research, 104, 10,439–10,460.CrossRefGoogle Scholar
Kelley, D. S. and Früh-Green, G. (2001). Volatile lines of descent in submarine plutonic environments: Insights from stable isotope and fluid inclusion analyses. Geochimica et Cosmochimica Acta, 65, 3325–3346.CrossRefGoogle Scholar
Kelley, D. S. and Shank, T. M. (2010). Hydrothermal systems: A decade of discovery in slow-spreading centers. In Rona, P. A., Devey, C. W., Dyment, J. and Murton, B. J. (eds.), Diversity of Hydrothermal Systems on Slow-Spreading Ocean Ridges. Geophysical Monograph 188, Washington, DC: American Geophysical Union, pp. 369–407, doi: 10.1029/2010GM000945.CrossRefGoogle Scholar
Kelley, D. S., Baross, J. A. and Delaney, J. R. (2002). Volcanoes, fluids, and life in submarine environments. Annual Review of Earth and Planetary Sciences, 30, 385–491.CrossRefGoogle Scholar
Kelley, D. S., Carbotte, S. M, Clague, D. A.et al. (2012). Endeavour Segment of the Juan de Fuca Ridge: One of the most remarkable places on Earth. Oceanography, 25(1), 44–61, doi.org/10.5670/oceanog.2012.03.CrossRefGoogle Scholar
Kelley, D. S., Delaney, J. R. and Juniper, S. K. (2014). Establishing a new era of submarine volcanic observatories: Cabling Axial Seamount and the Endeavour Segment of the Juan de Fuca Ridge. Marine Geology, 352, 426–450.CrossRefGoogle Scholar
Kelley, D. S., Delaney, J. R. and Yoerger, D. A. (2001a). Geology and venting characteristics of the Mothra Hydrothermal Field, Endeavour Segment, Juan de Fuca Ridge. Geology, 29, 959–962.2.0.CO;2>CrossRefGoogle Scholar
Kelley, D. S., Früh-Green, G. L. and Lilley, M. D. (2004). Volatiles in mid-ocean ridge environments: Food for life. In Wilcock, W. D., Kelley, D. S., Baross, J. A., DeLong, E. and Cary, C. (eds.), The Subseafloor Biosphere at Mid-Ocean Ridges. Geophysical Monograph 144, Washington, DC: American Geophysical Union, pp. 167–189.CrossRefGoogle Scholar
Kelley, D. S., Früh-Green, G. L., Karson, J. A. and Ludwig, K. A. (2007). Lost City hydrothermal field revisited. Oceanography, 20(4), 90–99.CrossRefGoogle Scholar
Kelley, D. S., Gillis, K. M. and Thompson, G. (1993). Fluid evolution in submarine magma-hydrothermal systems at the Mid-Atlantic Ridge. Journal of Geophysical Research, 98, 19,579–19,596.CrossRefGoogle Scholar
Kelley, D. S., Karson, J. A., Blackman, D. K.et al. (2001b). An off-axis hydrothermal field discovered near the Mid-Atlantic Ridge at 30°N. Nature, 412, 145–149.CrossRefGoogle Scholar
Kelley, D. S., Karson, J. A., Blackman, D. K.et al. (2001c). Discovery of Lost City: An off-axis, peridotite-hosted, hydrothermal field near 30°N on the Mid-Atlantic Ridge. RIDGE Events Newsletter, 11, 3–9.Google Scholar
Kelley, D. S., Karson, J. A., Früh-Green, G. L.et al. (2005). A serpentinite-hosted ecosystem: The Lost City Hydrothermal Field. Science, 307, 1428–1434.CrossRefGoogle ScholarPubMed
Kelley, D. S., Lilley, M. D., Lupton, J. E. and Olson, E. J. (1998). Enriched H2, CH4 and 3He concentrations in hydrothermal plumes associated with the 1996 Gorda Ridge Eruptive Event. Deep Sea Research II, Event Detection and Response: Gorda Ridge, 45, 2665–2682.CrossRefGoogle Scholar
Kelley, D. S., Robinson, P. T. and Malpas, J. (1992). Processes of brine generation and circulation in the oceanic crust: Fluid inclusion evidence from Troodos ophiolite, Cyprus. Journal of Geophysical Research, 97, 9307–9322.CrossRefGoogle Scholar
Kelley, D. S., Vanko, D. A. and Gu, C. (1995). Fluid evolution in oceanic crustal layer 2: Fluid inclusion evidence from the sheeted dike complex, ODP Hole 504B, Costa Rica Rift. In Stokking, L., Dick, H. J. B. and Erzinger, J. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 148. College Station, TX: Ocean Drilling Program, pp. 191–197.Google Scholar
Kellogg, J. P. (2011). Temporal and Spatial Variability of Hydrothermal Fluxes Within a Mid-Ocean Ridge Segment. Ph.D. thesis, University of Washington.Google Scholar
Kellogg, J. P. and McDuff, R. E. (2010). A hydrographic transient above the Salty Dawg hydrothermal field, Endeavour segment, Juan de Fuca Ridge. Geochemistry, Geophysics, Geosystems, 11, doi:10.1029/2010GC003299.CrossRefGoogle Scholar
Kempton, P. D. and Casey, J. F. (1997). Petrology and geochemistry of crosscutting diabase dikes, Sites 920 and 921. In Karson, J. A., Cannat, M., Miller, D. J. and Elthon, D. (eds.), Proceedings of the Ocean Drilling Program, Scientific Reports, Leg 153. College Station, TX: Ocean Drilling Program, pp. 363–377.Google Scholar
Kendall, J.-M. (1994). Teleseismic arrivals at a mid-ocean ridge: Effects of mantle melt and anisotropy. Geophysical Research Letters, 21(4), 301, doi: 10.1029/93GL02791.CrossRefGoogle Scholar
Kent, G. M. (1998). Melt to mush variations in crustal magma properties along the ridge crest at the southern East Pacific Rise. Nature, 394, 874–878.Google 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°17′N Deval from forward modeling of common depth point data. Journal of Geophysical Research, 98, 13,945–13,969.CrossRefGoogle Scholar
Kent, G. M., Harding, A. J. and Orcutt, J. A. (2000). Evidence from three-dimensional seismic reflectivity images for enhanced melt supply beneath mid-ocean-ridge discontinuities. Nature, 406, 614–618.CrossRefGoogle ScholarPubMed
Kent, G. M., Harding, A. J., Orcutt, J. A.et al. (1994). Uniform accretion of oceanic crust south of the Garrett transform at 14°15′S on the East Pacific Rise. Journal of Geophysical Research, 99(B5), 9097–9116.CrossRefGoogle Scholar
Kent, G., Harding, A., Babcock, J.et al. (2003). A new view of 3-D magma chamber structure beneath Axial Seamount and Coaxial Segment: Preliminary results from the 2002 multichannel seismic survey of the Juan de Fuca Ridge. Eos, Transactions of the American Geophysical Union, 84(46), Fall Meeting Supplement, Abstract B12A0544.Google Scholar
Key, K., Constable, S., Liu, L. and Pommier, A. (2013). Electrical image of passive mantle upwelling beneath the northern East Pacific Rise. Nature, 495, 499–502, doi:10.1038/nature11932.CrossRefGoogle ScholarPubMed
Kidd, R. G. W. (1977). A model for the process of formation of the upper oceanic crust. Geophysical Journal of the Royal Astronomical Society, 50, 149–183.CrossRefGoogle Scholar
Kingsley, C. (1863). The Water-Babies, A Fairy Tale for a Land Baby. New York: Houghton Mifflin.Google Scholar
Kinzler, R. J. and Grove, T. L. (1992). Primary magmas of mid-ocean ridge basalts, 1. Experiments and methods. Journal of Geophysical Research, 97, 6885–6906.CrossRefGoogle Scholar
Kinzler, R. and Grove, T. (1993). Corrections and further discussion of the primary magmas of mid-ocean ridge basalts, 1 and 2. Journal of Geophysical Research, 98, 22,339–22,347.CrossRefGoogle Scholar
Kivelson, M. G., Khurana, K. K., Russell, C. T.et al. (2000). Galileo magnetometer measurements: A stronger case for a subsurface ocean at Europa. Science, 289, 1340–1343.CrossRefGoogle ScholarPubMed
Klausen, M. B. and Larsen, H. C. (2002). East Greenland coast-parallel dike swarm and its role in continental breakup. In Menzies, M. A., Klemperer, S. L., Ebinger, C. J. and Baker, J. (eds.), Volcanic Rifted Margins. Special Paper 362, Boulder, CO: Geological Society of America, pp. 133–158.CrossRefGoogle Scholar
Klein, E. M., White, S., Nunnery, A.et al. (2013). Seafloor photo-geology and sonar terrain modeling at the 9°N overlapping spreading center, East Pacific Rise. Geochemistry, Geophysics, Geosystems, 14(12), 5146–5170, doi:10.1002/2013GC004858.CrossRefGoogle Scholar
Kleinrock, M. C. and Hey, R. N. (1989). Migrating transform zone and lithospheric transfer at the Galápagos 95.5°W propagator. Journal of Geophysical Research, 94, 13,859–13,878.CrossRefGoogle Scholar
Kleinrock, M. C. and Humphris, S. E. (1996a). Structural asymmetry of the TAG rift valley: Evidence from a near-bottom survey for episodic spreading. Geophysical Research Letters, 23, 3439–3442.CrossRefGoogle Scholar
Kleinrock, M. C. and Humphris, S. E. (1996b). Structural control on sea-floor hydrothermal activity at the TAG active mound. Nature, 382, 149–153.CrossRefGoogle Scholar
Klinkhammer, G., Rona, P., Gleaver, M. and Elderfield, H. (1985). Hydrothermal manganese plumes in the Mid-Atlantic Ridge valley. Nature, 314, 727–731.CrossRefGoogle Scholar
Klitgord, K., Casey, J. F., Agar, S. and ,Cruise Participants (1990). Cruise Report of Russian Mir Dives at Kings Trough (unpublished report).
Koepke, J., Christie, D. M., Dziony, W.et al. (2008). Petrography of the dike-gabbro transition at IODP Site 1256 (equatorial Pacific): The evolution of the granoblastic dikes. Geochemistry, Geophysics, Geosystems, 9, Q07O09, doi:10.1029/2008GC001939.CrossRefGoogle Scholar
Koepke, J., France, L., Müller, T.et al. (2011). Gabbros from IODP Site 1256, equatorial Pacific: Insight into axial magma chamber processes at fast spreading ocean ridges. Geochemistry, Geophysics, Geosystems, 12, doi:10.1029/2011GC003655.CrossRefGoogle Scholar
Komai, T., Giere, O. and Segonzac, M. (2007). New record of alvinocaridid shrimps (Crustacea: Decapoda: Caridea) from hydrothermal vent fields on the southern Mid-Atlantic Ridge, including a new species of the genus Opaepele. Species Diversity, 12, 237–253.CrossRefGoogle Scholar
Kong, L. S., Solomon, S. C. and Purdy, G. M. (1992). Microearthquake characteristics of a mid-ocean ridge along-axis high. Journal of Geophysical Research, 97, 1659–1685.CrossRefGoogle Scholar
Kongsrud, J. A. and Rapp, H. T. (2012). Nicomache (Loxochona) lokii sp. nov. (Annelida: Polychaeta: Maldanidae) from the Loki's Castle vent field: An important structure builder in an Arctic vent system. Polar Biology, 35, 1612–170.CrossRefGoogle Scholar
Konn, C., Charlou, J. L., Donval, J. P.et al. (2009). Hydrocarbons and oxidized organic compounds in hydrothermal fluids from Rainbow and Lost City ultramafic-hosted vents. Chemical Geology, 258, 299–314.CrossRefGoogle Scholar
Korenaga, J. and Kelemen, P. B. (1997). Origin of gabbro sills in the Moho transition zone of the Oman ophiolite: Implications for magma transport in the lower oceanic crust. Journal of Geophysical Research, 102, 27,729–27,749.CrossRefGoogle Scholar
Koschinsky, A., Billings, A., Devey, C.et al. (2006). Discovery of new hydrothermal vents on the Southern Mid-Atlantic Ridge (4°–10°S) during cruise M68/1. InterRidge News, 15, 9–15.Google Scholar
Koschinsky, A., Garbe-Schonberg, D., Sander, S.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., Stepanova, T. V.et al. (1995). Detailed geological studies of hydrothermal fields in the North Atlantic. In Parson, J., Walker, C. L. and Dixon, D. R. (eds.), Hydrothermal Vents and Processes. Special Publication 87 London: Geological Society of London, pp. 43–64.Google Scholar
Kristall, B., Kelley, D. S., Hannington, M. D. and Delaney, J. R. (2006). Growth history of a diffusely venting sulfide structure from the Juan de Fuca Ridge: A petrological and geochemical study. Geochemistry, Geophysics, Geosystems, 7(7), Q07001, doi:10.1029/2005GC001166.CrossRefGoogle Scholar
Kristall, B., Nielsen, D., Hannington, M. D., Kelley, D. S. and Delaney, J. R. (2011). Chemical microenvironments within sulfide structures from the Mothra Hydrothermal Field: Evidence from high-resolution zoning of trace elements. Chemical Geology, 290, 12–30.CrossRefGoogle Scholar
Kuhn, T., Alexander, B., Augustin, N.et al. (2004). The Logatchev hydrothermal field revisited: Preliminary results of the R/V Meteor cruise Hydromar I (M60/3). InterRidge News, 13, 1–4.Google Scholar
Kuhn, T., Herzig, P. M., Hannington, M. D., Garbe-Schönberg, D. and Stoffer, P. (2003). Origin of fluids and anhydrite precipitation in the sediment-hosted Grimsey hydrothermal field north of Iceland. Chemical Geology, 202, 5–12.CrossRefGoogle Scholar
Kurras, G. J., Fornari, D. J., Edwards, M. H., Perfit, M. R. and Smith, M. C. (2000). Volcanic morphology of the East Pacific Rise crest 9°49′–52′: Implications for volcanic emplacement processes at fast-spreading mid-ocean ridges. Marine Geophysical Researches, 21, 23–41, doi:10.1023/A:1004792202764.CrossRefGoogle Scholar
Kuzenetsov, V., Cherkashev, G., Lein, A.et al. (2006). 230Th/U dating of massive sulfides from the Logatchev and Rainbow hydrothermal fields (Mid-Atlantic Ridge). Geochronometria, 25, 51–56.Google Scholar
Kvassnes, A. J. and Grove, T. L. (2008). How partial melts of mafic lower crust affect ascending magmas at oceanic ridges. Contributions to Mineralogy and Petrology, 156(1), 49–71.CrossRefGoogle Scholar
Kyuno, A., Shintaku, M., Fujita, Y.et al. (2009). Dispersal and differentiation of deep-sea mussels of the genus Bathymodiolus (Mytilidae, Bathymodiolinae). Journal of Marine Biology, doi:10.1155/2009/625672.CrossRef
Lagabrielle, Y. and Cannat, M. (1990). Alpine Jurassic ophiolites resemble the modern central Atlantic basement. Geology, 18, 319–322.2.3.CO;2>CrossRefGoogle Scholar
Lagabrielle, Y., Bideau, D., Cannat, M., Karson, J. A. and Mével, C. (1998). Ultramafic-mafic plutonic rock suites exposed along the Mid-Atlantic Ridge (10°N–30°N): Symmetrical-asymmetrical distribution and implications for seafloor spreading processes. In Buck, W. R., Delaney, P. T., Karson, J. A. and Lagabrielle, Y. (eds.), Faulting and Magmatism at Mid-Ocean Ridges. Geophysical Monograph 106, Washington, DC: American Geophysical Union, pp. 153–176.Google Scholar
Lagabrielle, Y., Le Moigne, J., Maury, R. C., Cotton, J. and Bourgois, J. (1994). Volcanic record of the subduction of an active spreading ridge, Taito Peninsula (southern Chile). Geology, 22, 515–518.2.3.CO;2>CrossRefGoogle Scholar
Lalou, C., Brichet, J. S., Krasnov, T.et al. (1996). Initial chronology of a recently discovered hydrothermal field at 14°45′N, Mid-Atlantic Ridge. Earth and Planetary Science Letters, 144, 483–490.CrossRefGoogle Scholar
Lalou, C., Reyss, J. L. and Brichet, E. (1998). Age of sub-bottom sulfide samples at the TAG active mound. In Herzig, P. M. et al. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 158. College Station, TX: Ocean Drilling Program, pp. 111–117.Google 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, 100, 17,855–17,862.CrossRefGoogle Scholar
Lang, S. Q., Butterfield, D. A., Schulte, M., Kelley, D. S. and Lilley, M. D. (2010). Elevated concentrations of formate, acetate and dissolved organic carbon found at the Lost City Hydrothermal Field. Geochimica et Cosmochimica Acta, 74, 941–952.CrossRefGoogle Scholar
Lang, S. Q., Früh-Green, G. L., Bernasconi, S. M.et al. (2012). Microbial utilization of abiogenic carbon and hydrogen in a serpentinite-hosted system. Geochimica et Cosmochimica Acta, 92, 82–99.CrossRefGoogle Scholar
Langmuir, C. H. and Bender, J. F. (1984). The geochemistry of oceanic basalts in the vicinity of transform faults: Observations and implications. Earth and Planetary Science Letters, 69(1), 107–127.CrossRefGoogle Scholar
Langmuir, C. H. and Broecker, W. (2012). How to Build a Habitable Planet: The Story of Earth from the Big Bang to Humankind. Princeton, NJ: Princeton University Press.Google Scholar
Langmuir, C. H., Bender, J. F. and Batiza, R. (1986). Petrologic and tectonic segmentation of the East Pacific Rise, 5°30′–14°30′N. Nature, 322, 422–426.CrossRefGoogle Scholar
Langmuir, C. H., Bender, J. F., Bence, A. E., Hanson, G. N. and Taylor, S. R. (1977). Petrogenesis of basalts from the FAMOUS area: Mid-Atlantic Ridge. Earth and Planetary Science Letters, 36, 133–156.CrossRefGoogle Scholar
