Skip to main content
Log in

Vanadium magnetite–melt oxybarometry of natural, silicic magmas: a comparison of various oxybarometers and thermometers

  • Original Paper
  • Published:
Contributions to Mineralogy and Petrology Aims and scope Submit manuscript

Abstract

To test a recently developed oxybarometer for silicic magmas based on partitioning of vanadium between magnetite and silicate melt, a comprehensive oxybarometry and thermometry study on 22 natural rhyolites to dacites was conducted. Investigated samples were either vitrophyres or holocrystalline rocks in which part of the mineral and melt assemblage was preserved only as inclusions within phenocrysts. Utilized methods include vanadium magnetite–melt oxybarometry, Fe–Ti oxide thermometry and -oxybarometry, zircon saturation thermometry, and two-feldspar thermometry, with all analyses conducted by laser-ablation ICP–MS. Based on the number of analyses, the reproducibility of the results and the certainty of contemporaneity of the analyzed minerals and silicate melts the samples were grouped into three classes of reliability. In the most reliable (n = 5) and medium reliable (n = 10) samples, all fO2 values determined via vanadium magnetite–melt oxybarometry agree within 0.5 log units with the fO2 values determined via Fe–Ti oxide oxybarometry, except for two samples of the medium reliable group. In the least reliable samples (n = 7), most of which show evidence for magma mixing, calculated fO2 values agree within 0.75 log units. Comparison of three different thermometers reveals that temperatures obtained via zircon saturation thermometry agree within the limits of uncertainty with those obtained via two-feldspar thermometry in most cases, whereas temperatures obtained via Fe–Ti oxide thermometry commonly deviate by ≥50 °C due to large uncertainties associated with the Fe–Ti oxide model at T-fO2 conditions typical of most silicic magmas. Another outcome of this study is that magma mixing is a common but easily overlooked phenomenon in silicic volcanic rocks, which means that great care has to be taken in the application and interpretation of thermometers and oxybarometers.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

References

  • Andersen DJ, Lindsley DH (1985) New (and final!) models for the Ti-magnetite/ilmenite geothermometer and oxygen barometer. In: Abstract AGU 1985 Spring Meeting Eos Transactions. American Geophysical Union 66(18):416

  • Andersen DJ, Lindsley DH (1988) Internally consistent solution models for Fe–Mg–Mn–Ti oxides: Fe–Ti oxides. Am Miner 73:714–726

    Google Scholar 

  • Andersen DJ, Lindsley DH, Davidson PM (1993) QUILF: a pascal program to assess equilibria among Fe–Mg–Mn–Ti oxides, pyroxenes, olivine, and quartz. Comput Geosci 19(9):1333–1350

    Article  Google Scholar 

  • Arató R, Audétat A (2017) Experimental calibration of a new oxybarometer for silicic magmas based on vanadium partitioning between magnetite and silicate melt. Geochim Cosmochim Acta 209:284–295

    Article  Google Scholar 

  • Audétat A (2013) Origin of Ti-rich rims in quartz phenocrysts from the Upper Bandelier Tuff and the Tunnel Spring Tuff, southwestern USA. Chem Geol 360–361:99–104

    Article  Google Scholar 

  • Audétat A (2015) Compositional evolution and formation conditions of magmas and fluids related to porphyry Mo mineralization at Climax, Colorado. J Petrol 56(8):1519–1546

    Article  Google Scholar 

  • Audétat A, Lowenstern JB (2014) Melt inclusions. In: Scott SD (ed) Geochemistry of mineral deposits. Treatise on Geochemistry, vol 13, 2nd edn. Elsevier, Amsterdam, pp 143–173

    Chapter  Google Scholar 

  • Audétat A, Pettke T (2003) The magmatic-hydrothermal evolution of two barren granites: a melt and fluid inclusion study of the Rito del Medio and Canada Pinabete plutons in northern New Mexico (USA). Geochim Cosmochim Acta 67(1):97–121

    Article  Google Scholar 

  • Audétat A, Pettke T (2006) Evolution of a Porphyry-Cu mineralized magma system at Santa Rita, New Mexico (USA). J Petrol 47(10):2021–2046

    Article  Google Scholar 

  • Audétat A, Günther D, Heinrich C (2000) Magmatic-hydrothermal evolution in a fractionating granite: a microchemical study of the Sn-WF-mineralized Mole Granite (Australia). Geochim Cosmochim Acta 64(19):3373–3393

    Article  Google Scholar 

  • Audétat A, Dolejs D, Lowenstern JB (2011) Molybdenite saturation in silicic magmas: occurrence and petrological implications. J Petrol 52(5):891–904

    Article  Google Scholar 

  • Bachmann O (2010) The petrologic evolution and pre-eruptive conditions of the rhyolitic Kos Plateau Tuff (Aegean Arc). Open Geosci 2(3):270–305

    Article  Google Scholar 

  • Bachmann O, Charlier BLA, Lowenstern JB (2007) Zircon crystallization and recycling in the magma chamber of the rhyolitic Kos Plateau Tuff (Aegean arc). Geology 35(1):73–76

    Article  Google Scholar 

  • Bachmann O, Wallace PJ, Bourquin J (2009) The melt inclusion record from the rhyolitic Kos Plateau Tuff (Aegean Arc). Contrib Mineral Petrol 159(2):187–202

    Article  Google Scholar 

  • Bacon CR, Hirschmann MM (1988) Mg/Mn partitioning as a test for equilibrium between coexisting Fe-Ti oxides. Am Mineral 73(1–2):57–61

