Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-25T14:52:52.855Z Has data issue: false hasContentIssue false
This chapter is part of a book that is no longer available to purchase from Cambridge Core

References

John Ridley
Affiliation:
Colorado State University
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
Ore Deposit Geology , pp. 369 - 391
Publisher: Cambridge University Press
Print publication year: 2013

Access options

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

References

Adamides, N. G. (2010). Mafic-dominated volcanogenic sulphide deposits in the Troodos ophiolite, Cyprus Part 2 – a review of genetic models and guides for exploration. Applied Earth Science (Transactions of the Institution of Mining and Metallurgy) B 119, 193–204.Google Scholar
Ahrens, L. H. (1954). The lognormal distribution of the elements (A fundamental law of geochemistry and its subsidiary). Geochimica et Cosmochimica Acta 5, 49–73.Google Scholar
Albarede, F. (2009). Geochemistry, An Introduction, 2nd edition. Cambridge, Cambridge University Press.
Alpers, C. N. and Brimhall, G. H. (1989). Paleohydrologic evolution and geochemical dynamics of cumulative supergene metal enrichment at La Escondida, Atacama Desert, northern Chile. Economic Geology 84, 299–255.Google Scholar
Anand, R. R. (1994). Regolith-Landform Evolution and Geochemical Dispersion from the Boddington Gold Deposit, Western Australia. CRC LEME Open File Report 3, Perth, CRC LEME,
Anand, R. R. and Paine, M. (2002). Regolith geology of the Yilgarn Craton, Western Australia: implications for exploration. Australian Journal of Earth Sciences 49, 3–162.Google Scholar
Annen, C., Blundy, J. D. and Sparks, R. S. J. (2006). The genesis of intermediate and silicic magmas in deep crustal hot zones. Journal of Petrology 47, 505–539.Google Scholar
Appold, M. S. and Garven, G. (1999). The hydrology of ore formation in the southeast Missouri district: numerical models of topography-driven fluid flow during the Ouachita orogeny. Economic Geology 94, 913–936.Google Scholar
Arai, S. (1997). Origin of podiform chromitites. Journal of Asian Earth Sciences 15, 303–310.Google Scholar
Arribas, A. J., Hedenquist, J. W., Itaya, T. et al. (1995). Contemporaneous formation of adjacent porphyry and epithermal Cu–Au deposits over 300 ka in northern Luzon, Philippines. Geology 23, 337–340.Google Scholar
Atkinson, W. W. and Einaudi, M. T. (1978). Skarn formation and mineralization in the contact aureole at Carr Fork, Bingham, Utah. Economic Geology 73, 1326–1365.Google Scholar
Audétat, A. and Pettke, T. (2003). The magmatic-hydrothermal evolution of two barren granites: a melt and fluid inclusion study of the Rito del Medio and Cañada Pinabete plutons in northern New Mexico (USA). Geochimica et Cosmochimica Acta 67, 97–121.Google Scholar
Audétat, A., Günther, D. and Heinrich, C. A. (2000). Causes for large-scale zonation around mineralized plutons: fluid inclusion LA-ICP-MS evidence from the Mole Granite, Australia. Economic Geology 95, 1563–1582.Google Scholar
Audétat, A., Pettke, T., Heinrich, C. A. and Bodnar, R. J. (2008). The compositions of magmatic-hydrothermal fluids in barren and mineralized intrusions. Economic Geology 103, 877–908.Google Scholar
Audétat, A., Dolejs, D. and Lowenstern, J. B. (2011). Molybdenite saturation in silicic magmas: occurrence and petrological implications. Journal of Petrology 52, 891–904.Google Scholar
Ayres, D. E., Wray, E. M., Farstad, J. and Ibrahim, H. (1983). Geology of the Midwest uranium deposit. Geological Survey of Canada Paper 82–11, 33–40.Google Scholar
Ballhaus, C. G. and Stumpfl, E. F. (1986). Sulfide and platinum mineralization in the Merensky Reef: evidence from hydrous silicates and fluid inclusions. Contributions to Mineralogy and Petrology 94, 193–204.Google Scholar
Barnicoat, A. C., Henderson, I. H. C., Knipe, R. J. et al. (1997). Hydrothermal gold mineralization in the Witwatersrand Basin. Nature 386, 820–824.Google Scholar
Barton, C. A., Zoback, M. D. and Moos, D. (1995). Fluid-flow along potentially active faults in crystalline rock. Geology 23, 683–686.Google Scholar
Barton, P. B. (1991). Ore textures: problems and opportunities. Mineralogical Magazine 55, 303–315.Google Scholar
Baturin, G. N. (1989). Origin of marine phosphorite. International Geology Review 31, 327–342.Google Scholar
Bekker, A., Slack, J. F., Planavsky, N. et al. (2010). Iron formations: the sedimentary product of a complex interplay among mantle, tectonic, oceanic, and biospheric process. Economic Geology 105, 467–508.Google Scholar
Bentley, H. W., Phillips, F. M., Davis, S. N. et al. (1986). Chlorine 36 dating of very old groundwater. The Great Artesian Basin, Australia. Water Resources Research 22, 1991–2001.Google Scholar
Berning, J., Cooke, R., Hiemstra, S. A. and Hoffman, U. (1976). The Rössing uranium deposit, South West Africa. Economic Geology 71, 351–368.Google Scholar
Berry, A. J., Harris, A. C., Kamenetsky, V. S., Newville, M. and Sutton, S. R. (2009). The speciation of copper in natural fluid inclusions at temperatures up to 700 °C. Chemical Geology 259, 2–7.Google Scholar
Bethke, P. M. (1988). The Creede, Colorado ore-forming system: a summary model. US Geological Survey Open-File Report 88–403.Google Scholar
Bethke, P. M., Rye, R. O., Stoffregen, R. E. and Vikre, P. G. (2005). Evolution of the magmato-hydrothermal acid-sulfate system at Summitville, Colorado: integration of geological, stable isotope, and fluid-inclusion evidence. Chemical Geology 215, 281–315.Google Scholar
Beukes, N. J. and Gutzmer, J. (1996). A volcanic-exhalative origin for the world’s largest (Kalahari) manganese field. A discussion of the paper by D. H. Cornell and S. S. Schütte. Mineralium Deposita 31, 242–245.Google Scholar
Binns, R. A., Barriga, F. J. A. S. and Miller, D. J. (2007). Leg 193 synthesis: anatomy of an active felsic-hosted hydrothermal system, Eastern Manus Basin, Papua New Guinea. In Proceedings ODP Scientific Results 193, Barriga, F. J. A. S., Binns, R. A., Miller, D. J. and Herzig, P. M. (eds.), College Station, TX, Ocean Drilling Program, pp. 1–71.
Birch, G. J. and Buchanan, D. L. (1989). Controls on the distribution of nickel sulphide mineralization associated with the Madziwa mafic intrusion, Zimbabwe. In Magmatic Sulphides – The Zimbabwe Volume, Prendergast, M. D. and Jones, M. J. (eds.), London, Institution of Mining and Metallurgy, pp. 21–42.
Blevin, P. L. (2004). Redox and compositional parameters for interpreting the granitoid metallogeny of eastern Australia: implications for gold-rich ore systems. Resource Geology 54, 241–252.Google Scholar
Blevin, P. L. (2010). Eastern Australian granites: origins and metallogenesis. Conference presentation, SMEDG, Wines and Mines, Mudgee, New South Wales, 24–25 September 2010. See: .
Böhlke, J. K. (1999). Mother Lode gold. Geological Society of America Special Paper 338, 55–67.Google Scholar
Bolton, B. R., Barents, H. W. and Frakes, L. A. (1990). Groote-Eylandt manganese deposit. In Geology and Mineral Deposits of Australia and Papua New Guinea, Hughes, F. E. (ed.), Melbourne, Australasian Institute of Mining and Metallurgy, Monograph 14, pp.1575–1579.
Boni, M. (1985). Les gisements de type Mississippi Valley du sud-ouest de le Sardaigne (Italie), une synthèse. Chronique de la Recherche Minière 479, 7–34.Google Scholar
Boudreau, A. (2008). Modeling of the Merensky Reef, Bushveld Complex, Republic of South Africa. Contributions to Mineralogy and Petrology 156, 431–437.Google Scholar
Boudreau, A. E., Mathez, E. A. and McCallum, I. S. (1986). Halogen geochemistry of the Stillwater and Bushveld complexes: evidence for transport of the platinum group elements by Cl-rich fluids. Journal of Petrology 27, 967–986.Google Scholar
Brauhart, C., Groves, D. I. and Morant, P. (1998). Regional alteration systems associated with volcanogenic massive sulfide mineralization at Panorama, Pilbara, Western Australia. Economic Geology 93, 292–302.Google Scholar
Brauhart, C. W., Huston, D. L., Groves, D. I., Mikucki, E. J. and Gardoll, S. J. (2001). Geochemical mass-transfer patterns as indicators of the architecture of a complete volcanic-hosted massive sulfide hydrothermal system, Panorama district, Pilbara, Western Australia. Economic Geology 96, 1263–1278.Google Scholar
Broadbent, G. C., Myers, R. E. and Wright, J. V. (1998). Geology and origin of shale-hosted Zn–Pb–Ag mineralization at the Century deposit, Northwest Queensland, Australia. Economic Geology 93, 1264–1294.Google Scholar
Brown, A. C. (1971). Zoning in the White Pine copper deposit, Ontonagon County, Michigan. Economic Geology 66, 543–573.Google Scholar
Brown, A. C. (2006). Genesis of native copper lodes in the Keweenaw district, northern Michigan: a hybrid evolved meteoric and metamorphogenic model. Economic Geology 101, 1437–1444.Google Scholar
Brown, K. L. (1986). Gold deposition from geothermal discharges in New Zealand. Economic Geology 81, 979–983.Google Scholar
Buck, S. G. and Minter, W. E. L. (1985). Placer formation by fluvial degradation of an alluvial fan sequence: the Proterozoic carbon leader placer, Witwatersrand Supergroup, South Africa. Geological Society of London Journal 142, 757–764.Google Scholar
Burrows, D. R., Spooner, E. T. C., Wood, P. C. and Jemielita, R. A. (1993). Structural controls on formation of the Hollinger-McIntyre Au quartz vein system in the Hollinger shear zone, Timmins, southern Abitibi greenstone-belt, Ontario. Economic Geology 88, 1643–1663.Google Scholar
Cagnioncle, A-M., Parmentier, E. M. and Elkins-Tanton, L. T. (2007). Effect of solid flow above a subducting slab on water distribution and melting at convergent plate boundaries. Journal of Geophysical Research 112, article B090402.Google Scholar
Calabrese, S., Aiuppa, A., Allard, P. et al. (2011). Atmospheric sources and sinks of volcanogenic elements in a basaltic volcano (Etna, Italy). Geochimica et Cosmochimica Acta 75, 7401–7425.Google Scholar
Cameron, E. (1990). Yeelirrie uranium deposit. In Geology and Mineral Deposits of Australia and Papua New Guinea, Hughes, F. E. (ed.), Melbourne, Australasian Institute of Mining and Metallurgy, Monograph 14, pp. 1625–1629.