Langmuir, C. H., Humphris, S. E., Fornari, D. J.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, 69–91.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 Phipps Morgan, J., Blackman, D. K. and Sinton, J. (eds.), Mantle Flow and Melt Generation at Mid-Ocean Ridges. Geophysical Monograph 71, Washington, DC: American Geophysical Union, pp. 183–280.Google Scholar
Larson, B. I., Olson, E. J. and Lilley, M. D. (2009). Parameters of subsurface brines and hydrothermal processes 12–15 months after the 1999 magmatic event at the Main Endeavor Field as inferred from in situ time series measurements of chloride and temperature. Journal of Geophysical Research, 114, doi:10.1029/2008JB005627.Google Scholar
Larson, R. L., Popham, C. T. and Pockalny, R. A. (2005). Lithologic and structural observations at Endeavor Deep and their implications for the accretion process at fast to ultra-fast spreading rates. Eos, Transactions of the America Geophysical Union, 85(52), T33D–5094.Google Scholar
Laukert, G., Von Der Handt, A., Hellebrand, E.et al. (2013). High-pressure reactive melt stagnation recorded in abyssal pyroxenites from the ultraslow-spreading Lena Trough, Arctic Ocean. Journal of Petrology, 55, 427–458.CrossRefGoogle Scholar
Lavier, L., Buck, W. R. and Poliakov, A. N. B. (1999). Self-consistent rolling-hinge model for the evolution of large-offset low-angle faults. Geology, 27, 1127–1130.2.3.CO;2>CrossRefGoogle Scholar
Lawrence, R. M., Karson, J. A. and Hurst, S. D. (1998). Dike orientations, fault-block rotations, and the construction of slow spreading oceanic crust at 22°40′N on the Mid-Atlantic Ridge. Journal of Geophysical Research, 103, 663–676.CrossRefGoogle Scholar
Lawver, L. A. and Williams, D. L. (1979). Heat flow in the central Gulf of California. Journal of Geophysical Research, 84, doi:10.1029/0JGREA0000840000B7003465000001.CrossRefGoogle Scholar
le Roux, P. J., Shirey, S. B., Hauri, E. H., Perfit, M. R. and Bender, J. F. (2006). The effects of variable sources, processes and contaminants on the composition of northern EPR MORB (8–10°N and 12–14°N): Evidence from volatiles (H2O, CO2, S) and halogens (F, Cl). Earth and Planetary Science Letters, 251, 209–231.CrossRefGoogle Scholar
Lein, A. Y., Peresypking, V. I. and Simoneit, B. R. T. (2003). Origin of hydrocarbons in hydrothermal sulfide ores in the Mid-Atlantic Ridge. Lithology and Mineral Resources, 38, 383–393.CrossRefGoogle Scholar
Leitch, E. C. (1984). Island arc elements and arc-related ophiolites. Tectonophysics, 106, 177–203.CrossRefGoogle Scholar
Lilley, M. D., Baross, J. A. and Gordon, L. I. (1983). Reduced gases and bacteria in hydrothermal fluids: The Galápagos spreading center and 21°N East Pacific Rise. In Rona, P. A., Widenfalk, L. and Bostrom, K. (eds.), Hydrothermal Processes at Seafloor Spreading Centers. New York: Plenum, pp. 411–447.CrossRefGoogle Scholar
Lilley, M. D., Butterfield, D. A., Lupton, J. E. and Olson, E. J. (2003). Magmatic events can produce rapid changes in hydrothermal vent chemistry. Nature, 422, 878–881.CrossRefGoogle ScholarPubMed
Lilley, M. D., Butterfield, D. A., Olsen, E. J.et al. (1993). Anomalous CH4 and NH4+ concentrations at an unsedimented mid-ocean ridge hydrothermal system. Nature, 364, 45–47.CrossRefGoogle Scholar
Lilley, M. D., Landsteiner, M. C., McLaughlin, E. A.et al. (1995). Real-time mapping of hydrothermal plumes on the Endeavour Segment of the Juan de Fuca. Eos, Transactions of the American Geophysical Union, 76, 420.Google Scholar
Lin, J. and Phipps Morgan, J. (1992). The spreading rate dependence of three-dimensional mid-ocean ridge gravity structure. Geophysical Research Letters, 19, 13–16.CrossRefGoogle Scholar
Lissenberg, C. J. and Dick, H. J. B. (2008). Melt–rock reaction in the lower oceanic crust and its implications for the genesis of mid-ocean ridge basalt. Earth and Planetary Science Letters, 271(1), 311–325.CrossRefGoogle Scholar
Lissenberg, C. J., MacLeod, C. J., Howard, K. A. and Goddard, M. (2013). Pervasive reactive melt migration through fast-spreading lower oceanic crust (Hess Deep, equatorial Pacific Ocean). Earth and Planetary Science Letters, 361, 436–447.CrossRefGoogle Scholar
Lister, C. R. B. (1974). On the penetration of water into hot rock. Geophysical Journal of the Royal Astronomical Society, 39, 465–509.CrossRefGoogle Scholar
Lister, C. R. B. (1983). The basic physics of water penetration into hot rock. In Rona, P. A., Widenfalk, L. and Bostrom, K. (eds.), Hydrothermal Processes at Seafloor Spreading Centers. New York: Plenum, pp. 141–176.CrossRefGoogle Scholar
Liu, C. Z., Snow, J. E., Hellebrand, E.et al. (2008). Ancient, highly heterogeneous mantle beneath Gakkel Ridge, Arctic Ocean. Nature, 452(7185), 311–316.CrossRefGoogle ScholarPubMed
Lizarralde, D., Soule, S. A., Seewald, J. S. and Proskurowski, G. (2011). Carbon release by off-axis magmatism in a young sedimented spreading centre. Nature Geoscience, 4, doi:10.1038/NGEO1006.CrossRefGoogle Scholar
Lonsdale, P. (1977a). Abyssal pahoehoe with lava coils at the Galápagos rift. Geology, 5, 147–152.2.0.CO;2>CrossRefGoogle Scholar
Lonsdale, P. (1977b). Clustering of suspension-feeding macrobenthos near abyssal hydrothermal vents at oceanic spreading centers. Deep-Sea Research, 24, 857–863.CrossRefGoogle Scholar
Lonsdale, P. (1977c). Structural geomorphology of a fast-spreading rise crest: The East Pacific Rise crest at 3°25′S. Marine Geophysical Researches, 3, 251–293.CrossRefGoogle Scholar
Lonsdale, P. (1985). Linear volcanoes along the Pacific-Cocos plate boundary, 9°N to the Galápagos triple junction. Tectonophysics, 116, 255–279.CrossRefGoogle Scholar
Lonsdale, P. and Spiess, F. N. (1980). Deep-tow observations at the East Pacific Rise, 8°45′N, and some interpretations. Initial Reports of the Deep Sea Drilling Project, Leg 54. Washington, DC: US Government Printing Office, pp. 43–62.Google Scholar
Lonsdale, P., Bischoff, J. L., Burns, V. M., Kastner, M. and Sweeney, R. E. (1980). A high temperature hydrothermal deposit on the sea bed at a Gulf of Mexico spreading center. Earth and Planetary Science Letters, 44, 463–472.Google Scholar
Lourenço, N., Miranda, J. M., Luis, J. F.et al. (1998). Morpho-tectonic analysis of the Azores volcanic plateau from a new bathymetric compilation of the area. Marine Geophysical Researches, 20, 141–156.CrossRefGoogle Scholar
Love, B. A., Resing, J. A., Cowen, J. P.et al. (2008). Methane, manganese, and helium in hydrothermal plumes following volcanic eruptions on the East Pacific Rise near 9°50′N. Geochemistry, Geophysics, Geosystems, 9(6), Q06T01, doi:10.1029/2008GC002104.CrossRefGoogle Scholar
Lowell, J. D. and Genik, G. J. (1972). Sea-floor spreading and structural evolution of southern Red Sea. American Association of Petroleum Geologists Bulletin, 56, 247–259.Google Scholar
Lowell, R. P., Rona, P. A. and Von Herzen, R. P. (1995). Seafloor hydrothermal systems. Journal of Geophysical Research, 100, 327–352.CrossRefGoogle Scholar
Ludwig, K. A., Kelley, D. S., Butterfield, D. A., Nelson, B. and Früh-Green, G. (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 Lawrence, R. L. (2011). U-Th systematics and 230Th ages of carbonate chimneys at the Lost City Hydrothermal Field. Geochimica et Cosmochimica Acta, 75, 1869–1888.CrossRefGoogle Scholar
Lupton, J. E. (1979). Helium3 in the Guaymas basin: Evidence for injection of mantle volatiles in the Gulf of California. Journal of Geophysical Research, 84, 7446–7452, doi:10.1029/JB084iB13p07446.CrossRefGoogle Scholar
Lupton, J. E. (1995). Hydrothermal plumes: Near and far field. In Humphris, S., Zierenberg, R., Mullineau, L. and Thomson, R. (eds.), Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions. Geophysical Monograph 91, Washington, DC: American Geophysical Union, pp. 317–346.Google Scholar
Lupton, J. E., Baker, E. T. and Massoth, G. J. (1999). Helium, heat, and the generation of hydrothermal event plumes at mid-ocean ridges. Earth and Planetary Science Letters, 171, 343–350.CrossRefGoogle Scholar
Lupton, J. E., Baker, E. T. and Massoth, G. J. (2000). Reply to comment by M. R. Palmer and G. G. J. Ernst on “Helium, heat, and the generation of hydrothermal event plumes at mid-ocean ridges” by J. E. Lupton, E. T. Baker, and G. J. Massoth. Earth and Planetary Science Letters, 180, 219–222.CrossRefGoogle Scholar
Lupton, J. E., Butterfield, D. A., Lilley, M. D.et al. (2006). Young, submarine venting of liquid carbon dioxide on a Mariana arc volcano. Geochemistry, Geophysics, Geosystems, 7(8), Q08007, doi:10.1029/2005GC001152.CrossRefGoogle Scholar
Luther, G. W., Rozan, T. F., Taillefert, M.et al. (2001). Chemical speciation drives hydrothermal vent ecology. Nature, 410, 813–816.CrossRefGoogle ScholarPubMed
Lyell, C.. (1871). Student's Elements of Geology, London: John Murray.
Macdonald, K. C. (1982). Mid-ocean ridges: Fine scale tectonic, volcanic and hydrothermal processes within the plate boundary zone. Annual Review of Earth and Planetary Sciences, 10, 155–190.CrossRefGoogle Scholar
Macdonald, K. C. (1998a). Exploring the global Mid-Ocean Ridge: A quarter century of discovery. Oceanus, 41, (1).Google Scholar
Macdonald, K. C. (1998b). Linkages between faulting, volcanism, hydrothermal activity and segmentation on fast spreading centers. In Buck, W. R., Delaney, P. T., Karson, J. A. and Lagabrielle, Y. (eds.), Faulting and Magmatism at Mid-Ocean Ridges, Geophysical Monograph 106. Washington, DC: American Geophysical Union, pp. 27–58.Google Scholar
Macdonald, K. C. and Fox, P. J. (1983). Overlapping spreading centers: 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. and Luyendyk, B. P. (1977). Deep-tow studies of the structure of the Mid-Atlantic Ridge crest near 37° N (FAMOUS). Geological Society of America Bulletin, 88, 621–636.2.0.CO;2>CrossRefGoogle Scholar
Macdonald, K. C. and Luyendyk, B. P. (1985). Investigation of faulting and abyssal hill formation on the flanks of the East Pacific Rise (21°N) using Alvin. Marine Geophysical Researches, 7, 515–535.CrossRefGoogle Scholar
Macdonald, K. C., Fox, P. J., Alexander, R. T., Pockalny, R. and Gente, P. (1996). Seafloor topography: A faulted or volcanic origin?Nature, 380, 125–129.CrossRefGoogle Scholar
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, 299–344.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., Luyendyk, B. P., Mudie, J. D. and Spiess, F. N. (1975). Near-bottom geophysical study of the Mid-Atlantic Ridge median valley near lat. 37°N: Preliminary observations. Geology, 3, 211–215.2.0.CO;2>CrossRefGoogle Scholar
Maclennan, J., Hulme, T. and Singh, S. C. (2004). Thermal models of oceanic crustal accretion: Linking geophysical, geological and petrological observations. Geochemistry, Geophysics, Geosystems, 5, doi:10.1029/2003GC000605.CrossRefGoogle Scholar
Maclennan, J., Hulme, T. and Singh, S. C. (2005). Cooling the lower oceanic crust. Geology, 5, 357–360, doi:10.1130/G21207.CrossRefGoogle 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, 357–373.CrossRefGoogle Scholar
MacLeod, C. J., Escartín, J., Banerjee, D.et al. (2002). Direct geological evidence for oceanic detachment faulting: The Mid-Atlantic Ridge, 15°45′N. Geology, 30(10), 879–882, doi: 10.1130/0091-7613(2002)030087.2.0.CO;2>CrossRefGoogle Scholar
MacLeod, C. J., Lissenberg, C. J. and Bibby, L. E. (2013). “Moist MORB” axial magmatism in the Oman ophiolite: The evidence against a mid-ocean ridge origin. Geology, 41(4), 459–462, doi:10:1130/G33904.1.CrossRefGoogle Scholar
MacLeod, C. J., Searle, R. C., Murton, B. J.et al. (2009). Life cycle of oceanic core complexes. Earth and Planetary Science Letters, 287(3–4), 333–344.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
Malinverno, A. and Cowie, P. (1993). Normal faulting and the topographic roughness of mid-ocean ridge flanks. Journal of Geophysical Research, 98, 17,921–17,940.CrossRefGoogle Scholar
Malinverno, A., Hildebrand, J., Tominaga, M. and Channell, J. E. T. (2012). M-sequence geomagnetic polarity time scale (MHTC12) that steadies global spreading rates and incorporates astrochronology constraints. Journal of Geophysical Research, 117, (B06104), doi:10.1029/2012JB009260.CrossRefGoogle Scholar
Malpas, J., Calon, T. and Xenophontos, C. (1987). Plutonic rocks of the Troodos Ophiolite. In Field Excursion Guidebook, Troodos'87 Symposium, Ophiolites and Oceanic Lithosphere, Nicosia, Cyprus: Geological Survey Department, pp. 158–181.Google Scholar
Manning, C. E., MacLeod, C. J. and Weston, P. E. (2000). Lower crustal cracking front at fast spreading ridges: Evidence from the East Pacific Rise and Oman ophiolite. In Dilek, Y., Moores, E., Elthon, D. and Nicolas, A. (eds.), Ophiolites and Oceanic Crust: New Insights from Field Studies and the Ocean Drilling Program. Special Paper 349, Boulder, CO: Geological Society of America, pp. 261–272.CrossRefGoogle Scholar
Manning, C. E., Weston, P. E. and Mahon, K. I. (1996). Rapid high temperature metamorphism of the East Pacific Rise gabbros from Hess Deep. Earth and Planetary Science Letters, 144, 123–132.CrossRefGoogle Scholar
Marcon, Y., Sahling, H., Borowski, C.et al. (2013). Megafaunal distribution and assessment of total methane and sulfide consumption by mussel beds at Menez Gwen hydrothermal vent, based on geo-referenced photomosaics. Deep-Sea Research Part 1, 175, 93–109.CrossRefGoogle Scholar
Marcus, J., Tunnicliffe, V. and Butterfield, D. A. (2009). Post-eruptive succession of macrofunal communities at diffuse flow hydrothermal vents on Axial Volcano, Juan de Fuca Ridge, Northeast Pacific. Deep-Sea Research Part II, 56, 1586–1598.CrossRefGoogle Scholar
Marques, A. F. A., Barriga, F., Chavagnac, V. and Fouquet, Y. (2006). Mineralogy, geochemistry, and Nd isotope composition of the Rainbow hydrothermal field, Mid-Atlantic Ridge. Mineralium Deposita, 41, 52–67.CrossRefGoogle Scholar
Marques, A. F. A., Barriga, F. J. A. S. and Scott, S. D. (2007). Sulfide mineralization in an ultramafic-rock hosted seafloor hydrothermal system: From serpentinization to the formation of Cu-Zn-(Co)-rich massive sulfides. Marine Geology, 245, 20–39.CrossRefGoogle Scholar
Marques, A. F. A., Scott, S. D., Gorton, M. P., Barriga, F. J. A. S. and Fouquet, Y. (2009). Pre-eruption history of enriched MORB from the Menez Gwen (37°50′N) and Lucky Strike (37°17′N) hydrothermal systems, Mid-Atlantic Ridge. Lithos, 112, 18–39.CrossRefGoogle Scholar
Martin, W., Baross, J. A., Kelley, D. S. and Russel, M. J. (2008). Hydrothermal vents and the origin of life. Nature Reviews Microbiology, 6, 805–814, doi:10.1038/nrmicro1991, 1–10.CrossRefGoogle ScholarPubMed
Martins, I., Colaco, A., Dando, P. R.et al. (2008). Size-dependent variations on the nutritional pathway of Bathymodiolus azoricus demonstrated by a C-flux model. Ecological Modeling, 217, 59–71.CrossRefGoogle Scholar
McBirney, A. R. (1975). Differentiation of the Skaergaard intrusion. Nature, 253, 691–694.CrossRefGoogle Scholar
McCaig, A. M., Cliff, R. A.Escartín, J.Fallick, A. E. and MacLeod, C. J. (2007). Oceanic detachment faults focus very large volumes of black smoker fluids. Geology, 35, 935–938.CrossRefGoogle Scholar
McCaig, A. M., Delacour, A.Fallick, A. E.Castelain, T. and Früh-Green, G. L. (2010). Detachment fault control on hydrothermal circulation systems: Interpreting the subsurface beneath the TAG Hydrothermal Field using the isotopic and geological evolution of oceanic core complexes in the Atlantic. In Rona, P. A., Devey, C. W., Dyment, J., Murton, B. J. (eds.), Diversity of Hydrothermal Systems on Slow Spreading Ocean Ridges. Geophysical Monograph 108, Washington, DC: American Geophysical Union, pp. 207–240.CrossRefGoogle Scholar