    Google Scholar 

  • Bailey R (1989) Geologic map of Long Valley Caldera, Mono-Inyo craters volcanic chain and vicinity, Eastern California, scale 1:62,500. US Geol Surv Geophys Invest Map I—1993:11

  • Bailey RA (2004) Eruptive history and chemical evolution of the precaldera and postcaldera basalt-dacite sequences, Long Valley, California: Implications for magma sources, current seismic unrest, and future volcanism. USGS Prof Pap 1692

  • Best M (2013) The 36–18 Ma Indian Peak-Caliente ignimbrite fi eld and calderas, southeastern Great Basin, USA: multicyclic super-eruptions. Geosphere 9(4):864–950

    Article  Google Scholar 

  • Best MG, Christiansen EH, Blank RH (1989) Oligocene caldera complex and calc-alkaline tuffs and lavas of the Indian Peak volcanic field, Nevada and Utah. Geol Soc Am Bull 101(8):1076–1090

    Article  Google Scholar 

  • Blevin PL, Chappell BW, Allen CM (1996) Intrusive metallogenic provinces in eastern Australia based on granite source and composition. Geol Soc Am Spec Pap 315:281–290

    Google Scholar 

  • Boehnke P, Watson EB, Trail D, Harrison TM, Schmitt AK (2013) Zircon saturation re-revisited. Chem Geol 351:324–334

    Article  Google Scholar 

  • Bray BA (2014) Mafic replenishment of multiple felsic reservoirs at the Mono domes and Mono Lake islands, California. MSc thesis, McGill University

  • Buddington A, Lindsley D (1964) Iron-titanium oxide minerals and synthetic equivalents. J Petrol 5(2):310–357

    Article  Google Scholar 

  • Candela PA, Bouton SL (1990) The influence of oxygen fugacity on tungsten and molybdenum partitioning between silicate melts and ilmenite. Econ Geol 85(3):633–640

    Article  Google Scholar 

  • Canil D (1999) Vanadium partitioning between orthopyroxene, spinel and silicate melt and the redox states of mantle source regions for primary magmas. Geochim Cosmochim Acta 63(3):557–572

    Article  Google Scholar 

  • Canil D (2002) Vanadium in peridotites, mantle redox and tectonic environments: archean to present. Earth Planet Sci Lett 195(1):75–90

    Article  Google Scholar 

  • Carmichael ISE (1967) The iron-titanium oxides of salic volcanic rocks and their associated ferromagnesian silicates. Contrib Miner Petrol 14:36–64

    Article  Google Scholar 

  • Christiansen EH, Sheridan MF, Burt DM (1986) The geology and geochemistry of Cenozoic topaz rhyolites from the western United States. Geol Soc Am Spec Pap 205:1–82

    Google Scholar 

  • Cordell L, Long CL, Jones DW (1985) Geophysical expression of the batholith beneath Questa caldera, New Mexico. J Geophys Res 90(B13):11263–11269

    Article  Google Scholar 

  • Ebadi A, Johannes W (1991) Beginning of melting and composition of first melts in the system Qz–Ab–Or–H2O–CO2. Contrib Mineral Petrol 106(3):286–295

    Article  Google Scholar 

  • Eichler C (2012) Petrogenesis of the East Fork Member Rhyolites, Valles Caldera, New Mexico, USA. MSc thesis, University of Nevada

  • Ferriz H, Mahood GA (1984) Eruption rates and compositional trends at Los Humeros Volcanic Center, Puebla. Mexico. J Geophys Res 89(B10):8511–8524

    Article  Google Scholar 

  • Flege RF (1959) Geology of Lordsburg Quadrangle, Hidalgo County, New Mexico. New Mexico Bur Mines and Mineral Resources Bull (62):36

  • Frost BR, Lindsley D (1992) Equilibria among Fe-Ti oxides, pyroxenes, olivine, and quartz: part II. Application. Am Miner 77:1004

    Google Scholar 

  • Gervasoni F, Klemme S, Rocha-Júnior ERV, Berndt J (2016) Zircon saturation in silicate melts: a new and improved model for aluminous and alkaline melts. Contrib Miner Petrol 171(3):1–12

    Article  Google Scholar 

  • Ghiorso MS, Evans BW (2008) Thermodynamics of rhombohedral oxide solid solutions and a revision of the Fe-Ti two-oxide geothermometer and oxygen-barometer. Am J Sci 308(9):957–1039

    Article  Google Scholar 

  • Ghiorso MS, Sack O (1991) Fe-Ti oxide geothermometry: thermodynamic formulation and the estimation of intensive variables in silicic magmas. Contrib Miner Petrol 108:485–510

    Article  Google Scholar 

  • Gilbert C, Christensen M, Al-Rawi Y, Lajoie K (1968) Structural and volcanic history of Mono basin, California-Nevada. Geol Soc Am Mem 116:275–330

    Article  Google Scholar 

  • Godwin L, Gaskill D (1964) Post-Paleocene West Elk laccolithic cluster, west-central Colorado. Geological Survey Research 1964, US Geol Surv Profess Pap 501-C:3

  • Goff F, Gardner J (2004) Late Cenozoic geochronology of volcanism and mineralization in the Jemez Mountains and Valles caldera, north central New Mexico. In: Mack GH, KA G (eds) The Geology of New Mexico. New Mexico Geological Society, Socorro, pp 295–312