Campbell, I. H., Naldrett, A. J. and Barnes, S. J. (1983). A model for the origin of platinum-rich sulfide horizons in the Bushveld and Stillwater complexes. Journal of Petrology 24, 133–165.Google Scholar
Carman, G. D. (2003). Geology, mineralization, and hydrothermal evolution of the Ladolam Gold Deposit, Lihir Island, Papua New Guinea. Society of Economic Geologists Special Publication 10, 247–284.Google Scholar
Carr, H. W., Groves, D. I., and Cawthorn, R. G. (1994). Controls on the distribution of Merensky Reef potholes at the Western Platinum Mine, Bushveld Complex, South Africa: implications for disruptions of layering and pothole formation in the complex. South African Journal of Geology 97, 431–441.Google Scholar
Castor, S. B. (2008). The Mountain Pass rare-earth carbonatite and associated ultrapotassic rocks, California. Canadian Mineralogist 46, 779–806.Google Scholar
Catchpole, H., Kouzmanov, K., Fontbote, L., Guillong, M. and Heinrich, C. A. (2011). Fluid evolution in zoned Cordilleran polymetallic veins – insights from microthermometry and LA-ICP-MS of fluid inclusions. Chemical Geology 281, 293–304.Google Scholar
Cathles, L. M. (1977). An analysis of the cooling of intrusives by ground water convection which includes boiling. Economic Geology 72, 804–826.Google Scholar
Cathles, L. M. (1991). The importance of vein salvaging in controlling the intensity and character of subsurface alteration in hydrothermal systems. Economic Geology 86, 466–471.Google Scholar
Cathles, L. M. (1993). Mass balance evaluation of the late diagenetic hypothesis for Kupferschiefer Cu mineralization in the Lubin Basin of southwestern Poland. Economic Geology 88, 948–956.Google Scholar
Cathles, L. M. and Adams, J. J. (2005). Fluid-flow and petroleum and mineral resources in the upper (< 20-km) continental crust. In Economic Geology, 100th Anniversary Volume, Hedenquist, J. W., Thompson, J. F. H., Goldfarb, R. J. and Richards, J. P. (eds.), Colorado, Society of Economic Geologists, pp. 77–110.
Cathles, L. M. and Shannon, R. (2007). How potassium silicate alteration suggests the formation of porphyry ore deposits begins with the nearly explosive but barren expulsion of large volumes of magmatic water. Earth and Planetary Science Letters 262, 92–108.Google Scholar
Cawthorn, R. G. (2011). Geological interpretations from the PGE distribution in the Bushveld Merensky and UG2 chromitite reefs. Journal of the Southern African Institute of Mining and Metallurgy 111, 67–79.Google Scholar
Cerny, P., Ercit, T. S. and Vanstone, P. T. (1996). Petrology and mineralization of the Tanco rare-element pegmatite, southeastern Manitoba. In Field Trip Guidebook A-4, Winnipeg, Geological Association of Canada–Mineralogical Association of Canada, 1996.
Chai, G. and Naldrett, A. J. (1992). Characteristics of Ni–Cu–PGE mineralization and genesis of the Jinchuan Deposit, Northwest China. Economic Geology 87, 1475–1495.Google Scholar
Chang, Z., Hedenquist, J. W., White, N. C. et al. (2011). Exploration tools for linked porphyry and epithermal deposits: example from the Mankayan intrusion-centered Cu–Au district, Luzon, Philippines. Economic Geology 106, 1365–1398.Google Scholar
Charlier, B., Duchesne, J.-C. and Vander Auwera, J. (2006). Magma chamber processes in the Tellnes ilmenite deposit (Rogaland Anorthosite Province, SW Norway) and the formation of Fe–Ti ores in massif-type anorthosites. Chemical Geology 234, 264–290.Google Scholar
Chen, J., Halls, C. and Stanley, C. J. (1992). Tin-bearing skarns of south China – geological setting and mineralogy. Ore Geology Reviews 7, 225–248.Google Scholar
Chi, R. (2008). Weathered Crust Elution Deposited Rare Earth Ores. New York, Nova Science Publications.
Chouinard, A., William-Jones, A. E., Leonardson, R. W. et al. (2005). Geology and genesis of the multistage high-sulfidation epithermal Pascua Au–Ag–Cu deposit, Chile and Argentina. Economic Geology 100, 463–490.Google Scholar
Cline, J. S. and Bodnar, R. J. (1991). Can economic porphyry copper mineralization be generated by a typical calc-alkaline melt? Journal of Geophysical Research B 96, 8113–8126.Google Scholar
Cloud, P. (1973). Paleoecological significance of banded iron-formation. Economic Geology 68, 1135–1143.Google Scholar
Clout, J. M. F. and Simonson, B. M. (2005). Precambrian iron formations and iron formation-hosted iron ore deposits. In Economic Geology, 100th Anniversary Volume, Hedenquist, J. W., Thompson, J. F. H., Goldfarb, R. J. and Richards, J. P. (eds.), Colorado, Society of Economic Geologists, pp. 643–679.
Clout, J. M. F., Cleghorn, J. H. and Eaton, P. C. (1990). Geology of the Kalgoorlie gold field. In Geology and Mineral Deposits of Australia and Papua New Guinea, Hughes, F. E. (ed.), Melbourne, Australasian Institute of Mining and Metallurgy, Monograph 14, pp. 411–431.
Conmou, D., Driesner, T. and Heinrich, C. A. (2008). The structure and dynamics of mid-ocean ridge hydrothermal systems. Science 321, 1825–1828.Google Scholar
Constantinou, G. and Govett, G. J. S. (1978). Geology, geochemistry, and genesis of Cyprus sulfide deposits. Economic Geology 68, 843–858.Google Scholar
Cooke, D. R. and Simmons, S. F. (2000). Characteristics and genesis of epithermal gold deposits. Reviews in Economic Geology 13, 221–244.Google Scholar
Coward, M. P., Spencer, R. M. and Spencer, C. E. (1995). Development of the Witwatersrand Basin, South Africa. Geological Society Special Publication 95, 243–269.Google Scholar
Cowden, A. and Roberts, D. E. (1990). Komatiite-hosted nickel sulphide deposits, Kambalda. In Geology of the Mineral Deposits of Australia and Papua New Guinea, Hughes, F. E. (ed.), Melbourne, Australasian Institute of Mining and Metallurgy, Monograph 14, pp. 567–581.
Cox, S. F. (2005). Coupling between deformation, fluid pressures, and fluid flow in ore-producing hydrothermal systems at depth in the crust. In Economic Geology 100th Anniversary Volume, Hedenquist, J. W., Thompson, J. F. H., Goldfarb, R. J. and Richards, J. P. (eds.), Colorado, Society of Economic Geologists, pp. 39–76.
Cox, S. F. and Ruming, K. (2004). The St Ives mesothermal gold system, Western Australia – a case of golden aftershocks? Journal of Structural Geology 26, 1109–1125.Google Scholar
Craig, J. R. and Vaughan, D. J. (1994). Ore Microscopy and Ore Mineralogy, 2nd Edition, New York, John Wiley.
Crerar, D., Wood, S., Brantley, S. and Bocarsly, A. (1985). Chemical controls on solubility of ore forming minerals in hydrothermal solutions. Canadian Mineralogist 23, 333–352.Google Scholar
Cronan, D. S. (2000). Handbook of Marine Mineral Deposits. Baton Rouge, CRC Press.
Dalton, J. A. and Presnall, D. C. (1998). Carbonatitic melts along the solidus of model lherzolite in the system CaO–MgO–Al2O3–SiO2–CO2 from 3 to 7 GPa. Contributions to Mineralogy and Petrology 131, 123–135.Google Scholar
Dasgupta, R., Hirschmann, M. M. and Smith, N. D. (2007). Partial melting experiments of peridotite + CO2 at 3 GPa and genesis of alkalic ocean island basalts. Journal of Petrology 48, 2093–2124.Google Scholar
Dasgupta, R., Hirschmann, M. M., McDonough, W. F., Spiegelman, M. and Withers, A. C. (2009). Trace element partitioning between garnet lherzolite and carbonatite at 6.6 and 8.6 GPa with applications to the geochemistry of the mantle and of mantle-derived melts. Chemical Geology 262, 57–77.Google Scholar
Davidson, G. J., Paterson, H., Meffre, S. and Berry, R. F. (2007). Characteristics and origin of the Oak Dam East breccia-hosted, iron oxide Cu–U–(Au) deposit: Olympic Dam region, Gawler Craton, South Australia. Economic Geology 102, 1471–1498.Google Scholar
Deb, M. and Goodfellow, W. D. (eds.) (2004). Sediment-Hosted Lead–Zinc Sulphide Deposits: Attributes and Models of Some Major Deposits in India, Australia and Canada, New Delhi, Narosa Publishing House, p. 367.
DeMatties, T. A. (1994). Early Proterozoic volcanogenic massive sulfide deposits in Wisconsin: an overview. Economic Geology 89, 1122–1151.Google Scholar
Deming, D. and Nunn, J. A. (1991). Numerical simulation of brine migration by topographically driven recharge. Journal of Geophysical Research 96, 2485–2499.Google Scholar
de Ronde, C. E. J., Spooner, E. T. C., de Wit, M. J. and Bray, C. J. (1992). Shear zone-related, Au quartz vein deposits in the Barberton greenstone belt, South Africa: field and petrographic characteristics, fluid properties, and light stable isotope geochemistry. Economic Geology 87, 366–402.Google Scholar
de Ronde, C. E. J., Faure, K., Bray, C. J. and Whitford, D. J. (2000). Round Hill shear zone-hosted gold deposit, Macraes Flat, Otago, New Zealand: evidence of a magmatic fluid. Economic Geology 95, 1025–1048.Google Scholar
Dickinson, W. R. (2004). Evolution of the North American Cordillera. Annual Review of Earth and Planetary Sciences 32, 13–45.Google Scholar
Dietz, R. S. (1964). Sudbury structure as an astrobleme. Journal of Geology 72, 412–434.Google Scholar
Dilles, J. H. and Einaudi, M. T. (1992). Wall-rock alteration and hydrothermal flow paths about the Ann-Mason porphyry copper deposit, Nevada – a 6-km vertical reconstruction. Economic Geology 87, 1963–2001.Google Scholar
Distler, V. V., Kryachko, V. V. and Yudovskaya, M. A. (2008). Ore petrology of chromite-PGE mineralization in the Kempirsai ophiolite complex. Mineralogy and Petrology 92, 31–58.Google Scholar
Dong, G., Morrison, G. and Jaireth, S, (1995). Quartz textures in epithermal veins, Queensland – classification, origin and implication. Economic Geology 90, 1841–1856.Google Scholar
Duke, J. M. (1983). Ore deposit models; 7, Magmatic segregation deposits of chromite. Geoscience Canada 10, 15–24.Google Scholar
Eastoe, C. J. and Gustin, M. M. (1996). Volcanogenic massive sulfide deposits and anoxia in Phanerozoic oceans. Ore Geology Reviews 13, 179–197.Google Scholar
Eckstrand, O. R. and Hulbert, L. (2007). Magmatic nickel–copper–platinum group element deposits. Geological Association of Canada Mineral Deposits Division Special Publication 5, 205–222.Google Scholar
Eilu, P. and Mikucki, E. J. (1998). Alteration and primary geochemical dispersion associated with the Bulletin lode-gold deposit, Wiluna, Western Australia. Journal of Geochemical Exploration 63, 73–103.Google Scholar
Einaudi, M. T. (1982). Descriptions of skarns associated with porphyry copper plutons. In Advances in Geology of the Porphyry Copper Deposits Southwestern North America, Titley, S. R. (ed.), Tucson, University of Arizona Press, pp. 139–183.