McCarthy, J., Mutter, J. C., Morton, J. L., Sleep, N. H. and Thompson, G. A. (1988). Relict magma chamber structures preserved within the Mesozoic North Atlantic crust?Geological Society of America Bulletin, 100, 1423–1436.2.3.CO;2>CrossRefGoogle Scholar
McClinton, T., White, S. M., Colman, A. and Sinton, J. M. (2013). Reconstructing lava flow emplacement processes at the hotspot-affected Galápagos Spreading Center, 95°W and 92°W. Geochemistry, Geophysics, Geosystems, 14, 2731–2756, doi 10.1002/ggge.20157.CrossRefGoogle Scholar
McCollom, T. M. (2000). Geochemical constraints on primary productivity in submarine hydrothermal plumes. Deep-Sea Research Part 1: Oceanographic Research Papers, 47, 85–101.CrossRefGoogle Scholar
McCollom, T. M. (2007a). Geochemical constraints on sources of metabolic energy for chemolithoautotrophy in ultramafic-hosted deep-sea hydrothermal systems. Astrobiology, 7, 933–950.CrossRefGoogle ScholarPubMed
McCollom, T. M. (2007b). Observational, experimental, and theoretical constraints on carbon cycling in mid-ocean ridge hydrothermal systems. In Lowell, R. P., Seewald, J. S., Metaxas, A. and Perfit, M. R. (eds.), Magma to Microbes: Modeling Hydrothermal Processes at Ocean Spreading Centers. Geophysical Monograph 178, Washington, DC: American Geophysical Union, pp. 193–213.Google Scholar
McCollom, T. M. and Bach, W. (2009). Thermodynamic constraints on hydrogen generation during serpentinization of ultramafic rocks. Geochimica et Cosmochimica Acta, 73, 856–875.CrossRefGoogle Scholar
McCollom, T. M. and Seewald, J. S. (2001). A reassessment of the potential for reduction of dissolved CO2 to hydrocarbons during serpentinization. Geochimica et Cosmochimica Acta, 65, 3769–3778.CrossRefGoogle Scholar
McCollom, T. M. and Seewald, J. S. (2003). Experimental constraints on the hydrothermal reactivity of organic acids and acid anions: I. Formic acid and formate. Geochimica et Cosmochimica Acta, 67, 3625–3644.CrossRefGoogle Scholar
McCollom, T. M. and Seewald, J. S. (2006). Carbon isotope composition of organic compounds produced by abiotic synthesis under hydrothermal conditions. Earth and Planetary Science Letters, 243, 74–84.CrossRefGoogle Scholar
McKay, L. J., MacGregor, B. J., Biddle, J. F.et al. (2012). Spatial heterogeneity and underlying geochemistry of phylogenetically diverse orange and white Beggiatoa mats in Guaymas Basin hydrothermal sediments. Deep-Sea Research Part 1, 67, 21–31.CrossRefGoogle Scholar
McKenzie, D. (1984). The generation and compaction of partially molten rock. Journal of Petrology, 25, 713–765.CrossRefGoogle Scholar
McKenzie, D. 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. and O'Nions, R. K. (1991). Partial melt distributions from inversion of rare earth element concentrations. Journal of Petrology, 32, 1021–1091.CrossRefGoogle Scholar
Melson, W. G., Hart, S. R. and Thompson, G. (1972). St. Paul's Rocks, Equatorial Atlantic: Petrogenesis, radiometric ages, and implications for sea-floor spreading. In Shagan, R. (ed.), Studies in Earth and Space Sciences: The Hess Volume. Boulder, CO: Geological Society of America, pp. 241–272.CrossRefGoogle Scholar
Melson, W. G., Thompson, G. and Van Andel, Tj. H. (1968). Volcanism and metamorphism in the mid-Atlantic Ridge, 22°N latitude. Journal of Geophysical Research, 73(18), 5925–5941, doi:10.1029/JB073i018p05925.CrossRefGoogle Scholar
,Melt Seismic Team (1998). Imaging deep seismic structure beneath the mid-ocean ridge: The MELT experiment. Science, 280, 1215–1218.CrossRefGoogle Scholar
Menard, H. W. (1964). Marine Geology of the Pacific. New York: McGraw Hill.Google Scholar
Menard, H. W. (1967). Extension of Northeastern-Pacific fracture zones. Science, 155, 72–74.CrossRefGoogle ScholarPubMed
Menard, H. W. and Atwater, T. (1969). Origin of fracture zone topography. Nature, 222, 1037–1040.CrossRefGoogle Scholar
Meurer, W. P. and Boudreau, A. E. (1998a). Compaction of igneous cumulates part I: Geochemical consequences for cumulates and liquid fractionation trends. Journal of Geology, 106(3), 281–292.CrossRefGoogle Scholar
Meurer, W. P. and Boudreau, A. E. (1998b). Compaction of igneous cumulates part II: Compaction and the development of igneous foliations. Journal of Geology, 106(3), 293–304.CrossRefGoogle Scholar
Mével, C., Cannat, M., Gente, P.et al. (1991). Emplacement of deep crustal and mantle rocks on the West Median Valley Wall of the MARK Area (M.A.R. 23°N). Tectonophysics, 190, 31–53.CrossRefGoogle Scholar
Mével, G., Faidy, C. and Prieur, D. (1996). Distribution, activity, and diversity of heterotrophic nitrifiers originating from East Pacific deep-sea hydrothermal vents. Canadian Journal of Microbiology, 42, 162–171.CrossRefGoogle Scholar
Meyer, J. L., Akerman, N. H., Proskurowski, G. and Huber, J. A. (2013). Microbial characterization of post-eruption “snowblower” vents at Axial Seamount, Juan de Fuca Ridge. Frontiers in Microbiology, 4, doi:10.3389/frmicb.2013.00153.CrossRefGoogle ScholarPubMed
Michael, P. J. (1995). Regionally distinctive sources of depleted MORB: Evidence from trace elements and H2O. Earth and Planetary Science Letters, 131(3), 301–320.CrossRefGoogle Scholar
Michael, P. J. and Bonatti, E. (1985). Peridotite compositions from the North Atlantic: Regional and tectonic variations and implications for partial melting. Earth and Planetary Science Letters, 73, 91–104.CrossRefGoogle Scholar
Michael, P. J. and Cornell, W. C. (1998). Influence of spreading rate and magma supply on crystallization and assimilation beneath mid-ocean ridges: Evidence from chlorine and major element chemistry of mid-ocean ridge basalts. Journal of Geophysical Research, 103(B8), 18,325–18,356.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, 956–961, doi:10:1038/nature01704.CrossRefGoogle ScholarPubMed
Michael, P. J., Perfit, M. R., Palke, A., Fornari, D. J. and Soule, S. A. (2008). Observations of CO2 exsolution in submarine lava flows: The 2005–06 eruption on the East Pacific Rise, 9°50′N. Eos, Transactions of the American Geophysical Union, 89, V21B–2106.Google Scholar
Mittelstaedt, E., Escartín, J., Gracias, N.et al. (2012). Quantifying diffuse and discrete venting at the Tour Eiffel vent site, Lucky Strike hydrothermal field. Geochemistry, Geophysics, Geosystems, 13(4), Q04008, doi:10.1029/2011GC003991.CrossRefGoogle Scholar
Miyashiro, A. (1972). Metamorphism and related magmatism in plate tectonics. American Journal of Science, 272, 629–656.CrossRefGoogle Scholar
Miyashiro, A. (1973a). Metamorphism and Metamorphic Belts. New York: John Wiley and Sons.CrossRefGoogle Scholar
Miyashiro, A. (1973b). The Troodos ophiolite complex was probably formed in a volcanic arc. Earth and Planetary Science Letters, 19, 218–224.CrossRefGoogle Scholar
Miyashiro, A. (1975a). Classification, characteristics, and origin of ophiolites. Journal of Geology, 83, 249–281.CrossRefGoogle Scholar
Miyashiro, A. (1975b). Discussion of “Origin of the Troodos and other ophiolites: A reply to Hynes”. Earth and Planetary Science Letters, 25, 217–222.CrossRefGoogle Scholar
Miyashiro, A. (1975c). Origin of the Troodos and other ophiolites: A reply to Moores. Earth and Planetary Science Letters, 25, 227–235.CrossRefGoogle Scholar
Miyashiro, A., Shido, F. and Ewing, M. (1969). Composition and origin of serpentinites from the Mid-Atlantic Ridge near 24° and 30°N lat. Contributions to Mineralogy and Petrology, 23, 117–127.CrossRefGoogle Scholar
Miyashiro, A., Shido, F. and Ewing, M. A. (1971). Metamorphism in the Mid-Atlantic Ridge near 24°N and 30°N. Philosophical Transactions of the Royal Society of London, Series A, 268, 589–603.CrossRefGoogle Scholar
Moore, J. G. (1975). Mechanism of formation of pillow lava. American Scientist, 63, 269–277.Google Scholar
Moore, J. G. and Fabbi, B. P. (1971). An estimate of the juvenile sulfur content of basalt. Contributions to Mineralogy and Petrology, 33, 118–127.CrossRefGoogle Scholar
Moores, E. M. (1969). Petrology and Structure of the Vourinos Ophiolite Complex, Northern Greece. Special Paper 118, Boulder, CO: Geological Society of America, pp. 3–66.CrossRefGoogle Scholar
Moores, E. M. (1975). Discussion of “Origin of the Troodos and other ophiolites: A reply to Hynes” by Akiho Miyashiro. Earth and Planetary Science Letters, 25, 223–226.CrossRefGoogle Scholar
Moores, E. M. (1982). Origin and emplacement of ophiolites. Reviews of Geophysics and Space Physics, 20, 735–760.CrossRefGoogle Scholar
Moores, E. M. and Vine, F. J. (1971). The Troodos massif, Cyprus, and other ophiolites as oceanic crust: Evaluation and implications. Philosophical Transactions of the Royal Society of London A, 268, 443–466.CrossRefGoogle Scholar
Moores, E. M., Kellogg, L. H. and Dilek, Y. (2000). Tethyan ophiolites, mantle convection, and tectonic “historical contingency:” A resolution of the “Ophiolite Conundrum”. In Dilek, Y., Moores, E. M., Elthon, D. and Nicolas, A. (eds.), Ophiolites and Oceanic Crust: New Insights from Field Studies and the Ocean Drilling Program. Boulder, CO: Geological Society of America, Special Paper 349, pp. 3–12.CrossRefGoogle Scholar
Morishita, T., Hara, K., Nakamura, K.et al. (2009). Igneous, alteration and exhumation processes recorded in abyssal peridotites and related fault rocks from an oceanic core complex along the Central Indian Ridge. Journal of Petrology, 50(7), 1299–1325.CrossRefGoogle 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, 287, 217–228.CrossRefGoogle Scholar
Morris, E., Detrick, R. S., Minshull, T. A.et al. (1993). Seismic structure of oceanic crust in the Western North Atlantic. Journal of Geophysical Research, 98, 13879–13903.CrossRefGoogle Scholar
Mottl, M. J. (1983). Metabasalts, axial hot springs, and the structure of hydrothermal systems at mid-ocean ridges. Geological Society of America Bulletin, 94, 161–180.2.0.CO;2>CrossRefGoogle Scholar
Mottl, M. J., Komor, S. C., Fryer, P. and Moyer, C. L. (2003). Deep-slab fluids fuel extremophilic Archaea on a Mariana forearc serpentinite mud volcano: Ocean Drilling Program Leg 195. Geochemistry, Geophysics, Geosystems, 4(11), doi:10/10.1029/2003GC000588.CrossRefGoogle Scholar
Mount, V. S. and Suppe, J. (1987). State of stress near the San Andreas fault: Implications for wrench tectonics. Geology, 15, 1143–1146.2.0.CO;2>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, Q04006, doi.org/10.1029/2007GC001743.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, 7829–7834.CrossRefGoogle ScholarPubMed
Mullineaux, L. S., Mills, S. W., Sweetman, A. K.et al. (2005). Spatial structure and temporal variation in larval abundance at hydrothermal vents on the East Pacific Rise. Marine Ecology Progress Series, 293, 1–16.CrossRefGoogle Scholar
Mullineaux, L. S., Peterson, C. H., Micheli, F. and Mills, S. W. (2003). Successional mechanism varies along a gradient in hydrothermal fluid flux at deep-sea vents. Ecological Monographs, 73, 523–542.CrossRefGoogle Scholar
Müntener, O. (2010). Serpentine and serpentinization: A link between planet formation and life. Geology, 38, 959–960.CrossRefGoogle Scholar
Murray, J. A. and Renard, F. (1891). Report on Deep-Sea Deposits, Based on Specimens Collected During the Voyage of H.M.S. Challenger in the Years 1872–76. London: HM Stationery Office.Google Scholar
Murton, B. J., Baker, E.T., Sands, C.M., and German, C.R., (2006) Detection of an unusually large hydrothermal event plume above the slow-spreading Carlsberg Ridge: NW Indian Ocean. Geophysical Research Letters, doi:10.1029/2006GL026048.CrossRef
Mutter, C. Z. and Mutter, J. C. (1993). Variation in thickness of layer 3 dominates oceanic crustal structure. Earth and Planetary Science Letters, 117, 295–317.CrossRefGoogle Scholar
Mutter, J. C. and Carton, H. D. (2013). The Mohorovičić discontinuity in ocean basins: Some observations from seismic data. Tectonophysics, 609, 314–330.CrossRefGoogle Scholar
Mutter, J. C. and Karson, J. A. (1992). Structural processes at slow-spreading ridges. Science, 257, 627–634.CrossRefGoogle ScholarPubMed
Mutter, J. C., Carbotte, S. M., Nedimović, M., Canales, J. P. and Carton, H. (2009). Seismic imaging in three dimensions on the East Pacific Rise. Eos, Transactions of the American Geophysical Union, 90(42), 374–375.CrossRefGoogle Scholar
Naidoo, D. D. (1998). Accretion of the Upper Oceanic Crust. Ph.D. dissertation, University of Washington.Google Scholar
Nakamura, K., Morishita, T., Bach, W.et al. (2009). Serpentinized troctolites exposed near the Kairei hydrothermal field, Central Indian Ridge: Insights into the origin of the Kairei hydrothermal fluid supporting a unique microbial ecosystem. Earth and Planetary Science Letters, 280, 128–136.CrossRefGoogle Scholar
,National Research Council (1999). From monsoons to microbes: Understanding the ocean's role in human health. In A Review of the Ocean Research Priorities Plan and Implementation Strategy, Division of Earth and Life Studies, National Research Council, Ocean Studies Board, National Academies Press.Google Scholar
Natland, J. H. (1989). Partial melting of a lithologically heterogeneous mantle: Inferences from crystallization histories of magnesian abyssal tholeiites from the Siqueiros Fracture Zone. In Saunders, A. D. and Norry, M. J. (eds.), Magmatism in the Ocean Basins. Special Publication 42, London:Geological Society of London, pp. 41–70.Google Scholar
Natland, J. H. and Dick, H. J. B. (1996). Melt migration through high-level gabbro cumulates of the East Pacific Rise at Hess Deep: The origin of magma lenses and the deep crustal structure of fast-spreading ridges. In Mével, C., Gillis, K. M. and Allen, J. F. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 148. College Station, TX: Ocean Drilling Program, pp. 21–58.Google Scholar
Natland, J. H. and Dick, H. J. B. (2001). Formation of the lower ocean crust and the crystallization of gabbroic cumulates at a very slowly spreading ridge. Journal of Volcanology and Geothermal Research, 110(3–4), 191–233.CrossRefGoogle Scholar
Natland, J. H. and Dick, H. J. B. (2009). Paired melt lenses at the East Pacific Rise and the pattern of melt flow through the gabbroic layer at a fast-spreading ridge. Lithos, 112(1), 73–86.CrossRefGoogle Scholar
Neal, C. and Stranger, G. (1983). Hydrogen generation from mantle source rocks in Oman. Earth and Planetary Science Letters, 66, 315–320.CrossRefGoogle Scholar
Nehlig, P. and Juteau, T. (1988). Flow porosities, permeabilities and preliminary data on fluid inclusions and fossil thermal gradients in the crustal sequence of the Sumail ophiolite (Oman). Tectonophysics, 151, 199–221.CrossRefGoogle Scholar
Nehlig, P., Juteau, T., Bendel, T. and Cotton, J. (1994). The root zones of oceanic hydrothermal systems: Constraints from the Samail ophiolite (Oman). Journal of Geophysical Research, 99, 4703–4713.CrossRefGoogle Scholar
Nelson, D., Haymon, R. M. and Lilley, M. (1991). Rapid growth of unusual hydrothermal bacteria observed at new vents during the ADVENTURE dive program to the EPR at 9°45′–52′N. Eos, Transactions of the American Geophysical Union, 72, 481.Google Scholar
Neo, N., Yamazaki, S. and Miyashita, S. (2009). Data report: Bulk rock compositions of samples from the IODP Expedition 309/312 sample pool, ODP Hole 1256D. In Teagle, D.A.H., Alt, J.C., Umino, S.et al. (eds.), Proceedings of the Integrated Drilling Program, Scientific Results, Leg 312. College Station, TX: Integrated Ocean Drilling Program, pp. 309–312.Google Scholar