  • Hakim A, Sufni Hall R (1991) Tertiary volcanic rocks from the Halmahera Arc, Eastern Indonesia. J Southeast Asian Earth Sci 6:271–287

    Article  Google Scholar 

  • Hora JM, Singer BS, Wörner G (2007) Volcano evolution and eruptive flux on the thick crust of the Andean Central Volcanic Zone: 40Ar/39Ar constraints from Volcan Parinacota, Chile. Geol Soc Am Bull 119(3–4):343–362

    Article  Google Scholar 

  • Hora JM, Kronz A, Möller-McNett S, Wörner G (2013) An Excel-based tool for evaluating and visualizing geothermobarometry data. Comput Geosci 56:178–185

    Article  Google Scholar 

  • Horn I, Foley SF, Jackson SE, Jenner GA (1994) Experimentally determined partitioning of high field strength- and selected transition elements between spinel and basaltic melt. Chem Geol 117(1):193–218

    Article  Google Scholar 

  • Irving AJ (1978) A review of experimental studies of crystal/liquid trace element partitioning. Geochim Cosmochim Acta 42(6):743–770

    Article  Google Scholar 

  • Ishihara S (1981) The granitoid series and mineralization. Econ Geol 75th Anniv Vol, 458–484

  • Janssen A, Putnis A, Geisler T, Putnis CV (2010) The experimental replacement of ilmenite by rutile in HCl solutions. Miner Mag 74(4):633–644

    Article  Google Scholar 

  • Jochum KP, Weis U, Stoll B, Kuzmin D, Yang Q, Raczek I, Jacob DE, Stracke A, Birbaum K, Frick DA, Günther D, Enzweiler J (2011) Determination of reference values for NIST SRM 610-617 glasses following ISO guidelines. Geostand Geoanal Res 35(4):397–429

    Article  Google Scholar 

  • Johnsen RL, Smith EI, Biek RF (2010) Subalkaline volcanism in the Black Rock Desert and Markagunt Plateau volcanic fields of south-central Utah. In: Carney SM, Tabet DE, Johnson CL (eds) Geology of South-Central Utah. Utah Geological Association,  Salt Lake City, pp 109–150

  • Johnson CM, Lipman PW (1988) Origin of metaluminous and alkaline volcanic rocks of the Latir volcanic field, northern Rio Grande rift, New Mexico. Contrib Miner Petrol 100(1):107–128

    Article  Google Scholar 

  • Jones WR, Hernon RM, Moore SL (1967) General geology of Santa Rita quadrangle, Grant County, New Mexico. Geol Surv Prof Pap 555:144

  • Jugo P, Candela P, Piccoli P (1999) Magmatic sulfides and Au: cu ratios in porphyry deposits: an experimental study of copper and gold partitioning at 850 °C, 100 MPa in a haplogranitic melt–pyrrhotite–intermediate solid solution–gold metal assemblage, at gas saturation. Lithos 46(3):573–589

    Article  Google Scholar 

  • Jugo PJ, Wilke M, Botcharnikov RE (2010) Sulfur K-edge XANES analysis of natural and synthetic basaltic glasses: implications for S speciation and S content as function of oxygen fugacity. Geochim Cosmochim Acta 74(20):5926–5938

    Article  Google Scholar 

  • Keith J, Shanks W (1988) Chemical evolution and volatile fugacities of the Pine Grove porphyry molybdenum and ash-flow tuff system, southwestern Utah. In: Taylor RP, Strong DF (eds) Recent advances in the geology of granite-related mineral deposits. Canadian Institute of Mining and Metallurgy, Montréal, pp 402–423

  • Keith JD, Shanks WC, Archibald DA, Farrar E (1986) Volcanic and intrusive history of the Pine Grove porphyry molybdenum system, southwestern Utah. Econ Geol 81(3):553–577

    Article  Google Scholar 

  • Keller J (1969) Origin of rhyolites by anatectic melting of granitic crustal rocks. B Volcanol 33(3):942–959

    Article  Google Scholar 

  • Keppler H (1993) Influence of fluorine on the enrichment of high field strength trace elements in granitic rocks. Contrib Miner Petrol 114(4):479–488

    Article  Google Scholar 

  • Knott JR, Sarna-Wojcicki AM (2001) Stop C3 Late Pliocene tephrostratigraphy and geomorphic development of the Artists Drive structural block In: Machette MN, Johnson ML, Slate JL (eds) Quaternary and Late Pliocene Geology of the Death Valley Region: Recent Observations on Tectonics, Stratigraphy, and Lake Cycles. Guidebook for the 2001 Pacific Cell—Friends of the Pleistocene Fieldtrip, 105–111

  • Lasky SG (1938) Geology and ore deposits of the Lordsburg mining district, Hidalgo County, New Mexico. US Geol Surv Bull 885:62

  • Lattard D, Sauerzapf U, Käsemann M (2005) New calibration data for the Fe-Ti oxide thermo-oxybarometers from experiments in the Fe-Ti-O system at 1 bar, 1,000–1,300 °C and a large range of oxygen fugacities. Contrib Miner Petrol 149(6):735–754

    Article  Google Scholar 

  • Lehmann B (1990) Metallogeny of Tin. Springer, Berlin

  • Lepage LD (2003) ILMAT: an Excel worksheet for ilmenite–magnetite geothermometry and geobarometry. Comput Geosci 29(5):673–678