Einaudi, M. T., Hedenquist, J. W. and Esra Inan, E. (2003). Sulfidation state of fluids in active and extinct hydrothermal systems: transitions from porphyry to epithermal environments. Society of Economic Geologists Special Publication 10, 285–313.Google Scholar
Els, B. G. (1991). Placer formation during progradational fluvial degradation: the late Archean Middlevlei gold placer, Witwatersrand, South Africa. Economic Geology 86, 261–277.Google Scholar
Els, B. G. (2000). Unconformities of the auriferous, Neoarchaean Central Rand Group of South Africa: application to stratigraphy. Journal of African Earth Sciences 30, 47–62.Google Scholar
Emsbo, P., Hofstra, A. H., Lauha, E. A., Griffin, G. L. and Hutchinson, R. W. (2003). Origin of high-grade gold ore, source of ore fluid components, and genesis of the Meikle and neighboring Carlin-type deposits, northern Carlin trend, Nevada. Economic Geology 98, 1069–1106.Google Scholar
England, G. L., Rasmussen, B. Krapez, B., and Groves, D. I. (2002a). Archaean oil migration in the Witwatersrand Basin of South Africa. Journal of the Geological Society London 159, 189–201.Google Scholar
England, G. L., Rasmussen, B., Krapez, B. and Groves, D. I. (2002b). Palaeoenvironmental significance of rounded pyrite in siliciclastic sequences of the late Archaean Witwatersrand Basin: oxygen-deficient atmosphere or hydrothermal alteration? Sedimentology 49, 1133–1156.Google Scholar
Ericksen, G. E. (1981). Geology and origin of the Chilean nitrate deposits. US Geological Survey Professional Paper 1188.Google Scholar
Farquhar, J. and Wing, B. A. (2003). Multiple sulfur isotopes and the evolution of the atmosphere. Earth and Planetary Science Letters 213, 1–13.Google Scholar
Farquhar, J., Bau, H. M. and Thiemens, M. (2000). Atmospheric influence of Earth’s earliest sulfur cycle. Science 289, 756–758.Google Scholar
Farquhar, J., Wu, N., Canfield, D. E. and Oduro, H. (2010). Connections between sulfur cycle evolution, sulfur isotopes, sediments, and base metal sulfide deposits. Economic Geology 100, 509–533.Google Scholar
Farrow, C. E. G. and Lightfoot, P. C. (2002). Sudbury PGE revisited: toward an integrated model. Canadian Institute of Mining and Metallurgy, Special Volume 54, 273–297.Google Scholar
Field, M., Stiefenhofer, J., Robey, J. and Kurszlaukis, S. (2008). Kimberlite-hosted diamond deposits of southern Africa: a review. Ore Geology Reviews 34, 33–75.Google Scholar
Fleet, M. E. (1998). Detrital pyrite in Witwatersrand gold reefs: X-ray diffraction evidence and implications for atmospheric evolution. Terra Nova 10, 302–306.Google Scholar
Forster, C. and Smith, L. (1990). Fluid flow in tectonic regimes. In Crustal Fluids, Nesbitt, B. E. (ed.), MAC Short Course Handbook 18, Québec, Mineralogical Association of Canada, pp. 1–47.
Fournier, R. O. (1985). The behaviour of silica in hydrothermal solution. Reviews in Economic Geology 2, 45–62.Google Scholar
Francheteau, J., Needham, H. D., Choukroune, P. et al. (1979). Massive deep-sea sulphide ore deposits discovered on the East Pacific Rise. Nature 277, 523–528.Google Scholar
Garven, G. (1995). Continental scale groundwater flow and geologic processes. Annual Review of Earth and Planetary Sciences 23, 89–117.Google Scholar
Garven, G., Ge, S., Person, M. A. and Sverjensky, D. A. (1993). Genesis of stratabound ore deposits in the midcontinent basin of North America I. The role of regional groundwater flow. American Journal of Science 293, 497–568.Google Scholar
Gauthier-Lafaye, F. and Weber, F. (1989). The Francevillian (Lower Proterozoic) uranium ore deposits of Gabon. Economic Geology 84, 2267–2285.Google Scholar
Giggenbach, W. F. (1992). Magma degassing and mineral deposition in hydrothermal systems along convergent plate boundaries. Economic Geology 87, 1927–1944.Google Scholar
Gill, J. (1981). Orogenic Andesites and Plate Tectonics, New York, Springer.
Godel, B., Seat, Z., Maier, W. D. and Barnes, S-J. (2011). The Nebo-Babel Ni–Cu–PGE sulfide deposit (West Musgrave Block, Australia): Part 2. Constraints on parental magma and processes, with implications for exploration. Economic Geology 106, 557–584.Google Scholar
Goff, F., Stimac, J. A., Larocque, A. C. L. et al. (1994). Gold degassing and deposition at Galeras Volcano, Colombia. GSA Today 4, 243–247.Google Scholar
Golightly, J. P. (1981). Nickeliferous laterite deposits. Economic Geology, 75th Anniversary Volume, Skinner, B. (ed.), Lancaster, PA, Economic Geology Publishing, pp. 710–735.
Grant, J. N., Halls, C., Avila Salinas, B. W. and Avila, G. (1977). Igneous geology and the evolution of hydrothermal systems in some sub-volcanic tin deposits of Bolivia. In Volcanic Processes in Ore Genesis, London, Institution of Mining and Metallurgy, pp. 117–126.
Grauch, V. J. S., Rodriguez, B. D. and Wooden, J. L. (2003). Geophysical and isotopic constraints on crustal structure related to mineral trends in north-central Nevada and implications for tectonic history. Economic Geology 98, 26–286.Google Scholar
Gray, J. E. and Coolbaugh, M. F. (1994). Geology and geochemistry of Summitville, Colorado: an epithermal acid-sulfate deposit in a volcanic dome. Economic Geology 89, 1906–1923.Google Scholar
Gresham, J. J. and Loftus-Hills, G. D. (1981). The geology of the Kambalda nickel field, Western Australia. Economic Geology 76, 1373–1416.Google Scholar
Grove, D., and Harris, C. (2010). O- and H-isotope study of the carbon leader reef at the Tau Tona and Savuka mines (western deep levels), South Africa: implications of the origin and evolution of Witwatersrand basin fluids. South African Journal of Geology 113, 75–88.Google Scholar
Groves, D. I. and Vielreicher, N. M. (2001). The Phalabowra (Palabora) carbonatite-hosted magnetite–copper sulfide deposit, South Africa: an end-member of the iron oxide–copper–gold–rare earth element deposit group? Mineralium Deposita 36, 189–194.Google Scholar
Groves, D. I., Bierlein, F. P., Meinert, L. D. and Hitzman, M. W. (2010). Iron oxide copper–gold (IOCG) deposits through Earth history: implications for origin, lithospheric setting, and distribution from other epigenetic iron oxide deposits. Economic Geology 105, 641–654.Google Scholar
Gruen, G., Heinrich, C. A. and Schroeder, K. (2010). The Bingham Canyon porphyry Cu–Au–Mo deposit. II. Vein geometry and ore shell formation by pressure driven rock extension. Economic Geology 105, 69–90.Google Scholar
Grunder, A. L., Klemetti, E. W., Feeley, T. C. and McKee, C. M. (2008). Eleven million years of arc volcanism at the Aucanquilcha volcanic cluster, northern Chilean Andes: implications for the life span and emplacement of plutons. Transactions of the Royal Society of Edinburgh: Earth Sciences 97, 415–436.Google Scholar
Gustafson, L. B. and Hunt, J. P. (1975). The porphyry copper deposit at El Salvador, Chile. Economic Geology 70, 857–912.Google Scholar
Gutzmer, J. and Beukes, N. J. (1996). Mineral paragenesis of the Kalahari manganese field, South Africa. Ore Geology Reviews 11, 405–428.Google Scholar
Hagemann, S. G., Groves, D. I., Ridley, J. R. and Vearncombe, J. R. (1992). The Archean lode gold deposits at Wiluna, Western Australia: high-level brittle-style mineralization in a strike-slip regime. Economic Geology 87, 1022–1053.Google Scholar
Hamilton, P. J. (1979). Sr isotope and trace element studies of the Great Dyke and Bushveld mafic phases and their relation to Proterozoic magma genesis in southern Africa. Journal of Petrology 18, 24–52.Google Scholar
Hannington, M. T. and Barrie, C. T. (eds.) (1999). The Giant Kidd Creek Volcanogenic Massive Sulfide Deposit, Western Abitibi Subprovince, Canada. Economic Geology Monograph 10.Google Scholar
Hannington, M., Jamieson, J., Monecke, T., Petersen, S. and Beaulieu, S. (2011). The abundance of seafloor massive sulfide deposits. Geology 39, 1155–1158.Google Scholar
Hanor, J. S. (1994). Origin of saline fluids in sedimentary basins. Special Publications of the Geological Society 78, 151–174.Google Scholar
Harris, A. C., Golding, S. D. and White, N. C. (2005). Bajo de la Alumbrera copper–gold deposit: stable isotope evidence for a porphyry-related hydrothermal system dominated by magmatic aqueous fluids. Economic Geology 100, 863–886.Google Scholar
Harshman, E. N. (1972). Geology and uranium deposits, Shirley Basin area, Wyoming. US Geological Survey Professional Paper 745.Google Scholar
Haxel, G. B. (2005). Ultrapotassic mafic dikes and rare earth element- and barium-rich carbonatite at Mountain Pass, Mojave Desert, Southern California: summary and field trip localities. US Geological Survey Open-File Report 2005–1219.Google Scholar
Haydon, R. C. and McConachy, G. W. (1987). The stratigraphic setting of Pb–Zn–Ag mineralization at Broken Hill. Economic Geology 82, 826–856.Google Scholar
Hedenquist, J. W. and Lowenstern, J. B. (1994). The role of magmas in the formation of hydrothermal ore deposits. Nature 370, 519–527.Google Scholar
Hedenquist, J. W., Simmons, S. F., Giggenbach, W. F. and Eldridge, C. S. (1993). White Island, New Zealand, volcanic-hydrothermal system represents the geochemical environment of high-sulfidation Cu and Au ore deposition. Geology 21, 731–734.Google Scholar
Hedenquist, J. W., Matsuhisa, Y., Izawa, E. et al. (1994). Geology, geochemistry, and origin of high-sulfidation Cu–Au mineralization in the Nansatsu district, Japan. Economic Geology 89, 1–30.Google Scholar
Hedenquist, J. W., Arribas, A. and Reynolds, T. J. (1998). Evolution of an intrusion-centered hydrothermal system; far Southeast–Lepanto porphyry and epithermal Cu–Au deposits, Philippines. Economic Geology 93, 373–404.Google Scholar
Hedenquist, J. W., Arribas, A. R. and Gonzalez-Urien, E. (2000). Exploration for epithermal gold deposit. Reviews in Economic Geology 13, 245–277.Google Scholar
Heidrick, T. L. and Titley, S. R. (1982). Fracture and dike patterns in Laramide plutons and their structural and tectonic implications. In Advances in Geology of the Porphyry Copper Deposits Southwestern North America, Titley, S. R. (ed.), Tucson, University of Arizona Press, pp. 73–91.
Hein, J. R., Yen, H.-W., Gunn, S. H., Gibbs, A. E. and Wang, C.-H. (1994). Composition and origin of hydrothermal ironstones from central Pacific seamounts. Geochimica et Cosmochimica Acta 58, 179–189.Google Scholar
Hellston, K., Lewis, C. R. and Denn, S. (1998). Cawse nickel–cobalt deposit. In Geology of Australian and Papua New Guinean Mineral Deposits, Berkman, D. A. and Mackenzie, D. H. (eds.), Australasian Institute of Mining and Metallurgy, Monograph 22, pp. 335–338.
