Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-dnltx Total loading time: 0 Render date: 2024-04-24T15:01:17.338Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  08 March 2019

Alan Boudreau
Affiliation:
Duke University, North Carolina
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
Publisher: Cambridge University Press
Print publication year: 2019

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

Ahmed, A.H. and Arai, S. (2002). Unexpectedly high-PGE chromitite from the deeper mantle section of the northern Oman ophiolite and its tectonic implications. Contributions to Mineralogy and Petrology, 143, 263278.CrossRefGoogle Scholar
Aird, H.M. and Boudreau, A.E. (2013). High-temperature carbonate minerals in the Stillwater Complex, Montana, USA. Contributions to Mineralogy and Petrology, 166, 11431160.Google Scholar
Aird, H.M., Ferguson, K.M., Lehrer, M.L. and Boudreau, A.E. (2017). A study of the trace sulfide mineral assemblages in the Stillwater Complex, Montana. Mineralium Deposita, 52(3), 361382.CrossRefGoogle Scholar
Algar, C.K., Boudreau, B.P. and Barry, M.A. (2011). Initial rise of bubbles in cohesive sediments by a process of viscoelastic fracture. Journal of Geophysical Research, 116, B04207. DOI:10.1029/2010JB008133.Google Scholar
Almeev, R.R., Holtz, F., Koepke, J., Parat, F. and Botcharnikov, R.E. (2007). The effect of H2O on olivine crystallization in MORB: Experimental calibration at 200 MPa. American Mineralogist, 92, 670674.Google Scholar
Almeev, R.R., Holtz, F., Koepke, J. and Parat, F. (2012). Experimental calibration of the effect of H2O on plagioclase crystallization in basaltic melt at 200 MPa. American Mineralogist, 97, 12341240.Google Scholar
Anderson, A.T. Jr, (1974). Chlorine, sulfur and water in magmas and oceans. Geological Society of America Bulletin, 85, 14851492.Google Scholar
Andersen, J.C.Ø. (2006). Postmagmatic sulphur loss in the Skaergaard Intrusion: Implications for the formation of the Platinova Reef. Lithos, 92, 198221.Google Scholar
Andersen, J.C.Ø., Rasmussen, H., Nielsen, T.F.D. and Ronsbo, J.G. (1998). The triple group and the Platinova gold and palladium reefs in the Skaergaard intrusion: Stratigraphic and petrographic relations. Economic Geology, 93, 488509.Google Scholar
Appold, M.S. and Nunn, J.A. (2002). Numerical models of petroleum migration via buoyancy porosity waves in viscously deformable sediments. Geofluids, 2, 233247.Google Scholar
Arai, S. and Akizawa, N. (2014). Precipitation and dissolution of chromite by hydrothermal solutions in the Oman ophiolite, New behavior of Cr and chromite. American Mineralogist, 99, 2834.CrossRefGoogle Scholar
Arai, S., Prichard, H.M., Matsumoto, I. and Fisher, P.C. (1999). Platinum-group minerals in podiform chromitite from the Kamuikotan zone, Hokkaido, northern Japan. Resource Geology, 49, 3947.CrossRefGoogle Scholar
Aranovich, L.Ya., Zakirov, I.V., Sretenskaya, N.G. and Gerya, T.V. (2010). Ternary system H2O–CO2–NaCl at high T–P parameters: An empirical mixing model. Geochemistry International, 48(5), 446455.CrossRefGoogle Scholar
Armienti, P., Francalanci, L. and Landi, P. (2007). Textural effects of steady state behaviour of the Stromboli feeding system. Journal of Volcanology and Geothermal Research, 160, 8698.CrossRefGoogle Scholar
Armitage, G.J. (1992). The petrogenesis of potholes in the UG2 chromitite layer, Crocodile River Mine, Western Bushveld complex. Ph.D. dissertation, University of Natal, Pietermaritzburg, South Africa.Google Scholar
Armstrong, R. and Wilson, A.H. (2000). A SHRIMP U-Pb study of zircons from the layered sequence of the Great Dyke, Zimbabwe, and a granitoid anatectic dyke. Earth and Planetary Science Letters, 180, 112.CrossRefGoogle Scholar
Atkins, F.B. (1969). Pyroxenes of the Bushveld Intrusion, South Africa. Journal of Petrology, 10(2), 222249.CrossRefGoogle Scholar
Augé, T. (1988). Platinum-group minerals in the Tiébaghi and Vourinos ophiolitic complexes: Genetic implications. Canadian Mineralogist, 26, 177192.Google Scholar
Azimov, P.Ya. and Bushmin, S.A. (2007). Solubility of minerals of metamorphic and metasomatic rocks in hydrothermal solutions of varying acidity: Thermodynamic modeling at 400–800 °C and 1–5 kbar. Geochemistry International, 45(12), 12101234.Google Scholar
Bachmann, O. and Bergantz, G.W. (2006). Gas percolation in upper-crustal silicic crystal mushes as a mechanism for upward heat advection and rejuvenation of near-solidus magma bodies. Journal of Volcanology and Geothermal Research, 149, 85102.Google Scholar
Bagdassarov, N., Dorfman, A. and Dingwell, D.B. (2000). Effect of alkalis, phosphorus, and water on the surface tension of haplogranite melt. American Mineralogist, 85, 3340.Google Scholar
Baker, D.R. and Moretti, R. (2011). Modeling the Solubility of Sulfur in Magmas: A 50-Year Old Geochemical Challenge. In Sulfur in Magmas and Melts: Its Importance for Natural and Technical Processes (Behrens, H. and Websater, J.D., eds.). Reviews in Mineralogy and Geochemistry, 73, 167–213.CrossRefGoogle Scholar
Ballhaus, C.G. (1988). Potholes of the Merensky Reef at Brakspruit Shaft, Rustenburg Platinum Mines: Primary disturbances in the magmatic stratigraphy. Economic Geology, 83, 11401158.Google Scholar
Ballhaus, C.G. and Stumpfl, E.F. (1985). Occurrence and petrological significance of graphite in the Upper Critical Zone, Western Bushveld Complex, South Africa. Earth and Planetary Science Letters, 74, 5868.Google Scholar
Ballhaus, C.G. and Stumpfl, E.F. (1986). Sulphide and platinum mineralization in the Merensky Reef: Evidence from hydrous silicates and fluid inclusions. Contributions to Mineralogy and Petrology, 94, 193204.Google Scholar
Ballhaus, C.G. and Ulmer, P. (1995). Platinum-group elements in the Merensky Reef: II. Experimental solubilities of platinum and palladium in Fe1-XS from 950 to 450°C under controlled f(S2) and f(H2). Geochimica et Cosmochimica Acta, 59(23), 48814888.Google Scholar
Ballhaus, C.G., Ryan, C.G., Mernagh, T.P. and Green, D.H. (1994). The partitioning of Fe, Ni, Cu, Pt, and Au between sulfide, metal, and fluid phases: A pilot study. Geochimica et Cosmochimica Acta, 58(2), 811826.Google Scholar
Barnes, S.J. (1986). The effect of trapped liquid crystallization on cumulus mineral compositions in layered intrusions. Contributions to Mineralogy and Petrology, 93(4), 524531.Google Scholar
Barnes, S.J. (1993). Partitioning of the platinum group elements and gold between silicate and sulphide magmas in the Munni Munni Complex, Western Australia. Geochimica et Cosmochimica Acta, 57, 12771290.Google Scholar
Barnes, S.J. and Campbell, I.H. (1988). Role of late magmatic fluids in Merensky-type platinum deposits: A discussion. Geology, 16, 488491.2.3.CO;2>CrossRefGoogle Scholar
Barnes, S.J. and Fiorentini, M.L. (2008). Iridium, ruthenium and rhodium in komatiites: Evidence for iridium alloy saturation. Chemical Geology, 257, 4458.CrossRefGoogle Scholar
Barnes, S.J. and Hoatson, D.M. (1994). The Munni Munni Complex, Western Australia: Stratigraphy, structure, and petrogenesis. Journal of Petrology, 35, 715751.Google Scholar
Barnes, S.J. and Naldrett, A.J. (1986). Geochemistry of the J-M Reef of the Stillwater Complex, Minneapolis Adit Area II. Silicate Mineral Chemistry and Petrogenesis. Journal of Petrology, 27(4), 791825.CrossRefGoogle Scholar
Barnes, S.J., McIntyre, J.R., Nisbet, B.W. and Williams, C.R. (1990). Platinum group element mineralisation in the Munni Munni Complex, Western Australia. Mineralogy and Petrology, 42, 141164.Google Scholar
Barnes, S.J., Keays, R.R. and Hoatson, D.M. (1992). Distribution of sulphides and PGE within the porphyritic websterite zone of the Munni Munni Complex, Western Australia. Australian Journal of Earth Sciences, 39(3), 289302.Google Scholar
Barnes, S.-J. and Fiorentini, M.L. (2008). Iridium, ruthenium and rhodium in komatiites: Evidence for iridium alloy saturation. Chemical Geology, 257, 4458.CrossRefGoogle Scholar
Barnes, S.-J. and Gomwe, T.S. (2011). The Pd deposit of the Lac des Iles Complex, Northwestern Ontario. Reviews in Economic Geology, 17, 351370.Google Scholar
Barnes, S-J. and Maier, W.D. (1999). The fractionation of Ni, Cu and the noble metals in silicate and sulfide liquids. In Dynamic Processes in Magmatic Ore Deposits and Their Application in Mineral Exploration (Keays, R.R., Lesher, C.M. Lightfoot, P.C. and Farrow, C.E.G., eds). Geological Association of Canada, Short Course Volume 13, 69106.Google Scholar
Barnes, S.-J. and Maier, W.D. (2002a). Platinum-group element distributions in the Rustenberg Layered Suite of the Bushveld Complex, South Africa. In The Geology, Geochemistry, Mineralogy and Mineral Beneficiation of Platinum-Group Elements (Cabri, L.J., ed). Canadian Institute of Mining, Metallurgy and Petroleum Special Volume, 54, 431458.Google Scholar
Barnes, S.-J. and Maier, W.D. (2002b). Platinum-group elements and microstructures of Normal Merensky Reef from Impala Platinum Mines, Bushveld Complex. Journal of Petrology, 43, 103128.Google Scholar
Barnes, S.-J., Naldrett, A.J. and Gorton, M.P. (1985). The origin of the fractionation of platinum-group elements in terrestrial magmas. Chemical Geology, 53, 303323.Google Scholar
Barnes, S.-J., Savard, D., Bédard, L.P. and Maier, W.D. (2009). Selenium and sulfur concentrations in the Bushveld Complex of South Africa and implications for formation of the platinum-group element deposits. Mineralium Deposita, 44, 647664.CrossRefGoogle Scholar
Barnes, S.-J., Maier, W.D. and Curl, E.A. (2010). Composition of the marginal rocks and sills of the Rustenburg Layered Suite, Bushveld Complex, South Africa: Implications for the formation of the platinum-group element deposits. Economic Geology, 105, 14911511.Google Scholar
Barnes, S.-J., Pagé, P., Prichard, H.M., Zientek, M.L. and Fisher, P.C. (2016). Chalcophile and platinum-group element distribution in the Ultramafic series of the Stillwater Complex, MT, USA—Implications for processes enriching chromite layers in Os, Ir, Ru, and Rh. Mineralium Deposita, 51(1), 2547.Google Scholar
Barry, M.A., Boudreau, B.P., Johnson, B.D. and Reed, A.H. (2010). First‐order description of the mechanical fracture behavior of fine‐grained surficial marine sediments during gas bubble growth. Journal of Geophysical Research, 115, F04029, doi:10.1029/2010JF001833.Google Scholar
Batanova, V.G. and Savelieva, G.N (2009). Melt migration in the mantle beneath spreading zones and formation of replacive dunites: A review. Russian Geology and Geophysics, 50, 763778.Google Scholar
Bazarkina, E.F., Pokrovski, G.S. and Hazemann, J.-L. (2014). Structure, stability and geochemical role of palladium chloride complexes in hydrothermal fluids. Geochimica et Cosmochimica Acta, 146, 107131.CrossRefGoogle Scholar
Bédard, J.H. (1994). A procedure for calculating the equilibrium distribution of trace elements among the minerals of cumulate rocks, and the concentration of trace elements in the coexisting liquids. Chemical Geology, 118, 143153.Google Scholar
Bédard, J.H., Sparks, R.S.J., Renner, R., Cheadle, M.J. and Hallworth, M.A. (1988). Peridotite sills and metasomatic gabbros in the Eastern Layered Series of the Rhum Complex. Journal of the Geological Society, London, 145, 207224.Google Scholar
Bédard, J.H., Marsh, B.D., Hershum, T.G., Naslund, H.R. and Musaka, S.B. (2007). Large scale mechanical redistribution of orthopyroxene and plagioclase in the Basement Sill, Ferrar Dolerites, Antarctica. Journal of Petrology, 48, 22892326.Google Scholar
Behrens, H. and Webster, J.D., eds. (2011). Sulfur in magmas and melts: Its importance for natural and technical processes. Mineralogical Society of America, Reviews in Mineralogy, 30.Google Scholar
Belien, I.B., Cashman, K.V. and Rempel, A.W. (2010). Gas accumulation in particle-rich suspensions and implications for bubble populations in crystal-rich magma. Earth and Planetary Science Letters, 297, 133140.Google Scholar
Bell, A.S., Siman, A. and Guilling, M. (2009). Experimental constraints on Pt, Pd and Au partitioning and fractionation in silicate melt–sulfide–oxide–aqueous fluid systems at 800 C, 150 MPa and variable sulfur fugacity. Geochimica et Cosmochimica Acta, 73, 57785792.Google Scholar
Berkowitz, B. and Ewing, R.P. (1998). Percolation theory and network modeling applications in soil physics. Surveys in Geophysics, 19, 2372.Google Scholar
Bezman, N.I., Gorbachev, P.N., Shalynin, A.I., Asif, M. and Naldrett, A.J. (2008). Solubility of platinum and palladium in silicate melts under high water pressure as a function of redox conditions. Petrology, 16(2), 161176.Google Scholar
Bezos, A., Lorand, J.-P., Humler, E. and Gros, M. (2005). Platinum-group element systematics in Mid-Oceanic Ridge basaltic glasses from the Pacific, Atlantic, and Indian Oceans. Geochimica et Cosmochimica Acta, 69, 26132627.Google Scholar
Bird, D.K., Arnason, J.G., Brandriss, M.E., Nevle, R.J, Radford, G., Bernstein, S., Gannicott, R.A. and Kelemen, P.B. (1995). A gold-bearing horizon in the Kap Edvard Holm Complex, East Greenland. Economic Geology, 90(5), 12881300.Google Scholar
Bird, D.K., Brooks, C.K., Gannicott, R.A. and Turner, P.A. (1991). A gold-bearing horizon in the Skaergaard intrusion. Economic Geology, 56, 157166.Google Scholar
Blander, M. and Katz, J.L (1975). Bubble nucleation in liquids. American Institute of Chemical Engineering Journal, 21(5), 833848.Google Scholar
Blaine, F.A. (2010). The effect of volatiles (H2O, Cl and CO2) on the solubility and partitioning of platinum and iridium in fluid-melt systems. Ph.D. Disseration, University of Waterloo, Ontario, Canada.Google Scholar
Blaine, F.A., Linnen, R.L., Holtz, F. and Brügmann, G.E. (2005). Platinum solubility in a haplobasaltic melt at 1250°C and 0.2 GPa: The effect of water content and oxygen fugacity. Geochimica et Cosmochimica Acta, 69(5), 12651273.Google Scholar
Blaine, F.A., Linnen, R.L., Holtz, F. and Brügmann, G.E. (2011). The effect of Cl on Pt solubility in haplobasaltic melt: Implications for micronugget formation and evidence for fluid transport of PGEs. Geochimica et Cosmochimica Acta, 75, 77927805.Google Scholar
Bodnar, R. J. (1994). Synthetic fluid inclusions; XII, The system H2O-NaCl; experimental determination of the halite liquidus and isochores for a 40 wt. % NaCl solution. Geochimica et Cosmochimica Acta, 53, 725733.Google Scholar
Bodnar, R.J., Burnham, C.W. and Sterner, S.M. (1985). Synthetic fluid inclusions in natural quartz. III. Determination of phase equilibrium properties in the system H2O-NaCl to 1000 C and 1500 bars. Geochimica et Cosmochimica Acta, 49(9), 18611873.CrossRefGoogle Scholar
Boorman, S.L., McGuire, J.B., Boudreau, A.E. and Kruger, F.J. (2003). Fluid overpressure in layered intrusions: Formation of a breccia pipe in the Eastern Bushveld Complex, Republic of South Africa. Mineralium Deposita, 38, 356369.Google Scholar
Boorman, S., Boudreau, A.E. and Kruger, F.J. (2004). The Lower Zone – Critical Zone transition of the Bushveld Complex: A quantitative textural study. Journal of Petrology, 45, 12091235.Google Scholar
Borisov, A. and Danyushevsky, L. (2011). The effect of silica contents on Pd, Pt and Rh solubilities in silicate melts: An experimental study. European Journal of Mineralogy, 23(3), 355367.Google Scholar
Borisov, S. and Palme, H. (1995). The solubility of iridium in silicate melts: New data from experiments with lr10Pt90 alloys. Geochimica et Cosmochimica Acta, 59(3), 481485.Google Scholar
Borisov, S. and Palme, H. (2000). Solubilities of noble metals in Fe-containing silicate melts as derived from experiments in Fe-free systems. American Mineralogist, 85(11–12), 16651673.Google Scholar
Boudreau, A.E. (1988). Investigations of the Stillwater Complex; IV, The role of volatiles in the petrogenesis of the J-M Reef, Minneapolis adit section. Canadian Mineralogist, 26, 193208.Google Scholar
Boudreau, A.E. (1992). Volatile fluid overpressure in layered intrusions and the formation of potholes. Australian Journal of Earth Sciences, 39, 277287.Google Scholar
Boudreau, A.E. (1993). Chlorine as an exploration guide for the Platinum-group elements in layered intrusions. Journal of Geochemical Exploration, 48, 2137.Google Scholar
Boudreau, A.E. (1994). Pattern formation during crystallization in two crystal, two component systems. South African Journal of Geology, 97, 473485.Google Scholar
Boudreau, A.E. (1995). Crystal aging and the formation of fine-scale layering. Mineralogy and Petrology, 54, 5569.CrossRefGoogle Scholar
Boudreau, A.E. (1999a). Fluid fluxing of cumulates: The J-M Reef and associated rocks of the Stillwater Complex, Montana. Journal of Petrology, 40(5), 755772.CrossRefGoogle Scholar
Boudreau, A.E. (1999b). PELE – A version of the MELTS software program for the PC platform. Computers and Geosciences, 25, 201203.Google Scholar
Boudreau, A.E. (2004). PALLADIUM – A program to model the chromatographic separation of the platinum-group elements, base metals and sulfur in a solidifying igneous crystal pile. Canadian Mineralogist, 42, 393403.CrossRefGoogle Scholar
Boudreau, A.E. (2008). Modeling the Merensky Reef, Bushveld Complex, South Africa. Contributions to Mineralogy and Petrology, 156, 431437.CrossRefGoogle Scholar
Boudreau, A.E. (2009). Transport of the platinum-group elements by igneous fluids in layered intrusions. In New Developments in Magmatic Ni-Cu and PGE Deposits (Li, C. and Ripley, E.M., eds.). Beijing: Geological Publishing House, 229249.Google Scholar