    Article  Google Scholar 

  • Lindsley DH, Frost BR (1992) Equilibria among Fe-Ti oxides, pyroxenes, olivine, and quartz; Part I. Theory. Am Miner 77(9–10):987–1003

    Google Scholar 

  • Linnen RL, Pichavant M, Holtz F (1996) The combined effects of fO2 and melt composition on SnO2 solubility and tin diffusivity in haplogranitic melts. Geochim Cosmochim Acta 60(24):4965–4976

    Article  Google Scholar 

  • Lipman PW, Mehnert HH, Naeser CW (1986) Evolution of the Latir Volcanic Field, Northern New Mexico, and Its Relation to the Rio Grande Rift, as Indicated by Potassium-Argon and Fission Track Dating. J Geophys Res 91(B6):6329–6345

    Article  Google Scholar 

  • Lowenstern JB, Bacon CR, Calk LC, Hervig RL, Aines RD (1994) Major-element, trace-element, and volatile concentrations in silicate melt inclusions from the tuff of Pine Grove, Wah Wah Mountains, Utah. US Geological Survey Open-File Report 94–242:20

  • Mallmann G, O’Neill HSC (2009) The crystal/melt partitioning of V during mantle melting as a function of oxygen fugacity compared with some other elements (Al, P, Ca, Sc, Ti, Cr, Fe, Ga, Y, Zr and Nb). J Petrol 50(9):1765–1794

    Article  Google Scholar 

  • Mercer CN, Hofstra AH, Todorov TI, Roberge J, Burgisser A, Adams DT, Cosca M (2015) Pre-eruptive conditions of the hideaway park topaz rhyolite: insights into metal source and evolution of magma parental to the Henderson porphyry molybdenum deposit, Colorado. J Petrol 56(4):645–679

    Article  Google Scholar 

  • Mutschler F, Ernst D, Gaskill D, Billings P (1981) Igneous rocks of the Elk Mountains and vicinity, Colorado-chemistry and related ore deposits. Western Slope Colorado: New Mexico Geological Society 32nd Field Conference Guide Book, 317–324

  • Obradovich JD, Mutschler FE, Bryant B (1969) Potassium-argon ages bearing on the igneous and tectonic history of the Elk Mountains and Vicinity, Colorado: a preliminary report. Geol Soc Am Bull 80(9):1749–1756

    Article  Google Scholar 

  • Ogoh K, Toyoda S, Ikeda S, Ikeya M, Goff F (1993) Cooling history of the Valles caldera, New Mexico using ESR dating method. Appl Radiat Isot 44(1):233–237

    Article  Google Scholar 

  • Peiffert C, Cuney M, Nguyen-Trung C (1994) Uranium in granitic magmas: part 1. Experimental determination of uranium solubility and fluid-melt partition coefficients in the uranium oxide-haplogranite-H2O-Na2CO3 system at 720–770 °C, 2 kbar. Geochim Cosmochim Acta 58(11):2495–2507

    Article  Google Scholar 

  • Phillips W, Poths J, Goff F, Reneau S, McDonald E (1997) Ne-21 surface exposure ages from the Banco Bonito obsidian flow, Valles caldera, New Mexico, USA. Geological Society of America, Abstracts with Programs 29(6):A-419

  • Pierce KL, Fosberg M, Scott WE, Lewis GC, Colman SM (1982) Loess deposits of southeastern Idaho: age and correlation of the upper two loess units. Idaho Bur Mines Geol Bull 26:717–725

    Google Scholar 

  • Poletski SJ (2010) The Nomlaki Tuff eruption: chemical correlation of a widespread Pliocene stratigraphic marker. MSc thesis, California State University

  • Putirka KD (2008) Thermometers and barometers for volcanic systems. Rev Miner Geochem 69(1):61–120

    Article  Google Scholar 

  • Putirka K (2016) Rates and styles of planetary cooling on Earth, Moon, Mars, and Vesta, using new models for oxygen fugacity, ferric-ferrous ratios, olivine-liquid Fe–Mg exchange, and mantle potential temperature. Am Mineral 101:819–840

    Article  Google Scholar 

  • Ridolfi F, Renzulli A, Puerini M (2009) Stability and chemical equilibrium of amphibole in calc-alkaline magmas: an overview, new thermobarometric formulations and application to subduction-related volcanoes. Contrib Miner Petrol 160(1):45–66

    Article  Google Scholar 

  • Righter K, Leeman WP, Hervig RL (2006a) Partitioning of Ni, Co and V between spinel-structured oxides and silicate melts: importance of spinel composition. Chem Geol 227(1–2):1–25

    Article  Google Scholar 

  • Righter K, Sutton SR, Newville M, Le L, Schwandt CS, Uchida H, Lavina B, Downs RT (2006b) An experimental study of the oxidation state of vanadium in spinel and basaltic melt with implications for the origin of planetary basalt. Am Mineral 91(10):1643–1656

    Article  Google Scholar 

  • Rose AW, Baltosser WW (1966) The porphyry copper deposit at Santa Rita, New Mexico. In: Titley SR, Hicks CL (eds) Geology of the porphyry copper deposits in southwestern North America. University of Arizona, Tucson, pp 205–220

    Google Scholar 

  • Sarna-Wojcicki AM, Bowman H, Meyer CE, Russell P, Woodward M, McCoy G, Rowe Jr J, Baedecker P, Asaro F, Michael H (1984) Chemical analyses, correlations, and ages of upper Pliocene and Pleistocene ash layers of east-central and southern California. Geol Surv Prof Pap 1293:40