Henry, C. D., Elson, H. B., McIntosh, W. C., Heizler, M. T. and Castor, S. B. (1997). Brief duration of hydrothermal activity at Round Mountain, Nevada, determined from 40Ar/39Ar geochronology. Economic Geology 92, 807–826.Google Scholar
Herrington, R. J. and Wilkinson, J. J. (1993). Colloidal gold and silica in mesothermal vein systems. Geology 21, 539–542.Google Scholar
Hewitt, W. P. (1968). Geology and mineralization of the main mineral zone of the Santa Eulalia district, Chihuahua, Mexico. American Institute of Mining Engineers Transactions 241, 228–260.Google Scholar
Hiatt, E and Budd, D. A. (2003). Extreme paleoceanographic conditions in a Paleozoic oceanic upwelling system; productivity and widespread phosphogenesis in the Permian Phosphoria sea. Geological Society of America Special Paper 370, 245–264.Google Scholar
Hitzman, M. W. and Beaty, D. W. (1996). The Irish Zn–Pb–(Ba) orefield. Society of Economic Geologists Special Publication 4, 112–143.Google Scholar
Hitzman, M. W., Oreskes, N. and Einaudi, M. T. (1992). Geological characteristics and tectonic setting of Proterozoic iron oxide (Cu–U–Au–REE) deposits. Precambrian Research 58, 241–287.Google Scholar
Hitzman, M., Kirkham, R., Broughton, D., Thorson, J. and Selley, D. (2005). The sediment-hosted stratiform copper-ore systems. Economic Geology, 100th Anniversary Volume, Hedenquist, J. W., Thompson, J. F. H., Goldfarb, R. J. and Richards, J. P. (eds.), Colorado, Society of Economic Geologists, pp. 609–642.
Hodgson, C. J. (1989). Uses (and abuses) of ore deposit models in mineral exploration. Ontario Geological Survey Special Volume 3, 31–45.
Hofstra, A. H. and Cline, J. S. (2000). Characteristics and models for Carlin-type gold deposits. Reviews in Economic Geology 13, 163–220.Google Scholar
Hofstra, A. H., John, D. A. and Theodore, T. G. (2003). Preface: a special issue devoted to gold deposits in northern Nevada: part 2. Carlin-type deposits. Economic Geology 98, 1063–1068.Google Scholar
Holland, H. D. (1984). The Chemical Evolution of the Atmosphere and Oceans, Princeton, Princeton University Press.
Holliday, J. R., Wilson, A. J., Blevin, P. L. et al. (2002). Porphyry gold–copper mineralisation in the Cadia district, eastern Lachlan Fold Belt, New South Wales, and its relationship to shoshonitic magmatism. Mineralium Deposita 37, 100–116.Google Scholar
Horan, M., Morgan, J. W., Walker, R. J. and Cooper, R. W. (2001). Re–Os isotopic constraints on magma mixing in the Peridotite Zone of the Stillwater Complex, Montana, USA. Contributions to Mineralogy and Petrology 141, 446–457.Google Scholar
Hronsky, J. M. A, Groves, D. I., Loucks, R. R. and Begg, G. C. (2012). A unified model for gold mineralisation in accretionary orogens and implications for regional-scale exploration targeting methods. Mineralium Deposita 47, 339–358.Google Scholar
Hsieh, P. A. and Bredehoeft, J. D. (1981). A reservoir analysis of the Denver earthquakes – a case of induced seismicity. Journal of Geophysical Research B 86, 903–920.Google Scholar
Hughes, M. J., Phillips, G. N. and Carey, S. P. (2004). Giant placers of the Victorian Gold Province. Society of Economic Geologists Newsletter 56, 1–18.Google Scholar
Ibaraki, K. and Suzuki, R. (1993). Gold–silver quartz-adularia veins of the Main, Yamada and Sanjin deposits, Hishikari gold mine; a comparative study of their geology and ore deposits. Resource Geology Special Issue 14, 1–11.Google Scholar
Ildefonse, B. and Mancktelow, N. S. (1993). Deformation around rigid particles – the influence of slip at the particle matrix interface. Tectonophysics 221, 345–359.Google Scholar
Ingebritsen, S., Sanford, W. and Neuzil, C. (2006). Groundwater in Geologic Processes, 2nd Edition, Cambridge, Cambridge University Press.
Ingebritsen, S. E., Geiger, S., Hurwitz, S. and Driesner, T. (2010). Numerical simulation of magmatic hydrothermal systems. Reviews of Geophysics 48, RG1002.Google Scholar
Ionov, D. A., Dupuy, C. and O’Reilly, S. Y. (1993). Carbonated peridotite xenoliths from Spitsbergen – implications for trace-element signature of mantle carbonate metasomatism. Earth and Planetary Science Letters 119, 283–297.Google Scholar
Irvine, T. N. (1977). Origin of chromitite layers in the Muskox intrusion and other stratiform intrusions: a new interpretation. Geology 5, 273–277.Google Scholar
Janse, A. J. A. (1994). Is Clifford’s Rule still valid? Affirmative examples from around the World. In Proceedings of the Fifth International Kimberlite Conference, Araxa, Brazil, 1991, Volume 2, Meyer, H. O. A. and Leonardos, O. H. (eds.), CPRM Special Publication 1B, Oxford, Blackwell, pp. 215–235.
Jefferson, C. W., Thomas, D. J., Gandhi, S. S. et al. (2007). Unconformity-associated uranium deposits of the Athabasca Basin, Saskatchewan and Alberta. Geological Association of Canada Mineral Deposits Division Special Publication 5, 273–306.Google Scholar
Jelsma, H., Barnett, W., Richards, S. and Lister, G. (2009). Tectonic setting of kimberlites. Lithos 112S, 155–165.Google Scholar
Jowett, E. C. (1992). Role of organics and methane in sulfide ore formation, exemplified by Kupferschiefer Cu–Ag deposits, Poland. Chemical Geology 99, 51–63.Google Scholar
Jupp, T. and Schultz, A. (2000). A thermodynamic explanation for black smoker temperatures. Nature 403, 880–883.Google Scholar
Kelley, K. D. and Jennings, S. (2004). A special issue devoted to barite and Zn–Pb–Ag deposits in the Red Dog district, Western Brooks Range, Northern Alaska, Preface. Economic Geology 99, 1267–1280.Google Scholar
Kesler, S. E. and Wilkinson, B. H. (2008). Earth’s copper resources estimated from tectonic diffusion of porphyry copper deposits. Geology 36, 255–258.Google Scholar
Kesler, S. E. and Wilkinson, B. H. (2009). Resources of gold in Phanerozoic epithermal deposits. Economic Geology 104, 623–633.Google Scholar
Khashgerel, B-E., Rye, R. O., Hedenquist, J. W. and Kavalierts, I. (2006). Geology and reconnaissance stable isotope study of the Oyu Tolgoi porphyry Cu–Au system, South Gobi, Mongolia. Economic Geology 101, 503–522.Google Scholar
King, C.-Y., Zhang, W. and Zhang, Z. (2006). Earthquake-induced groundwater and gas changes. Pure and Applied Geophysics 163, 633–645.Google Scholar
Kinnaird, J., Kruger, F. J., Nex, P. A. M. and Cawthorn, R. G. (2002). Chromitite formation – a key to understanding processes of platinum enrichment. Transactions of the Institution of Mining and Metallurgy B 111, 23–35.Google Scholar
Kinnaird, J. A., Hutchinson, D., Schurmann, L., Nex, P. A. M. and de Lange, R. (2005). Petrology and mineralization of the southern Platreef: northern limb of the Bushveld Complex, South Africa. Mineralium Deposita 40, 576–597.Google Scholar
Kirk, J., Ruiz, J., Chesley, J., Titley, S. and Walshe, J. (2001). A detrital model for the origin of gold and sulfides in the Witwatersrand Basin based on Re–Os isotopes. Geochimica et Cosmochimica Acta 65, 2149–2159.Google Scholar
Kitto, P. A., Evans, D. A. and Mrozcek, C. R. (1997). Renison – new advances in the geological understanding of a world-class ore deposit. Australasian Institute of Mining and Metallurgy Publication Series 97, 31–39.Google Scholar
Kogel, J. E., Trivedi, N. C., Barker, J. M. and Krukowski, S. T. (eds.) (2006). Industrial Minerals and Rocks, Commodities, Markets, and Uses, 7th Edition, Littleton, Society for Mining, Metallurgy and Exploration Inc.
Kogiso, T. and Hirschmann, M. H. (2006). Partial melting experiments of bimineralic eclogite and the role of recycled mafic oceanic crust in the genesis of ocean island basalts. Earth and Planetary Science Letters 249, 185–199.Google Scholar
Kruger, F. J. (1994). The Sr-isotopic stratigraphy of the Western Bushveld Complex. South African Journal of Geology 97, 393–398.Google Scholar
Kruger, F. J. (2005). Filling of the Bushveld Complex magma chamber: lateral expansion, roof and floor interaction, magmatic unconformities, and the formation of giant chromitite, PGE and Ti-V-magnetitite deposits. Mineralium Deposita 40, 451–472.Google Scholar
Kruger, F. J. and Marsh, J. S. (1982). The significance of 87Sr/86Sr ratios in the Merensky cyclic unit of the Bushveld Complex. Nature 298, 53–55.Google Scholar
Kuleshov, V. N. (2011). Manganese deposits: communication 1. Genetic models of manganese ore formation. Lithology and Mineral Resources 46, 473–493.Google Scholar
Kwak, T. A. P. and Tan, T. H. (1981). The geochemistry of zoning in skarn minerals at the King Island (Dolphin) mine. Economic Geology 76, 468–497.Google Scholar
Lalou, C., Munch, U., Halbach, P. and Reyss, J. L. (1998). Radiochronological investigation of hydrothermal deposits from the MESO zone, Central Indian Ridge. Marine Geology 149, 243–254.Google Scholar
Landtwing, M. R., Pettke, T., Halter, W. E. et al. (2005). Copper deposition during quartz dissolution by cooling magmatic-hydrothermal fluids: the Bingham porphyry. Earth and Planetary Science Letters 235, 229–243.Google Scholar
Landtwing, M. R., Furrer, C., Redmond, P. B. et al. (2010). The Bingham Canyon porphyry Cu–Au–Mo deposit. III. Zoned copper–gold ore deposition by magmatic vapor expansion. Economic Geology 105, 91–118.Google Scholar
Lange, I. M., Nokleberg, W. J., Plahuta, J. T., Krouse, H. R. and Doe, B. R. (1985). Geological setting, petrology, and geochemistry of stratiform sphalerite–galena–barite deposits: Red Dog Creek and Drenchwater Creek areas, Northwestern Brooks Range, Alaska. Economic Geology 80, 1896–1926.Google Scholar
Langmuir, D. (1997). Aqueous Environmental Geochemistry, Upper Saddle River, NJ, Prentice Hall.
Large, D. and Walcher, E. (1999). The Rammelsberg massive sulphide Cu–Zn–Pb–Ba-deposit, Germany: an example of sediment-hosted, massive sulphide mineralisation. Mineralium Deposita 34, 522–538.Google Scholar
Large, R. R. (1992). Australian volcanic-hosted massive sulfide deposits: features, styles, and genetic models. Economic Geology 87, 471–510.Google Scholar
Large, R. R., Bull, S. W., Cooke, D. R. and McGoldrick, P. J. (1998). A genetic model for the HYC deposit, Australia: based on regional sedimentology, geochemistry and sulfide-sediment relationships. Economic Geology 93, 1345–1368.Google Scholar
Large, R., McGoldrick, P., Bull, S. and Cooke, D. (2004). Proterozoic stratiform sediment-hosted zinc–lead–silver deposits of Northern Australia. In Sediment-hosted Lead-Zinc Sulphide Deposits, Deb, M. and Goodfellow, W. D. (eds.), New Delhi, Narosa Publishing House, pp. 1–23.