Nicholson, R. and Pollard, D. D. (1985). Dilation and linkage of echelon cracks. Journal of Structural Geology, 20, 53–72.Google Scholar
Nicolas, A. (1989). Structure of Ophiolites and Dynamics of Oceanic Lithosphere. Dordrecht: Kluwer Academic Press.CrossRefGoogle Scholar
Nicolas, A. and Boudier, F. (2000). Large mantle upwellings and related variations in crustal thickness in the Oman ophiolite. In Dilek, Y., Moores, E., Elthon, D. and Nicolas, A. (eds.), Ophiolites and Oceanic Crust: New Insights from Field Studies and the Ocean Drilling Program. Special Paper 349. Boulder, CO: Geological Society of America, pp. 67–73.CrossRefGoogle Scholar
Nicolas, A., Boudier, F. and Ceuleneer, G. (1988). Mantle flow patterns and magma chambers at ocean ridges: Evidence from Oman ophiolite. Marine Geophysical Researches, 9, 293–310.CrossRefGoogle Scholar
Nicolas, A., Boudier, F. and France, L. (2009). Subsidence in magma chamber and development of magmatic foliation in Oman ophiolite gabbros. Earth and Planetary Science Letters, 284, 76–87.CrossRefGoogle Scholar
Nicolas, A., Boudier, F. and Meshi, A. (1999). Slow spreading accretion and mantle denudation in the Mirdita ophiolite (Albania). Journal of Geophysical Research, 104, 15,155–15,167.CrossRefGoogle Scholar
Niu, Y. (2004). Bulk-rock major and trace element compositions of abyssal peridotites: Implications for mantle melting, melt extraction and post-melting processes beneath mid-ocean ridges. Journal of Petrology, 45(12), doi:10.1093/petrology/egh068.CrossRefGoogle Scholar
Niu, Y. and Hékinian, R. (1997). Basaltic liquids and harzburgitic residues in the Garrett Transform: a case study at fast-spreading ridges. Earth and Planetary Science Letters, 146(1–2), 243–258.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(2), Q02002, doi:1029/2008GC002315.CrossRefGoogle Scholar
Norton, D. (1984). Theory of hydrothermal systems. Annual Review of Earth and Planetary Sciences, 12, 155–157.CrossRefGoogle Scholar
Nygaard, T. E., Bjerkgaard, T., Kelley, D., Thorseth, I. and Pedersen, R. B. (2003). Hydrothermal chimneys and sulphide mineralized breccias from the Kolbeinsey and Mohns Ridge. Geophysical Research Abstracts, 5, 11,863.Google Scholar
O'Hanley, D. S. (1996). Serpentinites: Records of Tectonic and Petrological History. New York: Oxford University Press.Google Scholar
Okal, E. A. and Stewart, L. M. (1982). Slow earthquakes along oceanic fracture zones: Evidence for asthenospheric flow away from hotspots?Earth and Planetary Science Letters, 57, 75–87.CrossRefGoogle Scholar
Olafsson, J. S., Thors, K., Dtefansson, U.et al. (1990). Geochemical observations of a boiling hydrothermal site on the Kolbeinsey Ridge. Eos, Transactions of the American Geophysical Union, 71, 1650.Google Scholar
Ondréas, H., Cannat, M., Fouguet, Y.et al. (2009). Recent volcanic events and the distribution of hydrothermal venting at the Lucky Strike hydrothermal field, Mid-Atlantic Ridge. Geochemistry, Geophysics, Geosystems, 10(2), Q02006, doi:10.1029/2008GC002171.CrossRefGoogle Scholar
Ondréas, H., Fouquet, Y., Voisset, M. and Radford-Knoery, J. (1997). Detailed study of three contiguous segments of the Mid-Atlantic Ridge, south of the Azores (37°N to 38°30′N), using acoustic imaging coupled with submersible observations. Marine Geophysical Researches, 19, 231–255.CrossRefGoogle Scholar
Oosting, S. E. and Von Damm, K. L. (1996). Bromide/chloride fractionation in seafloor hydrothermal fluids from 9–10° N East Pacific Rise. Earth and Planetary Science Letters, 144, 133–145.CrossRefGoogle Scholar
Opatkiewicz, A. D., Butterfield, D. A. and Baross, J. A. (2009). Individual hydrothermal vents at Axial Seamount harbor distinct subseafloor microbial communities. FEMS Microbial Ecology, 70, 413–424.CrossRefGoogle ScholarPubMed
Oreskes, N. (ed.) 2001. Plate Tectonics: An Insider's History of the Modern Theory of the Earth. Cambridge, MA: Westview Press.Google Scholar
,OTTER (1984). The geology of the Oceanographer transform: The ridge-transform intersection. Marine Geophysical Researches, 6, 109–141.CrossRefGoogle Scholar
,OTTER (1985). The geology of the Oceanographer Transform: The transform domain. Marine Geophysical Researches, 7, 329–358.CrossRefGoogle Scholar
Pace, N. R. (1997). A molecular view of microbial diversity and the biosphere. Science, 276, 734–740.CrossRefGoogle ScholarPubMed
Pallister, J. S. and Hopson, C. A. (1981). Samail ophiolite plutonic suite: Field relations, phase variation and layering and a model of a spreading ridge magma chamber. Journal of Geophysical Research, 86, 2593–2644.CrossRefGoogle Scholar
Paonita, A. and Martelli, M. (2007). A new view of the He–Ar–CO2 degassing at mid-ocean ridges: Homogeneous composition of magmas from the upper mantle. Geochimica et Cosmochimica Acta, 7(7), 1747–1763.CrossRefGoogle Scholar
Paonita, A. and Martelli, M. (2013). Comment on “CO2 variability in mid-ocean ridge basalts from syn-emplacement degassing: Constraints on eruption dynamics” by Soule et al. Earth and Planetary Science Letters, 374, 251–253.CrossRefGoogle Scholar
Paquet, F., Dauteuil, O., Hallot, E. and Moreau, F. (2007). Tectonics and magma dynamics coupling in a dyke swarm of Iceland. Journal of Structural Geology, 29, 1477–1493.CrossRefGoogle Scholar
Pariso, J. E. and Johnson, H. P. (1989). Magnetic properties and oxide petrography of the sheeted dike complex in Hole 504B. In Proceedings of the Ocean Drilling Program, Scientific Results Leg 148. College Station TX: Ocean Drilling Program, pp. 159–166.Google Scholar
Paulick, H., Bach, W., Godard, M.et al. (2006). Geochemistry of abyssal peridotites (Mid-Atlantic Ridge, 15°20′N, ODP Leg 209): Implications for fluid/rock interaction in slow spreading environments. Chemical Geology, 234, 179–210.CrossRefGoogle Scholar
Pearce, J. A. (2003). Suprasubduction zone ophiolites: The search for modern analogs. In Dilek, Y. and Newcomb, S. (eds.), Ophiolite Concept and the Evolution of Geologic Thought. Special Paper 373, Boulder, CO: Geological Society of America, pp. 269–293.Google Scholar
Pearce, J. A. and Cann, J. R. (1971). Ophiolite origin investigated by discriminant analysis using Ti, Zr, and Y. Earth and Planetary Science Letters, 12, 339–349.CrossRefGoogle Scholar
Pearce, J. A. and Cann, J. R. (1973). Tectonic setting of basic volcanic rocks determined using trace element analyses. Earth and Planetary Science Letters, 19, 290–300.CrossRefGoogle Scholar
Pearce, J. A., Ernewein, M., Bloomer, S. H.et al. (1994). Geochemistry of Lau Basin volcanic rocks: Influence of ridge segmentation and arc proximity. In Smellie, J. L. (ed.), Volcanism Associated with Extension at Consuming Plate Margins. Special Publication 81. London: Geological Society of London, pp. 53–75.Google Scholar
Pearce, J. A., Lippard, S. J. and Roberts, S. (1984). Characteristics and tectonic significance of supra-subduction zone ophiolites. In Kokelaar, B. P. and Howells, M. F. (eds.), Marginal Basin Geology: Volcanic and Associated Sedimentary and Tectonic Processes in Modern and Ancient Marginal Basins. Special Publication 16. London: Geological Society of London, 77–94.Google Scholar
Pedersen, R. B., Malpas, J. and Falloon, T. (1996). Petrology and geochemistry of gabbroic and related rocks from Site 894 Hess Deep. In Mével, C., Gillis, K. M. and Allen, J. F. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 147. College Station, TX: Ocean Drilling Program, pp. 3–20, doi:10.2973/odp.proc.sr.147.001.1996.Google Scholar
Pedersen, R. B., Rapp, H. T., Thorseth, I. H.et al. (2010a). Discovery of a black smoker vent field and vent fauna at the Arctic Mid-Ocean Ridge. Nature, doi:10.1038/ncomms1124.CrossRef
Pedersen, R. B., Thorseth, I. H., Hellevang, B.et al. (2005). Two vent fields discovered at the ultraslow spreading Arctic Ridge System. Eos, Transactions of the American Geophysical Union, 86(52), OS21C-01.Google Scholar
Pedersen, R. B., Thorseth, I. H., Lilley, M. D.et al. (2009). Discovery of the Loki's Castle vent field at the ultra-slow spreading Arctic mid-ocean ridge. Geochimica et Cosmochimica Acta, 73, Supplement 1, A1008.Google Scholar
Pedersen, R., Thorseth, I. H., Nygaard, T.-E., Lilley, M. D. and Kelley, D. S. (2010b). Hydrothermal activity at the Arctic Mid-Ocean Ridge. In Rona, P., Devey, C. and Murton, B. (eds.), Diversity of Hydrothermal Systems on Slow Spreading Ocean Ridges. Geophysical Monograph 188. Washington, DC: American Geophysical Union, pp. 67–89.CrossRefGoogle Scholar
Perfit, M. R., Langmuir, C. H., Baekisapa, M.et al. (1987). Geochemistry and petrology of volcanic rocks from the Woodlark basin: Addressing questions of ridge subduction. In Marine Geology, Geophysics, and Geochemistry of the Woodlark Basin-Solomon Islands. Houston, TX: Circum-Pacific Council for Energy and Mineral Resources Earth Science, 7, pp. 113–154.Google Scholar
Perfit, M. R. (2001). Mid-ocean ridge geochemistry and petrology. In Steel, J., Thorpe, S. and Turekian, K. (eds.), Encylopedia of Ocean Sciences. San Diego: Academic Press, pp. 1778–1788.CrossRefGoogle Scholar
Perfit, M. R. and Chadwick, W. W. (1998). Magmatism at mid-ocean ridges: Constraints from volcanological and geochemical investigations. In Buck, W. R., Delaney, P. T., Karson, J. A. and Lagabrielle, Y. (eds.), Faulting and Magmatism at Mid-Ocean Ridges. Washington, DC: American Geophysical Union, pp. 59–115.Google Scholar
Perfit, M. R., Cann, J. R., Fornari, D. J.et al. (2003). Interaction of seawater and lava during submarine eruptions at mid-ocean ridges. Nature, 426, 62–64.CrossRefGoogle Scholar
Perfit, M. R., Fornari, D. J., Malahoff, A. and Embley, R. W. (1983). Geochemical studies of abyssal lavas recovered by DSRV Alvin from Eastern Galápagos Rift, Inca Transform, and Ecuador Rift 3: Trace element abundances and petrogenesis. Journal of Geophysical Research, 88, 10,551–10,572.CrossRefGoogle Scholar
Perfit, M. R., Fornari, D. J., Ridley, W. I.et al. (1996). Recent volcanism in the Siqueiros transform fault: Picritic basalts and implications for MORB magma genesis. Earth and Planetary Science Letters, 141, 91–108.CrossRefGoogle Scholar
Perfit, M. R., Fornari, D. J., Smith, M. C.et al. (1994). Small-scale spatial and temporal variations in mid-ocean ridge crest magmatic processes. Geology, 22, 375–379.2.3.CO;2>CrossRefGoogle Scholar
Perfit, M. R. (1977). Petrology and geochemistry of mafic rocks from the Cayman Trench: Evidence for spreading. Geology, 5(2), 105–110.2.0.CO;2>CrossRefGoogle Scholar
Perfit, M. R., Ridley, W. I. and Jonasson, I. (1999). Geologic, petrologic and geochemical relationships between magmatism and massive sulfide mineralization along the eastern Galápagos Spreading Center. Reviews in Economic Geology, 8, 75–100.Google Scholar
Perfit, M. R., Wanless, V. D., Ridley, W. I.et al. (2012). Lava geochemistry as a probe into crustal formation at the East Pacific Rise. Oceanography, 25(1), 89–93.CrossRefGoogle Scholar
Perk, N. W., Coogan, L. A., Karson, J. A. and Klein, E. M. (2007). Primitive gabbroic rocks from a tectonic window at Pito Deep: Implications for the accretion of plutonic rocks beneath the East Pacific Rise. Contributions to Mineralogy and Petrology, doi:10.1007/s00410-007-0210-z.CrossRef
Perner, M., Kuever, J., Seifert, R.et al. (2007). The influence of ultramafic rocks on microbial communities at the Logatchev hydrothermal field, located 15°N on the Mid-Atlantic Ridge. FEMS Microbial Ecology, doi:10.1111/j.1547–6941.2007.00325.x.CrossRef
Petersen, S., Kuhn, K., Kuhn, T.et al. (2009). The geological setting of the ultramafic-hosted Logatchev hydrothermal field (14°45′N, Mid-Atlantic Ridge) and its influence on massive sulfide formation. Lithos, 112, 40–56.CrossRefGoogle Scholar
Peterson, D. W. and Swanson, D. A. (1974). Observed formation of lava tubes during 1970–71 at Kilauea Volcano, Hawaii. Studies in Speleology, 2, 209–222.Google Scholar
Peterson, D. W. and Tilling, R. I. (1980). Transition of basaltic lava from pahoehoe to ‘a'a, Kilauea volcano Hawaii: Field observations and key factors. Journal of Volcanology and Geothermal Research, 7, 271–293.CrossRefGoogle Scholar
Peterson, J. J., Fox, P. J. and Schreiber, E. (1974). Newfoundland ophiolites and the geology of the oceanic layer. Nature, 247, 194–196.CrossRefGoogle Scholar
Petrecca, R. F. and Grassle, J. F. (1990). Notes on fauna from several deep-sea hydrothermal vent and cold seep soft sediment communities. In McMurray, G. R. (ed.), Gorda Ridge: A Seafloor Spreading Center in the United States Exclusive Economic Zone. New York: Springer-Verlag, pp. 279–283.CrossRefGoogle Scholar
Phipps Morgan, J. A. (1987). Melt migration beneath mid-ocean spreading centers. Geophysical Research Letters, 14, 1238–1241.CrossRefGoogle Scholar
Pinkerton, H. and Wilson, L. (1994). Factors controlling the lengths of channel-fed lava flows. Bulletin of Volcanology, 56, 108–120.CrossRefGoogle Scholar
Plüger, W. L. P., Herzig, M., Becker, K. P.et al. (1990). Discovery of hydrothermal fields at the Central Indian Ridge. Marine Mining, 9, 73–86.Google Scholar
Pockalny, R. A. and Larson, R. L. (2003). Implications for crustal accretion at fast-spreading ridges from observations in Jurassic oceanic crust in the western Pacific. Geochemistry, Geophysics, Geosystems, 4(1), 8903, doi:10.1029/2001GC000274.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, 3179–3194.CrossRefGoogle Scholar
Pockalny, R. A., Fox, P. J., Fornari, D. J., Macdonald, K. C. and Perfit, M. R. (1997). Tectonic reconstruction of the Clipperton and Siqueiros fracture zones: Evidence and consequences of plate motion change for the last 3 million years. Journal of Geophysical Research, 102, 3167–3181.CrossRefGoogle Scholar
Podowski, E. L., Moore, T. S., Zelnio, K. A., Luther, G. W. I. and Fisher, C. R. (2009). Distribution of diffuse flow megafauna in two sites on the Eastern Lau Spreading Center, Tonga. Deep-Sea Research I, Oceanographic Research Papers, 56, 2041–2056.CrossRefGoogle Scholar
Pollard, D. D. and Aydin, A. (1988). Progress in understanding jointing over the past century. Geological Society of America Bulletin, 100, 1181–1204.2.3.CO;2>CrossRefGoogle Scholar
Pollock, M. A., Klein, E. M., Karson, J. A. and Coleman, D. S. (2009). Compositions of dikes and lavas from the Pito Deep Rift: Implications for accretion at superfast spreading centers. Journal of Geophysical Research, 114(B3), B03207, doi:10.1029/2007JB005436.CrossRefGoogle Scholar
Pollock, M. A., Klein, E. M., Karson, J. A. and Tivey, M. A. (2005). Temporal and spatial variability in the composition of lavas exposed along the Western Blanco Transform Fault. Geochemistry, Geophysics, Geosystems, 6(11), Q11009, doi:10.1029/2005GC001026.CrossRefGoogle Scholar
Pontbriand, C. W. and Sohn, R. A. (2014). Microearthquake evidence for reaction-driven cracking within the Trans-Atlantic Geotraverse active hydrothermal deposit. Journal of Geophysical Research, 119, 822–839, doi:10.1002/2013JB010110.Google Scholar
Pontbriand, C. W., Soule, S. A., Sohn, R. A.et al. (2012). Effusive and explosive volcanism on the ultraslow-spreading Gakkel Ridge, 85° E. Geochemistry, Geophysics, Geosystems, 13(10), doi:10.1029/2012GC004187.CrossRefGoogle Scholar
Proskurowski, G., Kelley, D. S., Delaney, J. R.et al. (2012). Post-eruptive time series hydrothermal contribution to the water column above Axial Seamount. EOS Proceedings of the American Geophysical Union, OS22A–06.