  • Seghedi I (2011) Permian rhyolitic volcanism, changing from subaqueous to subaerial in post-Variscan intra-continental Sirinia Basin (SW Romania-Eastern Europe). J Volcanol Geotherm Res 201(1–4):312–324

    Article  Google Scholar 

  • Smith P, York D, Chen Y, Evensen N (1996) Single crystal 40Ar-39Ar dating of a Late Quaternary paroxysm on Kos, Greece: concordance of terrestrial and marine ages. Geophys Res Lett 23(21):3047–3050

    Article  Google Scholar 

  • Steven TA (1989) Geologic map of the Crystal Peak caldera, west-central Utah. US Geol Surv Misc Inv Ser Map I-2002

  • Stormer JC (1983) The effects of recalculation on estimates of temperature and oxygen fugacity from analyses of multicomponent iron-titanium oxides. Am Miner 68(5–6):586–594

    Google Scholar 

  • Taylor JR, Wall VJ (1992) The behavior of tin in granitoid magmas. Econ Geol 87(2):403–420

    Article  Google Scholar 

  • Thorman C, Drewes H (1978) Cretaceous–early Tertiary history of the northern Pyramid Mountains, southwestern New Mexico. In: Callender JF, Wilt J, Clemons RE, James HL (eds), New Mexico Geological Society 29th Annual Fall Field Conference Guidebook, 215–218

  • Turley C, Nash W (1980) Petrology of late Tertiary and Quaternary volcanism in western Juab and Millard Counties, Utah. Utah Geol Min Surv Spec Stud 52:1–33

    Google Scholar 

  • Varga RJ, Bailey RA, Suemnicht GA (1990) Evidence for 600 year-old basalt and magma mixing at Inyo Craters Volcanic Chain, Long Valley Caldera, California. J Geophys Res 95(B13):21441–21450

    Article  Google Scholar 

  • Watson EB, Harrison TM (1983) Zircon saturation revisited: temperature and composition effects in a variety of crustal magma types. Earth Planet Sci Lett 64(2):295–304

    Article  Google Scholar 

  • Wones D (1981) Mafic silicates as indicators of intensive variables in granitic magmas. Min Geol 31(4):191–212

    Google Scholar 

  • Wones D, Eugster H (1965) Stability of biotite–experiment theory and application. Am Miner 50(9):1228–1272

    Google Scholar 

  • Wörner G, Harmon R, Davidson J, Moorbath S, Turner D, McMillan N, Nyes C, Lopez-Escobar L, Moreno H (1988) The Nevados de Payachata volcanic region (18 S/69 W, N. Chile). Bull Volcanol 50(5):287–303

    Article  Google Scholar 

  • Xirouchakis D, Lindsley DH, Frost BR (2001) Assemblages with titanite (CaTiOSiO4), Ca–Mg–Fe olivine and pyroxenes, Fe–Mg–Ti oxides, and quartz: Part II. Application. Am Miner 86(2–3):254–264

    Article  Google Scholar 

  • Yang Q-C, Jochum KP, Stoll B, Weis U, Kuzmin D, Wiedenbeck M, Traub H, Andreae MO (2012) BAM-S005 type A and B: new silicate reference glasses for microanalysis. Geostand Geoanal Res 36(3):301–313

    Article  Google Scholar 

  • Zajacz Z, Halter WE, Pettke T, Guillong M (2008) Determination of fluid/melt partition coefficients by LA-ICPMS analysis of co-existing fluid and silicate melt inclusions: controls on element partitioning. Geochim Cosmochim Acta 72(8):2169–2197

    Article  Google Scholar 

Download references

Acknowledgements

This research was supported by the German Science Foundation grant AU 314/5-1. We thank the Smithsonian Institution for providing samples of vitrophyres, as well as John Hora, Thomas Pettke and Sorin Silviu Udubasa for providing samples of the Parinacota volcano, the Kos Granite inclusion and the Oravita hyalodacite, respectively. We are also grateful to Raphael Njul for the preparation of polished sections. Detailed and constructive reviews by Keith D. Putirka and an anonymous reviewer significantly contributed to the final version of the manuscript and are gratefully appreciated.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Róbert Arató.

Additional information

Communicated by Mark S Ghiorso.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (XLSX 341 kb)

Appendix

Appendix

Samples

Four samples were kindly provided by the Smithsonian Institution (NMNH 117450-58 Los Humeros volcano—Mexico; NMNH 117455-27 Mount Rano—Indonesia; NMNH 117462-1 Blackfoot Lava Field—USA; NMNH 116377-38a Nomlaki Tuff—USA), one by Sorin Silviu Udubasa (Oravita area—Romania) and one by John Hora (Parinacota volcano—Chile). Additional vitrophyres (Inyo-Mono craters, Banco Bonito volcano—both USA), silicic tuffs (Winter Park, Smelter Knolls, Guaje, Pine Grove, Crystal Peak, Lordsburg, Chalk Mountain, Amalia and Cottonwood—all USA) and fine-grained porphyritic rocks (Santa Rita rhyodacite dikes, Lordsburg rhyolite and granodiorite, Topaz Mountain, The Dyke—all USA) were collected by Andreas Audétat, whereas a sample from Kos (Greece) was kindly donated by Thomas Pettke.