Lasky, S. G. (1950). How tonnages and grade relations help predict ore reserves. Engineering Mining Journal 151, 81–85.Google Scholar
Lawrance, L. M. (1990). Supergene gold mineralization. In Gold Deposits of the Archaean Yilgarn Block, Western Australia: Nature, Genesis and Exploration Guides, Ho, S. E., Groves, D. I. and Bennett, J. M. (eds.), Crawley, University of Western Australia, Publication 20, pp. 299–314.
Leach, D. L., Viets, J. G., Kozlowski, A. and Kibitlewski, S. (1996). Geology, geochemistry, and genesis of the Silesia–Cracow zinc–lead district, southern Poland. Society of Economic Geologists Special Publication 4, 144–170.Google Scholar
LeBlanc, M. and Nicolas, A. (1992). Ophiolitic chromites. International Geology Review 34, 653–686.Google Scholar
Lesher, C. M. (1989). Komatiite-associated nickel sulfide deposits, ore deposits associated with magmas. Reviews in Economic Geology 4, 45–102.Google Scholar
Lesher, C. M. (2007). Ni–Cu–(PGE) deposits in the Raglan area, Cape Smith belt, New Quebec. Geological Association of Canada Mineral Deposits Division Special Publication 5, 351–386.Google Scholar
Lewis, B. L. and Landing, W. M. (1992). The investigation of dissolved and suspended particulate trace-metal fractionation in the Black Sea. Marine Chemistry 40, 105–141.Google Scholar
Li, C., Maier, W. D. and de Waal, S. A. (2001). The role of magma mixing in the genesis of PGE mineralization in the Bushveld Complex: thermodynamic calculation and new interpretations. Economic Geology 96, 653–662.Google Scholar
Loen, J. S. (1992). Mass balance constraints on gold placers: possible solutions to “source area problems”. Economic Geology 87, 1624–1634.Google Scholar
Logan, R. G., Murray, W. J. and Williams, N. (1990). HYC silver–lead–zinc deposit, McArthur River. In Geology and Mineral Deposits of Australia and Papua New Guinea, Hughes, F. E. (ed.), Melbourne, Australasian Institute of Mining and Metallurgy, Monograph 14, pp. 907–911.
Long, K. R., Van Gosen, B. S., Foley, N. F. and Cordier, D. (2010). The principal rare earth elements deposits of the United States – a summary of domestic deposits and a global perspective. US Geological Survey Scientific Investigations Report 2010–5220.Google Scholar
Lott, D. A., Coveney, R. M., Murowchick, J. B. and Grauch, R. I. (1999). Sedimentary exhalative nickel–molybdenum ores in south China. Economic Geology 94, 1051–1066.Google Scholar
Lottermoser, B. G. (1995). Ore minerals of the Mt Weld rare-earth element deposit, Western Australia. Transactions of the Institute of Mining and Metallurgy B 104, 203–209.Google Scholar
Lowell, J. D. and Guilbert, J. M. (1970). Lateral and vertical alteration-mineralization zoning in porphyry ore deposits. Economic Geology 65, 373–408.Google Scholar
Lowenstein, T. K., Timofeeff, M. N., Kovalevych, V. M. and Horita, J. (2005). The major ion composition of Permian seawater. Geochimica et Cosmochimica Acta 69, 1701–1719.Google Scholar
Lydon, J. W. (1988). Ore deposits models; volcanogenic massive sulphide deposits. Geosciences Canada 15, 43–65.Google Scholar
McCracken, S. R., Etminan, H., Connor, A. G. and Williams, V. A. (1996). Geology of the Admiral Bay carbonate-hosted zinc–lead deposit, Canning Basin, Western Australia. Society of Economic Geologists Special Publication 4, 330–349.Google Scholar
McFarlane, M. J. (1976). Laterite and Landscape. London, Academic Press.
Macumber, P. G. (1992). Hydrological processes in the Tyrrell Basin, southeastern Australia. Chemical Geology 96, 1–18.Google Scholar
Maier, W. D., Arndt, N. T. and Curl, E. A. (2000). Progressive crustal contamination of the Bushveld Complex: evidence from Nd isotopic analyses of the cumulate rocks. Contributions to Mineralogy and Petrology 140, 316–327.Google Scholar
Maier, W. D., Barnes, S.-J., Chinyepi, G. et al. (2008). The composition of magmatic Ni–Cu–(PGE) sulfide deposits in the Tati and Selebi-Phikwe belts of eastern Botswana. Mineralium Deposita 43, 37–60.Google Scholar
Mandal, N., Misra, S. and Samanta, S. K. (2004). Role of weak flaws in nucleation of shear zones: an experimental and theoretical study. Journal of Structural Geology 26, 1391–1400.Google Scholar
Marschik, R. and Fontboté, L. (2001). The Candelaria–Punta del Cobre iron oxide Cu–Au(–Zn–Ag) deposits, Chile. Economic Geology 96, 1799–1826.Google Scholar
Martin, H. J. (1963). The Bikita Tinfield. Southern Rhodesia Geological Survey Bulletin 58.Google Scholar
Mathieson, G. A. and Clark, A. H. (1984). The Cantung E-zone scheelite skarn orebody, tungsten, Northwest Territories – a revised genetic model. Economic Geology 79, 883–901.Google Scholar
Matthai, S. K., Heinrich, C. A. and Driesner, T. (2004). Is the Mount Isa copper deposit the product of forced brine convection in the footwall of a major reverse fault? Geology 32, 357–360.Google Scholar
Mavrogenes, J. A. and O’Neill, H. S. (1999). The relative effects of pressure, temperature and oxygen fugacity on the solubility of sulfide in mafic magmas. Geochimica et Cosmochimica Acta 63, 1173–1180.Google Scholar
Maynard, J. B. (2010). The chemistry of manganese ores through time: a signal of increasing diversity of Earth-surface environments. Economic Geology 100, 535–552.Google Scholar
Megaw, P. K. M., Ruiz, J. R. and Titley, S. R. (1988). High-temperature, carbonate-hosted Ag–Pb–Zn(Cu) deposits of northern Mexico. Economic Geology 83, 1856–1885.Google Scholar
Meier, D. L., Heinrich, C. A. and Watts, M. A. (2009). Mafic dikes displacing Witwatersrand gold reefs: evidence against metamorphic-hydrothermal ore formation. Geology 37, 607–610.Google Scholar
Meinert, L. D., Dipple, G. M. and Nicoescu, S. (2005). World skarn deposits. Economic Geology 100th Anniversary Volume, Hedenquist, J. W., Thompson, J. F. H., Goldfarb, R. J. and Richards, J. P. (eds.), Colorado, Society of Economic Geologists, pp. 299–336.
Melcher, F., Grum, W., Thalhammer, T. V. and Thalhammer, O. A. R. (1999). The giant chromite deposit at Kempirsai, Urals: constraints from trace element (PGE, REE) and isotope data. Mineralium Deposita 34, 250–272.Google Scholar
Melchiorre, E. B. and Enders, M. S. (2003). Stable isotope geochemistry of copper carbonates at the Northwest Extension Deposit, Morenci District, Arizona: implications for conditions of supergene oxidation and related mineralization. Economic Geology 98, 607–622.Google Scholar
Menendez, A. and Sarmentero, A. (1984). Geology of the Los Pijiguaos bauxite deposit, Venezuela. In Proceedings of the 1984 Bauxite Symposium, Los Angeles, Jacobs, L. (ed.), New York, American Institute of Mining, Metallurgy and Petroleum Engineering, pp. 387–407.
Mertig, H. J., Rubin, J. N. and Kyle, J. R. (1994). Skarn Cu–Au orebodies of the Gunung Bijih (Ertsberg) district, Irian Jaya, Indonesia. Journal of Geochemical Exploration 50, 179–202.Google Scholar
Meyer, C., Shea, E., Goddard, C. et al. (1968). Ore deposits at Butte, Montana. In Ore Deposits of the United States 1933–1967 (Graton-Sales Volume), Ridge, J. D., (ed.), New York, American Institute of Mining, Metallurgy and Petroleum Engineers, pp. 1363–1416.
Meyer, F. M., Happel, U., Hausberg, J. and Wiechowski, A. (2002). The geometry and anatomy of the Los Pijiguaos bauxite deposit, Venezuela, Ore Geology Reviews 20, 27–54.Google Scholar
Milburn, D. and Wilcock, S. (1998). Kunwarara magnesite deposit. In Geology of Australian and Papua New Guinean Mineral Deposits, Berkman, D. A. and Mackenzie, D. H. (eds.), Australasian Institute of Mining and Metallurgy, Monograph 22, pp. 815–819.
Miller, A. R., Densmore, C. D., Degens, E. T. et al. (1966). Hot brines and recent iron deposits in the deeps of the Red Sea. Geochimica et Cosmochimica Acta 30, 341–359.Google Scholar
Millero, F. J. and Söhn, M. L. (1992). Chemical Oceanography, Caldwell, NJ, Telford Press.
Minter, W. E. L. (1978). A sedimentological synthesis of placer gold, uranium and pyrite concentrations in Proterozoic Witwatersrand sediments. Canadian Society of Petroleum Geology Memoir 5, 801–829.Google Scholar
Minter, W. E. L. (1999). Irrefutable detrital origin of Witwatersrand gold and evidence of eolian signatures. Economic Geology 94, 665–670.Google Scholar
Mitchell, R. H. (1986). Kimberlites: Mineralogy, Geochemistry and Geology. New York, Plenum Press.
Moore, D. W., Young, L. E., Modene, J. S. and Plahuta, J. T. (1986). Geologic setting and genesis of the Red Dog zinc–lead–silver deposit, Western Brooks, Range, Alaska. Economic Geology 81, 1696–1727.Google Scholar
Morris, R. C. (1993). Genetic modelling for banded iron-formation of the Hamersley Group, Pilbara Craton, Western Australia. Precambrian Research 60, 243–286.Google Scholar
Morris, R. C. and Ramanaidou, E. R. (2007). Genesis of the channel iron deposits (CID) of the Pilbara region, Western Australia. Australian Journal of Earth Sciences 54, 733–756.Google Scholar
Morris, R. C. and Kneeshaw, M. (2011). Genesis modelling for the Hamersley BIF-hosted iron ores of Western Australia: a critical review. Australian Journal of Earth Sciences 58, 417–451.Google Scholar
Mossman, D. J., Minter, W. E. L., Dutkiewicz, A. et al. (2008). The indigenous origin of Witwatersrand “carbon”. Precambrian Research 164, 173–186.Google Scholar
Mote, T. I., Becker, T. A., Renne, P. and Brimhall, G. H. (2001a). Chronology of exotic mineralization at El Salvador, Chile, by 40Ar/39Ar dating of copper wad and supergene alunite. Economic Geology 96, 351–366.Google Scholar
Mote, T. I., Brimhall, G. H., Tidy-Finch, E., Muller, G. and Carrasco, P. (2001b). Application of mass-balance modeling of sources, pathways and sinks of supergene enrichment to exploration and discovery of the Quebrada Turquesa exotic copper orebody, El Salvador District, Chile. Economic Geology 96, 367–386.Google Scholar
Mudd, G. M. (2007). The Sustainability of Mining in Australia: Key Production Trends And Their Environmental Impacts for the Future. Research Report no. RR5, Department of Civil Engineering, Monash University and Mineral Policy Institute.