Boudreau, A.E. (2011). The evolution of texture and layering in layered intrusions. International Geology Review, 53(3–4), 330353.Google Scholar
Boudreau, A.E. (2016a). Bubble migration in a compacting crystal-liquid mush. Contributions to Mineralogy and Petrology, 171, 32. doi: 10.1007/s00410-016-1237-9.Google Scholar
Boudreau, A.E. (2016b). The Stillwater Complex, Montana – Overview and the significance of volatiles. Mineralogical Magazine, 80, 585637.Google Scholar
Boudreau, A.E. (2016c). The Bushveld complex as an analog to subduction zone hydrothermal systems. Penrose Conference on Layered Intrusions, Red Lodge Montana, 6–10 August, 2016 (abstract).Google Scholar
Boudreau, A.E. and Hoatson, D.M. (2004). Halogen variations in the paleoProterozoic layered mafic-ultramafic Intrusions of the East Kimberly, Western Australia: Implications for platinum-group element mineralization. Economic Geology, 99, 10151026.Google Scholar
Boudreau, A.E. and Kruger, F.J. (1990). Variations in the composition of apatite through the Merensky cyclic unit on the Western Bushveld Complex. Economic Geology, 85, 737745.Google Scholar
Boudreau, A.E. and McBirney, A.R. (1997). The Skaergaard Layered Series. Part III. In Situ Layering. Journal of Petrology, 38, 10031020.CrossRefGoogle Scholar
Boudreau, A.E. and McCallum, I.S. (1985). Features of the Picket Pin Pt-Pd deposit). In Stillwater Complex, Montana: Geology and Guide (Czamanske, C.K. and Zientek, M.L., eds.). Montana Bureau of Mines and Geology, Special Publication 92, 346357.Google Scholar
Boudreau, A.E. and McCallum, I.S. (1986). Investigations of the Stillwater Complex. Part III. The Picket Pin Pt-Pd deposit. Economic Geology, 81, 19531975.Google Scholar
Boudreau, A.E. and McCallum, I.S. (1989). Investigations of the Stillwater Complex: Part V. Apatites as indicators of evolving fluid composition. Contributions to Mineralogy and Petrology, 102, 138153.Google Scholar
Boudreau, A.E. and McCallum, I.S. (1990). Low temperature alteration of REE-rich chlorapatite from the Stillwater Complex Montana. American Mineralogist, 75, 87693.Google Scholar
Boudreau, A.E. and McCallum, I.S. (1992a). Concentration of platinum-group elements by magmatic fluids in layered intrusions. Economic Geology, 87, 18301848.CrossRefGoogle Scholar
Boudreau, A.E. and McCallum, I.S. (1992b). Infiltration metasomatism in layered intrusions—an example from the Stillwater Complex, Montana. Journal of Volcanology and Geothermal Research, 52, 171183.Google Scholar
Boudreau, A.E. and Meurer, W.P. (1999). Chromatographic separation of the platinum-group elements, gold, base metals and sulfur during degassing of a compacting and solidifying igneous crystal pile. Contributions to Mineralogy and Petrology, 134, 174185.Google Scholar
Boudreau, A.E. and Philpotts, A.J. (2002). Quantitative modeling of compaction in the Holyoke flood basalt flow, Hartford Basin, Connecticut. Contributions to Mineralogy and Petrology, 144, 176184.Google Scholar
Boudreau, A.E. and Simon, A. (2007). Halogen variations and degassing in the Basement Ferrar sill, Antarctica. Journal of Petrology, 48, 13691386. DOI:10.1093/petrology/egm022.Google Scholar
Boudreau, A.E., Mathez, E.A. and McCallum, I.S. (1986). Halogen geochemistry of the Stillwater and Bushveld Complexes: Evidence for t transport of the platinum-group elements by Cl-rich Fluids. Journal of Petrology, 27, 967986.Google Scholar
Boudreau, A.E., Love, C. and Hoatson, D. (1993). Halogen geochemistry of the Munni Munni Complex and associated intrusions of the Pilbara Block, W. Australia. Geochimica et Cosmochimica Acta, 57, 44674477.Google Scholar
Boudreau, A.E., Love, C. and Prendergast, M.D. (1995). Halogen geochemistry of the Great Dyke, Zimbabwe. Contributions to Mineralogy and Petrology, 122, 289300.CrossRefGoogle Scholar
Boudreau, A.E., Stewart, M.A. and Spivack, A.J. (1997). Stable Cl isotopes and origin of high-Cl magmas of the Stillwater Complex, Montana. Geology, 25(9), 791794.Google Scholar
Boudreau, A.E., Djon, L., Ychalikian, A. and Corkery, J. (2014). The Lac Des Iles Palladium Deposit, Ontario, Canada. Part I. The effect of variable alteration on the Offset Zone. Mineralium Deposita, 49, 625654.Google Scholar
Boudreau, A.E., Butak, K.C., Geraghty, E.P., Holick, P.A. and Koski, M.S. (2019). Mineral Deposits of the Stillwater Complex. In Geology of Montana). Montana Bureau of Mines and Geology (In press).Google Scholar
Boudreau, B.P. (2012). The physics of bubbles in surficial, soft cohesive sediments. Marine and Petroleum Geology, 38, 118.Google Scholar
Boudreau, B.P., Algar, C., Johnson, B.D., Croudace, I., Reed, A., Furukawa, Y., Dorgan, K.M., Jumar, P.A., Grader, A.S. and Gardiner, B.S. (2005). Bubble growth and rise in soft sediments. Geology, 33, 517520.Google Scholar
Boyce, J.W., Tomlinson, S.M., McCubbin, F.M., Greenwood, J.P. and Treiman, A.H. (2014). The lunar apatite paradox. Science, 344, 400402.Google Scholar
Bragagni, A., Luguet, A., Fonseca, R.O.C., Pearson, D.G., Lorand, J.-P.,Nowell, G.M. and Kjarsgaard, B.A. (2017). The geological record of base metal sulfides in the cratonic mantle: A microscale 187Os/188Os study of peridotite xenoliths from Somerset Island, Rae Craton (Canada). Geochimica et Cosmochimica Acta, 216, 264285.Google Scholar
Brandeis, G. and Jaupart, C. (1986). On the interaction between convection and crystallization in cooling magma chambers. Earth and Planetary Science Letters, 77, 345361.Google Scholar
Braun, K., Meurer, W. and Boudreau, A.E. (1994). Compositions of pegmatoids beneath the J-M Reef of the Stillwater Complex, Montana, U.S.A. Chemical Geology, 113, 245257.Google Scholar
Brearley, M. and Scarfe, C.M. (1986). Dissolution rates of upper mantle minerals in an alkali basalt melt at high pressure—An experimental study and implications for ultramafic xenolith survival. Journal of Petrology, 27, 11571182.Google Scholar
Brenan, J.M. (1993). Partitioning of fluorine and chlorine between apatite and aqueous fluids at high pressure and temperature: Implications for the F and Cl content of high P-T fluids. Earth and Planetary Science Letters, 117, 251263.Google Scholar
Brenan, J.M. (1994). Kinetics of fluorine, chlorine and hydroxyl exchange in fluorapatite. Chemical Geology, 110, 195210.Google Scholar
Brenan, J.M. and Andrews, D. (2001). High temperature stability of laurite and Ru-Os-Ir alloy and their role in PGE fractionation in mafic magmas. Canadian Mineralogist, 39, 341360.Google Scholar
Brenan, J.M., McDonough, W.F. and Dalpé, C. (2003). Experimental constraints on the partitioning of rhenium and some platinum-group elements between olivine and silicate melt. Earth and Planetary Science Letters, 212, 135150.Google Scholar
Brenan, J.M., McDonough, W.F., and Ash, R. (2005). An experimental study of the solubility and partitioning of iridium, osmium and gold between olivine and silicate melt. Earth and Planetary Science Letters, 237, 855872.Google Scholar
Brenan, J.M., Finnigan, C.S., McDonough, W.F. and Homolova, V., (2012). Experimental constraints on the partitioning of Ru, Rh, Ir, Pt and Pd between chromite and silicate melt: The importance of ferric iron. Chemical Geology, 302–303, 1632.Google Scholar
Broek, D. (1982). Elementary Engineering Fracture Mechanics, Fourth Edition. Dordrecht: Kluwer Academic Publishers.Google Scholar
Brown, G. M. and Peckett, A. (1977). Fluorapatites from the Skaergaard intrusion, East Greenland. Mineralogy Magazine 41, 227232.Google Scholar
Büchl, A., Brügmann, G. and Batanov, V.G. (2004). Formation of podiform chromitite deposits: Implications from PGE abundances and Os isotopic compositions of chromites from the Troodos complex, Cyprus. Chemical Geology, 208, 217232.Google Scholar
Buick, I.S., Gibson, R.L., Cartwright, I., Maas, R., Wallmach, T. and Uken, R. (2000) Fluid flow in metacarbonates associated with emplacement of the Bushveld Complex, South Africa. Journal of Geochemical Exploration, 69–70, 391395.Google Scholar
Buntin, T.J., Grandstaff, D.E., Ulmer, G.C. and Gold, D.P. (1985). A pilot study of geochemical and redox relationships between potholes and adjacent normal Merensky Reef of the Bushveld Complex. Economic Geology, 80, 975987.Google Scholar
Burnham, C.W. (1979). Magmas and hydrothermal fluids. In Geochemistry of Hydrothermal Ore Deposits, 2nd Edition (Barnes, H.L., ed.). New York: Wiley, 71136.Google Scholar
Burton, K.W., Schiano, P., Birck, J.-L. and Allègre, C. J. (1999) Osmium isotope disequilibrium between mantle minerals in a spinel-lherzolite. Earth and Planetary Science Letters, 172, 311322.Google Scholar
Cabri, L.J. (1981). The platinum-group minerals. In Platinum-Group Elements: Mineralogy, Geology, Recovery (Cabri, L.J., ed.). Canadian Institute of Mining, Metallurgy and Petroleum Special Volume 23, 13129.Google Scholar
Caltabiano, T., Burton, M., Giammanov, S., Allard, P., Bruno, N., Murè, F. and Romano, R. (2004). Volcanic gas emissions from the summit craters and flanks of Mt. Etna, 1987–2000. American Geophysical Union Geophysical Monograph doi:10.1029/143GM08.Google Scholar
Cameron, E.N. (1978). The Lower Zone of the Eastern Bushveld Complex in the Olifants River Trough. Journal of Petrology, 19(3), 437462.Google Scholar
Cameron, E.N. and Desborough, G.A. (1964). Origin of certain magnetite-bearing pegmatites in the eastern part of the Bushveld Complex, South Africa. Economic Geology, 59, 197225.Google Scholar
Campbell, I.H. (1978). Some problems with the cumulus theory. Lithos, 11, 311323.Google Scholar
Campbell, I.H. (1986). A fluid dynamic model for the potholes of the Merensky reef. Economic Geology, 81, 11181125.Google Scholar
Campbell, I.H. (1987). Distribution of orthocumulate textures in the Jimberlana intrusion. Journal of Geology, 95, 3554.Google Scholar
Campbell, I.H. and Murck, B.W. (1993). Petrology of the G and H Chromitite Zones in the Mountain View Area of the Stillwater Complex, Montana. Journal of Petrology, 34(2), 291316.Google Scholar
Campbell, I.H. and Naldrett, A.J. (1979). The influence of silicate: Sulfide ratios on the geochemistry of magmatic sulfides. Economic Geology, 74(6), 15031506.Google Scholar
Campbell, I.H., Roeder, P.L. and Dixon, J.M. (1978). Plagioclase buoyancy in basaltic liquids as determined with a centrifuge furnace. Contributions to Mineralogy Petrology, 67, 369377.Google Scholar
Campbell, I.H., Naldrett, A.J. and Barnes, S.J. (1983). A model for the origin of the platinum-rich sulphide horizons in the Bushveld and Stillwater complexes. Journal of Petrology, 24, 33165.Google Scholar
Candela, P.A. (1986). Toward a thermodynamic model for the halogens in magmatic systems: An application to melt-vapor-apatite equilibria. Chemical Geology, 57, 289301.Google Scholar
Candela, P.A. (1991). Physics of aqueous phase evolution in plutonic environments. American Mineralogist, 76, 10811091.Google Scholar
Capobianco, C.H. and Drake, M. (1990). Partitioning of ruthenium, rhodium, and palladium between spinel and silicate melt and implications for platinum-group element fractionation trends. Geochimica et Cosmochimica Acta, 54, 869874.Google Scholar
Carr, H.W., Groves, D.I. and Cawthorn, R.G. (1994). The importance of synmagmatic deformation in the formation of Merensky Reef potholes in the Bushveld Complex. Economic Geology, 89, 13981410.Google Scholar
Cashman, K.V. (1993). Relationship between plagioclase crystallization and cooling rate in basaltic melts. Contributions to Mineralogy Petrology, 113, 126142.Google Scholar
Cashman, K.V. and Ferry, J.M. (1988). Crystal size distribution (CSD) in rocks and the kinetics and dynamics of crystallization. III. Metamorphic crystallization. Contributions to Mineralogy Petrology, 99, 401415.Google Scholar
Cawthorn, R.G. (1976). Some chemical controls in igneous amphibole compositions. Geochimica et Cosmochimica Acta, 40, 13191328.Google Scholar
Cawthorn, R.G. (1982). An origin for the small scale fluctuations in orthopyroxene composition in the lower and critical zones of the Bushveld Complex, South Africa. Chemical Geology, 36, 227236.Google Scholar
Cawthorn, R.G. (1994). Formation of chlor- and fluor-apatite in layered intrusions. Mineralogical Magazine, 58, 299306.Google Scholar
Cawthorn, R.G. (2002). Delayed accumulation of plagioclase in the Bushveld Complex. Mineralogical Magazine, 66(6), 881893.Google Scholar
Cawthorn, R.G. and Poulton, K.L. (1988). Evidence for fluid in the footwall beneath potholes in the Merensky Reef of the Bushveld Complex. In Geo-Platinum 87 (Prichard, H.M., Potts, P.J., Bowles, J.F.W. and Cribbs, S.J., eds.), London: Elsevier, 343356.Google Scholar
Cawthorn, R.G. and Walsh, K.L. (1988). The use of phosphorus contents in yielding estimates of the proportion of trapped liquid in cumulates of the Upper Zone of the Bushveld Complex. Mineralogical Magazine, 52, 8189.Google Scholar
Cawthorn, R.G. and Lee, C. (1998). Field Excursion Guide to the Bushveld Complex. 8th International Platinum Symposium. The Geological Society of South Africa and The South African Institute of Mining and Metallurgy.Google Scholar
Cawthorn, R.G. and Boerst, K. (2006). Origin of the pegmatitic pyroxenite in the Merensky Unit, Bushveld Complex, South Africa. Journal of Petrology, 47, 15091530.Google Scholar
Cawthorn, R.G. and Webb, S.J. (2013). Cooling of the Bushveld Complex, South Africa: Implications for paleomagnetic reversals. Geology, 41(6), 687690.Google Scholar
Cawthorn, R.G., Harris, C. and Kruger, F.J. (2000). Discordant ultramafic pegmatoidal pipes in the Bushveld Complex. Contributions to Mineralogy and Petrology, 140, 119133.Google Scholar
Cawthorn, R.G., Lee, C., Schouwstra, R.P. and Mellowship, P. (2002). Relationship between PGE and PGM in the Bushveld Complex. The Canadian Mineralogist, 40, 311328.Google Scholar
Cawthorn, R.G., Latipov, R., Klemd, R. and Vuthuza, A. (2017). Origin of discordant ultramafic pegmatites in the Bushveld Complex, South Africa. Contributions to Mineralogy and Petrology (In review).Google Scholar
Chauveau, B. and Kaminski, E. (2008). Porous compaction in transient creep regime and implications for melt, petroleum, and CO2 circulation. Journal of Geophysical Research – Solid Earth, 113, B09406. doi:10.1029/2007JB005088CrossRefGoogle Scholar
Chitiyo, G., Schweitzer, J., de Waal, S., Lambert, P. and Olgilvie, P. (2008). Predictability of pothole characteristics and their spatial distribution at Rustenburg Platinum Mine. The Journal of The Southern African Institute of Mining and Metallurgy, 108, 733740.Google Scholar
Chou, I.-M. (1987). Phase relations in the system NaCl-KCl-H2O. III: Solubilities of halite in vapor-saturated liquids above 445°C and redetermination of phase equilibrium properties in the system NaCl-H2O to 1000°C and 1500 bars. Geochimica et Cosmochimica Acta, 51(7), 19651975.Google Scholar
Chutas, N.I., Bates, E., Prevec, S.A., Coleman, D.S. and Boudreau, A.E. (2012). Sr and Pb isotopic disequilibrium between coexisting plagioclase and orthopyroxene in the Bushveld Complex, South Africa: Microdrilling and progressive leaching evidence for sub-liquidus contamination within a crystal mush. Contributions to Mineralogy and Petrology, 163(4), 653668.Google Scholar
Clennell, M.B., Judd, A. and Hovland, M., (2000). Movement and accumulation of methane in marine sediments: Relation to gas hydrate systems. In Natural Gas Hydrate in Oceanic and Permafrost Environments, (Max, M.D., ed.). Dordrecht: Kluwer Academic Publishers, 105122.Google Scholar
Connolly, J.A.D. (2010). The mechanics of metamorphic fluid expulsion. Elements, 6(3), 165172. doi:10.2113/gselements.6.3.165Google Scholar
Connolly, J.A.D. and Podladchikov, Y.Y. (1998). Compaction – Driven fluid flow in viscoelastic rock. Geodinamica Acta, 11(2/3), 5584.Google Scholar
Connolly, J.A.D. and Podladchikov, Y.Y. (2007). Decompaction weakening and channeling instability in ductile porous media: Implications for asthenospheric melt segregation. Journal of Geophysical Research – Solid Earth, 112(B10), 15. doi:10.1029/2005jb004213Google Scholar
Coombs, M.L., Sisson, T.W. and Kimura, J.-I. (2004). Ultra-high chlorine in submarine Kilauea glasses: Evidence for direct assimilation of brine by magma. Earth and Planetary Science Letters, 217, 297313.Google Scholar
Crawford, A.J., Fallon, T.J. and Green, D.H. (1989). Classification, petrogenesis and tectonic setting of boninites. In Boninites and Related Rocks (Crawford, A.J., ed.). London: Unwin Hyman, 149.Google Scholar
Criscenti, L.J. (1984). The Origin of Macrorhythmic Units in the Stillwater Complex. M.S. thesis, University of Washington, Seattle, Washington.Google Scholar
Czamanske, G.K., Zientek, M.L. and Manning, C.E. (1991). Low-K granophyres of the Stillwater Complex, Montana. American Mineralogist, 76, 16461661.Google Scholar
Damian, S. Smith, D.S., Bassson, I.J. and Reid, D.L. (2003). Normal reef subfacies of the Merensky Reef at Northam Platinum Mine, Zwartklip Facies, Western Bushveld complex, Sourth Africa. The Canadian Mineralogist, 42, 243260.Google Scholar
Davidson, J.P. and Tepley, F.J. (1997). Recharge in volcanic systems: Evidence from isotope profiles of phenocrysts. Science, 275, 826829.Google Scholar
Davies, G. and Tredoux, M. (1985). The platinum-group element and gold contents of the marginal rocks and sills of the Bushveld Complex. Economic Geology, 80, 838848.Google Scholar
Davydov, M.N. (2012). Nucleation and growth of a gas bubble in magma. Journal of Applied Mechanics and Technical Physics, 53(3), 324332.Google Scholar