Proskurowski, G., Lilley, M. D. and Brown, T. A. (2004). Isotopic evidence of magmatism and seawater bicarbonate removal at the Endeavour hydrothermal system. Earth and Planetary Science Letters, 225, 53–61.CrossRefGoogle Scholar
Proskurowski, G., Lilley, M. D., Kelley, D. S. and Olson, E. J. (2006). Low temperature volatile production at the Lost City Hydrothermal Field: Evidence from a hydrogen stable isotope geothermometer. Chemical Geology, 229, 331–343.CrossRefGoogle Scholar
Proskurowski, G., Lilley, M. D., Olson, E. J.et al. (2008). Abiogenic hydrocarbon production at Lost City Hydrothermal Field. Science, 319, 604–607.CrossRefGoogle ScholarPubMed
Purdy, G. M. and Rabinowitz, P. D. (1978). The Kane fracture zone at the Mid-Atlantic Ridge: IPOD Sites 395 and 396. In Melson, W. G., Rabinowitz, P.et al. (eds.), Initial Reports of the Deep Sea Drilling Project, Leg 45. Washington, DC: US Govenment Printing Office, map supplement.Google Scholar
Putirka, K. D., Perfit, M. R., Ryerson, F. J. and Jackson, M. G. (2007). Ambient and excess mantle temperatures, olivine thermometry, and active vs. passive upwelling. Chemical Geology, 241(3), 177–206.CrossRefGoogle Scholar
Putirka, K., Ryerson, F. J., Perfit, M. and Ridley, W. I. (2011). Mineralogy and composition of the oceanic mantle. Journal of Petrology, 52(2), 279–313.CrossRefGoogle Scholar
Quick, J. E. and Denlinger, R. P. (1993). Ductile deformation and the origin of layered gabbro in ophiolites. Journal of Geophysical Research, 98, 14,015–14,027.CrossRefGoogle Scholar
Radford-Knoery, J., Charlou, J.-L.Donval, J.-P.et al. (1998). Distribution of dissolved sulfide, methane, and manganese near the seafloor at the Lucky Strike (37°17′N) and Menez Gwen (37°50′N) hydrothermal vent sites on the mid-Atlantic RidgeDeep-Sea Research Part I, 45, 367–386.CrossRefGoogle Scholar
Ranero, C. and Reston, T. J. (1999). Detachment faulting at oceanic core complexes. Geology, 27, 983–986.2.3.CO;2>CrossRefGoogle Scholar
Rautenschlein, M., Jenner, G. A., Hertogen, J.et al. (1985). Isotopic and trace element composition of volcanic glasses from the Akaki Canyon, Cyprus: Implications for the origin of the Troodos ophiolite. Earth and Planetary Science Letters, 75, 369–383.CrossRefGoogle Scholar
Renninger, G. H., Kass, L., Gleeson, R. A.et al. (1995). Sulfide as a chemical stimulus for deep-sea hydrothermal vent shrimp. Biological Bulletin, 189, 69–76.CrossRefGoogle ScholarPubMed
Resing, J. A., Rubin, K. H., Embley, R. W.et al. (2011). Active submarine eruption of boninite in the northeastern Lau Basin. Nature Geoscience, 4, 799–806, doi:10.1038/NGEO1275.CrossRefGoogle Scholar
Reston, T. J. and Ranero, C. R. (2011). The 3-D geometry of detachment faulting at mid-ocean ridges. Geochemistry, Geophysics, Geosystems, 12(7), Q0AG05, doi:10.1029/2011GC003666.CrossRefGoogle Scholar
Reston, T. J., Ranero, C. and Belykh, I. (1999). The structure of Cretaceous oceanic crust of the NW Pacific: Constraints on processes at fast spreading centers. Journal of Geophysical Research, 104(B1), 629–644.CrossRefGoogle Scholar
Reynolds, T. J. and Beane, R. E. (1985). Evolution of hydrothermal fluid characteristics at the Santa Rita, New Mexico, porphyry copper deposit. Economic Geology, 80, 1328–1347.CrossRefGoogle Scholar
Richards, J. G., Cann, J. R. and Jensenius, J. (1989). Mineralogical zonation and metasomatism of the alteration pipes of Cyprus sulfide deposits. Economic Geology and the Bulletin of the Society of Economic Geologists, 84(1), 91–115.CrossRefGoogle Scholar
Riddihough, R. (1984). Recent movements of the Juan de Fuca Plate system. Journal of Geophysical Research, 89, 6980–6994.CrossRefGoogle Scholar
Ridley, W. I., Perfit, M. R., Fornari, D. J. and Smith, M. C. (2006). Magmatic processes in developing oceanic crust revealed in a cumulate xenolith collected at the East Pacific Rise, 9°50′N. Geochemistry, Geophysics, and Geosystems, 7, Q12O04, doi:10.1029/2006GC001316.CrossRefGoogle Scholar
Riedel, C., Schmidt, M., Botz, R. and Theilen, F. (2001). The Grimsey hydrothermal field offshore North Iceland: Crustal structure, faulting and related gas venting. Earth and Planetary Science Letters, 193, 409–421.CrossRefGoogle Scholar
Riou, V., Duperron, S., Halary, S.et al. (2010). Variation in physiological indicators in Bathymodiolus azoricus (Bivalvia: Mytilidae) at the Menez Gwen Mid-Atlantic Ridge deep-sea hydrothermal vent site within a year. Marine Environmental Research, 70, 264–271.CrossRefGoogle ScholarPubMed
Robigou, V. R., Delaney, J. R. and Stakes, D. S. (1993). Large massive sulfide deposits in a newly discovered active hydrothermal system, the High Rise Field, Endeavour Segment, Juan de Fuca Ridge. Geophysical Research Letters, 20, 1887–1890.CrossRefGoogle Scholar
Robinson, P. T., Malpas, J., Dilek, Y. and Zhou, M.-F. (2008). The significance of sheeted dike complexes in ophiolites. GSA Today, 18(11), doi:10.1130/GSATG22A.1, 4–10.CrossRefGoogle Scholar
Robinson, P. T., Melson, W. G., O'Hearn, T. and Schmincke, H.-U. (1983). Volcanic glass compositions of the Troodos ophiolite, Cyprus. Geology, 11, 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), doi:10.1371/journal.pbio.1001234.CrossRefGoogle ScholarPubMed
Rona, P. A. (1980). TAG hydrothermal field: Mid-Atlantic Ridge crest at latitude 26°N. Journal of the Geological Society London, 137(4), 365–402.CrossRefGoogle Scholar
Rona, P. A., Bougault, H., Charlou, J. L.et al. (1992). Hydrothermal circulation, serpentinization, and degassing at a rift valley–fracture zone intersection: Mid-Atlantic Ridge near 15°N, 45°W. Geology, 20, 783–786.2.3.CO;2>CrossRefGoogle Scholar
Rona, P. A., Hannington, M. D., Raman, C. V.et al. (1993). Active and relict sea-floor hydrothermal mineralization at the TAG hydrothermal field, Mid-Atlantic Ridge. Economic Geology, 88, 1989–2017.CrossRefGoogle Scholar
Rona, P. A., Klinkhammer, G., Nelsen, T. A., Trefry, J. H. and Elderfield, H. (1986). Black smokers, massive sulfides, and vent biota at the Mid-Atlantic Ridge. Nature, 321, 33–37.CrossRefGoogle Scholar
Rona, P. A., Thomson, G., Mottl, M. J.et al. (1984). Hydrothermal activity at the TAG hydrothermal field, Mid-Atlantic Ridge crest at 26°N. Journal of Geophysical Research, 89, 11,365–11,377.CrossRefGoogle Scholar
Rona, P. A., Widenfalk, L. and Bostrom, K. (1987). Serpentinized ultramafic and hydrothermal activity at the Mid-Atlantic Ridge crest near 15°N. Journal of Geophysical Research, 92, 1417–1427.CrossRefGoogle Scholar
Rouxel, O., Fouquet, Y. and Ludden, J. N. (2004). Subsurface processes at the Lucky Strike hydrothermal field, Mid-Atlantic Ridge: Evidence from sulfur, selenium, and iron isotopes. Geochimica et Cosmochimica Acta, 68, 2295–2311.CrossRefGoogle Scholar
Rowland, S. K. and Walker, G. P. L. (1990). Pahoehoe and a'a in Hawaii: Volumetric flow rate controls the lava structure. Bulletin of Volcanology, 52(8), 615–628, doi:10.1007/BF00301212.CrossRefGoogle Scholar
Rubin, A. M. (1992). Dike-induced faulting and graben subsidence in volcanic rift zones. Journal of Geophysical Research, 97, 1839–1858.CrossRefGoogle Scholar
Rubin, A. M. (1995). Propagation of magma-filled cracks. Annual Review of Earth and Planetary Sciences, 23, 287–336.CrossRefGoogle Scholar
Rubin, A. M. and Pollard, D. D. (1988). Dike-induced faulting in rift zones of Iceland and Afar. Geology, 16, 413–417.2.3.CO;2>CrossRefGoogle Scholar
Rubin, K. H. and Sinton, J. M. (2007). Inferences on mid-ocean ridge thermal and magmatic structure from MORB compositions. Earth and Planetary Science Letters, 260, 257–276.CrossRefGoogle Scholar
Rubin, K. H., Embley, R. W., Clague, D. A.et al. (2009a). Lavas from active boninite and very recent basalt eruptions at two submarine NE Lau Basin sites. Eos, Transactions of the American Geophysical Union, 90(52), V43I–05.Google Scholar
Rubin, K. H., Macdougall, J. D. and Perfit, M. R. (1994). 210Po – 210Pb dating of recent volcanic eruptions on the seafloor. Nature, 369, 841–844.CrossRefGoogle Scholar
Rubin, K. H., Sinton, J. M., Maclennan, J. and Hellebrand, E. (2009b). Magmatic filtering of mantle compositions at mid-ocean-ridge volcanoes. Nature Geoscience, 2(5), 321–328.CrossRefGoogle Scholar
Rubin, K. H., Smith, M. C., Bergmanis, E. C.et al. (2001a). Geochemical heterogeneity within mid-ocean ridge lava flows: Insights into eruption, emplacement and global variations in magma generation. Earth and Planetary Science Letters, 188, 349–367.CrossRefGoogle Scholar
Rubin, K. H., Smith, M. C., Bergmanis, E. C.et al. (2001b). Magmatic history and volcanological insights from individual lava flows erupted on the seafloor. Earth and Planetary Science Letters, 188, 349–367.CrossRefGoogle Scholar
Rubin, K. H., Soule, S. A., Chadwick, W. W.et al. (2012). Volcanic eruptions in the deep sea. Oceanography, 25(1), 142–157, doi:org/10.5670/oceanog.2012.12.CrossRefGoogle Scholar
Rudnicki, M. D., and German, C. R. (2002). Temporal variability of the hydrothermal plume above the Kairei vent field, 25°S, Central Indian Ridge. Geochemistry, Geophysics, Geosystems, 3(2), doi:10.1029/2001GC000240.CrossRefGoogle Scholar
Ruelas-Inzunza, J., Soto, L. A. and Páez-Osuna, F. (2003). Heavy-metal accumulation in the hydrothermal vent clam Vesicomya gigas from Guaymas basin, Gulf of California. Deep-Sea Research Part I, 50(3), 757–761.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, 355–371.CrossRefGoogle ScholarPubMed
Ryan, M. P. (1993). Neutral buoyancy and the structure of mid-ocean ridge magma reservoirs. Journal of Geophysical Research, 98, 22,321–22,338.CrossRefGoogle Scholar
Rzhnanov, Y., Mayer, L., Beaulieu, S.et al. (2006). Deep-sea geo-referenced video mosaics. IEEE Oceans 2006, doi:10.1109/OCEANS.2006.307018.CrossRef
Saal, A. E. and Van Orman, J. A. (2004). The 226Ra enrichment in oceanic basalts: Evidence for melt-cumulate diffusive interaction processes within the oceanic lithosphere. Geochemistry, Geophysics, Geosystems, 5, Q02008, doi:10.1029/2003GC000620.CrossRefGoogle Scholar
Sager, W. W., Zhang, J., Korenaga, J.et al. (2013). An immense shield volcano within the Shatsky Rise oceanic plateau, northwest Pacific Ocean. Nature Geoscience, 6, 976–981, doi:10.1038/ngeo1934.CrossRefGoogle Scholar
Sakai, H., Des Marias, D. J., Ueda, A. and Moore, J. G. (1984). Concentrations and isotope ratios of carbon, nitrogen and sulfur in ocean-floor basalts. Geochimica et Cosmochimica Acta, 48, 2433–2441.CrossRefGoogle ScholarPubMed
Salters, V. J. and Stracke, A. (2004). Composition of the depleted mantle. Geochemistry, Geophysics, Geosystems, 5(5), Q05B07, doi:10.1029/2003GC000597.CrossRefGoogle Scholar
Sano, T., Sakuyama, T., Ingle, S., Rodriguez, S. and Yamasaki, T. (2011). Petrological relationships among lavas, dikes, and gabbros from Integrated Ocean Drilling Program Hole 1256D: Insight into the magma plumbing system beneath the East Pacific Rise. Geochemistry, Geophysics, Geosystems, 12(6), Q06013, doi:10.1029/2011GC003548.CrossRefGoogle Scholar
Sansone, F. J., Resing, J. A., Tribble, G. W.et al. (1991). Lava–seawater interactions at shallow-water submarine lava flows. Geophysical Research Letters, 18, 1731–1734.CrossRefGoogle Scholar
Sarrazin, J. and Juniper, S. K. (1999). Biological characteristics of a hydrothermal edifice mosaic community. Marine Ecology Progress Series, 185, 1–19, doi:10.3354/meps185001.CrossRefGoogle Scholar
Sarrazin, J., Robigou, V., Juniper, S. K. and Delaney, J. R. (1997). Biological and geological dynamics over four years on a high temperature sulfide structure at the Juan de Fuca Ridge hydrothermal observatory. Marine Ecology Program Series, 152, 5–24.CrossRefGoogle Scholar
Sarrazin, J., Sarradin, P. M. and ,the MoMARETO cruise participants (2006). MoMARETO: A cruise dedicated to the spatio-temporal dynamics and the adaptations of hydrothermal vent fauna on the Mid-Atlantic Ridge. InterRidge News, 15, 24–33.Google Scholar
Sarrazin, J., Walter, C., Sarradin, P.-M.et al. (2006). Community structure and temperature dynamics within a mussel assemblage on the southern East Pacific Rise. In Felbeck, H.et al. (eds.), Third International Symposium on Hydrothermal Vent and Seep Biology, La Jolla, USA, September 12–16, 2005. Cahiers de Biologie Marine, 47(4), pp. 483–490.Google Scholar
Satake, K. and Atwater, B. F. (2007). Long-term perspectives on giant earthquakes and tsunamis at subduction zones. Annual Review of Earth and Planetary Sciences, 35, 349–374, doi:10.1146/annurev.earth.35.031306.140302.CrossRefGoogle Scholar
Schander, C., Rapp, H. T., Kongsrud, J. A.et al. (2010). The fauna of hydrothermal vents on the Mohns Ridge (North Atlantic). Marine Biology Research, 6, 155–171.CrossRefGoogle Scholar
Scheirer, D. S., Fornari, D. J., Humphris, S. E. and Lerner, S. (2000). High-resolution seafloor mapping using the DSL-120 sonar system: Quantitative assessment of sidescan and phase-bathymetry data from the Lucky Strike segment of the Mid-Atlantic Ridge. Marine Geophysical Researches, 21, 121–142.CrossRefGoogle Scholar
Scheirer, D. S., Shank, T. M. and Fornari, D. J. (2006). Temperature variations at diffuse and focused flow hydrothermal vent sites along the northern East Pacific Rise. Geochemistry, Geophysics, Geosystems, 7(3), Q03002, doi:10.1029/2005GC001094CrossRefGoogle Scholar
Schilling, J.-G., Zajac, M., Evans, R.et al. (1983). Petrologic and geochemical variations along the Mid-Atlantic Ridge from 29°N to 73°N. American Journal of Science, 283, 510–586.CrossRefGoogle Scholar
Schmidt, A. K., Perfit, M. R., Rubin, K. H.et al. (2011). Rapid cooling rates at an active mid-ocean ridge from zircon thermochronology. Earth and Planetary Science Letters, 302, 349–358, doi:10.1016/j.epsl.2010.12.022.CrossRefGoogle Scholar
Schmidt, C., Vuillemin, R., Le Gall, C., Gaill, F. and Le Bris, N. (2008). Geochemical energy sources from microbial primary production in the environment of hydrothermal vent shrimps. Marine Chemistry, 108, 18–31.CrossRefGoogle Scholar
Schmidt, K., Garbe-Schönberg, D., Koschinsky, A.et al. (2011b). Fluid elemental and stable isotope composition of the Nibelungen hydrothermal field (8°18′S, Mid-Atlantic Ridge): Constraints on fluid–rock interaction in heterogeneous lithosphere. Chemical Geology, 280, 1–18.CrossRefGoogle Scholar
Schmidt, K., Koschinsky, A., Garbe-Schonberg, D., de Carvalho, L. M. and Seifert, R. (2007). Geochemistry of hydrothermal fluids from the ultramafic-hosted Logatchev hydrothermal field, 15°N on the Mid-Atlantic Ridge: Temporal and spatial investigation. Chemical Geology, 242, 1–21.CrossRefGoogle Scholar
Schouten, H. and Denham, C. R. (2000). Comparison of volcanic construction in the Troodos ophiolite and oceanic crust using paleomagnetic inclinations from Cyprus Crustal Study Project (CCSP) CY-1 and CY-1A and Ocean Drilling Program (ODP) 504B drill cores. In Dilek, Y., Moores, E. M., Elthon, D. and Nicolas, A. (eds.), Ophiolites and Oceanic Crust: New Insights from Field Studies and the Ocean Drilling Program. Boulder, CO: Geological Society of America, pp. 181–194.CrossRefGoogle Scholar
Schouten, H., Klitgord, K. and Gallo, D. (1993). Edge-driven microplate kinematics. Journal of Geophysical Research, 98, 6689–6702.CrossRefGoogle Scholar
Schrenk, M. O., Kelley, D. S., Bolton, S. and Baross, J. D. (2004). Low archaeal diversity linked to sub-seafloor geochemical processes at the Lost City Hydrothermal Field, Mid-Atlantic Ridge. Environmental Microbiology, 6(10), 1086–1095.CrossRefGoogle Scholar
Schrenk, M. O., Kelley, D. S., Delaney, J. R. and Baross, J. A. (2003). Incidence and diversity of microorganisms within the walls of an active deep-sea sulfide chimney. Applied Environmental Microbiology, 69, 3580–3592.CrossRefGoogle Scholar
Schroeder, T. and John, B. E. (2004). Strain localization on an oceanic detachment fault system, Atlantis Massif, 30°N, Mid-Atlantic Ridge. Geochemistry, Geophysics, Geosystems, 5, Q11007, doi:10.1029/2004GC000728.CrossRefGoogle Scholar
Scott, R. B., Rona, P. A., McGregor, B. A. and Scott, M. R. (1974). The TAG hydrothermal field. Nature, 251, 301–302.CrossRefGoogle Scholar
Searle, R. C. and Francheteau, J. (1986). Morphology and tectonics of the Galápagos triple junction. Marine Geophysical Researches, 8, 95–129.Google 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
Searle, R. C., Rusby, R. I., Engeln, J.et al. (1989). Comprehensive sonar imaging of the Easter microplate. Nature, 341, 701–705.CrossRefGoogle Scholar
Segonzac, M., de Saint Laurent, M. and Casanova, B. (1993). L'énigme du comportement trophique des crevettes Alvinocarididae des sites hydrothermaux de la dorsale médio-atlantique. Cahiers de Biologie Marine, 34, 535–571.Google Scholar
Seher, T., Crawford, W. C., Singh, S. C. and Cannat, M. (2010). Seismic layer 2A variations in the Lucky Strike segment at the Mid-Atlantic Ridge from reflection measurements. Journal of Geophysical Research, 115, B07107, doi:10.1029/2009JB006783.Google Scholar
Seyfried, W. E. (1987). Experimental and theoretical constraints on hydrothermal alteration processes at mid-ocean ridges. Annual Review of Earth and Planetary Science Letters, 15, 317–335.CrossRefGoogle Scholar
Seyfried, W. E., Berndt, M. E. and Seewald, J. S. (1988). Hydrothermal alteration processes at mid-ocean ridges: Constraints from diabase alteration experiments, hot spring fluids and composition of the oceanic crust. Canadian Mineralogist, 26, 787–804.Google Scholar
Seyfried, W. E., Foustoukos, D. I. and Allen, D. E. (2004). Ultramafic-hosted hydrothermal systems at mid-ocean ridges: Chemical and physical controls on pH, redox, and carbon reduction reactions. In German, C. R., Lin, J. and Parsons, M. (eds.), Mid-Ocean Ridges: Hydrothermal Interactions between the Lithosphere and Oceans. Geophysical Monograph 148, Washington, DC: American Geophysical Union, pp. 267–284.Google Scholar
Seyfried, W. E., Pester, N. J., Ding, K. and Rough, M. (2011). Vent fluid chemistry of the Rainbow hydrothermal system (39°N): Phase equilibria and in situ pH controls on subseafloor alteration processes. Geochimica et Cosmochimica Acta, 75, 1574–1593.CrossRefGoogle Scholar
Seyfried, W. E., Pester, N. and Fu, Q. (2010). Phase equilibria controls on the chemistry of vent fluids from hydrothermal systems on slow spreading ridges: Reactivity of plagioclase and olivine solid solutions and the pH-silica connection. In Rona, P., Devey, C. and Murton, B. (eds.), Diversity of Hydrothermal Systems on Slow Spreading Ocean Ridges. Geophysical Monograph 188, Washington, DC: American Geophysical Union, pp. 297–320.CrossRefGoogle Scholar
Shand, S. J. (1949). Rocks of the Mid-Atlantic Ridge. Journal of Geology, 57, 89–92.CrossRefGoogle Scholar
Shank, T. M. (2004). The evolutionary puzzle of seafloor life. Oceanus, 42(2), 1–8.Google Scholar
Shank, T. M. (2006). Preliminary biological characterization of vent sites and the evolutionary relationships of vent fauna on the southern Atlantic Ridge. Exploration and biogeography of deep-water chemosynthetic ecosystems on the Atlantic Equatorial Belt region. First Results and Planning for Future Research Workshop, Barcelona, Spain.Google Scholar
Shank, T. M., Fornari, D. J., Von Damm, K. L.et al. (1998). Temporal and spatial patterns of biological community development at nascent deep-sea hydrothermal vents (9°50′N, East Pacific Rise). Deep-Sea Research Part II, 45, 465–515.CrossRefGoogle Scholar
Shank, T. M., Fornari, D. and Yoerger, D.et al. (2003). Deep submergence synergy: Alvin and ABE explore the Galápagos Rift at 86°W. Eos, Transactions of the American Geophysical Union, 84, 425–433, doi:10.1029/2003EO410001.CrossRefGoogle Scholar
Shanks, W. C.., Böhlke, J. K. and Seal, R. R. (1995) Stable isotopes in mid-ocean ridge hydrothermal systems: Interactions between fluids, minerals, and organisms. In Humphris, S., Zierenberg, R., Mullineau, L. and Thomson, R. (eds.), Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions. Geophysical Monograph 91, Washington, DC: American Geophysical Union, pp. 194–221.Google Scholar
Shervais, J. W. (2001). Birth, death, and resurrection: The life cycle of suprasubduction zone ophiolites. Geochemistry, Geophysics, Geosystems, 2, 1010, doi:10.1029/2000GC000080.CrossRefGoogle Scholar
Shih, J. S.-F. (1980). The Nature and Origin of Fine-Scale Sea-Floor Relief. Ph.D. dissertation, Massachusetts Institute of Technology.Google Scholar
,Shipboard Scientific Party (2005). Expedition 305: Oceanic Core Complex Formation, Atlantis Massif 2. USIO Preliminary Reports. College Station, TX: Integrated Ocean Drilling Program.Google Scholar
Siler, D. L. and Karson, J. A. (2009). Three-dimensional structure of inclined sheet swarms: Implications for crustal thickening and subsidence in the volcanic rift zones of Iceland. Journal of Volcanology and Geothermal Research, 188, 333–346.CrossRefGoogle Scholar
Sims, K. W. W., Blichert-Toft, J., Perfit, M. R.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, doi:101029/2002GC000443.CrossRefGoogle Scholar
Singh, S. C., Crawford, W. C., Carton, H.et al. (2006). Discovery of a magma chamber and faults beneath a Mid-Atlantic Ridge hydrothermal field. Nature, 442, 1029–1032, doi:10:1038/nature05105.CrossRefGoogle ScholarPubMed
Sinha, M. C., Navin, D. A., Macgregor, L. M.et al. (1997). Evidence for accumulated melt beneath the slow spreading Mid-Atlantic Ridge. In Cann, J. R., Elderfield, H. and Laughton, A. (eds.), Mid-Ocean Ridges: Dynamics of Processes Associated with Creation of New Ocean Crust. Cambridge: Cambridge University Press, pp. 233–253.Google 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., 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, 301–314.CrossRefGoogle Scholar
Sinton, J. M., White, S. M., Colman, A., Rubin, K. H. and Bowles, J. A. (2010). Volcanic eruptions on the western Galápagos Spreading Center: Connecting magma supply at depth to eruption rate on the surface. Eos, Transactions of the American Geophysical Union, Fall Meeting Supplement, V52A–06.