Mukasa, S. B., Wilson, A. H. and Carlson, R. W. (1998). A multielement geochronological study of the Great Dyke, Zimbabawe: significance of the robust and reset ages. Earth and Planetary Science Letters 164, 353–369.Google Scholar
Naldrett, A. J. (1973). Nickel sulphide deposits – their classification and genesis with special emphasis on deposits of volcanic association. Canadian Institute of Mining and Metallurgy, Bulletin 66, 45–63.Google Scholar
Naldrett, A. J. (2004). Magmatic Sulfide Deposits: Geology, Geochemistry and Exploration. Berlin, Springer.
Naldrett, A. J. (2010). From the mantle to the bank: the life of a Ni–Cu–(PGE) sulfide deposit. South African Journal of Geology 113, 1–32.Google Scholar
Naldrett, A. J. and Li, C. (2007). The Voisey’s Bay deposit, Labrador, Canada. Geological Association of Canada Mineral Deposits Division Special Publication 5, 387–408.Google Scholar
Naldrett, A. J., Lightfoot, P. C., Fedorenko, V., Doherty, W. and Gorbachev, N. S. (1992). Geology and geochemistry of intrusions and flood basalts of the Noril’sk region, USSR, with implications for the origin of the Ni-Cu ores. Economic Geology 87, 975–1004.Google Scholar
Naldrett, A. J., Wilson, A., Kinnaird, J. and Chunnett, G. (2009). PGE tenor and metal ratios within and below the Merensky Reef, Bushveld Complex: implications for its genesis. Journal of Petrology 50, 625–659.Google Scholar
Naranjo, J. A., Henriquez, F. and Nystrom, J. O. (2010). Subvolcanic contact metasomatism at El Laco volcanic complex, Central Andes. Andean Geology 37, 110–120.Google Scholar
Nash, J. T., Granger, H. C. and Adams, S. S. (1981). Geology and concepts of genesis of important types of uranium deposits. Economic Geology, 75th Anniversary Volume, Skinner, B. (ed.), Lancaster, PA, Economic Geology Publishing, pp. 63–116.
Nelson, J. (1997). The quiet counter-revolution: structural control of syngenetic deposits. Geoscience Canada 24, 91–98.Google Scholar
Northrop, H. R. and Goldhaber, M. B. (eds.) (1990). Genesis of the tabular-type vanadium-uranium deposits of the Henry Basin, Utah. Economic Geology 85, 215–269.Google Scholar
Nyström, J. O. and Henriquez, F. (1994). Magmatic features of iron ores of the Kiruna-type in Chile and Sweden: ore textures and magnetite geochemistry. Economic Geology 89, 820–839.Google Scholar
Oberthür, T., Weiser, T., Amanor, J. A. and Chryssoulis, S. L. (1997). Mineralogical siting and distribution of gold in quartz veins and sulfide ores of the Ashanti mine and other deposits of the Ashanti belt of Ghana. Mineralium Deposita 32, 2–15.Google Scholar
Ohle, E. L. (1996). Significant events in the geological understanding of the southeast Missouri lead district. Society of Economic Geologists Special Publication 4, 1–7.Google Scholar
Okamoto, A., Saishu, H., Hirano, N. and Tsuchiya, N. (2010). Mineralogical and textural variation of silica minerals in hydrothermal flow-through experiments: implications for quartz vein formation. Geochimica et Cosmochimica Acta 74, 3693–3706.Google Scholar
O’Neill, C. J., Moresi, L. and Jaques, A. L. (2005). Geodynamic controls on diamond deposits: implications for Australian occurrences. Tectonophysics 404, 217–236.Google Scholar
Oszczepalski, S. (1999). Origin of the Kupferschiefer polymetallic mineralization in Poland. Mineralium Deposita 34, 599–613.Google Scholar
Palabora Mining Company Limited Mine Geological and Mineralogical Staff (1976). The geology and economic deposits of copper, iron, and vermiculite in the Palabora Igneous Complex: a brief review. Economic Geology 71, 177–192.Google Scholar
Partington, G. A. (1990). Environment and structural controls on the intrusion of the giant rare metal Greenbushes pegmatite, Western Australia. Economic Geology 85, 437–456.Google Scholar
Pavlov, A. A. and Kasting, J. F. (2002). Mass-independent fractionation of sulfur isotopes in Archean sediments: strong evidence for an anoxic Archean atmosphere. Astrobiology 2, 27–41.Google Scholar
Percival, T. J. and Radtke, A. S. (1994). Sedimentary-rock hosted disseminated gold mineralization in the Alsar district, Macedonia. Canadian Mineralogist 32, 649–665.Google Scholar
Peters, S. G., Jianzhan, H., Zhiping, L. and Chenggui, J. (2007). Sedimentary rock-hosted Au deposits of the Dian-Qian-Gui area, Guizhou, and Yunnan Provinces, and Guangxi District, China. Ore Geology Reviews 31, 170–204.Google Scholar
Petersen, S., Herzig, P. M. and Hannington, M. D. (2000). Third dimension of a presently forming VMS deposit: TAG hydrothermal mound, Mid-Atlantic Ridge, 26° N. Mineralium Deposita 35, 233–259.Google Scholar
Petrov, S. V. (2004). Economic deposits associated with the alkaline and ultrabasic complexes of the Kola Peninsula. In Phoscorites and Carbonatites from Mantle to Mine: the Key Example of the Kola Alkaline Province, Wall, V. and Zaitsev, A. N. (eds.), London, Mineralogical Society, pp. 469–490.
Pettke, T., Diamond, L. W. and Kramers, J. D. (2000). Mesothermal gold lodes in the north-western Alps: a review of genetic constraints from radiogenic isotopes. European Journal of Mineralogy 12, 213–230.Google Scholar
Pettke, T., Oberli, F. and Heinrich, C. A. (2010). The magma and metal source of giant porphyry-type ore deposits based on lead isotope microanalysis of individual fluid inclusions. Earth and Planetary Science Letters 296, 267–277.Google Scholar
Pettke, T., Oberli, F., Audetat, A. et al. (2012). Recent developments in element concentration and isotope ratio analysis of individual fluid inclusion by laser ablation single and multiple collector ICP-MS. Ore Geology Reviews 44, 10–38.Google Scholar
Phillips, G. N. and Hughes, M. J. (1996). The geology and gold deposits of the Victorian gold province. Ore Geology Reviews 11, 255–302.Google Scholar
Phillips, G. N. and Law, J. D. M. (2000). Witwatersrand gold fields: geology, genesis, and exploration. Reviews in Economic Geology 13, 439–500.Google Scholar
Phillips, G. N., Meyers, F. M. and Palmer, J. A. (1987). Problems with the placer model for Witwatersrand gold. Geology 15, 1027–1030.Google Scholar
Pieczonka, J., Piestrzynski, A., Mucha, J. et al. (2008). The red-bed-type precious metal deposit in the Sieroszowice–Polkowice copper mining district, SW Poland. Annales Societas Geologorum Poloniae 78, 151–280.Google Scholar
Pollard, P. J. and Taylor, R. G. (2002). Paragenesis of the Grasberg Cu–Au deposit, Irian Jaya, Indonesia: results from logging section 13. Mineralium Deposita 37, 117–136.Google Scholar
Porter, T. M. (ed.) (2000). Hydrothermal Iron Oxide Copper–Gold and Related Deposits: A Global Perspective, PGC Publishing, Adelaide.
Poujol, M., Robb, L. J. and Respaut, J. P. (1999). U–Pb and Pb–Pb isotopic studies relating to the origin of gold mineralization in the Evander Goldfield, Witwatersrand Basin, South Africa. Precambrian Research 95, 167–185.Google Scholar
Prendergast, M. D. (2003). The nickeliferous late Archean komatiitic event in the Zimbabwe Craton – magmatic architecture, physical volcanology, and ore genesis. Economic Geology 98, 865–891.Google Scholar
Prendergast, M. D. and Wilson, A. H. (1989). The Great Dyke of Zimbabwe – II: mineralization and mineral deposits. In Magmatic Sulphides – The Zimbabwe Volume, Prendergast, M. D. and Jones, M. J. (eds.), London, Institution of Mining and Metallurgy, pp. 21–42.
Prescott, J. R. and Habermehl, M. A. (2008). Luminescence dating of spring mound deposits in the southwestern Great Artesian Basin, northern South Australia. Australian Journal of Earth Sciences 55, 167–181.Google Scholar
Profett, J. M. (2003). Geology of the Bajo de la Alumbrera porphyry copper-gold deposit, Argentina. Economic Geology 98, 1535–1574.Google Scholar
Rackley, R. I. (1976). Origin of Western-States type uranium mineralization. In Handbook of Strata Bound and Stratiform Ore Deposits. Part II. Regional Studies and Specific Deposits, Wolf, K. H. (ed.), Amsetrdam, Elsevier, pp. 89–156.
Radtke, A. S. (1985). Geology of the Carlin gold deposit, Nevada, US Geological Survey Professional Paper 1267.Google Scholar
Raffensperger, J. P. and Garven, G. (1995). The formation of unconformity-type uranium ore deposits. I: coupled groundwater flow and heat transport modeling. American Journal of Science 295, 581–636.Google Scholar
Ramdohr, P. (1980). The Ore Minerals and Their Intergrowths. Oxford, Pergamon Press.
Reeve, J. S., Cross, K. C., Smith, R. N. and Oreskes, N. (1990). Olympic Dam copper–uranium–gold–silver deposit, In Geology and Mineral Deposits of Australia and Papua New Guinea, Hughes, F. E. (ed.), Melbourne, Australasian Institute of Mining and Metallurgy, Monograph 14, pp. 1009–1035.
Rentzsch, J. (1974). The Kupferschiefer in comparison with the deposits of the Zambian copperbelt. In Gisements Stratiforme et Provinces, Cuprifères, Bartholomé, P. (ed.), Liège, Société Géologique de Belgique, pp. 395–418.
Ressel, M. W. and Henry, C. D. (2006). Igneous geology of the Carlin trend, Nevada: development of the Eocene plutonic complex and significance for Carlin-type gold deposit. Economic Geology 101, 347–383.Google Scholar
Reynolds, L. J. (2001). Geology of the Olympic Dam Cu–U–Au–Ag–REE deposit. MESA Journal 23, 4–11.Google Scholar
Rice, C. M., Harmon, R. S. and Shepherd, T. J. (1985). Central City, Colorado: the upper part of an alkaline porphyry molybdenum system. Economic Geology 80, 1769–1796.Google Scholar
Richards, J. P. (2009). Post subduction porphyry Cu–Au and epithermal Au deposits: products of remelting of subduction modified lithosphere. Geology 37, 247–250.Google Scholar
Richards, J. P. (2011). High Sr/Y arc magmas and porphyry Cu±Mo±Au deposits: just add water. Economic Geology 100, 1075–1082.Google Scholar
Richards, J. P. and Tosdal, R. M. (eds.) (2001). Structural controls on ore genesis. Reviews in Economic Geology 14.Google Scholar
Ridley, J. R. (1993). The relations between mean rock stress and fluid flow in the crust with reference to vein- and lode-style deposits. Ore Geology Reviews 8, 23–37.Google Scholar
Ridley, J. and Mengler, F. (2000). Lithological and structural control on the form and setting of vein stockwork orebodies at the Mount Charlotte gold deposit, Kalgoorlie. Economic Geology 95, 85–98.Google Scholar
Risacher, F. and Fritz, B. (2009). Origins of salts and brine evolution of Bolivian and Chilean salars. Aquatic Geochemistry 15, 123–157.Google Scholar
Robert, F. (1990). Structural setting and control on gold–quartz veins of the Val d’Or area, southeastern Abitibi Subprovince. In Gold and Base Metal Mineralization in the Abitibi Subprovince, Canada, with Emphasis on the Quebec Segment, Ho, S. E., Robert, F. and Groves, D. I. (eds.), University of Western Australia, Geology Key Centre and University Extension Publication 24, pp. 164–209.