DePaolo, D.J. and Wasserburg, G.J. (1979). Sm-Nd age of the Stillwater complex and the mantle evolution curve for neodymium. Geochimica et Cosmochimica Acta, 43(7), 9991008.Google Scholar
De Vivo, B., Torok, K., Ayuso, R.A., Lima, A. and Lirer, L. (1995). Fluid inclusion evidence for magmatic silicate/saline/CO2 immiscibility and geochemistry of alkaline xenoliths from Ventotene Island, Italy. Geochimica et Cosmochimica Acta, 59(14), 29412953.Google Scholar
De Waal, S.A. (1977). Carbon dioxide and water from metamorphic reactions as agents for sulphide and spinel precipitation in mafic magmas. Geological Society of South Africa Transactions, 80, 198196.Google Scholar
Dixon, J. E., Clague, D.A. and Stolper, E.M. (1991). Degassing history of water, sulfur and carbon in submarine lavas from Kilaueau Volcano, Hawaii. Journal of Geology, 99, 371394.Google Scholar
Doherty, A.L., Webster, J.D., Goldoff, B.A. and Piccoli, P.M. (2014). Partitioning behavior of chlorine and fluorine in felsic melt–fluid(s)–apatite systems at 50 MPa and 850–950 °C. Chemical Geology, 384, 94109.Google Scholar
Donaldson, M.J. (1974). Petrology of the Munni Munni Complex, Roebourne, Western Australia. Journal of the Geological Society of Australia, 21, 116.Google Scholar
Dreibus, G., Palme, H., Spettel, B., Zipfel, J. and Wanke, H. (1995). Sulphur and selenium in chondritic meteorites. Meteoritics, 30, 439445.Google Scholar
Drinkwater, J.L., Czamanske, G.K. and Ford, A.B. (1990). Apatite of the Dufek intrusion: Disribution, paragenesis, and chemistry. Canadian Mineralogist, 28, 835854.Google Scholar
Duran, C.J., Barnes, S.-J. and Corkery, J.T. (2015). Chalcophile and platinum-group element distribution in pyrites from the sulfide-rich pods of the Lac des Iles Pd deposits, Western Ontario, Canada: Implications for post-cumulus re-equilibration of the ore and the use of pyrite compositions in exploration. Journal of Geochemical Exploration, 158, 223242.Google Scholar
Duran, C.J., Barnes, S.-J. and Corkery, J.T. (2016). Geology, petrography, geochemistry, and genesis of sulfide-rich pods in the Lac des Iles palladium deposits, western Ontario, Canada. Mineralium Deposita, 51(4), 509532.Google Scholar
Eales, H.V. and Cawthorn, R.G. (1996). The Bushveld Complex. In Layered Intrusions (Cawthorn, R.G., ed.). Amsterdam: Elsevier, 181229.Google Scholar
Eales, H.V., De Klerk, W.J. and Teigler, B. (1990). Evidence for magma mixing processes within the Critical and Lower Zones of the northwestern Bushveld Complex. Chemical Geology, 88, 261278.Google Scholar
Eales, H.V., De Klerk, W.J., Teigler, B. and Maier, W.D. (1994) Nature and origin of orthopyroxenites in the western Bushveld Complex, in the light of compositional data. South African Journal of Geology, 97(4), 399407.Google Scholar
Eales, H.V., Field, M., de Klerk, W.J. and Scoon, R.N. (1988). Regional trends of chemical variation and thermal erosion in the Upper Critical Zone, western Bushveld Complex. Mineralogical Magazine, 52: 6379.Google Scholar
Economou, M.I. (1986). Platinum group elements (PGE) in chromite and sulphide ores within the ultramafic zone of some Greek ophiolite complexes. In Metallogeny of Basic and Ultrabasic Rocks (Gallagher, M.J., Ixer, R.A., Neary, C.R. and Prichard, H.M., eds.). London: The Institution of Mining and Metallurgy, 441453.Google Scholar
Eggler, D.H. and Burnham, C.W. (1973). Crystallization and fractionation trends in the system andesite-H2O-CO2-O2 at pressures to 10 kilobars. Geological Society of America Bulletin 84, 25172532.Google Scholar
Elliot, W.C., Grandstaff, D.E., Ulmer, G.C. and Gold, D.P. (1982). An intrinsic oxygen fugacity study of platinum-carbon associations in layered intrusions. Economic Geology, 77, 14931510.Google Scholar
Eriksson, P.G., Hattingh, P.J. and Altermann, W. (1995). An overview of the geology of the Transvaal Sequence and Bushveld Complex. Mineralium Deposita, 30, 98111.Google Scholar
Eriksson, P.G., Schweitzer, K., Bosch, P.J.A., Schereiber, U.M., Van Deventer, J. and Hatton, C.J. (1993). The Transvaal Sequence: an overview. Journal of African Earth Sciences, 16(1/2), 2551.Google Scholar
Ertel, W., O’Neill, H.St.C., Sylvester, P.J. and Dingwell, D.B. (1999). Solubilities of Pt and Rh in a haplobasaltic silicate melt at 1300°C. Geochimica et Cosmochimica Acta, 63(16), 24392449.Google Scholar
Fahlquist, L.S. and Popp, R.K. (1989). The effect of NaCl on bunsenite solubility and Ni complexing in supercritical aqueous fluids. Geochimica et Cosmochimica Acta, 53, 989995.Google Scholar
Fehlhaber, K. and Bird, D. K. (1991). Oxygen-isotope and mineral alteration in gabbros of the Lower Layered Series, Kap Edvard Holm Complex, East Greenland. Geology, 19, 819822.Google Scholar
Feig, S.T, Koepke, J. and Snow, J.E. (2010). Effect of oxygen fugacity and water on phase equilibria of a hydrous tholeiitic basalt. Contributions to Mineralogy and Petrology, 160, 551568Google Scholar
Ferguson, J. and McCarthy, T.S. (1970). Origin of an ultramafic pegmatoid in the eastern part of the Bushveld Complex. Geological Society of South Africa Special Pub, 1, 7479.Google Scholar
Ferris, J., Johnson, A. and Strey, B. (1998). Form and extent of the Dufek intrusion, Antarctica, from newly compiled aeromagnetic data. Science Letters, 154, 185202.Google Scholar
Fiege, A., Behrens, H., Holtz, F. and Adams, F. (2014). Kinetic vs. thermodynamic control of degassing of H2O-S ± Cl-bearing andesitic melts. Geochimica et Cosmochimica Acta, 125, 241264Google Scholar
Finnigan, C.S., Brenan, J.M., Mungall, J.E. and McDonough, W.F. (2008). Experiments and models bearing on the role of chromite as a collector of platinum group minerals by local reduction. Journal of Petrology, 49, 16471665.Google Scholar
Fiorentini, M.L, Stone, W.E., Beresford, S.W. and Barley, M.E. (2004). Platinum-group element alloy inclusions in chromites from Archaean mafic-ultramafic units: Evidence from the Abitibi and the Agnew-Wiluna Greenstone Belts. Mineralogy and Petrology, 82, 341355.Google Scholar
Fleet, M.E. and Wu, T.-W. (1993). Volatile transport of platinum-group elements in sulfide-chloride assemblages at 1000 °C. Geochimica et Cosmochimica Acta, 57, 35193531.Google Scholar
Fleet, M.E. and Wu, T.-W. (1995). Volatile transport of precious metals at 1000 °C: Speciation, fractionation, and effect of base metal sulfide. Geochimica et Cosmochimica Acta, 59, 487495.Google Scholar
Fleet, M.E., Crocket, J.H. and Stone, W.E. (1996). Partitioning of platinum-group elements (Os, Ir, Ru, Pt, Pd) and gold between sulfide liquid and basalt melt. Geochimica et Cosmochimica Acta, 60, 23972412.Google Scholar
Fonseca, R.O.C., Campbell, I.H., O’Neill, H.S.C. and Allen, C.M. (2009). Solubility of Pt in sulfide mattes: Implications for the genesis of PGE-rich horizons in layered intrusions. Geochimica et Cosmochimica Acta, 73(19), 57645777.Google Scholar
Font, L., Davidson, J.P., Pearson, D.G., Nowell, G.M., Jerram, D.A. and Ottley, C.J. (2008). Sr and Pb isotope micro-analysis of Plagioclase Crystals from Skye Lavas: An insight into open-system processes in a Flood Basalt Province. Journal of Petrology, 49(8), 14491471.Google Scholar
Ford, C.E., Biggar, G.M., Humphries, D.J., Wilson, G., Dixon, D. and O’Hara, M.J. (1972). Role of water in the evolution of the lunar crust; an experimental study of sample 14310; an indication of lunar calc-alkaline volcanism. Proceedings of the Third Lunar Science Conference. Geochimica et Cosmochimica Acta, Suppl. 3, 207–229.Google Scholar
Frost, B.R. and Touret, J.L.R. (1989). Magmatic CO2 and saline melts from the Sybille monzosyenite, Laramie anorthosite complex, Wyoming. Contributions to Mineralogy and Petrology, 103, 178186.Google Scholar
Fowler, A.C., Rust, A.C. and Vynnycky, M. (2015). The formation of vesicular cylinders in pahoehoe lava flows. Geophysical and Astrophysical Fluid Dynamics, 109, 3961.Google Scholar
Gaetani, G.A. and Grove, T.L. (1998). The influence of water on melting of mantle peridotite. Contributions to Mineralogy and Petrology, 131, 323346.Google Scholar
Gaetani, G.A., Grove, T.L. and Bryan, E.B. (1994). Experimental phase relations of basaltic andesite from Hole 839B under hydrous and anhydrous conditions). In Proceedings of the Ocean Drilling Program, Scientific Results, 135 (Hawkins, J., et al., eds). College Station, Ocean Drilling Program, 557564.Google Scholar
Gafeira, J., Dolan, M.F.J. and Monteys, X. (2018). Geomorphic characterization of pockmarks by using a GIS-based semi-automatic toolbox. Geosciences, 8. doi:10.3390/geosciences8050154Google Scholar
Gain, S.B. (1985). The geologic setting of the platiniferous UG2 chromitite layer on the Farm Maandagshoek, eastern Bushveld Complex. Economic Geology, 80, 925943.Google Scholar
Gammons, C.H., Bloom, M.S. and Yu, Y. (1992). Experimental investigation of the hydrothermal geochemistry of platinum and palladium: I. Solubility of platinum and palladium sulfide minerals in NaCl/H2SO4 solutions at 300°C. Geochimica et Cosmochimica Acta, 56, 38813894.Google Scholar
Gardner, J.E., Hilton, M., M. and Carroll, M.R. (1999). Experimental constraints on degassing of magma: Isothermal bubble growth during continuous decompression from high pressure. Earth and Planetary Science Letters, 168, 201218.Google Scholar
Garuti, G., Bea, F., Zaccarini, F. and Montero, P. (2001). Age, geochemistry and petrogenesis of the ultramafic pipes in the Ivrea Zone, NW Italy. Journal of Petrology, 42, 433457.Google Scholar
Garuti, G., Zaccarini, F., Cabella, R., Fershtater, G., (1997). Occurrence of unknown Ru–Os–Ir–Fe oxides in the chromitites of the Nurali ultramafic complex, Southern Urals, Russia. Canadian Mineralogist, 35, 14311439.Google Scholar
Garuti, G., Zaccarini, F., Moiloshaq, V. and Alimov, V. (1999). Platinum-group minerals as indicators of sulfur fugacity in ophiolitic upper mantle; an example from chromitites of the Ray-Iz ultramafic complex, Polar Urals, Russia. Canadian Mineralogist, 37, 10991115.Google Scholar
Gee, J.S., Cheadle, M.J., Meurer, W.P. and Grimes, C.B. (2016). Magnetic Constraints on the Thermal History of the Dufek Intrusion. Geological Society of America Penrose Conference on Layered Mafic Intrusions and Associated Economic Deposits, Red Lodge, Montana, August 8–12, 2016 (Abstract).Google Scholar
Gerlach, T.M. and Graber, E.J. (1985). Volatile budget of Kilauea volcano. Nature, 313, 273277.Google Scholar
Ghiorso, M.S. and Sack, R.O. (1995). Chemical mass transfer in magmatic processes. IV. A revised and internally consistent thermodynamic model for the interpretation and extrapolation of liquid-solid equilibria in magmatic systems at elevated temperatures and pressures. Contributions to Mineralogy and Petrology, 119, 197212.Google Scholar
Ghiorso, M.S., Hirschman, M. and Sack, R.O. (1994). MELTS: Software for thermodynamic modeling of magmatic systems. EOS, 75, 571.Google Scholar
Gijbels, R.H., Millard, H.T. Jr., Desborough, G.A. and Bartel, A.J. (1974). Osmium, ruthenium, iridium and uranium in silicates and chromite from the eastern Bushveld Complex, South Africa. Geochimica et Cosmochimica Acta, 38, 319337.Google Scholar
Gleason, J.D., Gutzmer, J., Kesler, S.E. and Zwingmann, H. (2011). 2.05-Ga Isotopic Ages for Transvaal Mississippi Valley–Type Deposits: Evidence for Large-Scale Hydrothermal Circulation around the Bushveld Igneous Complex, South Africa. Journal of Geology, 119, 6980.Google Scholar
Glebovitsky, V.A., Semenov, V.S., Belyatsky, B.V., Koptev-Dvornikov, E.V., Pchelintseva, N.F., Kireev, B.S. and Koltsov, A.B. (2001). The structure of the Lukkulaisvaara Intrusion, Oulanka Group, Northern Karelia: Petrological Implications. Canadian Mineralogist, 39, 607637.Google Scholar
Godard, M., Bodinier, J.-L. and Vasseur, G. (1995). Effects of mineralogical reactions on trace element redistributions in mantle rocks during percolation processes: A chromatographic approach. Earth and Planetary Science Letters, 133, 449461.Google Scholar
Godel, B. (2015). Platinum-Group Element Deposits. In Layered Intrusions: Recent Advances in the Understanding of the Ore Forming Processes (Charlier, B., Namur, O., Latypov, R. and Tegner, C., eds.). Dordrecht: Springer Geology, 379432.Google Scholar
Godel, B., Barnes, S.-J. and Maier, W.D. (2006). 3-D distribution of sulphide minerals in the Merensky Reef (Bushveld Complex, South Africa) and the J-M Reef (Stillwater Complex, USA) and their relationship to microstrucures using X-ray computed tomography. Journal of Petrology, 47, 18531872.Google Scholar
González-Jiménez, J.M., Griffin, W.L., Gervilla, F., Proenza, J.A., O’Reilly, S.Y. and Pearson, N.J. (2013). Chromitites in ophiolites: How, where, when, why? Part I. A review and new ideas on the origin and significance of platinum-group minerals. Lithos, 189, 127139.Google Scholar
Graham, C.M., Valley, J.W., Eiler, J.M. and Wada, H. (1998). Timescales and mechanisms of fluid infiltration in a marble: An ion microprobe study. Contributions to Mineralogy and Petrology, 132, 371-389.Google Scholar
Grant, N.K. and Moiling, P.A. (1981). A strontium isotope and trace element profile through the Partridge River tractolite, Duluth Complex, Minnesota. Contributions to Mineralogy and Petrology, 77, 296305.Google Scholar
Green, D.H., Schmidt, M.W. and Hibberson, W.O. (2004). Island-arc Ankaramites: Primitive Melts from Fluxed Refractory Lherzolitic Mantle. Journal of Petrology, 45, 392403.Google Scholar
Grove, T.L., Till, C.B. and Krawczynski, M.J. (2012). The role of H2O in subduction zone magmatism. Annual Review of Earth and Planetary Sciences, 40, 413439.Google Scholar
Gurioli, L., Harris, A.J.L., Houghton, B.F., Polacci, M. and Ripepe, M. (2008). Textural and geophysical characterization of explosive basaltic activity at Villarrica volcano. Journal of Geophysical Research, 113, B08206. doi:10.1029/2007JB005328.Google Scholar
Halkoaho, T.A. and Alapieti, T.T. (1993). Variation in F and Cl-contents of apatite in the Penikat layered intrusion, Northern Finland. In Symposium on layering in igneous complexes, September, 1993, Johannesburg, South Africa, Abstracts Volume, 18–19.Google Scholar
Hamilton, D.L. and Anderson, G.M. (1967). Effects of water and oxygen pressure on the crystallization of basaltic magmas, In Basalts (Hess, H.H. and Poldervaart, A., eds.). New York: Wiley Interscience, 445482.Google Scholar
Hamilton, J. (1977). Sr isotope and trace element studies of the Great Dyke and Bushveld mafic phase and their relation to early Proterozoic magma genesis in southern Africa. Journal of Petrology, 18, 2452.Google Scholar
Hamlyn, P.R. and Keays, R.R. (1986). Sulfur saturation and second-stage melts: Application to the Bushveld platinum metal deposits. Economic Geology, 81, 14311445.Google Scholar
Hanley, J.J. and Gladney, E.R. (2011). The presence of carbonic-dominant volatiles during the crystallization of sulfide-bearing mafic pegmatites in the North Roby Zone, Lac des Iles Complex, Ontario. Economic Geology, 106, 3354.Google Scholar
Hanley, J., Pettke, T., Mungall, J. and Spooner, E.T.C. (2005a) Fluid and melt inclusion evidence for platinum-group element transport by high salinity fluids and halide melts below the J-M Reef, Stillwater Complex, Montana, U.S.A., 10th Platinum Symposium (abstract)Google Scholar
Hanley, J.J., Mungall, J.E., Pettke, T., Spooner, E.T.C. and Bray, C.J. (2005b). Ore metal redistribution by hydrocarbon–brine and hydrocarbon–halide melt phases, North Range footwall of the Sudbury Igneous Complex, Ontario, Canada. Mineralium Deposita, 40, 237256.Google Scholar
Hanley, J.J., Pettke, T., Mungall, J.E. and Spooner, E.T.C. (2005c). The solubility of platinum and gold in NaCl brines at 1.5 kbar, 600 to 800°C: A laser ablation ICP-MS pilot study of synthetic fluid inclusions. Geochimica et Cosmochimica Acta, 69(10), 25932611.Google Scholar
Hanley, J.J., Mungall, J.E., Pettke, T., Spooner, E.T.C. and Bray, C.J. (2008). Fluid and halide melt inclusions of magmatic origin in the Ultramafic and Lower Banded series, Stillwater Complex, Montana, USA. Journal of Petrology, 49(6), 11331160.Google Scholar
Hanley, J.J., Ames, D., Barnes, J., Sharp, Z. and Guillong, M. (2011). Interaction of magmatic fluids and silicate melt residues with saline groundwater in the footwall of the Sudbury Igneous Complex, Ontario, Canada: New evidence from bulk rock geochemistry, fluid inclusions and stable isotopes. Chemical Geology, 281, 125.Google Scholar
Hardy, S.C. and Voorhees, P.W. (1988). Ostwald ripening in a system with a high volume fraction of coarsening phase. Metallurgical and Materials Transactions A, 19A, 27132721.Google Scholar
Harper, M.P. (2004). Platinum Group Element Mineralization in “Ballrooms” of the J-M Reef of the Stillwater Complex, Montana. M.S. thesis, Brigham Young University, Provo.Google Scholar
Harmer, R.E. and Sharpe, M.R. (1985). Field relations and Sr isotope systematics of the marginal rocks of the eastern Bushveld Complex. Economic Geology, 80, 813837.Google Scholar
Harmer, R.E., Auret, J.M. and Eglington, B.M. (1995). Lead isotope variations within the Bushveld complex, Southern Africa: A reconnaissance study. Journal of African Earth Sciences, 21(4), 595606.Google Scholar
Harris, C. and Chaumba, J.B. (2001). Crustal contamination and the fluid-rock interaction during the formation of the Platreef, Northern limb of the Bushveld Complex, South Africa. Journal of Petrology, 42, 13211347.Google Scholar