Sample description

Oraviţa hyalodacite

Unfortunately, there is no detailed description (such as exact locality, rock unit, etc.) available for this specimen. Most probably it stems from one of the Permian rhyolitic flow units of the Oraviţa area in Romania. Publications about these Permian volcanics in the area are mainly in Romanian language and/or not accessible for the public, but a recent paper (Seghedi 2011) describes the Lower Permian volcanic occurrences of the neighboring Sirinia basin. The exclusively rhyolitic composition (74–80% SiO2) and the glass-rich, dominantly hyaline texture of Sirinia basin rhyolites—similarly to our sample—let us to suggest that they are genetically related.

Mount Rano

The Mount Rano volcano is located on the southwestern part of the North Maluku Island, Indonesia. The volcano forms part of the Late Cretaceous-Eocene Oha Formation (Hakim et al. 1991), consisting of basalts and andesites. However, our vitrophyre sample, which stems from the southwestern flank of Mount Rano and was kindly provided by the Smithsonian institution (NMNH 117455-27), has a rhyolitic composition.

Parinacota

The Parinacota volcano is located in Chile, Central Andes, and its evolution can be subdivided into five main stages, spanning from 163 ka until 5 ka (K–Ar; Hora et al. 2007). A detailed description of the geochemistry, petrography and evolution history can be found in Wörner et al. (1988) and in Hora et al. (2007). Our sample stems from the Rhyodacite Dome Plateau sequence (47–40 ka), which represents the second stage of the volcano’s evolution history. In contrast to most other magmas produced at this stage, the rhyodacite sample originates from one of the non-mixing endmember reservoirs. This sample was thoroughly investigated in a case study involving a test of multiple thermometers (Hora et al. 2013). Despite some high amphibole–plagioclase temperatures obtained from glomerocrysts, the temperature estimates obtained by Hora et al. (2013) from phenocrysts/microphenocrysts provide convincing evidence that this unit is homogenous and free of magma mixing.

Hideaway Park tuff

The Hideaway Park tuff (also referred to as Winter Park tuff) is an extrusive volcanic unit of Oligocene age (27.77 ± 0.34 Ma, sanidine K–Ar; Mercer et al. 2015) that is genetically related to the Red Mountain intrusive complex in Colorado, USA, which hosts the giant Urad-Henderson porphyry Mo deposit. Its mineralogy and geochemistry is described in detail by Mercer et al. (2015), who provided also an extensive set of mineral– and melt inclusion analyses. In order to calculate the composition of the crystallized melt inclusions, the average Al2O3 content from all homogenized melt inclusions and pumice glasses analyzed by Mercer et al. (2015) was used as internal standard (13.30 wt% Al2O3).

Cottonwood Wash tuff

The Cottonwood Wash tuff belongs to the Indian Peak volcanic field that was active from ca. 30 to 29 Ma (plagioclase K–Ar; Best 2013) and comprises tens of calderas and related ash-flow sheets that cover about 50,000 km2 across the Utah-Nevada state line, USA. The Cottonwood Wash tuff formed during a super-eruption 31.13 million years ago (plagioclase K–Ar; Best 2013), producing about 2000 km3 of crystal-rich, dacitic tuff, with a fairly homogenous composition over its entire extent. A detailed description about the geochemistry, petrography and mineralogy of the volcanic units of the Indian Peak volcanic field can be found in Best et al. (1989) and Best (2013). The rapidly quenched, fresh, glassy matrix and the presence of abundant Fe–Ti oxide microphenocrysts allow easy application of the mgt–ilm oxybarometer and the V partitioning oxybarometer. At the same time, the latter method could be applied also to inclusions, as both magnetite inclusions (in quartz) and melt inclusions (in quartz and feldspar) were available.

Lordsburg rhyolite

The Lordsburg mining district is located in the Pyramid Mountains, a north-trending range of Lower Cretaceous to middle Tertiary volcanic and plutonic rocks in the Basin and Range province of southwestern New Mexico, southwest of the city of Lordsburg. The Paleocene intrusive rocks of the area include a granodiorite stock (58.8 ± 2 Ma, biotite K–Ar; Thorman and Drewes 1978), rhyolitic vents, breccia pipes and dikes, the first two formations of which were investigated in this study. Although some early studies have been conducted on these rock units (Lasky 1938; Flege 1959; Thorman and Drewes 1978), nothing has been published about their major- and trace element composition.

Lordsburg granodiorite

This sample was collected from the most widespread intrusive unit of the Pyramid Mountains—the granodiorite—which intruded also the previously described Lordsburg rhyolite. The lack of quartz phenocrysts precluded any inclusion-based thermometry and oxybarometry. However, due to the relatively unaltered nature of this rock (unlike the Lordsburg rhyolite) the relevant information was obtained from the composition of fresh-looking microphenocrysts and the fine-grained rock matrix.

Smelter Knolls rhyolite

The Smelter Knolls rhyolite belongs to a bimodal association of silica-rich, often topaz-bearing rhyolites and contemporaneous basalts and basaltic andesites of Cenozoic age that are widespread in the Basin and Range province and along the Rio Grande Rift in Western USA and Mexico (Christiansen et al. 1986). Thorough geological, geochemical and age data about these topaz rhyolites can be found in the comprehensive study of Christiansen et al. (1986), whereas the detailed description of the Smelter Knolls complex was published by Turley and Nash (1980). The Smelter Knolls represent a single rhyolite flow-dome complex measuring 5 km in diameter and 2.2 km3 in volume, located north of the city of Delta in Utah, USA. The dome complex formed at 3.4 ± 0.1 Ma based on sanidine K–Ar dating (Turley and Nash 1980). A recent study on the Black Rock Desert volcanic field (Johnsen et al. 2010)—which includes the Smelter Knolls—contains a large amount of whole rock geochemical data, from which the average Al2O3 content of Smelter Knolls rhyolite was used as an internal standard for our melt inclusion analyses.