Robert, F. and Brown, A. C. (1986). Archean gold-bearing quartz veins at the Sigma mine, Abitibi greenstone belt, Quebec. Economic Geology 81, 578–616.Google Scholar
Roedder, E. (1984). Fluid inclusions. Reviews in Mineralogy 12.Google Scholar
Rojstaczer, S. S. and Wolf, S. (1994). Hydrologic changes associated with the earthquake in the San Lorenzo and Pescadero drainage basins. US Geological Survey Professional Paper 1551-E, 51–64.Google Scholar
Rollinson, H. (1993). Using Geochemical Data. Oxford, Blackwell.
Roscoe, S. M. and Minter, W. E. L. (1993). Pyritic paleoplacer gold and uranium deposits. Special Paper, Geological Association of Canada 40, 103–124.Google Scholar
Rousell, D. H., Fedorowich, J. S. and Driessler, B. O. (2003). Sudbury Breccia (Canada): a product of the 1850 Ma Sudbury Event and hosts to footwall Cu–Ni–PGE deposits. Earth-Science Reviews 60, 147–174.Google Scholar
Rowe, M. C., Kent, A. J. R. and Nielsen, R. L. (2009). Subduction influence on oxygen fugacity and trace and volatile contents in basalts across the Cascade Volcanic arc. Journal of Petrology 50, 61–91.Google Scholar
Rowland, J. V. and Simmons, S. F. (2012). Hydrologic, magmatic, and tectonic controls on hydrothermal fluid flow, Taupo Volcanic Zone, New Zealand: implications for the formation of epithermal vein deposits. Economic Geology 102, 427–459.Google Scholar
Roy, P. S. (1999). Heavy mineral beach placers in southeastern Australia: their nature and genesis. Economic Geology 94, 567–588.Google Scholar
Rubin, J. N. and Kyle, J. R. (1997). Precious metal mineralogy in porphyry-, skarn-, and replacement-type ore deposits of the Ertsberg (Gunung Bijih) district, Irian Jaya, Indonesia. Economic Geology 92, 535–550.Google Scholar
Rubright, R. D. and Hart, O. J. (1968). Non-porphyry ores of the Bingham district, Utah. In Ore Deposits of the United States, 1933–67 (Graton-Sales Volume), Ridge, J. D. (ed.), New York, AIME, pp. 886–907.
Rusk, B. G., Reed, M. H. and Dilles, J. H. (2008). Fluid inclusion evidence for magmatic-hydrothermal fluid evolution in the porphyry copper–molybdenum deposit at Butte, Montana. Economic Geology 103, 307–334.Google Scholar
Sanderson, D. J. and Zhang, X. (1999). Critical stress localization of flow associated with deformation of well-fractured rock masses, with implications for mineral deposits. Geological Society of London, Special Publication 155, 69–81.Google Scholar
Sanford, R. F. (1994). A quantitative model for ground-water flow during formation of tabular uranium sandstone uranium deposits. Economic Geology 89, 341–360.Google Scholar
Sangster, A. L. and Smith, P. K. (2007) Metallogenic summary of the Meguma gold deposits, Nova Scotia. Geological Association of Canada Mineral Deposits Division Special Publication 5, 723–732.Google Scholar
Sangster, D. F. (2002). The role of dense brines in the formation of vent-distal sedimentary-exhalative (SEDEX) lead–zinc deposits: field and laboratory evidence. Mineralium Deposita 37, 149–157.Google Scholar
Sato, T. (1977). Kuroko deposits: their geology, geochemistry and origin. In Volcanic Processes in Ore Genesis, Anon, (ed.), London, Institution of Mining and Metallurgy, pp. 153–161.
Saunders, J. A. (1990). Colloidal transport of gold and silica in epithermal precious-metal systems: evidence from the Sleeper deposit, Nevada. Geology 18, 757–760.Google Scholar
Saunders, J. A. and Brueseke, M. E. (2012). Volatility of Se and Te during subduction-related distillation and the geochemistry of epithermal ores of the western United States. Economic Geology 107, 165–171.Google Scholar
Scherba, G. N. (1970). Greisens. International Geology Review 12, 114–151, 239–254.Google Scholar
Schmitt, A. K., Stockli, D. F., Lindsay, J. M. et al. (2010). Episodic growth and homogenization of plutonic roots in arc volcanoes from combined U–Th and (U–Th)/He zircon dating. Earth and Planetary Science Letters 295, 91–103.Google Scholar
Schoenberg, R., Nagler, T. F., Gnos, E., Kramers, J. D. and Kamber, B. S. (2003). The source of the Great Dyke, Zimbabwe, and its tectonic significance: evidence from Re–Os isotopes. Journal of Geology 111, 565–578.Google Scholar
Schouwstra, R. P., Kinloch, E. D. and Lee, C. A. (2000). A short geological review of the Bushveld Complex. Platinum Metals Review 44, 33–39.Google Scholar
Scott, S. D. (1997). Submarine hydrothermal systems and deposits. In Geochemistry of Hydrothermal Ore Deposits, Barnes, H. L. (ed.), New York, John Wiley, pp. 797–875.
Seedorff, E. and Einaudi, M. T. (2004). Henderson porphyry molybdenum system, Colorado: I. Sequence and abundance of hydrothermal mineral assemblages, flow paths of evolving fluid, and evolutionary style. Economic Geology 99, 3–37.Google Scholar
Seeger, C. M., Nuelle, L. M., Day, W. C. et al. (2001). Geologic maps and cross sections of mine levels at the Pea Ridge iron mine, Washington County, Missouri. US Geological Survey Miscellaneous Field Studies Map MF-2353.Google Scholar
Selley, D., Broughton, D., Scott, R. et al. (2005). A new look at the geology of the Zambian Copperbelt. Economic Geology 100th Anniversary Volume, Hedenquist, J. W., Thompson, J. F. H., Goldfarb, R. J. and Richards, J. P. (eds.), Colorado, Society of Economic Geologists, pp. 965–1000.
Shannon, J. R., Nelson, E. P. and Golden, R. J. (2004). Surface and underground geology of the world-class Henderson molybdenum porphyry mine, Colorado. Geological Society of America Field Guide 5, 207–218.Google Scholar
Sharp, Z. (2007). Principles of Stable Isotope Geology, New York, Pearson/Prentice Hall.
Sheldon, R. P. (1980). Episodicity of phosphate deposition and deep ocean circulation: a hypothesis. Society of Economic Paleontologists and Mineralogists Special Publication 29, 239–247.Google Scholar
Shepherd, M. S. (1990). Eneabba heavy mineral sand placers. In Geology and Mineral Deposits of Australia and Papua New Guinea, Hughes, F. E. (ed.), Melbourne, Australasian Institute of Mining and Metallurgy, Monograph 14, pp. 1591–1594.
Shepherd, T. J., Rankin, A. H. and Alderton, D. H. M. (1985). A Practical Guide to Fluid Inclusion Studies. Glasgow, Blackie.
Sibson, R. H. (1987). Earthquake rupturing as a mineralizing agent in hydrothermal systems. Geology 15, 710–714.Google Scholar
Sibson, R. H. (1996). Structural permeability of fluid-driven fault fracture meshes. Journal of Structural Geology 18, 1031–1042.Google Scholar
Sibson, R. H., Robert, F. and Poulsen, K. H. (1988). High-angle reverse faults, fluid-pressure cycling, and mesothermal gold deposits. Geology 16, 551–555.Google Scholar
Sidder, G. B., Day, W. C., Nuelle, L. M. et al. (1993). Mineralogic and fluid-inclusion studies of the Pea Ridge iron–rare-earth-element deposit, southeast Missouri. US Geological Survey Bulletin 2093, 205–216.Google Scholar
Siebold, E. and Berger, W. H. (1996). An Introduction to Marine Geology. Heidelberg, Springer.
Sillitoe, R. H. (2003). Iron oxide–copper–gold deposits: an Andean view. Mineralium Deposita 38, 787–812.Google Scholar
Sillitoe, R. H. (2008). Major gold deposits and belts of the North and South American Cordillera: distribution, tectonomagmatic settings and metallogenic considerations. Economic Geology 103, 663–687.Google Scholar
Sillitoe, R. H. (2010). Porphyry copper systems. Economic Geology 105, 3–41.Google Scholar
Sillitoe, R. H. and Burrows, D. R. (2002). New field evidence bearing on the origin of the El Laco magnetite deposit, northern Chile. Economic Geology 97, 1101–1109.Google Scholar
Simmons, S. F. and Brown, K. L. (2006). Gold in magmatic hydrothermal solutions and the rapid formation of a giant ore deposit. Science 314, 288–291.Google Scholar
Simmons, S. F. and Brown, K. L. (2007). The flux of gold and related metals through a volcanic arc, Taupo Volcanic Zone, New Zealand. Geology 35, 1099–1102.Google Scholar
Simmons, S. F. and Browne, P. R. L. (2000). Hydrothermal minerals and precious metals in the Broadlands–Ohaaki geothermal system: implications for understanding low-sulfidation epithermal environments. Economic Geology 95, 971–999.Google Scholar
Simmons, S. F., White, N. C. and John, D. A. (2005). Geological characteristics of epithermal precious and base metal deposit. Economic Geology, 100th Anniversary Volume, Hedenquist, J. W., Thompson, J. F. H., Goldfarb, R. J. and Richards, J. P. (eds.), Colorado, Society of Economic Geologists, pp. 485–522.
Simon, A. C., Pettke, T., Candela, P. A., Piccoli, P. M. and Heinrich, C. A. (2006). Copper partitioning in sulphur bearing magmatic systems. Geochimica et Cosmochimica Acta 70, 5583–5600.Google Scholar
Simpson, M. P. and Mauk, J. L. (2007). The Favona epithermal gold–silver deposit, Waihi, New Zealand. Economic Geology 102, 817–840.Google Scholar
Simpson, M. P., Mauk, J. L. and Simmons, S. F. (2001). Hydrothermal alteration and hydrologic evolution of the Golden Cross epithermal Au–Ag deposit, New Zealand. Economic Geology 96, 773–796.Google Scholar
Skinner, B. J. (1976). A second iron age ahead? American Scientist 64, 258–269.Google Scholar
Smurthwaite, A. J. (1990). Alumina, geology and mineral resources of Western Australia. Geological Survey of Western Australia Memoir 61, 5–624.Google Scholar
Solomon, M. (1990). Subduction, arc reversal, and the origin of porphyry copper–gold deposits in island arcs. Geology 18, 630–633.Google Scholar
Solomon, M. and Groves, D. I. (2000a). Volcanic-hosted massive sulphide deposits of the Tasman Fold Belt System. In The Geology and Origin of Australia’s Mineral Deposits, Hobart, Centre for Ore Deposit Research, pp. 580–723.
Solomon, M. and Groves, D. I. (2000b). Proterozoic sediment-hosted, stratiform (sedex), lead-zinc deposits. In The Geology and Origin of Australia’s Mineral Deposits, Hobart, Centre for Ore Deposit Research, pp. 168–239.