Harris, C., Pronost, J.J.M., Ashwal, L.D. and Cawthorn, R.G. (2005). Oxygen and Hydrogen Isotope Stratigraphy of the Rustenburg Layered Suite, Bushveld Complex: Constraints on Crustal Contamination. Journal of Petrology, 46(3), 579601.Google Scholar
Harris, N., McMillan, A., Holness, M., Uken, R., Watkeys, M., Rogers, N. and Fallick, A. (2003). Melt generation and fluid flow in the thermal aureole of the Bushveld Complex. Journal of Petrology, 44(6), 10311054.Google Scholar
Hart, S.R. and Ravizza, G.E. (1996). Os partitioning between phases in lherzolite and basalt. Geophysical Monographs, 95, 123134.Google Scholar
Hart, S.R. and Kinloch, E.D. (1989). Osmium isotope systematics in Witwatersrand and Bushveld ore deposits. Economic Geology, 84, 16511655.Google Scholar
Harvey, J., Warren, J.M. and Shirey, S.B. (2016). Mantle sulfides and their role in Re–Os and Pb isotope geochronology. Reviews in Mineralogy and Geochemistry, 81, 579649.Google Scholar
Haskin, L.A. and Salpas, P.A. (1992). Genesis of compositional characteristics of Stillwater AN-I and AN-II thick anorthosite units. Geochimica et Cosmochimica Acta, 56, 11871212.Google Scholar
Hatch, F.H., Wells, A.K. and Wells, M.K. (1974). Petrology of the Igneous Rocks. London: Thomas Murphy and Company.Google Scholar
Hauri, E.H., Gaetani, G.A. and Green, T.H. (2006). Partitioning of water during melting of the Earth’s upper mantle at H2O-undersaturated conditions. Earth and Planetary Science Letters, 248, 715734.Google Scholar
Haughton, D.R., Roeder, P.L. and Skinner, B.J. (1974). Solubility of sulfur in mafic magmas. Economic Geology, 69, 451467.Google Scholar
Hawkesworth, C.j., Gallagher, K., Hergt, J.M. and McDermott, F. (1993). Manlte and slab contributions in arc magmas. Annual Review of Earth and Planetary Sciences, 21, 175204.Google Scholar
Helz, R.T. (1976). Phase relations of basalts in their melting ranges at PH2O = 5 kb. Part II. Melt compositions. Journal of Petrology, 17(2), 139193.Google Scholar
Helz, R.T. (1985). Compositions of fine-grained mafic rocks from sills and dikes associated with the Stillwater complex. In Stillwater complex, Geology and Guide (Czamanske, G.K. and Zientek, M.L., eds.). Montana Bureau of Mines and Geology Special Publication 92, pp.97–117.Google Scholar
Henderson, P. (1970). The significance of mesostasis of basic layered igneous rocks. Journal of Petrology, 11(3), 463473.Google Scholar
Hermann, J., Spandler, C. Hack, A. and Korsakov, A.V. (2006). Aqueous fluids and hydrous melts in high-pressure and ultra-high pressure rocks: Implications for element transfer in subduction zones. Lithos, 92, 399417.Google Scholar
Hess, H.H. (1960). Stillwater igneous complex, Montana – A quantitative mineralogical study. Geological Society of America, Memoir 80.Google Scholar
Hiemstra, S. (1979). The role of collectors in the formation of the platinum deposits in the Bushveld Complex. Canadian Mineralogist, 17, 469482.Google Scholar
Higgins, M.D. (1991). The origin of laminated and massive anorthosite, Sept Iles Layered Intrusion, Québec, Canada. Contributions to Mineralogy and Petrology, 106, 340354.Google Scholar
Higgins, M.D. (1998). Origin of anorthosite by textural coarsening: Quantitative measurements of a natural sequence of textural development. Journal of Petrology, 39, 13071325.Google Scholar
Higgins, M.D. (2002). The role of textural coarsening in the development of the Kiglapait layered mafic intrusion, Labrador, Canada: A crystal size distribution study. Contributions to Mineralogy and Petrology, 144, 314330.Google Scholar
Higgins, M.D. (2006). Quantitative Textural Measurements in Igneous and Metamorphic Petrology: Cambridge, Cambridge University Press, 275 p.Google Scholar
Higgins, M.D. (2011). Textural coarsening in igneous rocks: International Geology Review, 53(3–4), 354376.Google Scholar
Hoatson, D.M. (1991). The petrology and platinum-group element geochemistry of the Munni Munni and Mount Sholl layered mafic-ultramafic intrusions of the West Pilbara Block, Western Australia. Unpublished PhD Dissertation, University of Melbourne.Google Scholar
Hoatson, D.M. and Keays, R.R., (1989). Formation of platiniferous sulfide horizons by crystal fractionation and magma mixing in the Munni Munni layered intrusion, West Pilbara Block, Western Australia. Economic Geology, 84, 17751804.Google Scholar
Hoatson, D.M., Sun, S.-S. and Keays, R.R. (1992). Petrogenesis of the west Pilbara layered mafic-ultramafic intrusions. In Petrology and Platinum-Group-Element Geochemistry of Archaean Layered Mafic-Ultramafic Intrusions, West Pilbara Block, Western Australia (Hoatson, D.M. et al., eds.). Australian Geological Survey Organization Bulletin, 242, 150–66.Google Scholar
Holland, H.D. (1972). Granites, solutions, and base metal deposits. Economic Geology, 67, 281301.Google Scholar
Holland, T. and Blundy, J. (1994). Non-ideal interactions in calcic amphiboles and their bearing on amphibole-plagioclase thermometry. Contributions to Mineralogy and Petrology, 116, 433447.Google Scholar
Holloway, J.R. and Burham, C.W. (1972). Melting relations of basalt with equilibrium water pressure less than total pressure. Journal of Petrology, 13, 129.Google Scholar
Holness, M.B., Hallworth, M.A., Woods, A. and Sides, R.E. (2007). Infiltration Metasomatism of Cumulates by Intrusive Magma Replenishment: The Wavy Horizon, Isle of Rum, Scotland. Journal of Petrology, 48, 563587.Google Scholar
Holtz, F., Pichavant, M., Barbey, P. and Johannes, W. (1992). Effects of H2O on liquidus phase relations in the haplogranite system at 2 and 5 kbar. American Mineralogist, 77, 12231241.Google Scholar
Holzheid, A. and Grove, T.L. (2002) Sulfur saturation limits in silicate melts and their implications for core formation scenarios for terrestrial planets. American Mineralogist, 87, 227237.Google Scholar
Hoshide, T. and Obata, M. (2010). Zoning and resorption of plagioclase in a layered gabbro, as a petrographic indicator of magmatic differentiation. Earth and Environmental Science Transactions of the Royal Society of Edinburgh, 100, 235249.Google Scholar
Hovis, G.L. and Harlov, D.E. (2010). Solution calorimetric investigation of fluor-chlorapatite crystalline solutions. American Mineralogist, 95, 946952.Google Scholar
Hovland, M. and Judd, A.G. (1988). Seabed Pockmarks and Seepages: Impact on Geology, Biology and the Marine Environment. London: Graham and Trotman (Kluwer).Google Scholar
Hsu, L.C., Lechler, P.J. and Nelson, J.H. (1991). Hydrothermal solubility of palladium in chloride solutions from 300° to 700°C: Preliminary results. Economic Geology, 86, 422427.Google Scholar
Huber, C., Bachmann, O., Vigneresse, J.L., Dufek, J. and Parmigiani, A. (2012). A physical model for metal extraction and transport in shallow magmatic systems. Geochemistry, Geophysics, Geosystems, 13, Q08003. doi:10.1029/2012GC004042.Google Scholar
Huber, C. and Parmigiani, A. (2018). A physical model for three-phase compaction in Silicic Magma Reservoirs. Journal of Geophysical research: Solid Earth, 123. doi.org/10.1002/2017JB015224Google Scholar
Hulbert, L.J., Carne, R.C., Gregoire, C.D. and Paktunc, D. (1992). Sedimentary nickel, zinc and platinum-group-element mineralization in Devonian black shales at the Nick Property, Yukon, Canada: A new deposit type. Exploration and Mining Geology, 1, 3962.Google Scholar
Hulbert, L.J. and Von Gruenewaldt, G. (1985) Textural and compositional features of chromite in the Lower and Critical zones of the Bushveld complex south of Potgietersrus. Economic Geology, 80, 872895.Google Scholar
Hunter, R.H., (1987). Textural equilibrium in layered igneous rocks. In Origins of Igneous Layering (Parsons, I., ed.). NATO ASI Series C, 196, 473503.Google Scholar
Huntington, H. D. (1979). Kiglapait mineralogy I: apatite, biotite, and volatiles. Journal of Petrology, 20, 625652.Google Scholar
Hurwitz, S. and Navon, O. (1994). Bubble nucleation in rhyolitic melts: Experiments at high pressure, temperature, and water content. Earth and Planetary Science Letters, 122, 267280.Google Scholar
Ireland, R.H.P. and Penniston-Dorland, S.C. (2015). Chemical interactions between a sedimentary diapir and surrounding mafic magma: Evidence from the Phepane Dome and Bushveld Complex, South Africa. American Mineralogist, 100, 19852000.Google Scholar
Irvine, T.N. (1965). Chromian spinel as a petrogenetic indicator. Part I: Theory. Canadian Journal of Earth Sciences, 2(6), 648672.Google Scholar
Irvine, T.N. (1967). Chromian spinel as a petrogenetic indicator. Part II: Petrologic implications. Canadian Journal of Earth Sciences, 4, 71103.Google Scholar
Irvine, T.N. (1974). Petrology of the Duke Island Ultramafic Complex, Southeastern Alaska. Geological Society of America Memoir 128, 240 p.Google Scholar
Irvine, T.N. (1975). Crystallization sequences in the Muskox intrusion and other layered intrusions. II. Origin of chromite layers and similar deposits of other magmatic ores. Geochimica et Cosmochimica Acta, 39, 9911020.Google Scholar
Irvine, T.N. (1977). Origin of chromite layers in the Muskox intrusion and other stratiform intrusions: A new interpretation. Geology, 5, 273277.Google Scholar
Irvine, T.N. (1980). Magmatic infiltration metasomatism, double diffusive fractional crystallization and adcumulus growth in the Muskox Intrusion and other layered intrusions. In Physics of Magmatic Processes (Hargraves, R.B., ed.). Princeton: Princeton University Press, 325384.Google Scholar
Irvine, T.N., Keith, D.W. and Todd, S.G. (1983). The J-M platinum-palladium reef of the Stillwater Complex, Montana: II. Origin by double-diffusive convective magma mixing and implications for the Bushveld Complex. Economic Geology, 78, 12871334.Google Scholar
Jackson, E.D. (1961). Primary Textures and Mineral Associations in the Ultramafic Zone of the Stillwater Complex: Montana. United States Geological Survey Professional Paper 358.Google Scholar
Jackson, E.D. (1963) Stratigraphic and lateral variation of chromite composition in the Stillwater Complex). Mineralogical Society of America Special Paper, 1, 46–54.Google Scholar
Jain, A. and Juanes, R. (2009). Preferential mode of gas invasion in sediments: Grain-scale mechanistic model of coupled multiphase fluid flow and sediment mechanics. Journal of Geophysical Research – Solid Earth, 114, B08101. http://dx.doi.org/10.1029/2008JB006002.Google Scholar
Jaupart, C. and Brandeis, G. (1986). The stagnant bottom layer of convecting magma chambers. Earth and Planetary Sciences Letters, 80, 183199.Google Scholar
Joachim, B., Pawley, A., Lyon, I.C., Marquardt (neé Hartmann), K., Henkel, T., Clay, P.L., Ruzié, L., Burgess, R. and Ballentine, C.J. (2015). Experimental partitioning of F and Cl between olivine, orthopyroxene and silicate melt at Earth’s mantle conditions. Chemical Geology, 416, 6578.Google Scholar
Johnson, J.W., Oelkers, E.H. and Helgeson, H.C. (1992). SUPCRT92: A software package for calculating the standard molal thermodynamic properties of minerals, gases, aqueous species and reactions from 1 to 5000 bars and 0° to 1000 °C. Computers and Geosciences, 18, 899947.Google Scholar
Johnson, B.D., Boudreau, B.P., Gardiner, B.S. and Maass, R. (2002). Mechanical response of sediments to bubble growth. Marine Geology, 187, 347363.Google Scholar
Jones, M. (1994). Mechanical principles of sediment deformation. In The Geological Deformation of Sediments (Maltman, A., ed.). London: Chapman and Hall, 3771.Google Scholar
Kamenetsky, V.S. (2006). Melt inclusion record of magmatic immiscibility in crustal and mantle magmas. In Melt Inclusions in Plutonic Rocks (Webster, J.D., ed.). Mineralogical Association of Canada Short Course Series, 36, 8198.Google Scholar
Kanitpanyacharoen, W. and Boudreau, A. (2013). Sulfide-associated mineral assemblages in the Bushveld Complex, South Africa: Platinum-group element enrichment by vapor refining by chloride–carbonate fluids. Mineralium Deposita, 48, 193210.Google Scholar
Keays, R.R. (1982). Palladium and iridium in komatiites and associated rocks: Application to petrogenetic problems. In Komatiites (Arndt, N.T. and Nisbet, E.G., eds.). Berlin: Springer, 435457.Google Scholar
Keays, R.R. (1989). Formation of platiniferous sulfide horizons by crystal fractionation and magma mixing in the Munni Munni layered intrusion, West Pilbara Block, Western Australia. Economic Geology, 84, 17751804.Google Scholar
Keays, R.R. (1995). The role of komatiitic and picritic magmatism and S-saturation in the formation of ore deposits. Lithos, 43(1), 118.Google Scholar
Keays, R.R. and Lightfoot, P.C. (2010). Crustal sulfur is required to form magmatic Ni–Cu sulfide deposits: Evidence from chalcophile element signatures of Siberian and Deccan Trap basalts. Mineralium Deposita, 45(3), 241257.Google Scholar
Keays, R.R. and Tegner, C. (2016). Magma Chamber Processes in the Formation of the Low-sulphide Magmatic Au–PGE Mineralization of the Platinova Reef in the Skaergaard Intrusion, East Greenland. Journal of Petrology, 56, 23192340.Google Scholar
Keays, R.R., Lightfoot, P.C. and Hamlyn, P.R. (2012). Sulfide saturation history of the Stillwater Complex, Montana: Chemostratigraphic variation in platinum group elements. Mineralium Deposita, 47, 151173.Google Scholar
Kelemen, P.B., Dick, H.J.B. and Quick, J.E. (1992). Formation of harzburgite by pervasive melt/rock reaction in the upper mantle. Nature, 358, 635641.Google Scholar
Kelemen, P., Hirth, G., Shimizu, N., Spiegelman, M. and Dick, H. (1997). A review of melt migration processes in the adiabatically upwelling mantle beneath oceanic spreading ridges. Philosophical Transactions of the Royal Society of London, Series A, 355, 283318.Google Scholar
Kelemen, P.B., Shimizu, N. and Salters, V.J.M. (1995). Extraction of mid-ocean-ridge basalt from the upwelling mantle by focused flow of melt in dunite channels. Nature, 375, 747753.Google Scholar
Keller, T. and Katz, R.F. (2016). The role of volatiles in reactive melt transport in the asthenosphere. Journal of Petrology, 57(6), 10731108.Google Scholar
Keller, T., Katz, R.F. and Hirschmann, M.M. (2017). Volatiles beneath mid-ocean ridges: Deep melting, channelised transport, focusing, and metasomatism. Earth and Planetary Science Letters, 464, 5568.CrossRefGoogle Scholar
Khitarov, N.I., Lebedev, Ye.B., Dorfman, A.M. and Bagdasarov, N.Sh. (1979). Effects of temperature, pressure, and volatiles on the surface tension of molten basalt. Geochemistry International, 16(5), 7886.Google Scholar
Kinloch, E.D. (1982). Regional trends in the platinum-group mineralogy of the Critical zone of the Bushveld Complex, South Africa. Economic Geology, 77, 13281347.Google Scholar
Kinloch, E.D. and Peyerl, W. (1990). Platinum-group minerals in various rock types of the Merensky Reef: Genetic implications. Economic Geology, 85, 537555.Google Scholar
Konnerup-Madsen, J. and Rosew-Hansen, J. (1982). Volatiles associated with alkaline igneous rift activity: Fluid inclusions in the Ilimaussaq intrusion and Gardar granite complexes (South Greenland). Chemical Geology, 37, 7993.Google Scholar
Korzhinskii, D.S. (1959). Physicochemical Basis of the Analysis of the Paragenesis of Minerals. London, Chapman and Hall.Google Scholar
Korzhinskii, D.S. (1968). The theory of metasomatic zoning. Mineralium Deposita, 3, 222231.Google Scholar
Kruger, F.J. (1990). The stratigraphy of the Bushveld Complex: A reappraisal and the relocation of the Main Zone boundaries. South African Journal of Geology, 93, 376381.Google Scholar
Kruger, F.J. (1994). The Sr-isotopic stratigraphy of the western Bushveld Complex. South African Journal of Geology, 97, 393398.Google Scholar
Kruger, F.J. and Marsh, J.S. (1982). Significance of 87Sr/86Sr ratios in the Merensky cyclic unit of the Bushveld Complex. Nature, 298, 5355.Google Scholar
Kusebauch, C., John, T., Whitehouse, M.J., Klemme, S. and Putnis, A. (2015). Distribution of halogens between fluid and apatite during fluid-mediated replacement processes. Geochimica et Cosmochimica Acta, 170, 225246.Google Scholar
Kushiro, I. (1975). On the nature of silicate melt and its significance in magma genesis: Regularities in the shift of the liquidus boundaries involving olivine, pyroxene, and silica minerals. American Journal of Science, 275, 411431.Google Scholar
Kushiro, I., Yoder, H.S. and Nishikawa, M. (1968). Effect of water on the melting of enstatite. Geological Society of America Bulletin, 79, 16851692.Google Scholar
Labotka, T.C. and Kath, R.L. (2001). Petrogenesis of the contact-metamorphic rocks beneath the Stillwater Complex, Montana. Geological Society of America Bulletin, 113, 13121323.Google Scholar
Lackey, J.S. and Valley, J.W. (2004). Complex patterns of fluid flow during wollastonite formation in calcareous sandstones at Laurel Mountain, Mt. Morrison Pendant, California. Geological Society of America Bulletin, 116(1/2), 7693.Google Scholar
Lake, E.T. (2013). Crystallization and saturation front propagation in silicic magma chambers. Earth and Planetary Science Letters, 383, 182193.Google Scholar
Lambert, D.D. and Simmons, E.C. (1988). Magma evolution in the Stillwater Complex, Montana: II. Rare earth element evidence for the formation of the J-M Reef. Economic Geology, 83, 11091126.Google Scholar
Langmuire, C.H. (1989). Geochemical consequences of in situ crystallization. Nature, 340, 199205.Google Scholar
Larsen, R.B., Brooks, C.K. and Bird, D. K. (1992). Methane-bearing, aqueous, saline solutions in the Skaergaard intrusion, east Greenland. Contributions to Mineralogy and Petrology, 112, 428437.Google Scholar