Amalia Tuff

The Questa caldera and the cogenetic volcanic and intrusive rocks of the Latir volcanic field in northern New Mexico, USA, were extensively studied by Lipman (1988), Johnson and Lipman (1988), Johnson et al. (1989), and Czamanske et al. (1990). Most of the Latir volcanic units and associated intrusive rocks were emplaced within a few million years (28–26 Ma), including the eruption of the only major ash-flow tuff—the Amalia Tuff—26.5 Ma ago (sanidine K–Ar; Lipman et al. 1986). This peralkaline, rhyolitic tuff can be subdivided into two subunits, and the upper sequence—according to similarities in composition and mineralogy—is thought to represent the erupted portion of resurgent peralkaline intrusions of the Questa caldera. Although plutons at exposed levels are texturally discrete and of contrasting composition, regional gravity data (Cordell et al. 1985) suggest that the entire Questa caldera is underlain by cogenetic batholithic rocks of 10 by 20 km size at shallow depth. The Amalia Tuff sample investigated in the present study was collected from an outflow sheet 40 km SW of the caldera rim near the town of Tres Piedras.

Banco Bonito vitrophyre

The Banco Bonito Flow forms part of the East Fork Member of the Valles Rhyolite in New Mexico, USA. It represents the youngest volcanic unit of the Valles caldera complex that was active between 45 ka and 35 ka (ESR ages, 21Ne exposure ages, regional constraints; (Goff and Gardner 2004; Ogoh et al. 1993; Phillips et al. 1997). Detailed descriptions of the mineralogy, petrography and geochemistry of the East Fork Member can be found in Eichler (2012). Based on the trace element geochemistry of the phenocrysts and the volcanic glass, the East Fork Member rhyolites had a complex evolution history, starting with fractional crystallization of a basalt, followed by magma ascent to lower crustal levels and crustal assimilation, and finally fractional crystallization in a granitic magma chamber in the upper crust.

Santa Rita rhyodacite dike

The porphyry-copper deposit at Santa Rita (Chino Mine) in southwestern New Mexico formed during the Laramide orogeny (45–75 Ma) as a result of subduction along an Andean-type continental margin. There were several stages of the igneous activity in the region, starting with dioritic to quartz–dioritic sills intruded into the Precambrian basement rocks, followed by the eruption of basaltic–andesitic to andesitic magma and the formation of mafic dikes, and subsequently the intrusion of granodioritic to quartz–monzodioritic magma. The last stage of magmatic activity is represented by dikes of rhyodacitic to rhyolitic composition, which cut across all other lithologies. Details on the geology and petrography of the complex can be found in Rose and Baltosser (1966) and Jones et al. (1967), whereas a more recent study (Audétat and Pettke 2006) focuses on the chemical analysis of mineral and melt inclusions. We investigated two samples (SR15, SR9) from two rhyodacite dikes.

The Dyke

The West Elk laccolite cluster occupies the northern part of the West Elk Mountains, northwestern Gunnison County, Colorado. The laccoliths are located along a dike swarm related to a NNE-SSW trending, at least 40 km long fracture zone (Godwin and Gaskill 1964). The mafic, early stage of each stock is crosscut by SiO2-rich dikes or stock-internal granodiorites (Mutschler et al. 1981). “The Dyke” represents this second stage and forms a prominent outcrop of ca. 2.5 km length south of Ruby Peak. There are no age data available from The Dyke, but considering that it intrudes the Early Eocene Wasatch formation and the genetically related Crested Butte laccolith and Paradise stock, which yielded biotite K–Ar ages of 29.0 ± 1.1 and 29.1 ± 1.0 Ma, respectively (Obradovich et al. 1969), it is probable that The Dyke granite formed in the Oligocene.

Nomlaki Tuff

This sample was kindly provided by the Smithsonian Institution (NMNH 116377-38a). The Nomlaki Tuff (4.2–3.6 Ma) formed by a Plinian eruption in the Pliocene and covers areas in California, Nevada, Utah, Arizona and New Mexico. The deposits consist of widespread ash-fall and proximal ash-flow units and are commonly used as marker horizons in the region (Knott and Sarna-Wojcicki 2001). Information about the distribution and regional correlation of the Late Cenozoic tuffs of the Central Coast Ranges of California can be found in Sarna-Wojcicki et al. (1984), in Knott and Sarna-Wojcicki (2001) and in Poletski (2010). Whole rock geochemical data can be found in Knott and Sarna-Wojcicki (2001), whereas an extensive dataset regarding the geochemistry of glass fragments and pumices, as well as regarding mgt–ilm thermometry results—both from the occurrences in the Sacramento Valley—are published in Poletski (2010). Poletski (2010) distinguished two chemo-types within the tuff based on the major element composition of glass shards and magnetite–ilmenite microphenocryst thermometry results, and explained the phenomenon by a zoned magma chamber, containing a more evolved and cooler magma in its upper part, and a less evolved, hotter magma at its bottom.