Solomon, M. and Groves, D. I. (2000c). Proterozoic uranium–platinum group element–gold deposits of the Pine Creek Inlier and the Murphy Inlier. In The Geology and Origin of Australia’s Mineral Deposits, Hobart, Centre for Ore Deposit Research, pp. 287–343.
Spandler, C., Mavrogenes, J. and Arculus, R. (2005). Origin of chromitites in layered intrusions: evidence from chromite-hosted melt inclusions from the Stillwater Complex. Geology 33, 893–896.Google Scholar
Spiridonov, E. M. (2010). Ore-magmatic systems of the Noril’sk ore field. Russian Geology and Geophysics 51, 1059–1077.Google Scholar
Stimac, J. and Hickmott, D. (1996). Ore metal partitioning in intermediate-to-silicic magmas: PIXE results on natural mineral assemblages. In Giant Ore Deposits – II: Controls on the Scale of Orogenic Magmatic-Hydrothermal Mineralization, Clark, A. H. (ed.), Ontario, Kingston, pp. 197–235.
Strom, S., Shane, P., Schmitt, A. K. and Lindsay, J. M. (2012). Decoupled crystallization and eruption histories of rhyolite magmatic systems at Tarawera volcano revealed by zircon ages and growth rates. Contributions to Mineralogy and Petrology 163, 505–519.Google Scholar
Sutton, S. J., Ritger, S. D. and Maynard, J. B. (1990). Stratigraphic control of chemistry and mineralogy in metamorphosed Witwatersrand quartzite. Journal of Geology 98, 329–341.Google Scholar
Sverjensky, D. A. (1986). Genesis of Mississippi Valley-type lead–zinc deposits. Annual Reviews of Earth and Planetary Sciences 14, 177–199.Google Scholar
Sverjensky, D. A. (1989). Chemical evolution of basinal brines that formed sediment-hosted Cu–Pb–Zn deposits. Geological Association of Canada Special Paper 36, 127–134.Google Scholar
Tankard, A. J., Jackson, M. P. A., Eriksson, K. A. et al. (1982). Crustal Evolution of southern Africa. New York, Springer-Verlag.
Taylor, D., Dalstra, H. J., Harding, A. E. et al. (2001). Genesis of high-grade hematite orebodies of the Hamersley province, Western Australia. Economic Geology 96, 837–873.Google Scholar
Taylor, D. H. and Gentle, L. V. (2002). Evolution of deep-lead palaeodrainages and gold exploration at Ballarat, Australia. Australian Journal of Earth Sciences 49, 869–878.Google Scholar
Teal, L. and Benavides, A. (2010). History and geologic overview of the Yanacocha Mining District, Cajamarca, Peru. Economic Geology 105, 1173–1190.Google Scholar
Thompson, T. B. and Arehart, G. B. (1990). Geology and origin of ore deposits in the Leadville district, Colorado: part I. Geological studies of orebodies and wall rocks. Economic Geology Monograph 7, 130–155.Google Scholar
Thompson, T. B., Trippel, A. D. and Dwelley, P. C. (1985). Mineralized veins and breccias in the Cripple Creek district, Colorado. Economic Geology 80, 1669–1688.Google Scholar
Thomson, J., Calvert, S. E., Mukherjee, S., Burnett, W. C. and Bremner, J. M. (1984). Further studies of the nature, composition and ages of contemporary phosphorite from the Namibian shelf. Earth and Planetary Science Letters 69, 341–353.Google Scholar
Thorne, W. S., Hagemann, S. G. and Barley, M. (2004). Petrographic and geochemical evidence for hydrothermal evolution of the North Deposit, Mt Tom Price, Western Australia. Mineralium Deposita 39, 766–783.Google Scholar
Titley, S. R. (ed.) (1982). Advances in Geology of the Porphyry Copper Deposits Southwestern North America, Tucson, University of Arizona Press.
Titley, S. R. (1993). Characteristics of porphyry copper occurrence in the American southwest. Geological Association of Canada Special Paper 40, 433–463.Google Scholar
Titley, S. R. (2001). Crustal affinities of metallogenesis in the American southwest. Economic Geology 96, 1323–1342.Google Scholar
Tompkins, L. A., Rayner, M. J. and Groves, D. I. (1994). Evaporites: in situ sulfur source for rhythmically banded ore in the Cadjebut Mississippi-Valley type Zn–Pb deposit, Western Australia. Economic Geology 89, 467–492.Google Scholar
Tompkins, L. A., Eisenlohr, B., Groves, D. I. and Raetz, M. (1997). Temporal changes in mineralization style at the Cadjebut Mississippi Valley-type deposit, Lennard Shelf, W. A. Economic Geology 92, 843–862.Google Scholar
Tornos, F. (2006). Environment of formation and styles of volcanogenic massive sulphides: the Iberian Pyrite Belt. Ore Geology Reviews 28, 259–306.Google Scholar
Tosdal, R. M. and Richards, J. P. (2001). Magmatic and structural controls on the development of porphyry Cu±Mo±Au deposits. Reviews in Economic Geology 14, 157−181.Google Scholar
Turner-Peterson, C. E. (1985). Lacustrine–humate model for primary uranium ore deposits, Grants Uranium Region, New Mexico. American Association of Petroleum Geologists Bulletin 69, 1999–2020.Google Scholar
Ulmer, P. (2001). Partial melting in the mantle wedge – the role of H2O in the genesis of mantle-derived “arc-related” magmas. Physics of Earth and Planetary Interiors 127, 215–232.Google Scholar
Ulrich, T. and Heinrich, C. A. (2002). Geology and alteration geochemistry of the porphyry Cu–Au deposit at Bajo de la Alumbrera, Argentina. Economic Geology 97, 1865–1888.Google Scholar
Ulrich, T., Günther, D. and Heinrich, C. A. (1999). Gold concentrations in magmatic brines and the metal budget of porphyry copper deposits. Nature 399, 676–679.Google Scholar
Urabe, T. and Sato, T. (1978). Kuroko deposits of the Kosaka Mine, Northeast Honshu, Japan – products of submarine hot springs on Miocene ocean floor. Economic Geology 73, 161–179.Google Scholar
Van Gosen, B. S. (2009). The Iron Hill (Powderhorn) carbonatite complex, Gunnison County, Colorado; a potential source of several uncommon mineral resources. US Geological Survey Open-File Report 2009–1005.Google Scholar
Van Houten, F. B. and Bhattacharyya, D. P. (1982). Phanerozoic oolitic ironstones – geological record and facies model. Annual Review of Earth and Planetary Sciences 10, 441–457.Google Scholar
Vearncombe, S., Barley, M. E., Groves, D. I. et al. (1995). 3.26 Ga black smoker-type mineralization in the Strelley Belt, Pilbara Craton, Western Australia. Journal of the Geological Society, London 152, 587–590.Google Scholar
Volkov, A. V., Serafimovski, T., Kochneva, N. T., Tomson, I. N. and Tasev, G. (2006). The Alshar epithermal Au–As–Sb–Tl deposit, southern Macedonia. Geology of Ore Deposits 48, 175–192.Google Scholar
von Quadt, A., Emiu, M., Martinek, K. et al. (2011). Zircon crystallization and the lifetimes of ore-forming magmatic hydrothermal systems. Geology 39, 731–734.Google Scholar
Wagner, P. A. (1929). The Platinum Deposits and Mines of South Africa. Capetown, C. Struik (Pty) Ltd.
Waite, K. A., Keith, J. D., Christiansen, E. H. et al. (1998). Petrogenesis of the volcanic and intrusive rocks associated with the Bingham Canyon porphyry Cu–Au–Mo deposit, Utah. Society of Economic Geologists Guidebook Series 29, 69–90.Google Scholar
Warren, J. K. (2000). Evaporites, brines and base metals: low-temperature ore emplacement controlled by evaporite diagenesis. Australian Journal of Earth Sciences 30, 179–208.Google Scholar
Webb, S. J., Cawthorn, R. G., Nguuri, T., and James, D. (2004). Gravity modeling of Bushveld Complex connectivity supported by Southern African Seismic Experiment results. South African Journal of Geology 107, 207–218.Google Scholar
White, W. S. (1968). The native copper deposits of northern Michigan. In Ore Deposits of the United States 1933–1967, Ridge, J. D. (ed.), New York, American Institute of Mining, Metallurgy and Petroleum Engineers, vol. 1, pp. 303–325.
Whitmeyer, S. J. and Karlstrom, K. E. (2007). Tectonic model for the Proterozoic growth of North America. Geosphere 3, 220–259.Google Scholar
Wilkinson, J. J. (2001). Fluid inclusions in hydrothermal ore deposits. Lithos 55, 229–272.Google Scholar
Williams, P. J., Barton, M. D., Johnson, D. A. et al. (2005). Iron oxide copper–gold deposits: geology, space-time distribution, and possible modes of origin. Economic Geology, 100th Anniversary Volume, Hedenquist, J. W., Thompson, J. F. H., Goldfarb, R. J. and Richards, J. P. (eds.), Colorado, Society of Economic Geologists, pp. 71–405.
Wilson, A. J., Cooke, D. R., Stein, H. J. et al. (2007). U–Pb and Re–Os geochronological evidence for two alkalic porphyry ore-forming events in the Cadia district, New South Wales, Australia. Economic Geology 102, 3–26.Google Scholar
Wodzicki, A. and Piestrzynski, A. (1994). An ore genetic model for the Lubin–Sieroszowice mining district, Poland. Mineralium Deposita 29, 30–43.Google Scholar
Yang, K-F., Fan, H-R., Santosh, M., Hu, F-F. and Wang, K-Y. (2011). Mesoproterozoic carbonatitic magmatism in the Bayan Obo deposit, Inner Mongolia, North China: constraints for the mechanism of super accumulation of rare earth elements. Ore Geology Reviews 40, 122–131.Google Scholar
Zajacz, Z., Hanley, J. J., Heinrich, C. A., Halter, W. E. and Guillong, M. (2009). Diffusive reequilibration of quartz-hosted silicate melt and fluid inclusions: are all metal concentrations unmodified? Geochimica et Cosmochimica Acta 73, 3013–3027.Google Scholar
Zaw, K., Peters, S. G., Cromie, P., Burrett, C. and Hou, Z. (2007). Nature, diversity of deposit types and metallogenic relations of South China. Ore Geology Reviews 31, 3–47.Google Scholar
Zientek, M. L., Cooper, R. W., Corson, S. R. and Geraghty, E. P. (2002). Platinum-group element mineralization in the Stillwater Complex, Montana. Canadian Institute of Mining, Metallurgy and Petroleum Special Volume 54, 459–481.Google Scholar
Zierenberg, R. A., Fouquet, Y., Miller, D. J. et al. (1998). The deep structure of a sea-floor hydrothermal deposit. Nature 392, 485–488.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • References
  • John Ridley, Colorado State University
  • Book: Ore Deposit Geology
  • Online publication: 05 June 2013
  • Chapter DOI: https://doi.org/10.1017/CBO9781139135528.009
Available formats
×

Save book to Dropbox

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

  • References
  • John Ridley, Colorado State University
  • Book: Ore Deposit Geology
  • Online publication: 05 June 2013
  • Chapter DOI: https://doi.org/10.1017/CBO9781139135528.009
Available formats
×

Save book to Google Drive

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

  • References
  • John Ridley, Colorado State University
  • Book: Ore Deposit Geology
  • Online publication: 05 June 2013
  • Chapter DOI: https://doi.org/10.1017/CBO9781139135528.009
Available formats
×