Latypov, R., O’Driscoll, B. and Lavrenchuk, A. (2013). Towards a model for the in situ origin of PGE reefs in layered intrusions: Insights from chromitite seams of the Rum Eastern Layered Intrusion, Scotland. Contributions to Mineralogy and Petrology, 166, 309327.Google Scholar
Latypov, R. and 16 others (2015). A fundamental dispute: A discussion of “On some fundamentals of igneous petrology” by Bruce D. Marsh, Contributions to Mineralogy and Petrology (2013) 166: 665–690. Contributions to Mineralogy and Petrology, 169, DOI 10.1007/s00410-015-1108-9Google Scholar
Latypov, R., Costin, G., Chistyakova, S., Hunt, E.J., Mukherjee, R. and Naldrett, A. (2018). Platinum-bearing chromite layers are caused by pressure reduction during magma ascent. Nature Communications. DOI: 10.1038/s41467-017-02773-wGoogle Scholar
Laurenz, V., Fonseca, R.O.C., Ballhaus, C. and Sylvester, P.J. (2010). The solubility of palladium in picritic melts: 1. The effect of iron. Geochimica et Cosmochimica Acta, 74, 29892998.Google Scholar
Laurenz, V., Fonseca, R.O.C., Ballhaus, C., Jochum, K.P., Heuser, A. and Sylvester, P.J. (2013). The solubility of palladium and ruthenium in picritic melts: 2. The effect of sulfur. Geochimica et Cosmochimica Acta, 108, 172183.Google Scholar
Le Guen, Y., Renard, F., Hellmann, R., Brosse, E., Collombet, M., Tisserand, D. and Gratier, J.P. (2007). Enhanced deformation of limestone and sandstone in the presence of high PCO2 fluids. Journal of Geophysical Research – Solid Earth, 112(B5), B05421.Google Scholar
Le Roux, V., Bodinier, J.-L. Alard, O., O;Reilly, S.Y. and Griffin, W.L. (2009). Isotopic decoupling during porous melt flow: A case-study in the Lherz peridotite. Earth and Planetary Science Letters, 279, 7685.Google Scholar
Lee, C.A. and Butcher, A.R. (1990). Cyclicity in the Sr isotope stratigraphy through the Merensky and Bastard Reef units, Atok section, eastern Bushveld Complex. Economic Geology, 85, 877883.Google Scholar
Lee, C.-T.A. and Morton, D.M. (2015). High silica granites: Terminal porosity and crystal settling in shallow mama chambers. Earth and Planetary Science Letters, 409, 2331.Google Scholar
Leeman, W.P. and Dasch, E.J. (1978). Strontium, lead, and oxygen isotopic investigation of the Skaergaard Intrusion, East Greenland. Earth and Planetary Science Letters, 41, 4759.Google Scholar
Lenormand, R., Touboul, E. and Zarcone, C. (1988). Numerical models and experiments on immiscible displacements in porous media. Journal of Fluid Mechanics, 189, 165187.Google Scholar
Li, C. and Boudreau, A.E. (2017). The origin of high Cu/S sulfides by shallow level degassing in the Skaergaard Intrusion, East Greenland. Geology, 45(12), 10751078.Google Scholar
Li, C. and Ripley, E.M. (2005). Empirical equations to predict the sulfur content of mafic magmas at sulfide saturation and applications to magmatic sulfide deposits. Mineralium Deposita, 40, 218230.Google Scholar
Li, C. and Ripley, E.M. (2009). Sulfur contents at sulfide-liquid or anhydrite saturation in silicate melts: Empirical equations and example applications. Economic Geology, 104, 405412.Google Scholar
Li, C., Ripley, E.M., Oberthür, T, Miller, J.D. and Joslin, G.D. (2008). Textural, mineralogical and stable isotope studies of hydrothermal alteration in the main sulfide zone of the Great Dyke, Zimbabwe and the precious metals zone of the Sonju Lake Intrusion, Minnesota, USA. Mineralium Deposita, 43, 97110.Google Scholar
Li, C., Ripley, E.M., Sarkar, A., Shin, D. and Maier, W.D. (2005). Origin of phlogopite-orthopyroxene inclusions in chromites from the Merensky Reef of the Bushveld Complex, South Africa. Contributions to Mineralogy Petrology. 150, 119130.Google Scholar
Li, H. and Hermann, J. (2015). Apatite as an indicator of fluid salinity: An experimental study of chlorine and fluorine partitioning in subducted sediments. Geochimica et Cosmochimica Acta, 166, 267297.Google Scholar
Liebscher, A., Barnes, J.D. and Sharp, Z.D. (2006). Chlorine isotope vapor-liquid fractionation during experimental fluid-phase separation at 400 °C/23 MPa to 450 °C/42 MPa. Chemical Geology 234, 340345.Google Scholar
Lin, S. and Popp, R.K. (1984). Solubility and complexing of Ni in the system NiO-H2O-HCl. Geochimica et Cosmochimica Acta, 48, 27132722.Google Scholar
Lindsley, D.H. (1983). Pyroxene thermometry. American Mineralogist, 68, 477493.Google Scholar
Lipin, B.R. (1993) Pressure increases, the formation of chromite seams, and the development of the Ultramafic series in the Stillwater Complex, Montana. Journal of Petrology, 34, 955976.Google Scholar
Lipman, P.W., Banks, N.G. and Rhodes, J.M. (1985). Degassing-induced crystallization of basaltic magma and effects on lava rheology. Nature, 317, 604607Google Scholar
Liu, Y., Samaha, N.-T. and Baker, D.R. (2007). Sulfur concentration at sulfide saturation (SCSS) in magmatic silicate melts. Geochimica et Cosmochimica Acta, 71, 17831799.Google Scholar
Locmelis, M., Pearson, N.J., Barnes, S.J. and Fiorentini, M.L. (2011). Ruthenium in komatiitic chromite. Geochimica et Cosmochimica Acta, 75 (13), 36453661.Google Scholar
Locmelis, M., Fiorentini, M.L., Barnes, S.J. and Pearson, N.J. (2013). Ruthenium variation in chromite from komatiites and komatiitic basalts – A potential mineralogical indicator for nickel sulfide mineralization. Economic Geology, 108, 355364.Google Scholar
Locmelis, M., Fiorentini, M.L., Rushmer, T., Arevalo, R. Jr., Adamc, J. and Denyszyn, S.W. (2016). Sulfur and metal fertilization of the lower continental crust. Lithos, 244, 7493.Google Scholar
Loferski, P.J. and Arculus, R.J. (1993). Multiphase inclusions in plagioclase from anorthosites in the Stillwater Complex, Montana: Implications for the origin of the anorthosites. Contributions to Mineralogy and Petrology, 114, 6378.Google Scholar
Lofersky, P.J., Lipin, B.R. and Cooper, R.W. (1990). Petrology of chromite-bearing rocks from the lowermost cyclic units in the Stillwater Complex, Montana. U.S. Geological Survey Bulletin, 1674E, E1E25.Google Scholar
Longhi, J. (1982). Effects of fractional crystallization and cumulus processes on mineral composition trends of some lunar and terrestrial rock series. Proceedings of the 13th Lunar Planet Science Conference, Journal of Geophysical Research 87, A54A64.Google Scholar
Lorand, J.-P. and Alard, O. (2001). Platinum-group element abundances in the upper mantle: New constraints from in situ and whole-rock analyses of massif central xenoliths (France). Geochimica et Cosmochimica Acta, 65, 27892805.Google Scholar
Lorenz, V. (1975). Formation of phreatomagmatic maar-diatreme volcanoes and its relevance to kimberlite diatremes. Physical Chemistry of the Earth, 9, 1727.Google Scholar
Løseth, H., Wensaas, L., Arntsen, B., Hanken, N.-M., Basire, C. and Graue, K. (2011). 1000 m long gas blow-out pipes. Marine and Petroleum Geology, 28, 10471060.Google Scholar
Løvoll, G., Méheust, Y., Toussaint, R., Schmittbuhl, J. and Måløy, K.J. (2004). Growth activity during fingering in a porous Hele-Shaw cell. Physical Review E, 70, 02630–1. doi:10.1103/PhysRevE.70.026301Google Scholar
Lowenstein, J.B. (1995). Application of silicate-melt inclusions to the study of magmatic volatiles. In Magmas, Fluids, and Ore Deposits (Thompson, J.F.H., ed.). Mineralogical Association of Canada, Short Course Series 23, 7199.Google Scholar
Lubetkin, S.D. (2003). Why is it much easier to nucleate gas bubbles than theory predicts? Langmuir, 19, 25752587.Google Scholar
Luguet, A., Lorand, J.-P., Alard, O. and Cottin, J.-Y. (2004). A multi-technique study of platinum-group elements systematic in some ligurian ophiolitic peridotites, Italy. Chemical Geology, 208, 175194.Google Scholar
Luguet, A., Shirey, S.B., Lorand, J.-P., Horan, M.F. and Carlson, R.W. (2007). Residual platinum-group minerals from highly depleted harzburgites of the Lherz massif (France) and their role in HSE fractionation of the mantle. Geochimica et Cosmochimica Acta, 71, 30823097.Google Scholar
Maier, D.W., 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(3), 316327.Google Scholar
Maier, W.D. and Barnes, S.-J. (1998). Concentrations of rare earth elements in silicate rocks of the Lower, Critical and Main Zones of the Bushveld Complex. Chemical Geology, 150, 85103.Google Scholar
Maier, W.D. and Barnes, S.-J. (2004). Pt/Pd and Pt/Ir Ratios in mantle-derived magmas: A possible role for mantle metasomatism. South African Journal of Geology, 107, 333340.Google Scholar
Maier, W.D. and Barnes, S.-J. (2008). Platinum-group elements in the UG1 and UG2 chromitites, and the Bastard reef, at Impala platinum mine, western Bushveld Complex, South Africa: Evidence for late magmatic cumulate instability and reef constitution. South African Journal of Geology, 111, 159176.Google Scholar
Maier, W.D., Barnes, S.-J. and Groves, D.I. (2013). The Bushveld Complex, South Africa: Formation of platinum-palladium, chrome and vanadium-rich layers via hydrodynamic sorting of a mobilized cumulate slurry in a large, relatively slowly cooling, subsiding magma chamber. Mineralium Deposita, 48, 156.Google Scholar
Maier, W.D., Karykowski, B.T. and Yang, S.-H. (2016). Formation of transgressive anorthosite seams in the Bushveld Complex via tectonically induced mobilisation of plagioclase-rich crystal mushes. Geoscience Frontiers, 7, 875889.Google Scholar
Maier, W.D., Prichard, H.M., Barnes, S.J. and Fisher, P.C. (1999). Compositional variation of laurite at Union Section in the western Bushveld Complex. South African Journal of Geology, 102, 286292.Google Scholar
Makovicky, E. and Karup-Møller, S. (2000). Phase relations in the metal-rich portions of the phase system Pt-Ir-Fe-S at 1000°C and 1100°C. Mineralogical Magazine, 64(6), 10471056.Google Scholar
Mangan, M. and Sisson, T. (2004). Delayed, disequilibrium degassing in rhyolite magma: Decompression experiments and implications for explosive volcanism. Earth and Planetary Science Letters, 183, 441455.Google Scholar
Mangan, M. and Sisson, T. (2005). Evolution of melt-vapor surface tension in silicic volcanic systems: Experiments with hydrous melts. Journal of Geophysical Research, 110, B01202. doi:10.1029/2004JB003215.Google Scholar
Mangan, M., Mastin, L. and Sisson, T. (2004). Gas evolution in eruptive conduits: Combining insights from high temperature and pressure decompression experiments with steady-state flow modeling. Journal of Volcanology and Geothermal Research, 129, 2336.Google Scholar
Magenheima, A., Spivack, A.J., Michael, P.J. and Gieskes, J.M. (1995). Chlorine stable isotope composition of the oceanic crust: Implications for Earth’s distribution of chlorine. Earth and Planetary Science Letters, 131(3–4), 427432.Google Scholar
Manning, C. E. (1994). The solubility of quartz in the lower crust and upper mantle. Geochimica et Cosmochimica Acta, 58(22), 48314839.Google Scholar
Marsh, B.D. (2004). A magmatic mush column Rosetta stone: The McMurdo Dry Valleys of Antarctica. EOS, 85, 497502.Google Scholar
Marsh, B.D. (2013). On some fundamentals of igneous petrology. Contributions to Mineralogy and Petrology, 166(3), 665690.Google Scholar
Mather, T.A., Witt, M.L.I., Pyle, D.M., Quayle, B.M., Aiuppa, A., Bagnato, E., Martin, R.S., Sims, K.W.W., Edmonds, M., Sutton, A.J. and Ilyinskaya, E. (2012). Halogens and trace metal emissions from the ongoing 2008 summit eruption of Kīlauea volcano, Hawai`i. Geochemica et Cosmochimica Acta, 83, 292323.Google Scholar
Mathez, E.A. (1995). Magmatic metasomatism and formation of the Merensky reef, Bushveld Complex. Contributions to Mineralogy and Petrology, 119, 277286.Google Scholar
Mathez, E.A. and Marcantonio, F. (1995). Sr isotopes and magma mixing in the Bushveld Complex. EOS Transactions, American Geophysical Union, 76, F641.Google Scholar
Mathez, E.A. and Waight, T.E. (2003). Lead isotopic disequilibrium between sulfide and plagioclase in the Bushveld Complex and the chemical evolution of large layered intrusions. Geochimica et Cosmochimica Acta, 67, 18751888.Google Scholar
Mathez, E.A. and Webster, J. D. (2005). Partitioning behavior of chlorine and fluorine in the system apatite-silicate melt-fluid. Geochimica et Cosmochimica Acta, 69, 12751286.Google Scholar
Mathez, E.A. and Kent, A.J.R. (2007). Variable initial Pb isotopic compositions of rocks associated with the UG2 chromitite, eastern Bushveld Complex. Geochimica et Cosmochimica Acta, 71, 55145527.Google Scholar
Mathez, E.A. and Kinzler, R.J. (2017). Metasomatic chromitite seams in the Bushveld and Rum Layered Intrusions. Elements, 13(6), 397402.Google Scholar
Mathez, E.A., Dietrich, V.J., Holloway, J.R. and Boudreau, A.E. (1989). Carbon distribution in the Stillwater Complex and evolution of vapour during crystallization of Stillwater and Bushveld magmas. Journal of Petrology, 30(1), 153173.Google Scholar
Mathez, E.A., Agrinier, P. and Hutchinson, R. (1994). Hydrogen isotopic composition of the Merensky reef and related rocks, Atok section, Bushveld Complex. Economic Geology, 89, 791802.Google Scholar
Mazzini, A., Svensen, H.H., Forsberg, C.F., Linge, H., Lauritzen, S.-E., Haflidason, H., Hammer, Ø., Planke, S. and Tjelta, T.I. (2017). A climatic trigger for the giant Troll pockmark field in the northern North Sea. Earth and Planetary Science Letters, 464, 2434.Google Scholar
McBirney, A.R. (1975). Differentiation of the Skaergaard Intrusion. Nature, 253, 691694.Google Scholar
McBirney, A.R. (1987). Constitutional zone refining of layered intrusions. In Origins of Igneous Layering (Parsons, I., ed.). NATO ASI Series C, 196, 437452.Google Scholar
McBirney, A.R., (1996). The Skaergaard Intrusion. In Developments in Petrology 15 – Layered Intrusions (Cawthorn, R.G., ed.). Amsterdam: Elsevier Science, 147180.Google Scholar
McBirney, A.R. (2009). Factors governing the textural development of Skaergaard gabbros: A review. Lithos, 111, 15.Google Scholar
McBirney, A.R. and Creaser, R.A. (2003). The Skaergaard Layered Series, Part VII: Sr and Nd isotopes. Journal of Petrology, 44(4), 757771.Google Scholar
McBirney, A.R. and Hunter, R.H. (1995). The cumulate paradigm reconsidered. The Journal of Geology, 103, 114122.Google Scholar
McBirney, A.R. and Noyes, R.M. (1979). Crystallization and layering of the Skaergaard Intrusion. Journal of Petrology, 20, 487564.Google Scholar
McBirney, A.R. and Sonnenthal, E.L. (1990). Metasomatic replacement in the Skaergaard Intrusion, East Greenland: Preliminary observations. Chemical Geology, 88, 245260.Google Scholar
McCallum, I.S. (1996). The Stillwater Complex. In Developments in Petrology 15 – Layered Intrusions (Cawthorn, R.G., ed.). Amserdam: Elsevier Science, 441484.Google Scholar
McCallum, I.S., Raedeke, L.D. and Mathez, E.A. (1977). Stratigraphy and petrology of the Banded zone of the Stillwater Complex, Montana. EOS, 58, 1245.Google Scholar
McCallum, I.S., Raedeke, L.D. and Mathez, E.A. (1980). Investigations in the Stillwater Complex: Part I. Stratigraphy and structure of the Banded zone. In The Jackson Volume, vol. 280-A of American Journal of Science (Irving, A. and Dungan, M., eds.). New Haven, Connecticut, Kline Geology Laboratory, Yale University, 5987.Google Scholar
McCallum, I.S., Thurber, D.W., O’Brien, H.E. and Nelson, B.K. (1999). Lead isotopes in sulfides from the Stillwater Complex, Montana: Evidence for subsolidus remobilization. Contributions to Mineralogy and Petrology, 137, 206219.Google Scholar
McCallum, M.E. (1985). Experimental evidence for fluidization processes in breccia pipe formation. Economic Geology, 80, 15231543.Google Scholar
McCandless, T.E. and Ruiz, J. (1991). Osmium isotopes and crustal sources for platinum-group element mineralization in the Bushveld Complex, South Africa. Geology, 19, 12251228.Google Scholar
McCandless, T.E., Ruiz, J., Adair, B.I. and Freydier, C. (1999). Re–Os isotope and Pd/Ru variations in chromitites from the Critical zone, Bushveld Complex, South Africa. Geochimica et Cosmochimica Acta, 63, 911923.Google Scholar
McCubbin, F.M., Vander Kaaden, K.E., Tartese, R., Boyce, J.W., Mikhail, S., Whitson, E.S., Bell, A.S., Anand, M., Franchi, I.A., Wang, J. and Hauri, E.H. (2015). Experimental investigation of F, Cl and OH partitioning between apatite and Fe-rich basaltic melt at 1.0–1.2 GPa and 950–1000 °C. American Mineralogist, 100(8–9), 17901802.Google Scholar
McDonald, I., Holwell, D.A. and Armitage, P.E.B. (2005) Geochemistry and Mineralogy of the Platreef and “Critical Zone” of the Northern Lobe of the Bushveld Complex, South Africa: Implications for Bushveld Stratigraphy and the Development of PGE Mineralization. Mineralium Deposita, 40, 526549.Google Scholar
McDonough, W.F., Sun, S.-S. (1995). The composition of the Earth. Chemical Geology, 120, 223253.Google Scholar
Mcllveen, C.L. (1996). Anomalous platinum-group element occurrence below the JM Reef, Stillwater Complex, Montana. M.S. Thesis, University of Montana.Google Scholar
McInnes, B.I.A., McBride, J.S., Evans, N.J., Lambert, D.D. and Andrew, A.S. (1999). Osmium isotope constraints on ore metal recycling in subduction zones. Science, 286, 512516.Google Scholar
McKenzie, D. (1984). The generation and compaction of partially molten rock. Journal of Petrology, 25, 713765.Google Scholar
McKenzie, D. (2011). Compaction and crystallization in magma chambers: Towards a model of the Skaergaard Intrusion. Journal of Petrology, 52, 905930.Google Scholar
Médard, E. and Grove, T.L. (2007). Water in basaltic melts: Effect on the liquidus temperature, olivine-melt thermometry and mantle melting. In Workshop on Water in Planetary Basalts, 2627. LPI Contribution 373, Lunar and Planetary Institute, Houston.Google Scholar