Kos granite enclave

The geochemical evolution of the Kos Plateau Tuff and related magma chamber processes have been exceptionally well studied. Petrographic observations (Keller 1969), sanidine Ar–Ar dating (Smith et al. 1996), zircon U–Pb dating (Bachmann et al. 2007), melt inclusion analysis (Bachmann et al. 2009), and a recent comprehensive study on the Kos Plateau Tuff (Bachmann 2010) all contributed to a comprehensive understanding of the region’s magmatism.

The Kos Plateau Tuff formed 160,000 years ago (Smith et al. 1996), and represents one of the largest Quaternary explosive eruptions. The non-welded tuff consists mainly of juvenile ash, different types of pumice, and lithic fragments, and locally contains equigranular granitic enclaves. The rhyolitic magma most probably evolved from a more mafic parent dominantly by fractional crystallization, as shown by the lack of inherited zircons (Bachmann et al. 2007). The magma was probably in a crystal mush state before the eruption, with some completely crystallized units at the edge of the magma body having been entrained in the form of granitic enclaves. The system was partly reheated by injection of a more mafic magma batch, which triggered the eruption and resulted in the formation of andesitic bands within pumice, plagioclase overgrowths on K-feldspar, and inverse zonations in plagioclases (Bachmann 2010).

Los Humeros

This sample was obtained from the Smithsonian Institution (NMNH 117450-58). The Los Humeros volcanic center of Pleistocene age (K–Ar) is located 180 km east of Mexico City and represents the easternmost expression of the late Tertiary to Quaternary Mexican Neovolcanic Belt (Ferriz and Mahood 1984). The volcanic province can be characterized by a series of lava flows, large eruptions (resulting in large scale ignimbrite deposition) and subsequent caldera collapse and dome-forming events. The most common compositions are rhyolitic and rhyodacitic, however, andesites and basalts (the latter mainly at the last stage of volcanism) occur as well. A detailed description of the volcanics in the region can be found in Ferriz and Mahood (1984). Our sample was obtained from the Smithsonian Institution (NMNH 117450-58).

Samples from the Mono-Inyo volcanic field

The Long Valley Volcanic Field is located at the intersection of the Sierra Nevada and the Basin and Range tectonic province in east-central California. Volcanism in the region started about 4 million years ago (Gilbert et al. 1968) and continued in multiple phases until recently. It can be separated into a pre-caldera and a post-caldera episode, the first of which peaked at 0.76 ka with the eruption of the Bishop Tuff, followed by a caldera collapse and the formation of multiple craters along the N–S trending Mono-Inyo fissure system (Bailey 2004). The composition of magmas erupted from the Mono-Inyo crater system ranges from basaltic to rhyolitic, with a compositional gap between trachyandesite and dacite, and includes several high-silica varieties containing andesitic inclusions. These andesitic inclusions lie on the mixing line between finely porphyritic rhyolites and typical post-caldera mafic magmas. Their xenocryst assemblage is identical to the phenocryst assemblage of the host rhyolites, which provides strong evidence for the involvement of mafic magma and magma mixing during silicic eruptions (Varga et al. 1990). Extensive geochemical datasets, detailed geological observations and interpretation of the complex geological evolution of the region can be found in Bailey (2004) and in Bray (2014).

Tunnel Spring Tuff

The Tunnel Spring Tuff was erupted 35.4 million years ago (K–Ar age) from a vent that probably was close to Crystal Peak, Utah, where the tuff forms a canyon fill of more than 400 m thickness (Steven 1989). The P–T–x history of the parental magma was investigated by Audétat (2013) based on zircon saturation thermometry of melt inclusions and Ti-in-quartz thermobarometry.

Blackfoot lava field vitrophyre

The Blackfoot lava field is located in southeastern Idaho, USA. It is characterized by bimodal volcanism, consisting of five rhyolitic domes located in a predominantly basaltic volcanic field. The rhyolites, which belong to the group of Cenozoic Topaz Rhyolites (Christiansen et al. 1986), contain inclusions of older basalts and andesites, but they are older than the basalts exposed on the surface. Our vitrophyre sample was kindly provided by the Smithsonian Institution (NMNH 117462-1) and was collected on the northern side of a rhyolite dome named China Hat (also referred to as China Cap or Middle Cone), the age of which was determined at 61 ± 6 ka (sanidine K–Ar; Pierce et al. 1982). The geology of the lava field, including whole rock data, is described in Christiansen et al. (1986).

Pine Grove Tuff

The Pine Grove intrusions and associated tuffs are located in the southern Wah Wah Mountains, southern Utah, USA. Whole rock and K-feldspar K–Ar data suggests that the Pine Grove system, which consists of rhyolitic (and partly dacitic) tuffs, and intrusions ranging from rhyolitic to mafic compositions, formed 23 to 22 m.y. ago (Keith et al. 1986). According to Keith et al. (1986) the tuffs were erupted from a magma chamber that was intruded by a trachyandesitic magma multiple times, and was compositionally zoned from dacite to rhyolite. Detailed petrography and K–Ar age data of the Pine Grove system can be found in Keith et al. (1986), an extensive dataset of melt inclusion analyses was published by Lowenstern et al. (1994), whereas thermometry and oxybarometry data can be found in Keith and Shanks (1988) and Audétat et al. (2011). Our sample was collected from the rhyolitic air-fall unit.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Arató, R., Audétat, A. Vanadium magnetite–melt oxybarometry of natural, silicic magmas: a comparison of various oxybarometers and thermometers. Contrib Mineral Petrol 172, 52 (2017). https://doi.org/10.1007/s00410-017-1369-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00410-017-1369-6

Keywords

Navigation