Sinton, J. M., Wilson, D. S., Christie, D. M., Hey, R. N. and Delaney, J. R. (1983). Petrologic consequences of rift propagation on oceanic spreading ridges. Earth and Planetary Science Letters, 62, 193–207.CrossRefGoogle Scholar
Smith, D.K., (2004). Ears in the Ocean, Oceanus, 42 (2), ISSN 1559-1263.
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, doi:10.1038/348152a0.CrossRefGoogle Scholar
Smith, D. K. and Cann, J. R. (1993). Building the crust at the Mid-Atlantic Ridge. Nature, 365, 707–715.CrossRefGoogle Scholar
Smith, D. K., Cann, J. R. and Escartín, J. (2006). Widespread active detachment faulting and core complex formation near 13°N on the Mid-Atlantic Ridge. Nature, doi:10.1038/nature04950.CrossRef
Smith, D. K., Escartín, 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, doi:10.1029/2002JB001964.CrossRefGoogle Scholar
Smith, D. K., Escartín, J., Schouten, H. and Cann, J. R. (2008). Fault rotation and core complex formation: Significant processes in seafloor formation at slow-spreading mid-ocean ridges (Mid-Atlantic Ridge, 13–15° N). Geochemistry, Geophysics, Geosystems, 9(3), Q03003, doi:10.1029/2007GC001699.CrossRefGoogle Scholar
Smith, D. K., Escartín, J., Schouten, H. and Cann, J. R. (2012). Active long-lived faults emerging along slow-spreading mid-ocean ridges. Oceanography, 25(1), 94–99, doi.org/10.5670/oceanog.2012.07.CrossRefGoogle Scholar
Smith, D. K., Exon, N., Barriga, F., and Tatsumi, Y. (2010). Forty years of successful international collaboration in scientific ocean drilling, Eos, Transactions of the American Geophysical Union, 91, 393–394.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 mid-ocean ridge basalt. Journal of Geophysical Research, 99(B3), 4787–4812, doi:10.1029/93JB02158.CrossRefGoogle Scholar
Smith, M. C., Perfit, M. R., Fornari, D. J. et al. (2001). Magmatic processes and segmentation at a fast spreading mid-ocean ridge: Detailed investigation of an axial discontinuity on the East Pacific Rise crest at 9°37′N. Geochemistry, Geophysics, Geosystems, 2, 1–32, doi:10.1029/2000GC000134.CrossRefGoogle Scholar
Smith, W. H. and Sandwell, D. T. (1997). Global sea floor topography from satellite altimetry and ship depth soundings. Science Magazine, 277, issue 5334.Google Scholar
Snow, J., Jokat, W., Hellebrand, E., Mühe, R. and Shipboard Scientific Party (2001). Magmatic and hydrothermal activity in Lena Trough, Arctic Ocean. Eos, Transactions of the American Geophysical Union, 85(47), T11G-01.Google Scholar
Sogin, M., Morrison, H. G., Huber, J. A.et al. (2006). Microbial diversity in the deep sea and the unexplored ‘rare biosphere’. Proceedings of the National Academy of Sciences, 103, 12,115–12,120.CrossRefGoogle Scholar
Sohn, R. A., Willis, C., Humphris, S. E.et al. (2008). Explosive volcanism on the ultraslow-spreading Gakkel ridge, Arctic Ocean. Nature, (453), 1236–1238.CrossRef
Solomon, S. C., Huang, P. Y. and Meinke, L. (1988). The seismic moment budget of slowly spreading ridges. Nature, 334, 58–60.CrossRefGoogle Scholar
Sontag, S. and Drew, C. (1998). Blind Man's Bluff: The Untold Story of American Submarine Espionage. New York: Harper Collins, ISBN 0-06-103004-X.
Soule, S. A. and Cashman, K. V. (2005). Shear rate dependence of the pahoehoe-to-'a'a transition: Analog experiments. Geology, 33, 361–364.CrossRefGoogle Scholar
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, Q08005, doi:10.1029/2005GC000912.CrossRefGoogle Scholar
Soule, S. A., Fornari, D. J., Perfit, M. R.et al. (2006). Incorporation of seawater into mid-ocean ridge lavas during emplacement. Earth and Planetary Science Letters, 252, 289–307.CrossRefGoogle Scholar
Soule, S. A., Fornari, D. J., Perfit, M. P. and Rubin, K. (2007a). New insights into mid-ocean ridge volcanic processes from the 2005–06 eruption of the East Pacific Rise, 9°46′–56′N. Geology, 35, 1079–1082, doi:10.1130/G23924A.1.CrossRefGoogle Scholar
Soule, S. A., Nakata, D. S., Fornari, D. J.et al. (2012). CO2 variability in mid-ocean ridge basalts from syn-emplacement degassing: Constraints on eruption dynamics. Earth and Planetary Science Letters, 327, 39–49, doi:10.1016/j.epsl.2012.01.034.CrossRefGoogle Scholar
Soule, S. A., White, S. M., Fornari, D. J., Ferrini, V. and Tivey, M. A. (2007b). Comparison of pre- and post- eruption surveys of the 2005–2006 East Pacific Rise Volcanic Event: Implications for fast-spreading mid-ocean ridge eruption processes. Eos, Transactions of the American Geophysical Union, 88(52), V21A–0383.Google 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, J. W. (1995). Geochemistry of the lower sheeted dike complex, Hole 504B, Leg 140. In Erzinger, J., Becker, K., Dick, H. J. B. and Stokking, L. B. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 140. College Station, TX: Ocean Drilling Program, pp. 81–97.Google Scholar
Sparks, R. S. J., Meyer, P. and Sigurdsson, H. (1980). Density variation amongst mid-ocean ridge basalts: Implications for magma mixing and the scarcity of primitive lavas. Earth and Planetary Science Letters, 46, 419–430.CrossRefGoogle Scholar
Spiegelman, M. (1993). Physics of melt extraction: Theory, implications and applications. Philosophical Transactions of the Royal Society of London, Series A, 342, 23–41.CrossRefGoogle Scholar
Spiegelman, M. (1996). Geochemical consequences of melt transport in 2D: The sensitivity of trace elements to mantle dynamics. Earth and Planetary Science Letters, 139, 115–132.CrossRefGoogle Scholar
Spiegelman, M. and Kelemen, P. B. (2003). Extreme chemical variability as a consequence of channelized melt transport. Geochemistry, Geophysics, Geosystems, 4(7), 1055, doi:10.1029/2002GC000336.CrossRefGoogle Scholar
Spiegelman, M. and McKenzie, D. (1987). Simple 2-D models for melt extraction at mid-ocean ridges and island arcs. Earth and Planetary Science Letters, 83, 137–152.CrossRefGoogle Scholar
Spiegelman, M. and Reynolds, J. R. (1999). Combined dynamic and geochemical evidence for convergent melt flow beneath the East Pacific Rise. Nature, 402, 282–285.CrossRefGoogle Scholar
Spiess, F. N. and Tyce, R. C. (1973). Marine Physical Laboratory Deep Tow Instrumentation System. La Jolla, CA: Scripps Institute of Oceanography Reference 73–4Google Scholar
Spiess, F. N., Macdonald, K. C., Atwater, T.et al. (1980). East Pacific Rise: Hot springs and geophysical experiments. Science, 207, 1421–1433.CrossRefGoogle ScholarPubMed
Spooner, E. T. C. and Fyfe, W. S. (1973). Sub-seafloor metamorphism, heat and mass transfer. Contributions to Mineralogy and Petrology, 42, 287–304.CrossRefGoogle Scholar
Squyres, S. W., Arvidson, R. E., Blaney, D. L.et al. (2006). Rocks of the Columbia Hills. Journal of Geophysical Research, 111, E02S11, doi:10.1029/2005JE002562.CrossRefGoogle 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, Q04003, doi:10.1029/2005GC001038.CrossRefGoogle Scholar
Stakes, D. S., Shervais, J. W. and Hopson, C. A. (1984). The volcanic-tectonic cycle of the FAMOUS and AMAR valleys, Mid-Atlantic Ridge (36°47′N): Evidence from basalt glass and phenocryst compositional variations for a steady state magma chamber beneath the valley midsections. Journal of Geophysical Research, 89, 6995–7028.CrossRefGoogle Scholar
Staudigel, H., Gee, J., Tauxe, L. and Varga, R. J. (1992). Shallow intrusive directions of sheeted dikes in the Troodos ophiolite: Anisotropy of magnetic susceptibility and structural data. Geology, 20, 841–844.2.3.CO;2>CrossRefGoogle Scholar
Staudigel, H., Koppers, A. A. P., Lavelle, J. W., Pitcher, T. J. and Shank, T. M. (2010). From the guest editors 2010. Oceanography, 23(1), 18–19, doi:org/10.5670/oceanog.2010.59.CrossRefGoogle Scholar
Steinmann, G. (1905). Geologische Beobachtungen in den Alpen, II. Die Schardtsche Ueberfaltungstheorie und die geologische Bedeutung der Tiefseeabsätze und der ophiolithischen Massengesteine. Berichte der Naturforschenden Gesellschaft zu Freiburg, 1(16), 44–67.Google Scholar
Stewart, M. A., Karson, J. A. and Klein, E. M. (2005). Four-dimensional upper crustal construction at fast-spreading mid-ocean ridges: A perspective from an upper crustal cross-section a the Hess Deep Rift. Journal of Volcanology and Geothermal Research, 144, 287–309.CrossRefGoogle Scholar
Stewart, M. S., Klein, E. M. and Karson, J. A. (2002). The geochemistry of dikes and lavas from the north wall of the Hess Deep Rift: Insights into the four-dimensional character of crustal construction at fast-spreading mid-ocean ridges. Journal of Geophysical Research, 107(B9), 2181, doi:10.1029/2000JB000051.CrossRefGoogle Scholar
Stewart, M. A., Klein, E. M., Karson, J. A. and Brophy, J. G. (2003). Geochemical relationships between dikes and lavas at the Hess Deep Rift: Implications for magma eruptibility. Journal of Geophysical Research, 108(B4), 2184, doi:10.1029/2001JB001622.CrossRefGoogle Scholar
Stracke, A. and Bourdon, B. (2009). The importance of melt extraction for tracing mantle heterogeneity. Geochimica et Cosmochimica Acta, 73, 218–238.CrossRefGoogle Scholar
Streams, M. E., Fisher, C. R. and Fiala-Médioni, A. (1997). Methanotrophic symbiont location and fate of carbon incorporated from methane in a hydrocarbon seep mussel. Marine Biology, 129, 465–476.CrossRefGoogle Scholar
Streckeisen, A. (1976). To each plutonic rock its proper name. Earth-Science Reviews, 12(1), 1–33.CrossRefGoogle Scholar
Stroup, J. B. and Fox, P. J. (1981). Geologic investigations in the Cayman Trough: Evidence for thin oceanic crust along the Mid-Cayman Rise. Journal of Geology, 89, 395–420.CrossRefGoogle Scholar
Sturm, M., Goldstein, S., Klein, E. M., Karson, J. A. and Murrell, M. T. (2000). Uranium-series age constraints on lavas from the axial valley of the Mid-Atlantic Ridge, MARK Area. Earth and Planetary Science Letters, 181, 61–71.CrossRefGoogle Scholar
Sturm, M., Klein, E. M., Karsten, J., Graham, D. and Karson, J. (2000). Evidence for subduction-related contamination of the mantle beneath the southern Chile Ridge: Implications for ambiguous ophiolite compositions. In Dilek, Y., Moores, E. M., Elthon, D. and Nicolas, A. (eds.), Ophiolites and Oceanic Crust: New Insights from Field Studies and Oceanic Drilling. Boulder, CO: Geological Society of America, pp. 13–20.CrossRefGoogle Scholar
Summit, M. (2000). Ecology, Physiology, and Phylogeny of Subseafloor Thermophiles from Mid-Ocean Ridge Environments. Ph.D. thesis, University of Washington.Google Scholar
Summit, M. and Baross, J. A. (1998). Thermophilic subseafloor microorganisms from the 1996 North Gorda Ridge eruption. Deep-Sea Research II, 45, 2751–2766.CrossRefGoogle Scholar
Sun, S. S. and McDonough, W. F. (1989). Chemical and isotopic systematics of oceanic basalts: Implications for mantle composition and processes. In Saunders, A. D. and Norry, M. J. (eds), Magmatism in the Ocean Basins. Special Publication 42, London: Geological Society of London, pp. 313–345.Google Scholar
Sun, S. S., Nesbitt, R. W. and Sharaskin, A. Y. (1979). Geochemical characteristics of mid-ocean ridge basalts. Earth and Planetary Science Letters, 44(1), 119–138.CrossRefGoogle Scholar
Swanson, D. A. (1973). Pahoehoe flows from the 1969–1971 Mauna Ulu eruption, Kilauea volcano, Hawaii. Geological Society of America Bulletin, 84, 615–626.2.0.CO;2>CrossRefGoogle Scholar
Sykes, L. R. (1967). Mechanism of earthquakes and nature of faulting on the Mid-Oceanic Ridges. Journal of Geophysical Research, 72(8), 2131–2153.CrossRefGoogle Scholar
Sykes, L. R. (1969). Seismicity of the mid-oceanic ridge system. In Hart, P. J. (ed.), The Earth's Crust and Upper Mantle. Geophysical Monograph 13. Washington, DC: American Geophysical Union, pp. 148–153.Google Scholar
Takai, K., Nakamura, K., Toki, T.et al. (2008). Cell proliferation at 122 °C and isotopically heavy CH4 production by a hyperthermophilic methanogen under high pressure cultivation. Proceedings of the National Academy of Sciences, 105, 10,949–10,954.CrossRefGoogle Scholar
Tandberg, A. H., Rapp, H. T., Schander, C.et al. (2012) Exitomelita sigynae gen. et sp. nov.: A new amphipod from the Arctic Loki Castle vent field with potential gill ectosymbionts. Polar Biology, 35, 705–716.CrossRefGoogle Scholar
Tartarotti, P., Hayman, N. W., Anma, R.et al. (2006). Structure of Hole 1256D: The role of mechanical deformation in superfast-spread crust. Eos, Transactions of the American Geophysical Union, 87(52), B31B–1091.Google Scholar
Tauxe, L., Gee, J. S. and Staudigel, H. (1998). Flow directions in dikes from anisotropy of magnetic susceptibility data: The bootstrap way. Journal of Geophysical Research, 103, 17,775–17,790.CrossRefGoogle Scholar
Taylor, C. D. and Wirsen, C. O. (1997). Microbiology and ecology of filamentous sulfur formation. Science, 277, 1483–1485.CrossRefGoogle Scholar
Teixera, S., Cambon-Bonavita, M.-A., Serrao, E. A., Desbruyères, D. and Arnaud-Haond, S. (2011). Recent population expansion and connectivity in the hydrothermal shrimp Rimicaris exoculata along the Mid-Atlantic Ridge. Journal of Biogeography, 38, 563–574.Google Scholar
Teske, A., Hinrichs, K.-U., Edgcomb, V.et al. (2007). Microbial diversity of hydrothermal sediments in the Guaymas Basin: Evidence for anerobic methanotrophic communities. Applied and Environmental Microbiology, 68(4), 1994–2007.CrossRefGoogle Scholar
Thompson, G., Bryan, W. B. and Humphris, S. E. (1989). Axial volcanism on the East Pacific Rise, 10–12°N. In Saunders, A. D. and Norry, M. J. (eds.), Magmatism in the Ocean Basins. Special Publication 42, London: Geological Society of London, pp. 181–200.Google Scholar
Thompson, G., Mottl, M. J. and Rona, P. A. (1985). Morphology, mineralogy, and chemistry of hydrothermal deposits from the TAG area, 26°N Mid-Atlantic Ridge. Chemical Geology, 49, 243–257.CrossRefGoogle Scholar
Thordarson, T. and Self, S. (2003). Atmospheric and environmental effects of the 1783–1784 Laki eruption: A review and reassessment. Journal of Geophysical Research, 108(D1), 4011, doi:10.1029/2001JD002042.CrossRefGoogle Scholar
Thordarson, T., Self, S., Miller, D. J., Larsen, G. and Vilmundardóttir, E. G. (2003). Sulphur release from flood lava eruptions in the Veidivötn, Grímsvötn and Katla volcanic systems, Iceland. In Oppenheimer, C., Pyle, D. M. and Barclay, J. (eds.), Volcanic Degassing. Special Publication 213, London: Geological Society of London, 103–121.Google Scholar
Thorseth, I. H., Pedersen, R. B., Kruber, C. and Kosler, J. (2007). Low-temperature hydrothermal deposits at the 71°N vent fields at the Arctic Mid-Ocean Ridge: Architecture, microstructures, and geochemistry. Eos, Transactions of the American Geophysical Union, 88(52), Fall Meeting Supplement, Abstract OS43A-0996.Google Scholar
Tiezzi, L. J. and Scott, R. B. (1980). Crystal fractionation in a cumulate gabbro, Mid-Atlantic Ridge, 26°N. Journal of Geophysical Research, 85, 5438–5454.CrossRefGoogle Scholar
Titus, T. N., Kieffer, H. H. and Christensen, P. R. (2003). Exposed water ice discovered near the south pole of Mars. Science, 299, 1048–1051.CrossRefGoogle ScholarPubMed
Tivey, M. A. and Johnson, H. P. (2002). Crustal magnetization reveals subsurface structure of Juan de Fuca Ridge hydrothermal vent fields. Geology, 30(11), 979–982.2.0.CO;2>CrossRefGoogle Scholar
Tivey, M. A., Johnson, H. P., Fleutelot, C.et al. (1998a). Direct measurement of magnetic reversal polarity boundaries in a cross-section of oceanic crust. Geophysical Research Letters, 25, 3631–3634.CrossRefGoogle Scholar