Melcher, F., Grum, W., Simon, G., Thalhammer, T.V. and Stumpfl, E.F. (1997). Petrogenesis of the Ophiolitic Giant Chromite Deposits of Kempirsai, Kazakhstan: A Study of Solid and Fluid Inclusions in Chromite. Journal of Petrology, 38, 14191458.Google Scholar
Mei, Y., Etschmann, B., Liu, W., Sherman, D.M., Barnes, S.J., Fiorentini, M.L., Seward, T.M., Tesemale, D. and Brugger, J. (2015). Palladium complexation in chloride- and bisulfide-rich fluids: Insights from ab initio molecular dynamics simulations and X-ray absorption spectroscopy. Geochimica et Cosmochimica Acta, 161, 128145.Google Scholar
Merkle, R.K.W. (1992). Platinum-group minerals in the middle group of chromitite layers at Marikana, western Bushveld Complex: Indications for collection mechanisms and post-magmatic modification. Canadian Journal of Earth Sciences, 29, 209221.Google Scholar
Métrich, N. and Wallace, P. (2008). Volatile abundances in basaltic magmas and their degassing paths tracked by melt inclusions. In Minerals, Inclusions and Volcanic Processes (Putirka, K. and Tepley, F., eds,). Mineralogical Society of America, Reviews in Mineralogy and Geochemistry, 69, 363402Google Scholar
Métrich, N. and Rutherford, M. J. (1992). Experimental study of chlorine behavior in hydrous silicic melts. Geochimica et Cosmochimica Acta, 57, 607616.Google Scholar
Métrich, N., Bertagnini, A., Landi, P. and Rosi, M. (2001). Crystallization driven by decompression and water loss at Stromboli volcano (Aeolian Islands, Italy). Journal of Petrology, 42(8), 14711490.Google Scholar
Meurer, W.P. and Boudreau, A.E. (1996). Compaction of density-stratified cumulates: Effect on trapped-liquid distribution. Journal of Geology, 104, 115120.Google 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, 281292.Google Scholar
Meurer, W.P. and Boudreau, A.E. (1998b). Compaction of igneous cumulates. Part II – Compaction and the development of igneous foliation. Journal of Geology, 106, 293304.Google Scholar
Meurer, W.P., Klaber, S.A. and Boudreau, A.E. (1997). Discordant bodies from Olivine-Bearing zones III and IV of the Stillwater complex, Montana – Evidence for post-cumulus fluid migration in layered intrusions. Contributions to Mineralogy and Petrology, 130, 8192.Google Scholar
Meurer, W.P. and Meurer, M.E.S. (2006). Using apatite to dispel the ‘‘trapped liquid’’ concept and to understand the loss of interstitial liquid by compaction in mafic cumulates: An example from the Stillwater Complex, Montana. Contributions to Mineralogy Petrology, 151, 187201.Google Scholar
Meurer, W.P., Willmore, C.C. and Boudreau, A. E. (1998). Metal redistribution during fluid exsolution and migration in the Middle Banded series of the Stillwater complex, Montana. Lithos, 47, 143156.Google Scholar
Modal, S.K. and Mathez, E.A. (2007). Origin of the UG2 chromitite layer, Bushveld Complex. Journal of Petrology, 48(3), 495510.Google Scholar
Morse, S.A. (1980). Basalts and Phase Diagrams. New York: Springer, 493.Google Scholar
Morse, S.A. (1983). Strontium isotope fractionation in the Kiglapait Intrusion. Science, 220, 193195.Google Scholar
Mourtada-Bonnefoi, C.C. and Laporte, D. (1999). Kinetics of bubble nucleation in a rhyolitic melt: An experimental study of the effect of ascent rate. Earth and Planetary Science Letters, 218, 521537.Google Scholar
Mukherjee, R., Latypov, R. and Balakrishna, A. (2017). An intrusive origin of some UG-1 chromitite layers in the Bushveld Igneous Complex, South Africa: Insights from field relationships. Ore Geology Reviews, 90, 94109.Google Scholar
Mungall, J.E. (2002). Kinetic controls on the partitioning of trace elements between silicate and sulfide liquids. Journal of Petrology, 43(5), 749768.Google Scholar
Mungall, J.E. (2015). Physical controls of nucleation, growth and migration of vapor bubbles in partially molten cumulates. In Layered Intrusions (Charlier, B., Namur, O., Latypov, R. and Tegner, C., eds.). Dordrecht: Springer Geology, 331378.Google Scholar
Munoz, J.L. and Swenson, A. (1981). Chloride-hydroxyl exchange in biotite and estimate of relative HCl/HF activities in hydrothermal fluids. Economic Geology, 76, 22122221.Google Scholar
Murase, T. and McBirney, A.R. (1973). Properties of some common igneous rocks and their melts at high temperatures. Geological Society of America Bulletin, 84, 35633592.Google Scholar
Mutanen, T., Törnroos, R. and Johanson, B. (1988). The Significance of Cumulus Chlorapatite and High-temperature Dashkesanite to the Genesis of PGE Mineralization in the Koitelainen and Keivitsa-Satovaara Complexes, Northern Finland. In Geo-Platinum 87 (Prichard, H.M., Potts, P.J., Bowles, J.F.W. and Cribb, S.J., eds.). Dordrecht: Springer,Google Scholar
Mysen, B.O. and Boettcher, A.L. (1975). Melting of a hydrous mantle: I. Phase relations of natural peridotite at high pressures and temperatures with controlled activities of water, hydrogen, and carbon dioxide. Journal of Petrology, 16, 520548.Google Scholar
Nakagawa, M. and Franco, H.E.A. (1997). An assessment of placer Os–Ir–Ru alloys and sulfides as an indicator of sulfur fugacity in a primitive stage of ophiolite mantle. Journal of Asian Earth Sciences, 15, 311315.Google Scholar
Naldrett, A.J. and Wilson, A.H. (1990). Horizontal and vertical variations in noble metals in the Great Dyke of Zimbabwe: A model for the origin of PGE mineralization by fractional segregation. Chemical Geology, 88, 279300.Google Scholar
Naldrett, A.J., Kinnaird, J., Wilson, A. and Chunnett, G. (2008). The concentration of PGE in the Earth’s crust with special reference to the Bushveld Complex. Earth Science Frontiers, 15(5), 264297.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, 625659.Google Scholar
Naldrett, A.J., Wilson, A., Kinnaird, J., Yudovskaya, M. and Chunnett, G. (2012). The origin of chromitites and related PGE mineralization in the Bushveld Complex: New mineralogical and petrological constraints. Mineralium Deposita, 47, 209232.Google Scholar
Nash, W.P. (1976). Fluorine chlorine and OH-bearing minerals in the Skaergaard Intrusion. American Journal of Science, 276, 546–57.Google Scholar
Naslund, H.R. (1986). Disequilibrium partial melting and rheomorphic layer formation in the contact aureole of the Basistoppen sill, East Greenland. Contributions to Mineralogy and Petrology, 93, 359367.Google Scholar
Navon, O. and Lyakhovsky, V. (1998). Vesiculation processes in silicic magmas. Geological Society Special Publication, 145, 2750.Google Scholar
Newton, R.C., Aranovich, L.Ya., Hansen, E.C. and Vandenheuvel, B.A. (1998). Hypersaline fluids in Precambrian deep-crustal metamorphism. Precambrain Research, 91(1–2), 4163.Google Scholar
Newton, R.C. and Manning, C.E. (2000). Quartz solubility in H2O-NaCl and H2O-CO2 solutions at deep crust-upper mantle pressures and temperatures: 2–15 kbar and 500–900°C. Geochimica et Cosmochimica Acta, 64(17), 29933005.Google Scholar
Newton, R.C. and Manning, C.E. (2002). Experimental determination of calcite solubility in H2O-NaCl solutions at deep crust/upper mantle pressures and temperatures: Implications for metasomatic processes in shear zones. American Mineralogist, 87, 14011409Google Scholar
Newton, R.S., Cunningham, R.C. and Schubert, C.E. (1980). Mud volcanoes and pockmarks: Seafloor engineering hazards or geological curiosities? Proceedings 12th Offshore Technical Conference, Houston Texas, 3729, 425–429.Google Scholar
Nicholson, D.M. and Mathez, E.A. (1991). Petrogenesis of the Merensky Reef in the Rustenburg section of the Bushveld Complex. Contributions to Mineralogy and Petrology, 107, 293309.Google Scholar
Nielsen, T.F.D. (2004). The shape and volume of the Skaergaard Intrusion, Greenland: Implications for mass balance and bulk composition. Journal of Petrology, 45, 507530.Google Scholar
Nielsen, T.F.D., Andersen, J.C.Ø., Holness, M.B., Keidin, J.K., Rudashevsky, N.S., Rudashevsky, V.N., Salmonsen, L.P., Tegner, C. and Veksler, I.V. (2015). The Skaergaard PGE and Gold Deposit: The Result of in situ Fractionation, Sulphide Saturation, and Magma Chamber-scale Precious Metal Redistribution by Immiscible Fe-rich Melt, Journal of Petrology, 56, 16431676.Google Scholar
O’Driscoll, B., Emeleus, C.H., Donaldson, C.H. and Daly, J.S. (2010). Cr-spinel seam petrogenesis in the Rum Layered Suite, NW Scotland: Cumulate assimilation and in situ crystallization in a deforming crystal mush. Journal of Petrology, 51(6), 11711201.Google Scholar
O’Driscoll, B. and González-Jiménez, J.M. (2016). Petrogenesis of the Platinum-Group Minerals. Reviews in Mineralogy and Geochemistry, 81, 489578.Google Scholar
O’Neill, H.St.C. and Mavrogenes, J.A. (2002). The sulfide capacity and the sulfur content at sulfide saturation of silicate melts at 1400 °C and 1 bar. Journal of Petrology, 43, 10491087.Google Scholar
Oberthür, T. (2011). Platinum-group element mineralization of the Main Sulfide Zone, Great Dyke, Zimbabwe. Reviews in Economic Geology, 17, 329349.Google Scholar
Oberthür, T., Cabri, L.J., Weiser, T.W., McMahon, G. and Müller, P. (1997). Pt, Pd and other trace elements in sulfides of the Main Sulfide Zone, Great Dyke, Zimbabwe – A reconnaissance study. Canadian Mineralogist, 35, 597609.Google Scholar
Oberthür, T., Weiser, T.W., Müller, P., Lodziak, J. and Cabri, L.J. (1998). New observations on the distribution of platinum group elements (PGE) and minerals (PGM) in the MSZ at Hartley Mine, Great Dyke, Zimbabwe. In 8th Intl Pt Symposium Abstracts, South African Institute of Mining and Metallurgy, Johannesburg, 293296.Google Scholar
Oberthür, T., Davis, D.W., Blenkinsop, T.G. and Höhndorf, A. (2002). Precise U–Pb mineral ages, Rb–Sr and Sm–Nd systematics for the Great Dyke, Zimbabwe – Constraints on late Archean events in the Zimbabwe Craton and Limpopo Belt. Precambrian Research, 113, 293305.Google Scholar
Oberthür, T., Weiser, T.W., Gast, L. and Kojonen, K. (2003). Geochemistry and mineralogy of platinum-group elements at Hartley Platinum Mine, Zimbabwe. Mineralium Deposita, 38, 327343.Google Scholar
O’Neill, H.St.C., Dingwell, D.B., Borisov, A., Spettle, B. and Palme, H. (1995). Experimental petrochemistry of some high siderophile elements at high temperatures and some implications for core formation and the mantle’s early history. Chemical Geology, 120, 255273.Google Scholar
Oppenheimer, J., Rust, A.C., Cashman, K.V. and Sandnes, B. (2015). Gas migration regimes and outgassing in particle-rich suspensions. Frontiers in Physics, 3: 115.Google Scholar
Orville, P.M. (1972). Plagioclase cation exchange equilibria with aqueous chloride solution: Results at 700 °C and 2000 bars in the presence of quartz. American Journal of Science, 272, 234272.Google Scholar
Page, N.J, Shimek, R. and Huffman, C. (1972). Grain-size variations within an olivine cumulate, Stillwater Complex, Montana. United State Geological Survey Professional Paper, 800-C, C29-C37.Google Scholar
Page, N.J. and Zientek, M.L. (1987). Composition of primary postcumulus amphibole and phlogopite within an olivine cumulate in the Stillwater Complex, Montana. United States Geological Survey Bulletin, 1674-A.Google Scholar
Pagé, P. and Barnes, S.-J. (2013). Improved in-situ determination of PGE concentration of chromite by LA-ICP-MS: Towards a better understanding. Ore deposits associated with mafic and ultramafic rocks, 12th SGA Biennial Meeting 2013. Proceedings, 3, 10501053.Google Scholar
Pagé, P. and Barnes, S.-J., (2016). The influence of chromite on osmium, iridium, ruthenium and rhodium distribution during early magmatic processes. Chemical Geology, 420, 5168.Google Scholar
Pagé, P., Barnes, S.J. and Zientek, M.L. (2011). Formation and evolution of the chromitites of the Stillwater Complex: A trace element study. In Let‘s Talk Ore Deposits: Proceedings of the 11th SGA Biennial Meeting, Antofagasta, Chile (Barra, F., ed.). Society for Geology Applied to Mineral Deposits, 678680.Google Scholar
Pagé, P., Barnes, S.-J., Bedard, J.H. and Zientek, M.L. (2012). In situ determination of Os, Ir and Ru in chromites formed from komatiite, tholeiite and boninite magmas: Implications for chromite control of Os, Ir and Ru during partial melting and crystal fractionation. Chemical Geology, 302–303, 315.Google Scholar
Pagé, P., Barnes, S.-J., Méric, J. and Houlé, M.G. (2015). Geochemical composition of chromite from Alexo komatiite in the western Abitibi greenstone belt: Implications for mineral exploration, In Targeted Geoscience Initiative 4: Canadian Nickel-Copper-Platinum Group Elements-Chromium Ore Systems — Fertility, Pathfinders, New and Revised Models, (Ames, D.E. and Houlé, M.G., eds.). Geological Survey of Canada, Open File 7856, 187195.Google Scholar
Palacz, Z.A. and Tait, S.R. (1985). Isotopic and geochemical investigation of unit 10 from the Eastern Layered Series of the Rhum Intrusion, northwest Scotland. Geological Magazine, 122, 485490.Google Scholar
Palme, H. and O’Neill, H.St.C. (2003). Cosmochemical Estimates of Mantle Composition. In Treatise on Geochemistry, Volume 2 (Carlson, R.W., ed). Oxford: Elsevier-Pergamon, 138.Google Scholar
Pan, P. and Wood, S.A. (1994). Solubility of Pt and Pd sulfides and Au metal in aqueous bisulfide solutions. Mineralium Deposita, 29(5), 373390.Google Scholar
Park, J.-W., Campbell, I.H. and Eggins, S.M. (2012). Enrichment of Rh, Ru, Ir and Os in Cr spinels from oxidized magmas: Evidence from the Ambae volcano, Vanuatu. Geochimica et Cosmochimica Acta, 78, 2850.Google Scholar
Parmigiani, A., Huber, C., Bachmann, O. and Chopard, B. (2011). Pore-scale mass and reactant transport in multiphase porous media flow. Journal of Fluid Mechanics, 686, 4076.Google Scholar
Patten, C., Barnes, S.-J. and Mathez, E.A. (2012). Textural variations in MORB sulfide droplets due to differences in crystallization history. Canadian Mineralogist, 50, 675692.Google Scholar
Peach, C.L, Mathez, E.A. and Keays, R.R. (1990). Sulfide melt-silicate melt distribution coefficients for noble metals and other chacophile elements as deduced from MORB: Implications for partial melting. Geochimica et Cosmochimica Acta, 54, 33793389.Google Scholar
Peach, C.L., Mathez, E.A., Keays, R.R. and Reeves, S.J. (1994). Experimentally determined sulfide melt‐silicate melt partition coefficients for iridium and palladium. Chemical Geology, 117, 361377.Google Scholar
Pebane, M. and Latypov, R. (2017). The significance of magmatic erosion for bifurcation of UG1 chromitite layers in the Bushveld Complex. Ore Geology Reviews, 90, 6593.Google Scholar
Peck, D.C. and Keays, R.R. (1990). Geology, geochemistry and origin of platinum-group element-chromitite occurrences in the Heazlewood River Complex, Tasmania. Economic Geology, 85, 765793.Google Scholar
Peck, D.C, Keays, R.R. and Ford, R.J. (1992). Direct crystallization of refractory platinum-group elements alloys from boninitic magmas: Evidence form western Tasmania. Australian Journal of Earth Sciences, 39, 373387.Google Scholar
Pedersen, A.K., Watt, M., Watt, W.S. and Larsen, L.M. (1997). Structure and stratigraphy of early Tertiary basalts of the Blosseville Kyst, East Greenland. Journal of the Geological Society, London, 154(3), 565570.Google Scholar
Pelch, M.A., Appold, M.S., Emsbo, P. and Bodnar, R.J. (2015). Constraints from fluid inclusion compositions on the origin of Mississippi valley-type mineralization in the Illinois-Kentucky District. Economic Geology, 110, 787808.Google Scholar
Peslier, A.H. (2010). A review of water contents of nominally anhydrous natural minerals in the mantles of Earth, Mars and the Moon. Journal of Volcanology and Geothermal Research, 197, 239258.Google Scholar
Petford, N. and Mirhadizadeh, S. (2017). Image-based modelling of lateral magma flow: The Basement Sill, Antarctica. Royal Society Open Science, 4, 161083). http://dx.doi.org/10.1098/rsos.161083.Google Scholar
Peyerl, W. (1982). The influence of the Driekop dunite pipe on the platinum-group mineralogy of the UG-2 chromtite in its vicinity. Economic Geology, 77, 14321438.Google Scholar
Philpotts, A.R., Brustman, C.M., Shi, J., Carlson, W.D. and Denison, C. (1999). Plagioclase-chain networks in slowly cooled basaltic magma. American Mineralogist, 84, 18191829.Google Scholar
Philpotts, A.R., Shi, J. and Brustman, C. (1998). Role of plagioclase crystal chains in the differentiation of partly crystallized basaltic magma. Nature, 395, 343346.Google Scholar
Philpotts, A.R., Carroll, M. and Hill, J.M. (1996). Crystal-mush compaction and the origin of pegmatitic segregation sheets in a thick flood-basalt flow in the Mesozoic Hartford Basin, Connecticut. Journal of Petrology, 37, 811836.Google Scholar
Piccoli, P.M. and Candela, P.A. (1994). Apatite in felsic rocks; a model for the estimation of initial halogen concentrations in the Bishop Tuff (Long Valley) and Tuolumne Intrusive Suite (Sierra Nevada Batholith) magmas. American Journal of Science, 294, 92135.Google Scholar
Piccoli, P.M. and Candela, P.A. (2002). Apatite in igneous systems. Reviews in Mineralogy and Geochemistry, 48, 255292.Google Scholar
Podmore, F. and Wilson, A.H. (1987). A reappraisal of the structure, geology and emplacement of the Great Dyke, Zimbabwe. In Mafic Dyke Swarms (Halls, H.C. and Fahrig, W.F., eds.). Geological Association of Canada Special Paper, 34, 317330.Google Scholar
Polacci, M., Corsaro, R.A. and Andronico, D. (2006). Coupled textural and compositional characterization of basaltic scoria: Insights into the transition from Strombolian to fire fountain activity at Mount Etna, Italy. Geology, 34(3), 201204.Google Scholar
Prendergast, M.D. (1988). The geology and economic potential of the PGE-rich Main Sulphide Zone of the Great Dyke, Zimbabwe. In Geo-Platinum’87 (Prichard, H.M., Potts, P.J., Bowles, J.F.W. and Cribb, S.J., eds). Barking, Essex, Elsevier Applied Science, 281302.Google Scholar
Prendergast, M.D. and Keays, R.R. (1989). Controls of platinum-group element mineralisation and the origin of the PGE-rich Main Sulphide Zone in the Wedza Subchamber of the Great Dyke, Zimbabwe: Implications for the genesis of, and exploration for, stratiform PGE mineralisation in layered intrusions. In Magmatic Sulphides – The Zimbabwe Volume (Prendergast, M.D. and Jones, M.J., eds). London, The Institution of Mining and Metallurgy, 4369.Google Scholar
Prendergast, M.D. and Wilson, A.H. (1989). The Great Dyke of Zimbabwe - II: Mineralisation and mineral deposits. In Magmatic Sulphides – The Zimbabwe Volume (Prendergast, M.D. and Jones, M.J., eds). London: The Institution of Mining and Metallurgy, 2142.Google Scholar
Prevec, S.A., Ashwal, L.D. and Makaza, M.S. (2005). Mineral disequilibrium in the Merensky Reef, western Bushveld Complex, South Africa: New Sm–Nd isotopic evidence. Contributions to Mineralogy and Petrology, 149(3), 306315.Google Scholar
Prichard, H.M., Potts, P.J. and Neary, C.R. (1981). Platinum group element minerals in the Unst chromite, Shetland Isles. Transactions of the Institute of Mining and Metallurgy, Section B, 90, 186188.Google Scholar
Prichard, H.M., Sa, J.H.S. and Fisher, P.C. (2001). Platinum-group mineral assemblages and chromite composition in the altered and deformed Bacuri complex, Amapa, northeastern Brazil. Canadian Mineralogist, 39, 377396.Google Scholar
Proenza, J.A., Zaccarini, F., Lewis, J.F., Longo, F. and Garuti, G. (2007). Chromian spinel composition and the platinum-group minerals of the PGE-rich Loma Peguera chromitites, Loma Caribe peridotite, Dominican Republic. The Canadian Mineralogist, 45, 631648.Google Scholar
Proussevitch, A.A. and Sahagian, D.L. (1998). Dynamics and energetics of bubble growth in magmas: Analytical formulation and numerical modeling. Journal of Geophysical Research, 103, 1822318251.Google Scholar
Puchtel, I.S. and Humayun, M. (2001). Platinum group element fractionation in a komatiitic basalt lava lake. Geochimica et Cosmochimica Acta, 65, 29792993.Google Scholar
Raedeke, L.D. (1982). Petrogenesis of the Stillwater Complex. Ph.D. Dissertation, University of Washington.Google Scholar
Raedeke, L.D. and McCallum, I.S. (1980). A comparison of fractionation trends in the lunar crust and the Stillwater Complex. In Proceedings of the conference on the lunar highlands crust (Merrill, R.B. and Papike, J.J. eds.). Geochimica et Cosmochimica Acta, supplement 12, 133153.Google Scholar
Raedeke, L.D. and McCallum, I.S. (1984). Investigations in the Stillwater Complex: Part II. Petrology and petrogenesis of the ultramafic series. Journal of Petrology, 25, 395420.Google Scholar
Raedeke, L.D. and Vian, R.W. (1986). A three dimensional view of mineralization in the Stillwater J-M Reef. Economic Geology, 81, 11871195.Google Scholar
Reid, D.L. and Basson, I.J. (2002). Iron-rich ultramafic pegmatite replacement bodies within the Upper Critical Zone, Rustenburg Layered Suite, Northam Platinum Mine, South Africa. Mineralogical Magazine, 66(6), 895914.Google Scholar
Reisberg, L., Tredoux, M., Harris, C., Coftier, A. and Chaumba, J. (2011). Re and Os distribution and Os isotope composition of the Platreef at the Sandsloot–Mogolakwena mine, Bushveld complex, South Africa. Chemical Geology, 24(3–4), 352363.Google Scholar
Ribe, N.M. (1985). The generation and composition of partial melts in the earth’s mantle. Earth and Planetary Science Letters, 73, 361376.Google Scholar
Righter, K., Campbell, A.J., Humayun, M. and Hervig, R.L. (2004). Partitioning of Ru, Rh, Pd, Ir and Au between Cr-bearing spinel, olivine, pyroxene and silicate melts. Geochimica et Cosmochimica Acta, 68, 867880.Google Scholar
Righter, K., Humayun, M. and Danielson, L. (2008). Partitioning of palladium at high pressures and temperatures during core formation. Nature Geoscience, 1, doi:10.1038/ngeo180.Google Scholar
Ripley, E.M. (2005). Re/Os isotopic and fluid inclusion studies of fluid-rock interaction in the contact aureole of the Duluth Complex, Minnesota. Geochimica et Cosmochimica Acta, 69(10) Supplement, 332.Google Scholar
Rivalta, E. and Dahm, T. (2006). Acceleration of buoyancy-driven fractures and magmatic dikes beneath the free surface. Geophysical Journal International, 166, 14241439.Google Scholar
Roelofse, F. and Ashwal, L.D. (2012). The Lower Main Zone in the northern limb of the Bushveld complex – A >1.3 km thick sequence of intruded and variably contaminated crystal mushes. Journal of Petrology, 53(7), 14491476.Google Scholar
Roelofse, F., Ashwal, L.D. and Romer, R.L (2015). Multiple, isotopically heterogeneous plagioclase populations in the Bushveld Complex suggest mush intrusion. Chemie der Erde, 75, 357364.Google Scholar
Rudashevsky, N.S., Avdontsev, S.N. and Dneprovskaya, M.B. (1992). Evolution of PGE mineralization in hortonolitic Ddunites of the Mooihoek and Onverwacht Pipes, Bushveld Complex. Mineralogy and Petrology, 47, 3754.Google Scholar
Rüpke, L.H., Morgan, J.P., Hort, M. and Connolly, J.A.D. (2004). Serpentine and the subduction zone water cycle. Earth and Planetary Science Letters, 223, 1734.Google Scholar
Rust, A.C. and Cashman, K.V. (2004). Permeability of vesicular silicic magma: Inertial and hysteresis effects. Earth and Planetary Science Letters, 228, 93107.Google Scholar
Ryder, G. and Spettel, B. (1985). The parental magma for some rocks from the Norite 1 subzone of the Stillwater Complex: A lunar analog study. Proceedings of the 15th Lunar and Planetary Science Conference, Part 2, Journal of Geophysical research 90 (supplement), C545C559.Google Scholar
Ryerson, F.J. and Watson, E.B. (1987). Rutile saturation in magmas, implications for Ti-Nb-Ta depletion in island-arc basalts. Earth Planetary Science Letters, 86, 225239.Google Scholar
Sahagian, D.L., Proussevitch, A.A. and Carlson, W.D. (2002). Analysis of vesicular basalts and lava emplacement processes for application as a paleobarometer/Ppaleoaltimeter. Journal of Geology, 110: 671685.Google Scholar
Salpas, P.A., Haskin, L.A. and McCallum, I.S. (1983). Stillwater Anorthosites: A lunar analog? Proceedings of the 14th Lunar and Planetary Science Conference, Part 1. Journal of Geophysical Research, 88(suppl), B27-B39.Google Scholar
Sassani, D.C. and Shock, E.L. (1998). Solubility and transport of platinum-group elements in supercritical fluids: Summary and estimates of thermodynamic properties for ruthenium, rhodium, palladium, and platinum solids, aqueous ions, and complexes to 1000 °C and 5 kbar. Geochimica et Cosmochimica Acta, 62, 26432671.Google Scholar
Sato, H., Holtz, F., Behrens, H., Botcharnikov, R. and Nakada, S. (2005). Experimental petrology of the 1991–1995 Unzen dacite, Part II: Cl/OH partitioning between hornblende and melt and its implications for the origin of oscillatory zoning of hornblende Phenocrysts. Journal of Petrology, 46, 339354.Google Scholar
Schannor, M., Veksler, I.V., Hecht, L., Harris, C. and Romer, R.L. (2018). Small-scale Sr and O isotope variations through the UG2 in the eastern Bushveld Complex: The role of crustal fluids. Chemical Geology, 485, 9099.Google Scholar
Schiffries, C.M. (1982). The petrogenesis of a platiniferous dunite pipe in the Bushveld complex: Chloride complexing and cation exchange metasomatism. Economic Geology, 77, 14391453.Google Scholar
Schiffries, C.M. (1990). Liquid absent aqueous fluid inclusions and phase equilibria in the system CaCl2-NaCl-H2O. Geochimica et Cosmochimica Acta, 54, 611619.Google Scholar
Schiffries, C.M. and Rye, D.M. (1989). Stable isotopic systematics of the Bushveld Complex: I. Constraints of magmatic processes in layered intrusions. American Journal of Science, 289, 841873.Google Scholar
Schiffries, C.M. and Rye, D.M. (1990). Stable isotopic systematics of the Bushveld Complex: II. Constraints on hydrothermal processes in layered intrusions. American Journal of Sciences, 290, 209245.Google Scholar
Schiffries, C.M. and Skinner, B.J. (1987). The Bushveld hydrothermal system: Field and petrologic evidence. American Journal of Sciences, 287, 566595.Google Scholar
Schisa, P., Boudreau, A.E., Djon, L., Ychalikian, A. and Corkery, J. (2015). The Lac des Iles palladium deposit, Ontario, Canada. Part II. Halogen variations in apatite. Mineralium Deposita, 50, 339355.Google Scholar
Schmidberger, S.S., Simonetti, A. and Francis, D. (2003). Small-scale Sr isotope investigation of clinopyroxenes from peridotite xenoliths by laser ablation MC-ICP-MS – Implications for mantle metasomatism. Chemical Geology, 199, 317329.Google Scholar
Schoenberg, R., Kruger, F.J., Nägler, T.F., Meisel, T. and Kramers, J.D. (1999). PGE enrichment in chromitite layers and the Merensky Reef of the western Bushveld Complex; a Re–Os and Rb–Sr isotope study. Earth and Planetary Science Letters, 172, 4964.Google Scholar
Scholten, L., Watenphulb, A., Beermanna, O., Testemale, D., Ames, D. and Schmidt, C. (2018). Nickel and platinum in high-temperature H2O + HCl fluids: Implications for hydrothermal mobilization. Geochimica et Cosmochimica Acta, 224, 187199.Google Scholar
Scoates, J.S. (2000). The plagioclase-magma density paradox re-examined and the crystallization of Proterozoic anorthosites. Journal of Petrology, 41, 627649.Google Scholar
Scoates, J.S. and Friedman, R.M. (2008). Precise age of the platiniferous Merensky reef, Bushveld Complex, South Africa, by the U–Pb zircon chemical abrasion ID–TIMS technique. Economic Geology, 103, 465471.Google Scholar
Scoon, R.N. (1987). Metasomatism of cumulus magnesian olivine by iron-rich postcumulus liquids in the upper Critical Zone of the Bushveld Complex. Mineralogical Magazine, 51, 389396.Google Scholar
Scoon, R.N. and Mitchell, A.A. (1994). Discordant Iron-rich ultramafic pegmatites and their relationship to iron-rich intercumulus and residual liquids. Journal of Petrology, 35, 881917.Google Scholar
Scoon, R.N. and Mitchell, A.A. (2004a). The platiniferous dunite pipes in the eastern limb of the Bushveld Complex: Review and comparison with unmineralized discordant ultramafic bodies. South African Journal of Geology, 107, 505520.Google Scholar
Scoon, R.N. and Mitchell, A.A., (2004b). Petrogenesis of discordant magnesian dunites from the central sector of the Eastern Bushveld Complex with emphasis on the Winnaarshoek pipe and disruption of the Merensky Reef. Economic Geology, 99, 517541.Google Scholar
Scoon, R.N. and Mitchell, A.A. (2009). A multi-stage orthomagmatic and partial melting hypothesis for the Driekop platiniferous dunite pipe, eastern limb of the Bushveld Complex, South Africa. South African Journal of Geology, 112, 187196.Google Scholar
Scoon, R.N. and Mitchell, A.A. (2010). The principle geologic features of the Onverwacht platiniferous dunite pipe, eastern limb of the Bushveld complex. South African Journal of Geology, 113, 155168.Google Scholar
Scoon, R.N. and Mitchell, A.A. (2011). The principle geologic features of the Mooihoek platiniferous dunite pipe, eastern limb of the Bushveld complex, and similaritites with replaced Merensky Reef at the Amandelbult Mine, South Africa. South African Journal of Geology, 114, 1540.Google Scholar
Scoon, R.N. and Teigler, B. (1994). Platinum-group element mineralization in the critical zone of the Western Bushveld Complex: 1. Sulfide poor chromitites below the UG2. Economic Geology, 89, 10941121.Google Scholar
Scoon, R.N. and Teigler, B. (1994). A new LG-6 chromite reserve at Eerste Geluk in the boundary zone between the central and southern sectors of the eastern Bushveld Complex. Economic Geology, 90, 969982.Google Scholar
Scott, D.R., and Stevenson, D.J. (1984) Magma solitons: Geophysical Research Letters, 11, 1161–1164.Google Scholar
Seabrook, C.L., Cawthorn, R.G. and Kruger, F.J. (2005). The Merensky Reef, Bushveld Complex: Mixing of minerals not mixing of magmas. Economic Geology, 100, 11911206.Google Scholar
Segerstrom, K. and Carlson, R.R., 1982. Geologic map of the banded upper zones of the Stillwater Complex and adjacent rocks, Stillwater, Sweet Grass, and Park Counties, Montana. United States Geological Survey Map I-1383.Google Scholar
Self, S., Blake, S., Sharma, K., Widdowson, M. and Sephton, S. (2008). Sulfur and chlorine in Late Cretaceous Deccan magmas and eruptive gas release. Science, 391(5870), 16541657,Google Scholar
Sessa, G., Moroni, M., Tumiati, S., Caruso, S. and Fiorentini, M.L. (2017). Ni-Fe-Cu-PGE ore deposition driven by metasomatic fluids and melt-rock reactions in the deep crust: The ultramafic pipe of Valmaggia, Ivrea-Verbano, Italy. Ore Geology Reviews, 90, 307321.Google Scholar
Sharpe, M.R. (1985). Strontium isotope evidence for preserved density stratification in the Main Zone of the Bushveld Complex, South Africa. Nature, 316, 119126.Google Scholar
Sharpe, M.R. and Irvine, T.N. (1983). Melting relations of two Bushveld chilled margin rocks and implications for the origin for the origin of chromitite. Carnegie Institute of Washington Yearbook, 82, 295300.Google Scholar
Shinohara, H. (2008). Excessive degassing from volcanoes and its role on eruptive and intrusive activity. Reviews of Geophysics, 46(4), RG4005. doi:10.1029/2007RG000244.Google Scholar
Shinohara, H., Ilyama, J.T. and Matsuo, S. (1989). Partition of chlorine compounds between silicate melt and hydrothermal solutions: I. Partition of NaCl-KCl. Geochimica et Cosmochimica Acta, 53, 26172630.Google Scholar
Shirley, D.N. (1986). Compaction of igneous cumulates. Journal of Geology, 94: 795809.Google Scholar
Shirley, D.N. (1987). Differentiation and compaction in the Palisades Sill, New Jersey. Journal of Petrology, 28, 835865.Google Scholar
Sisson, T.W. and Grove, T.L. (1993). Experimental investigations of the role of water in calc-alkaline differentiation and subduction zone magmatism. Contributions to Mineralogy and Petrology, 113, 143166.Google Scholar
Skelton, A.D.L. (1996). The timing and direction of metamorphic fluid flow in Vermont. Contributions to Mineralogy and Petrology, 125, 7584.Google Scholar
Skelton, A.D.L., Graham, C.M. and Bickle, M.J. (1995). Lithological and Structural Controls on Regional 3-D Fluid Flow Patterns during Greenschist Facies Metamorphism of the Dalradian of the SW Scottish Highlands. Journal of Petrology, 36, 563586.Google Scholar
Smith, D.S. and Basson, I.J. (2006). Shape and distribution analysis of Merensky Reef potholing, Northam Platinum Mine, western Bushveld Complex: Implications for pothole formation and growth. Mineralium Deposita, 41(3), 281295.Google Scholar
Smythe, D.J., Wood, B.J. and Kiseeva, E.S. (2017). The S content of silicate melts at sulfide saturation: New experiments and a model incorporating the effects of sulfide composition. American Mineralogist, 102, 795803.Google Scholar
Söhnge, P.G. (1963). Genetic problems of pipe deposits in South Africa. Proceedings of the Geological Society of South Africa, 66, 5972.Google Scholar
Somarin, A.K., Kissin, S.A., Heerema, D.D. and Bihari, D.J. (2009). Hydrothermal alteration, fluid inclusion and stable isotope studies of the North Roby zone, Lac des Iles PGE mine, Ontario, Canada. Resource Geology, 59, 107120.Google Scholar
Sonnenthal, E.L. (1992). Geochemistry of dendritic anorthosites and associated pegmatites in the Skaergaard Intrusion, East Greenland: Evidence for metasomatism by a chlorine-rich fluid. Journal of Volcanology and Geothermal Research, 52, 209230.Google Scholar
Sorensen, H.S. and Wilson, J.R. (1995). A strontium and neodymium isotopic investigation of the Fongen–Hyllingen Layered Intrusion, Norway. Journal of Petrology, 36, 161187.Google Scholar
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(11), 893896.Google Scholar
Stelling, J., Botcharnikov, R.E., Beerman, O. and Nowak, M. (2008). Solubility of H2O- and chlorine-bearing fluids in basaltic melt of Mount Etna at T = 1050–1250 °C and P = 200 MPa. Chemical Geology, 256(3–4), 102110.Google Scholar
Stewart, B.W. and DePaolo, D.J. (1990). Isotopic studies of processes in mafic magma chambers: II. The Skaergaard Intrusion, East Greenland. Contributions to Mineralogy and Petrology, 104, 125141.Google Scholar
Stormer, J.C. Jr., Pierson, M.L. and Tacker, R.C. (1993). Variation of F and Cl X-ray intensity due to anisotropic diffusion in apatite during electron microprobe analysis. American Mineralogist, 78, 641648.Google Scholar
Stumpfl, E.F. (1961). Some new platinoid-rich minerals, identified with the electron microanalyser. Mineralogical Magazine, 32, 833847.Google Scholar
Stumpfl, E.F. (1962). Some aspects of the genesis of platinum deposits. Economic Geology, 57, 619625.Google Scholar
Stumpfl, E.F. and Rucklidge, J.C. (1982). The platiniferous dunite pipes of the eastern Bushveld Complex. Economic Geology, 77, 14191431.Google Scholar
Stumpfl, E.F. and Tarkian, M. (1976). Platinum genesis: New mineralogical evidence. Economic Geology, 71, 14511460.Google Scholar
Sourirajan, S. and Kennedy, G.C. (1962). The system H2O-NaCl at elevated temperatures and pressures. American Journal of Science, 260, 115141.Google Scholar