Tivey, M. A., Rona, P. A. and Kleinrock, M. C. (1996). Reduced crustal magnetization beneath relict hydrothermal mounds: TAG hydrothermal field, Mid-Atlantic Ridge, 26°N. Geophysical Research Letters, 23, 3511–3514.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, 108(B5), 2277, doi:10.1029/2002JB001967.CrossRefGoogle Scholar
Tivey, M. K. (1995). Modeling chimney growth and associated fluid flow at seafloor hydrothermal vent sites. In Humphris, S., Zierenberg, R., Mullineaux, L. and Thomson, R. (eds.), Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions. Geophysical Monograph 91, Washington, DC: American Geophysical Union, pp. 158–177.Google Scholar
Tivey, M. K. (2004). Environmental conditions within active seafloor vent structures: Sensitivity to vent fluid composition and fluid flow. In Wilcock, W., Cary, C., DeLong, E., Kelley, D. and Baross, J. (eds.), Subseafloor Biosphere at Mid-Ocean Ridges. Geophysical Monograph 144, Washington, DC: American Geophysical Union, pp. 137–152.CrossRefGoogle Scholar
Tivey, M. K. (2007). Generation of seafloor hydrothermal vent fluids and associated mineral deposits. Oceanography, 20, 50–65.CrossRefGoogle Scholar
Tivey, M. K. and Delaney, J. R. (1986). Growth of large sulfide structures on the Endeavour Segment of the Juan de Fuca Ridge. Earth and Planetary Science Letters, 77, 303–317.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, 100, 12,527–12,555.CrossRefGoogle Scholar
Tivey, M. K., Mills, R. A. and Teagle, D. A. H. (1998b). Temperature and salinity of fluid inclusions in anhydrite as indicators of seawater entrainment and heating in the TAG active mound. In Herzig, P. M., Humphris, S. E.et al. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Leg 158. College Station TX: Ocean Drilling Program, pp. 179–190.Google Scholar
Tivey, M. K., Stakes, D. S., Cook, T. L., Hannington, M. D. and Petersen, S. (1999). A model for growth of steep-sided vent structures on the Endeavour Segment of the Juan de Fuca Ridge: Results of a petrologic and geochemical study. Journal of Geophysical Research, 104, 22,859–22,883.CrossRefGoogle Scholar
Tolstoy, M., Cowan, J. P., Baker, E. T.et al. (2006). A seafloor spreading event captured by seismometers. Science, 314, 1920–1922, doi:10.1126science.1133950.CrossRefGoogle Scholar
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, 181–184.CrossRefGoogle ScholarPubMed
Tominaga, M. and Sager, W. W. (2010). Revised Pacific M-anomaly geomagnetic polarity timescale. Geophysical Journal International, 182, 203–232, doi:10.1111/j.1365-246X.2010.04619.x.Google Scholar
Tominaga, M. and Umino, S. (2010). Lava deposition history in ODP Hole 1256D: Insights from log-based volcanostratigraphy. Geochemistry, Geophysics, Geosystems, 11, Q05003, doi:10.1029/2009GC002933.CrossRefGoogle Scholar
Tominaga, M., Teagle, D. A., Alt, J. C. and Umino, S. (2009). Determination of the volcanostratigraphy of oceanic crust formed at superfast spreading ridge: Electrofacies analyses of ODP/IODP Hole 1256D. Geochemistry, Geophysics, Geosystems, 10(1), doi:10.1029/2008GC002143.CrossRefGoogle Scholar
Toner, B. M., Fakra, S. C., Manganini, S. J.et al. (2009). Preservation of iron (II) by carbon-rich matrices in a hydrothermal plume. Nature Geosciences, 2, 197–201.CrossRefGoogle Scholar
Toomey, D. R. and Hooft, E. E. E. (2008). Mantle upwelling, magmatic differentiation, and the meaning of axial depth at fast-spreading ridges. Geology, 36, 679–682.CrossRefGoogle Scholar
Toomey, D. R., Jousselin, D., Dunn, R. A., Wilcock, W. S. D. and Detrick, R. S. (2007). Skew of mantle upwelling beneath the East Pacific Rise governs segmentation. Nature, 446, 409–414.CrossRefGoogle ScholarPubMed
Toomey, D. R., Purdy, G. M., Solomon, S. and Wilcox, W. (1990). The three dimensional seismic velocity structure of the East Pacific Rise near latitude 9°30′ N. Nature, 347, 639–644.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, 9093–9112.CrossRefGoogle Scholar
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. (1994). Tomographic imaging of the shallow crustal structure of the East Pacific Rise at 9°30′N. Journal of Geophysical Research, 99, 24,135–24,157.CrossRefGoogle Scholar
Trask, J. L. and Van Dover, C. L. (1999). Site-specific and ontogenetic variations in nutrition of mussels (Bathymodiolus sp.) from the Lucky Strike hydrothermal vent field, Mid-Atlantic Ridge. Limnology and Oceanography, 44, 334–343.CrossRefGoogle Scholar
Treves, B. and Harper, G. D. (1994). Exposure of serpentinites on the ocean floor: Sequence of faulting and hydrofracturing in the northern Apennine ophicalcites. Ofioliti, 19b, 435–466.Google Scholar
Tucholke, B. E. and Schouten, H. (1988). Kane fracture zone. Marine Geophysical Researches, 10, 1–39.CrossRefGoogle Scholar
Tucholke, B. E. and Vogt, P. R. (1979). Western North Atlantic: Sedimentary evolution and aspects of tectonic history In Tucholke, B. E., Vogt, P. R.et al., Initial Reports Deep Sea Drilling Project, Leg 63, Washington, DC: US Government Printing Office, pp. 791–825.Google 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(6), 455–458, doi:10.1130/G24639A.CrossRefGoogle Scholar
Tucholke, B. E., Lin, J. and Kleinrock, M. C. (1998). Megamullions and mullion structure defining oceanic metamorphic core complexes on the Mid-Atlantic Ridge. Journal of Geophysical Research, 103, 9857–9866.CrossRefGoogle Scholar
Tucholke, B. E., Lin, J., Kleinrock, M. C.et al. (1997). Segmentation and crustal structure of the western Mid-Atlantic Ridge flank, 25°25′–27°10′N and 0–29 m.y. Journal of Geophysical Research, 102, 10,203–10,224.CrossRefGoogle Scholar
Tunnicliffe, V. and Fowler, M. R. (1996). Influence of sea-floor spreading on the global hydrothermal vent fauna. Nature, 379, 531–533.CrossRefGoogle Scholar
Tunnicliffe, V. A., McArthur, G. and McHugh, D. (1998). A biogeographical perspective of the deep-sea hydrothermal vent fauna. Advances in Marine Biology, 34, 353–442.CrossRefGoogle Scholar
Twiss, R. J. and Moores, E. M. (1992). Structural Geology. New York: W. H. Freeman.Google Scholar
Umbgrove, J. H. F. (1947). The Pulse of the Earth. The Hague: Martinus Nijhoff.CrossRefGoogle Scholar
Umino, S., Crispini, L., Tartarotti, P.et al. (2008). Origin of the sheeted dike complex at superfast spread East Pacific Rise revealed by deep ocean crust drilling at Ocean Drilling Program Hole 1256D. Geochemistry, Geophysics, Geosystems, 9, Q06O08, doi:10.1029/2007GC001760.CrossRefGoogle Scholar
Umino, S., Nealson, K. and Wood, B. (2013). Drilling to Earth's mantle. Physics Today, 66(8), 36–41, doi:10.1063/PT.3.2082.CrossRefGoogle Scholar
,University National Laboratory System (UNOLS) (1994). The Global Abyss: An Assessment of Deep Submergence Science in the United States. Narragansett, RI: UNOLS Office, University of Rhode Island.Google Scholar
Van Ark, E. M., Detrick, R. S., Canales, J. P.et al. (2007). Seismic structure of the Endeavour Segment, Juan de Fuca Ridge: Correlations with seismicity and hydrothermal activity. Journal of Geophysical Research, 112, B02401, doi.org/10.1029/2005JB004210.CrossRefGoogle Scholar
Van Dover, C. L. (1995). Ecology of Mid-Atlantic Ridge hydrothermal vents. In Parson, L. M., Walker, C. L. and Dixon, D. R. (eds.), Hydrothermal Vents and Processes, Special Publication 87, London: Geological Society of London, pp. 257–294.Google Scholar
Van Dover, C. L. (2000). The Ecology of Deep-Sea Hydrothermal Vents. Princeton, NJ: Princeton University Press.Google Scholar
Van Dover, C. L. (2003). Variation in community structure within hydrothermal vent mussel beds of the East Pacific Rise. Marine Ecology-Progress Series, 253, 55–66.CrossRefGoogle Scholar
Van Dover, C. L. and Doerries, M. B. (2005). Community structure in mussel beds at Logatchev hydrothermal vents and a comparison of macrofaunal species richness on slow- and fast-spreading mid-ocean ridges. Marine Ecology, 26, 110–120.CrossRefGoogle Scholar
Van Dover, C. L. and Trask, J. L. (2001). Biodiversity in mussel beds at a deep-sea hydrothermal vent and a shallow-water intertidal site. Marine Ecology Program Series, 195, 169–178.CrossRefGoogle Scholar
Van Dover, C. L., Desbruyères, D., Segonzac, M.et al. (1996). Biology of the Lucky Strike hydrothermal field. Deep-Sea Research Part I, 43(9), 1509–1529.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, 95, 1253–1257.CrossRefGoogle Scholar
Van Dover, C. L., Humphris, S. E., Fornari, D.et al. (2001). Biogeography and ecological setting of Indian Ocean hydrothermal vents. Science, 294, 818–823.CrossRefGoogle ScholarPubMed
Van Dover, C. L., Szuts, E. A., Chamberlain, S. C and Cann, J. R. (1989). A novel eye in “eyeless” shrimp from hydrothermal vents of the Mid-Atlantic Ridge. Nature, 337, 458–460.CrossRefGoogle ScholarPubMed
Vanko, D. A. (1988). Temperature, pressure, and composition of hydrothermal fluids, with their bearing on the magnitude of tectonic uplift at mid-ocean ridges, inferred from fluid inclusions in oceanic layer 3 rocks. Journal of Geophysical Research, 93, 4595–4611.CrossRefGoogle Scholar
Vanko, D. A. and Batiza, R. (1982). Gabbroic rocks from the Mathematician Ridge failed rift. Nature, 300, 742–744.CrossRefGoogle Scholar
Varga, R. J. (1991). Modes of extension at mid-ocean ridge spreading centers: Evidence from the Solea graben, Troodos ophiolite Cyprus. Journal of Structural Geology, 13, 517–538.CrossRefGoogle Scholar
Varga, R. J., Gee, J. S., Staudigel, H. and Tauxe, L. (1998). Dike surface lineations as magma flow indicators within the sheeted dike complex of the Troodos Ophiolite, Cyprus. Journal of Geophysical Research, 103, 5241–5256.CrossRefGoogle Scholar
Varga, R. J., Gee, J., Bettison-Varga, L., Anderson, R. S. and Johnson, C. L. (1999). Early establishment of seafloor hydrothermal systems during structural extension: Paleomagnetic evidence from the Troodos ophiolite, Cyprus. Earth and Planetary Science Letters, 1999, 221–235.CrossRefGoogle Scholar
Varga, R. J., Horst, A., Gee, J. S. and Karson, J. A. (2008). Direct evidence from anisotropy of magnetic susceptibility for lateral melt migration at superfast spreading centers. Geochemistry, Geophysics, Geosystems, 9(8), Q08008, doi:10.1029/2008GC002075.CrossRefGoogle Scholar
Varga, R. J., Karson, J. A. and Gee, J. S. (2004). Paleomagnetic constraints on tilt and fault models for oceanic crust from oriented samples from the Hess Deep Rift, Equatorial Pacific Ocean. Journal of Geophysical Research, 109(2), 1–22, doi:10.1029/2003JB002486.Google Scholar
Veloso, E. E., Hayman, N. W., Anna, R.et al. (2014). Magma flow directions in the sheeted dike complex at superfast spreading mid-ocean ridges: Insights from IODP Hole 1256D, eastern Pacific. Geochemistry, Geophysics, Geosystems, 15, 1283–1295. doi:10.1002/2013GC004957.CrossRefGoogle Scholar
Ver Eecke, H. C., Akerman, N. H., Huber, J. A.et al. (2013) Growth kinetics and energetics of a deep-sea hyperthermophilic methanogen under varying environmental conditions. Environmental Microbial Reports, 5(5), 665–671.Google ScholarPubMed
Vera, E. E. and Diebold, J. B. (1994). Seismic imaging of oceanic layer 2A between 9°30′N and 10°N on the East Pacific Rise from two-ship wide-aperture profiles. Journal of Geophysical Research, 99, 3031–3041.CrossRefGoogle Scholar
Vetriani, C., Chew, Y. S., Miller, S. M.et al. (2004b). Mercury adaptation among bacteria from a deep-sea hydrothermal vent. Applied Environmental Microbiology, 71, 220–226.CrossRefGoogle Scholar
Vetriani, C., Speck, M. D., Ellor, S. V., Lutz, R. A. and Starovoytov, V. (2004a). Thermovibrio ammonificans sp. nov.: A thermophilic, chemolithotrophic, nitrate ammonifying bacterium from deep-sea hydrothermal vents. International Journal of Systematic and Evolutionary Biology, 54, 175–181.Google ScholarPubMed
Vine, F. J. and Matthews, D. H. (1963). Magnetic anomalies over oceanic ridges. Nature, 199, 947–949.CrossRefGoogle Scholar
Von Damm, K. L. (1995). Controls on the chemistry and temporal variability of seafloor hydrothermal fluids. In Humphris, S., Zierenberg, R., Mullineau, L. and Thomson, R. (eds.), Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions. Geophysical Monograph 91, Washington, DC: American Geophysical Union, pp. 222–247Google Scholar
Von Damm, K. L. (2000). Chemistry of hydrothermal vent fluids from 9°–10°N, East Pacific Rise: “Time zero,” the immediate post eruptive period. Journal of Geophysical Research, 105(B5), 11,203–11,222.CrossRefGoogle Scholar
Von Damm, K. L. and Lilley, M. D. (2004). Diffuse flow hydrothermal fluids from 9°50′N East Pacific Rise: Origin, evolution and biologeochemical controls. In Wilcock, W. D., Kelley, D. S., Baross, J. A., DeLong, E. and Cary, C. (eds.), The Subseafloor Biosphere at Mid-Ocean Ridges. Geophysical Monograph 144, Washington, DC: American Geophysical Union, pp. 245–268.CrossRefGoogle Scholar
Von Damm, K. L., Bray, A. M., Buttermore, L. G. and Oosting, S. E. (1998). The geochemical controls on vent fluids from the Lucky Strike vent field, Mid-Atlantic Ridge. Earth and Planetary Science Letters, 160, 521–536.CrossRefGoogle Scholar
Von Damm, K. L., Buttermore, L. G., Oosting, S. E.et al. (1997). Direct observation of the evolution of a seafloor ‘black smoker’ from vapor to brine. Earth and Planetary Science Letters, 149, 101–111.CrossRefGoogle Scholar
Von Damm, K. L., Edmond, J. M., Measures, C. I. and Grant, B. (1985). Chemistry of submarine hydrothermal solutions at Guaymas Basin, Gulf of California. Geochimica et Cosmochimica Acta, 49, 2221–2237.CrossRefGoogle Scholar
Von Damm, K. L., Lilley, M. D., Shanks, W. C.et al. (2003). Extraordinary phase separation and segregation in vent fluids from the southern East Pacific Rise. Earth and Planetary Science Letters, 206, 365–378.CrossRefGoogle Scholar
Von Damm, K. L., Oosting, S. E., Kozlowski, R.et al. (1995). Evolution of East Pacific Rise hydrothermal vent fluids following a volcanic eruption. Nature, 375, 47–50.CrossRefGoogle Scholar
Von Damm, K. L., Parker, C. M., Zierenberg, R. A.et al. (2005). The Escanaba Trough, Gorda Ridge hydrothermal system: Temporal stability and subseafloor complexity. Geochimica et Cosmochimica Acta, 69, 4971–4989.CrossRefGoogle Scholar
Vrijenhoek, R. C. (2010). Genetic diversity and connectivity of deep-sea hydrothermal vent metapopulations. Molecular Ecology, 19, 4391–4411.CrossRefGoogle ScholarPubMed
Wager, L. R. and Brown, G. M. (1968). Layered Igneous Rocks. Edinburgh: Oliver and Boyd.Google Scholar
Wager, L. R., Brown, G. M. and Wadsworth, W. J. (1960). Types of igneous cumulates. Journal of Petrology, 1(1), 73–85.CrossRefGoogle Scholar
Walker, G. P. L. (1986). Koolau dike complex, Oahu: Intensity and origin of a sheeted dike complex high in a Hawaiian volcanic edifice. Geology, 14, 310–313.2.0.CO;2>CrossRefGoogle Scholar
Walker, G. P. L. (1993). Basaltic-volcano systems. In Prichard, H. M., Alabaster, T., Harris, N. B. W. and Neary, C. R. (eds.), Magmatic Processes and Plate Tectonics. London: Blackwell Scientific, pp. 3–38.Google Scholar
Waller, R. G., Scanlon, K. M. and Robinson, L. F. (2011). Cold-water coral distributions from towed camera observations: Initial interpretations. PLoS ONE, 6(1), doi:10.1371/journal.pone.0016153.CrossRefGoogle ScholarPubMed
Wang, T., Tucholke, B. E. and Lin, J. (2014). Spatial and temporal variations in crustal production at the Mid-Atlantic Ridge, 25°N–27°30′N and 0–28 Ma. Geochemistry, Geophysics, Geosystems (submitted).