Sun, S.-S., Nesbitt, R.W. and McCulloch, M.T. (1989). Geochemistry and petrogenesis of Archaean and early Proterozoic siliceous high magnesian basalts. In Boninites and related rocks (Crawford, A., ed.). London: Allen and Unwin, 148173.Google Scholar
Tacker, R.C. and Stormer, J.C. (1989). A thermodynamic model for apatite solid solutions applicable to high temperature geologic problems. American Mineralogist, 74, 877888.Google Scholar
Tacker, R.C. and Stormer, J.C. (1993). Thermodynamics of mixing of liquids in the system Ca3(PO4)2 – CaF2 – CaCl2 – Ca(OH)2. Geochimica et Cosmochimica Acta, 57, 46634676.Google Scholar
Tagirov, B.R., Baranova, N.N., Zotov, A.V., Akinfiev, N.N., Polonyanko, N.A., Shikina, N.D., Koroleva, L. A., Shvarov, Y.V. and Basrakov, E.V. (2013). The speciation and transport of palladium in hydrothermal fluids: Experimental modeling and thermodynamic constraints. Geochimica et Cosmochimica Acta, 117, 348373.Google Scholar
Talkington, R.W. and Lipin, B.R. (1986). Platinum-group minerals in chromite seams of the Stillwater Complex, Montana. Economic Geology, 81, 11791186.Google Scholar
Talkington, R.W. and Watkinson, D.H. (1984). Trends in the distribution of the precious metals in the Lac-Des-Iles Complex, Northwestern Ontario. Canadian Mineralogist, 22, 125136Google Scholar
Talkington, R.W., Watkinson, D.H., Whittaker, P.J. and Jones, P.C. (1984). Platinum-Group Minerals and Other Solid Inclusions in Chromite of Ophiolitic Complexes: Occurrence and Petrological Significance. Tschermaks Mineralogische und Petrographische Mitteilungen, 32, 285301.Google Scholar
Tang, Y.-J., Zhang, H.-F., Nakamura, E. and Ying, J.-F. (2011). Multistage melt/fluid-peridotite interactions in the refertilized lithospheric mantle beneath the North China Craton: Constraints from the Li–Sr–Nd isotopic disequilibrium between minerals of peridotite xenoliths. Contributions to Mineralogy and Petrology, 161, 845861.Google Scholar
Taylor, H.P. Jr. and Forester, R. W. (1979). An oxygen and hydrogen isotope study of the Skaergaard Intrusion and its country rocks: A description of a 55-MY-old fossil hydrothermal system. Journal of Petrology, 20, 355419.Google Scholar
Teague, A.J., Kohn, S.C., Klimm, K. and Botcharnikov, R.E. (2008). Sulphur solubility in Mount Hood andesites and CO2 fluids: Implications for volcanic degassing. EOS Transactions of the American Geophysical Union, 89(53), Fall Meeting Suppl. Abstract V21B-2086.Google Scholar
Tegner, C., Thy, P., Holness, M.B., Jakobsen, J.K. and Lesher, C.E. (2009). Differentiation and compaction in the Skaergaard Intrusion. Journal of Petrology, 50(5), 813840.Google Scholar
Teigler, B. and Eales, H.V. (1993). Correlation between chromite composition and PGE mineralization in the Critical Zone of the western Bushveld Complex. Mineralium Deposita, 28, 291302.Google Scholar
Teigler, B. and Eales, H.V. (1996). The Lower and Critical Zones of the western limb of the Bushveld Complex as intersected by the Nooitgedacht boreholes. Geological Survey of South Africa Bulletin, 111, 126.Google Scholar
Tepley, F.J. III and Davidson, J.P. (2003). Mineral-scale Sr-isotope constraints on magma evolution and chamber dynamics in the Rum layered intrusion, Scotland. Contributions to Mineralogy and Petrology, 145, 621648.Google Scholar
Tepley, F.J. and Davidson, J.P. (2003). Mineral-scale Sr-isotope constraints on magma evolution and chamber dynamics in the Rum layered intrusion, Scotland. Contributions to Mineralogy and Petrology, 145(5), 628641.Google Scholar
Toramaru, A. (1995). Numerical study of nucleation and growth of bubbles in viscous magmas. Journal of Geophysical Research, 100, 19131931.Google Scholar
Tredoux, M., Lindsay, N.M., Davies, G. and Mcdonald, I. (1995). The fractionation of platinum-group elements in magmatic systems, with the suggestion of a novel causal mechanism. South African Journal of Geology, 98, 157167.Google Scholar
Tremaine, P.R. and Leblanc, J.C. (1980). The solubility of nickel oxide and hydrolysis of Ni2+ in water to 573 K. Journal of Chemical Thermodynamics, 12, 521538.Google Scholar
Tsuji, T., Jiang, F. and Christenson, K.Y (2016). Characterization of immiscible fluid displacement processes with various capillary numbers and viscosity ratios in 3D natural sandstone. Advances in Water Resources, 95, 315.Google Scholar
Tsujimori, T. and Ernst, W.G. (2014). Lawsonite blueschists and laalwsonite eclogites as proxies for palaeo-subduction zone processes: A review. Journal of Metamorphic Geology, 32, 437454.Google Scholar
van Haren, J.L.M., Ague, J.J. and Rye, D.M. (1996). Oxygen isotope record of fluid infiltration and mass transfer during regional metamorphism of pelitic schist, Connecticut, USA. Geochimica et Cosmochimica Acta, 60(18), 3487-3504.Google Scholar
Van Orman, J.A., Keshav, S. and Fei, Y. (2008). High-pressure solid/liquid partitioning of Os, Re and Pt in the Fe-S system. Earth and Planetary Science Letters, 274, 250257.Google Scholar
Vidal, O. and Durin, L. (1999). Aluminium mass transfer and diffusion in water at 400–550°C, 2 kbar in the K2O–Al2O3–SiO2–H2O system driven by a thermal gradient or by a variation of temperature with time. Mineralogical Magazine, 63(5), 633647.Google Scholar
Vigneresse, J.-L. (2015). Textures and melt-crystal-gas interactions in granites. Geoscience Frontiers, 6(5), 635663.Google Scholar
Viljoen, M.J. (1999). The nature and origin of the Merensky Reef of the western Bushveld Complex based on geological facies and geophysical data. South African Journal of Geology, 102, 221239Google Scholar
Viljoen, M.J. and Scoon, R.N. (1985). The distribution and main geologic features of discordant bodies of iron-rich ultramafic pegmatite in the Bushveld Complex. Economic Geology, 80, 11091128.Google Scholar
Volborth, A. and Housley, R.M. (1984). A preliminary description of complex graphite, sulfide, arsenide and platinum group element mineralization in a pegmatoid pyroxenite of the Stillwater Complex, Montana, U.S.A. Tschermaks Mineralogische und Petrographische Mitteilungen, 33, 213230.Google Scholar
Volfinger, M., Roberts, J.-L., Vielzeuf, D. and Neiva, A.M.R. (1985). Structural controls of the chlorine content of OH-bearing silicates (micas and amphiboles). Geochimica et Cosmochimica Acta, 49, 3748.Google Scholar
von Gruenewaldt, G., Hatton, C.J., Merkle, R.K.W. and Gain, S.B. (1986). Platinum-group element-chromite associations in the Bushveld Complex. Economic Geology, 81, 10671079Google Scholar
von Gruenewaldt, G., Hulbert, L.J. and Naldrett, A.J. (1989). Contrasting platinum-group element concentration patterns in cumulates of the Bushveld Complex. Mineralium Deposita, 24, 219229.Google Scholar
Voordouw, R., Gutzmer, J. and Beukes, N.J. (2009). Intrusive origin for Upper Group (UG1, UG2) stratiform chromitite seams in the Dwars River area, Bushveld Complex, South Africa. Mineralogy and Petrology, 97, 7594.Google Scholar
Wager, L.R. (1960). The major element variation of the Layered Series of the Skaergaard Intrusion and a re-estimation of the average composition of the Hidden Layered Series and of the successive residual magmas. Journal of Petrology, 1, 364398.Google Scholar
Wager, L.R. and Brown, G.M. (1968). Layered Igneous Rocks. Edinburgh: Oliver &BoydGoogle Scholar
Wager, L.R., Brown, G.M. and Wadsworth, W.J. (1960). Types of igneous cumulates. Journal of Petrology, 1, 7385.Google Scholar
Wager, L.R. and Deer, W.A. (1939). Geological investigations in east Greenland, Part III. Petrology of the Skaergaard intrusion, Kangerdlugssuaq, east Greenland. Meddelelser om Gronland, 105(4), 1352.Google Scholar
Wager, L.R., Vincent, E.A. and Smales, A.A. (1957). Sulphides in the Skaergaard Intrusion, East Greenland. Economic Geology, 52, 855903.Google Scholar
Wagner, P.A. (1929). The Platinum Deposits and Mines of South Africa: Edinburgh, Oliver and Boyd, 326 p.Google Scholar
Wall, C.J. and Scoates, J.S. (2016). High-precision U-Pb Zircon-Baddeleyite dating of the J-M Reef platinum group element deposit in the Stillwater Complex, Montana (USA). Economic Geology, 111(3), 771782.Google Scholar
Wall, C.J. and Scoates, J.S., Weis, D., Friedman, R.M., Amini, M. and Meurer, W.P. (2018). The Stillwater Complex: Integrating zircon geochronological and geochemical constraints on the age, emplacement history and crystallization of a large, open-system layered intrusion. Journal of Petrology, 59(1), 153190.Google Scholar
Wallace, P.J. and Anderson, A.T. Jr. (1998). Effects of eruption and lava drainback on the H2O contents of basaltic magmas at Kilauea Volcano. Bulletin of Volcanology, 59, 327344.Google Scholar
Wallace, P.J., Plank, T., Edmonds, M. and Hauri, E.H. (2015). Volatiles in Magmas. In Encyclopaedia of Volcanoes2nd ed. (Sigurdsson, H., Houghton, B., McNutt, S., Rymer, H. and Stix, J. eds). Amsterdam: Academic Press, 163184.Google Scholar
Walraven, F. (1988). Notes on the age and genetic relationships of the Makhutso granite, Bushveld Complex, South Africa. Chemical Geology, 72, 1728.Google Scholar
Watenphul, A., Schmidt, C. and Jahn, S. (2014). Cr(III) solubility in aqueous fluids at high pressures and temperatures. Geochimica et Cosmochimica Acta, 126, 212227.Google Scholar
Waters, C. and Boudreau, A.E. (1996). A re-evaluation of crystal-size distributions in chromite cumulates. American Mineralogist, 81, 14521459.Google Scholar
Webster, J.D. and Botcharnikov, R.E. (2011). Distribution of Sulfur between melt and fluid in S-O-H-C-Cl-bearing magmatic systems at shallow crustal pressures and temperatures. Reviews in Mineralogy and Geochemistry, 73, 247283.Google Scholar
Webster, J.D., Kinzler, R.J. and Mathez, E.A. (1999). Chloride and water solubility in basalt and andesite liquids and implications for magmatic degassing. Geochimica et Cosmochimica Acta, 63, 729738.Google Scholar
Webster, J.D. (2004). The exsolution of magmatic hydrosaline melts. Chemical Geology, 21, 3348.Google Scholar
Webster, J.D. and DeVivo, B. (2002). Experimental and modeled solubilities of chlorine in aluminosilicate melts, consequences of magma evolution, and implications for exsolution of hydrous chloride melt at Mt. Somma-Vesuvius, Italy. American Mineralogist, 87, 10461061.Google Scholar
Webster, J.D. and Piccoli, P.M. (2015). Magmatic apatite: A powerful, yet deceptive, mineral. Elements, 11, 177182Google Scholar
Webster, J.D., Kinzler, R.J. and Mathez, E.A. (1999). Chloride and water solubility in basalt and andesite melts and implications for magmatic degassing. Geochimica et Cosmochimica Acta, 63, 729738.Google Scholar
Webster, J.D., Thomas, R., Förster, H.-J., Seltmann, R. and Tappen, C. (2004). Geochemical evolution of halogen-enriched, granite magmas and mineralizing fluids of the Zinnwald Tin-Tungsten Mining District, Erzgebirge, Germany. Mineralium Deposita, 39, 452472.Google Scholar
Webster, J.D., Tappen, D. and Mandeville, C.W. (2009). Partitioning behavior of chlorine and fluorine in the system apatite-melt-fluid: II. Felsic silicate systems at 200 MPa. Geochimica et Cosmochimica Acta, 73, 559581.Google Scholar
Webster, J.D., Vetere, V., Botcharnikov, R.E., Goldoff, B. McBirney, A. and Doherty, A.L. (2015). Experimental and modeled chlorine solubilities in aluminosilicate melts at 1 to 7000 bars and 700 to 1250 °C: Applications to magmas of Augustine Volcano, Alaska. American Mineralogist, 100, 522535.Google Scholar
Weiss, D. and Morse, S.A. (1995). Mafic–felsic isotopic disparity in the Kiglapait Intrusion: textures vs. diffusion? Terra Abstracts, 7, 18.Google Scholar
Weiss, D. and Morse, S.A. (1993). Disparate U-Pb systematics of mafics and plagioclase in the Kiglapait Intrusion, Labrador. EOS Transactions, American Geophysical Union, 74, 623.Google Scholar
Wernette, B. and Boudreau, A.E. (2018). The Skaergaard Intrusion: Using Petrography to Characterize an Exsolved Magmatic Volatile Phase. Goldschmidt Conference, Boston, abstract.Google Scholar
Wheeler, S.J. (1988). A conceptual model for soils containing large gas bubbles. Géotechnique, 38, 389397.Google Scholar
Wheeler, S.J. (1990). Movement of large gas bubble in unsaturated fine-grained sediment. Marine Geotechnology, 9, 113129.Google Scholar
Wheeler, S.J. and Gardner, T.N. (1989). The elastic moduli of soils containing large gas bubbles. Géotechnique, 39, 333342.Google Scholar
White, J.A. (1994). The Potgietersrus prospect–geology and exploration history. 15th Council of Mining and Metallurgical Institutions Congress, South African Institute of Mining and Metallurgy, 3, 173181.Google Scholar
Willmore, C.C., Boudreau, A.E. and Kruger, F.J. (2000). The halogen geochemistry of the Bushveld Complex, Republic of South Africa: Implications for chalcophile element distribution in the Lower and Critical zones. Journal of Petrology, 41, 15171539Google Scholar
Willmore, C.C., Boudreau, A.E., Spivack, A. and Kruger, F.J. (2002). Halogens of the Bushveld Complex, South Africa: δ37Cl and Cl/F evidence for hydration melting of the source region in a back-arc setting. Chemical Geology, 182, 503511.Google Scholar
Wilson, A.H. (1982). The geology of the Great Dyke, Zimbabwe: The ultramafic rocks. Journal of Petrology, 23, 240292.Google Scholar
Wilson, A.H. (1992). The geology of the Great Dyke, Zimbabwe: Crystallization, layering, and cumulate formation in the PI pyroxenite of Cyclic Unit 1 of the Darwendale subchamber. Journal of Petrology, 33, 611663.Google Scholar
Wilson, A.H. (1996). The Great Dyke of Zimbabwe. In Layered Intrusions (Cawthorn, R.G., ed.). Amserdam: Elsevier, 365402.Google Scholar
Wilson, A.H. (2001). Compositional and lithological controls on the PGE-bearing sulphide zones in the Selukwe subchamber, Great Dyke: A combined equilibrium-Rayleigh fractionation model. Journal of Petrology, 42, 18451867.Google Scholar
Wilson, A.H. (2012). A Chill Sequence to the Bushveld Complex: Insight into the first stage of emplacement and Implications for the parental magmas. Journal of Petrology, 53(6), 11231168.Google Scholar
Wilson, A.H. and Prendergast, M.D. (1989). The Great Dyke of Zimbabwe-I: Tectonic setting, stratigraphy, petrology, structure, emplacement and crystallisation. In Magmatic sulphides – The Zimbabwe volume (Prendergast, M.D. and Jones, M.J., eds). London, Institute of Mining and Metallurgy, 120.Google Scholar
Wilson, A.H. and Prendergast, M.D. (2001). Platinum-group element mineralization in the Great Dyke, Zimbabwe, and its relationship to magma evolution and magma chamber structure. South African Journal of Geology, 104, 319342.Google Scholar
Wilson, A.H., Naldrett, A.J. and Tredoux, M. (1989). Distribution and controls of platinum-group element and base-metal mineralization in the Darwendale subchamber of the Great Dyke, Zimbabwe. Geology, 17, 649652.Google Scholar
Wilson, A.H., Murahwi, C.Z. and Coghill, B. (2000). Stratigraphy, geochemistry and platinum group element mineralisation of the central zone of the Selukwe Subchamber, Great Dyke, Zimbabwe. Journal of African Earth Sciences, 30, 833853.Google Scholar
Wilson, C.R., Spiegelman, M., van Keken, P.E. and Hacker, B.R. (2014). Fluid flow in subduction zones: The role of solid rheology and compaction pressure. Earth and Planetary Science Letters, 401, 261274.Google Scholar
Wilson, J.R., Robins, B., Nielsen, F.M., Duchesne, J.C. and Auwers, J.V. (1996). The Bjerkreim–Sokndal Layered Intrusion, southwest Norway. In Layered Intrusions (Cawthorn, R.G., ed.). Amsterdam: Elsevier, 231255.Google Scholar
Woods, W. and Norris, S. (2010). On the role of caprock and fracture zones in dispersing gas plumes in the subsurface. Water Resources Research, 46: W08522. doi:10.1029/2008WR007568.Google Scholar
Worst, B.G. (1960). The Great Dyke of Southern Rhodesia. Southern Rhodesia Geological Survey Bulletin, 47, 234 pp.Google Scholar
Yang, S.-H., Maier, W.D., Lahaye, Y. and O’Brien, H. (2013). Strontium isotope disequilibrium of plagioclase in the Upper Critical Zone of the Bushveld Complex: Evidence for mixing of crystal slurries. Contributions to Mineralogy and Petrology, 166(4), 959974.Google Scholar
Yardley, B.W.D. (2005). Metal Concentrations in Crustal Fluids and Their Relationship to Ore Formation. Economic Geology, 100, 613632.Google Scholar
Zhang, C., Oostrom, M. Wietsma, T.W., Grate, J.W. and Warner, M.G. (2011). Influence of viscous and capillary forces on immiscible fluid displacement: Pore-scale experimental study in a water-wet micromodel demonstrating viscous and capillary fingering. Energy and Fuels, 25, 34933505.Google Scholar
Zhu, C. and Sverjensky, D.A. (1992). F-Cl-OH partitioning between biotite and apatite. Geochimica et Cosmochimica Acta, 56(9), 34353468.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. In The Geology, Geochemistry, Mineralogy and Mineral Beneficiation of Platinum-Group Elements (Cabri, L.J., ed.). Canadian Institute of Mining and Metallurgy Special Volume 54, 459481.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
  • Alan Boudreau, Duke University, North Carolina
  • Book: Hydromagmatic Processes and Platinum-Group Element Deposits in Layered Intrusions
  • Online publication: 08 March 2019
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
  • Alan Boudreau, Duke University, North Carolina
  • Book: Hydromagmatic Processes and Platinum-Group Element Deposits in Layered Intrusions
  • Online publication: 08 March 2019
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
  • Alan Boudreau, Duke University, North Carolina
  • Book: Hydromagmatic Processes and Platinum-Group Element Deposits in Layered Intrusions
  • Online publication: 08 March 2019
Available formats
×