Wanless, V. D. and Shaw, A. M. (2012). Lower crustal crystallization and melt evolution at mid-ocean ridges. Nature Geoscience, 5(9), 651–655.CrossRefGoogle Scholar
Wanless, V. D., Perfit, M. R., Klein, E. M., White, S. and Ridley, W. I. (2012). Reconciling geochemical and geophysical observations of magma supply and melt distribution at the 9°N overlapping spreading center, East Pacific Rise. Geochemistry, Geophysics, Geosystems, 13, Q11005, doi:10.1029/2012GC004168.CrossRefGoogle Scholar
Wanless, V. D., Perfit, M. R., Ridley, W. I. and Klein, E. M. (2010). Dacite petrogenesis on mid-ocean ridges: Evidence for crustal melting and assimilation. Journal of Petrology, 51, 2377–2410.CrossRefGoogle Scholar
Waters, C. L., Sims, K. W., Perfit, M. R., Blichert-Toft, J. and Blusztajn, J. (2011). Perspective on the genesis of E-MORB from chemical and isotopic heterogeneity at 9–10° N East Pacific Rise. Journal of Petrology, 52, 565–602.CrossRefGoogle Scholar
Weekly, R. T., Wilcock, W. S. D., Hooft, E. E. E.et al. (2013). Termination of a 6 year ridge-spreading event observed using a seafloor seismic network on the Endeavour Segment, Juan de Fuca Ridge. Geochemistry, Geophysics, Geosystems, 14, 1375–1398, doi:10.1002/ggge.20105.CrossRefGoogle Scholar
Wegener, A. (1912). Die Entstehung der Kontinente. Geologisches Rundschau, 3, 276–292.CrossRefGoogle Scholar
Welhan, J. K., Simoneit, B. R. T. and Tarafa, M. (1988). C1–C8 hydrocarbons in sediments from Guaymas Basin, Gulf of California: Comparison to Peru Margin, Japan Trench and California Borderlands. Organic Geochemistry, 12, 171–194.Google Scholar
Welhan, J. A. (1988). Origins of methane in hydrothermal systems. Chemical Geology, 71, 183–198.CrossRefGoogle Scholar
Wendt, J. I., Regelous, M., Niu, Y., Hékinian, R. and Collerson, K. D. (1999). Geochemistry of lavas from the Garrett Transform Fault: Insights into mantle heterogeneity beneath the eastern Pacific. Earth and Planetary Science Letters, 173, 271–284CrossRefGoogle Scholar
Wentworth, C. K. and Macdonald, G. A. (1953). Structures and forms of basaltic rocks in Hawai'i. U.S. Geological Survey Bulletin, 994, 1–98.Google Scholar
West, M., Menke, W., Tolstoy, M., Webb, S. and Sohn, R. (2001). Magma storage beneath Axial Volcano on the Juan de Fuca mid-ocean ridge. Nature, 413, 833–836.CrossRefGoogle Scholar
Wetzel, L. R., Weins, D. A. and Kleinrock, M. C. (1993). Evidence from earthquakes for bookshelf faulting at large non-transform ridge offsets. Nature, 362, 235–237.CrossRefGoogle Scholar
Wheat, C. G. and Fisher, A. T. (2007). Seawater recharge along an eastern bounding fault in Middle Valley, northern Juan de Fuca Ridge. Geophysical Research Letters, 34, L20602, doi:10.1029/2007GL031347.CrossRefGoogle Scholar
White, R. S. and McKenzie, D. P. (1989). Magmatism in rift zones: The generation of volcanic continental margins and flood basalts. Journal of Geophysical Research, 94, 7685–7729.CrossRefGoogle Scholar
White, R. S., Bown, J. W. and Smallwood, J. R. (1995). The temperature of the Iceland plume and the origin of outward-propagating V-shaped ridges. Journal of the Geological Society, London, 152, 1039–1045.CrossRefGoogle Scholar
White, R. S., McKenzie, D. P. and O'Nions, R. K. (1992). Oceanic crustal thickness from seismic measurements and rare earth element inversions. Journal of Geophysical Research, 97, 19,683–19,715.CrossRefGoogle Scholar
White, R. S., Minshull, T. A., Bickle, M. J. and Robinson, C. J. (2001). Melt generation at very slow-spreading ocean ridges: Constraints from geochemical and geophysical data. Journal of Petrology, 42, 1171–1196.CrossRefGoogle Scholar
White, S. M., Macdonald, K. C. and Haymon, R. M. (2000). Basaltic lava domes, lava lakes, and volcanic segmentation on the southern East Pacific Rise. Journal of Geophysical Research, 105, 23,591–23,536.CrossRefGoogle Scholar
White, S. M., Mason, J. L., Macdonald, K. C.et al. (2009). Significance of widespread low effusion rate eruptions over the past two million years for delivery of magma to the overlapping spreading centers at 9°N East Pacific Rise. Earth and Planetary Science Letters, 280, 175–184, doi:10.1016/j.epsl.2009.01.030.CrossRefGoogle Scholar
White, S. M., Meyer, J. D., Haymond, R. M.et al. (2008). High-resolution surveys along the hotspot-affected Galápagos Spreading Center: 2. Influence of magma supply on volcanic morphology. Geochemistry, Geophysics, Geosystems, 9, Q09004, doi:10.1029/2008GC002036.CrossRefGoogle Scholar
White, S. N., Chave, A. D. and Reynolds, G. T. (2002). Investigations of ambient light emission at deep-sea hydrothermal vents. Journal of Geophysical Research, 107, doi:10.1029/2000JB000015.CrossRefGoogle Scholar
White, W. M. and Klein, E. M. (2014). Composition of the oceanic crust. In Holland, H. D. and Turekian, K. K. (eds.), Treatise on Geochemistry (Second Edition). Oxford: Elsevier, pp. 457–496, doi:org/10.1016/B978-0-08-095975-7.00315-6.CrossRefGoogle Scholar
Wilcock, W. S. D. (2004). Physical response of mid-ocean ridge hydrothermal systems to local earthquakes. Geochemistry, Geophysics, Geosystems, 5(11), Q11009, doi.org/10.1029/2004GC000701.CrossRefGoogle Scholar
Wilcock, W. S. D. and Delaney, J. R. (1996). Mid-ocean ridge sulfide deposits: Evidence for heat extraction from magma chambers or cracking fronts?Earth and Planetary Science Letters, 145, 49–64.CrossRefGoogle Scholar
Wilcock, W. S. D. and Fisher, A. T. (2004). Geophysical constraints on the subseafloor environment near mid-ocean ridges. In Wilcock, W. D., Kelley, D. S., Baross, J. A., DeLong, E. and Cary, C. (eds.), The Subseafloor Biosphere at Mid-Ocean Ridges. Geophysical Monograph 144, Washington, DC: American Geophysical Union, pp. 51–74.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, 107(B12), 2336, doi:org/10.1029/2001JB000505.CrossRefGoogle Scholar
Wilcock, W. S. D., DeLong, E. D., Kelley, D. S., Baross, J. A. and Cary, S. C. (2004). The Subseafloor Biosphere at Mid-Ocean Ridges. Geophysical Monograph 144, Washington DC: American Geophysical Union.CrossRefGoogle Scholar
Wilcock, W. S. D., Hooft, E. E., McGill, P. R.et al. (2007). Microearthquakes beneath the Endeavour hydrothermal vent fields: Insights into reaction zone processes. Eos, Transactions of the American Geophysical Union, 88(52), Fall Meeting Supplement, Abstract S13F-02.Google Scholar
Wilcock, W. S. D., Hooft, E. E. E., Toomey, D. R.et al. (2009). The role of magma injection in localizing black smoker activity. Nature Geoscience, 2, 509–513.CrossRefGoogle Scholar
Williams, D. L., Becker, K., Lawver, L. A. and Von Herzen, R. P. (1979). Heat flow at the spreading centers of the Guaymas Basin, Gulf of California. Journal of Geophysical Research, 84, 6757–6796.CrossRefGoogle Scholar
Wilson, D. S., Teagle, D. A. H., Alt, J. A.et al. (2006). Drilling to gabbro in intact oceanic crust. Science, 312, 1016–1020.CrossRefGoogle Scholar
Wilson, J. T. (1965). A new class of faults and their bearing on continental drift. Nature, 207, 343–347.CrossRefGoogle Scholar
Wilson, L. and Head, J. W. (1981). Ascent and eruption of basaltic magma on the Earth and Moon. Journal of Geophysical Research, 86(B4), 2971–3001.CrossRefGoogle Scholar
Wilson, L. and Head, J. W. (2007). Explosive volcanic eruptions on Mars: Tephra and accretionary lapilli formation, dispersal and recognition in the geologic record. Journal of Volcanology and Geothermal Research, 163, 83–97.CrossRefGoogle Scholar
Wilson, L. and Parfitt, E. A. (1993). The formation of perched lava ponds on basaltic volcanoes: The influence of flow geometry on cooling-limited lava flow lengths. Journal of Volcanology and Geothermal Research, 56(1), 113–123.CrossRefGoogle Scholar
Winn, C. D., Karl, D. M. and Massoth, G. J. (1986). Microorganisms in deep-sea hydrothermal plumes. Nature, 320, 644–746.CrossRefGoogle Scholar
Wirsen, C. O., Tuttle, J. H. and Jannasch, H. W. (1986). Activities of sulfur-oxidizing bacteria at the 21°N East Pacific Rise vent site. Marine Biology, 92, 449–456.CrossRefGoogle Scholar
Wolfe, C. J. and Solomon, S. C. (1998). Shear-wave splitting and implications for mantle flow beneath the MELT region of the East Pacific Rise. Science, 280, 1230–1232.CrossRefGoogle ScholarPubMed
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, 24,449–24,472.CrossRefGoogle Scholar
Workman, R. and Hart, S. R. (2005). Major and trace element composition of the depleted MORB mantle (DMM). Earth and Planetary Science Letters, 321, 53–72.CrossRefGoogle Scholar
Wright, T. J., Ebinger, C. J., Biggs, J.et al. (2006). Magma-maintained rift segmentation at continental rupture in the 2005 Afar dyking episode. Nature, 442, 291–294, doi:10.1038/nature04978.CrossRefGoogle ScholarPubMed
Wyllie, P. J. (ed.) (1967). Ultramafic and Related Rocks. New York: John Wiley and Sons.Google Scholar
Yang, H. J., Kinzler, R. J. and Grove, T. L. (1996). Experiments and models of anhydrous, basaltic olivine-plagioclase-augite saturated melts from 0.001 to 10 kbar. Contributions to Mineralogy and Petrology, 124(1), 1–18.CrossRefGoogle Scholar
Yaouancq, G. and MacLeod, C. J. (2000). Petrofabric investigation of gabbros from the Oman Ophiolite: Comparison between AMS and rock fabric. Marine Geophysical Researches, 21, 289–305.CrossRefGoogle Scholar
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, doi:10.1002/ggge.20082.CrossRefGoogle Scholar
Yeo, I., Searle, R. C., Achenbach, K., Lebas, T. and Murton, B. J. (2012). Eruptive hummocks: Building blocks of the upper ocean crust. Geology, 40, 91–94, doi:10.1130/G31892.CrossRefGoogle Scholar
Yoerger, D. R., Jakuba, M., Bradley, A. M. and Bingham, B. (2007). Techniques for deep sea near bottom survey using an autonomous underwater vehicle. International Journal of Robotics Research, 26(1), 41–54.CrossRefGoogle Scholar
Yoerger, D. R., Kelley, D. S. and Delaney, J. R. (1999). Fine-scale three dimensional mapping and imaging of the Mothra hydrothermal field, Juan de Fuca Ridge, Eos, Transactions of the American Geophysical Union, 80, 85.Google Scholar
Young, K. D., Jancin, M., Voight, B. and Orkan, N. I. (1985). Transform deformation of Tertiary rocks along the Tjörnes Fracture Zone, north central Iceland. Journal of Geophysical Research, 90, 9986–10,010.CrossRefGoogle Scholar
Zhao, M., Canales, J. P., Sohn, R. A. (2012). Three-dimensional seismic structure of a Mid-Atlantic Ridge segment characterized by active detachment faulting (Trans-Atlantic Geotraverse, 25°55′N-26°20′N. Geochemistry, Geophysics, Geosystems, 13, Q0AG13, doi:10.1029/2012GC004454.CrossRefGoogle Scholar
Zhu, W., Tivey, M. K., Gittings, H. and Craddock, P. R. (2007). Permeability–porosity relationships in seafloor vent deposits: Dependence on pore evolution processes. Journal of Geophysical Research Letters, 112, B05208, doi:10.1029/2006JB004716.Google Scholar
Zierenberg, R. A., Koski, R. A., Morton, J. L. and Bouse, R. M. (1993). Genesis of massive sulfide deposits on a sediment-covered spreading center, Escanaba Trough, southern Gorda Ridge. Economic Geology, 88(8), 2069–2098, doi:10.2113/gsecongeo.88.8.2069.CrossRefGoogle Scholar
Zierenberg, R. A., Shanks, W. C., Seyfried, W. E., Koski, R. A. and Strickler, M. D. (1988). Mineralization, alteration, and hydrothermal metamorphism of the ophiolite-hosted Turner-Albright sulfide deposit, southwestern Oregon. Journal of Geophysical Research, 93(B5), 4657–4674, doi:10.1029/JB093iB05p04657.CrossRefGoogle Scholar
Zonenshain, L. P., Kuzmin, M. I., Lisitsin, A. P., Bogdanov, Y. A. and Baranov, B. V. (1989). Tectonics of the Mid-Atlantic Rift Valley between the TAG and MARK Areas (26°–24°N): Evidence for vertical tectonism. Tectonophysics, 159, 1–23.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.

Available formats
×

Save book to Dropbox

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

Available formats
×

Save book to Google Drive

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

Available formats
×