Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-23T16:44:13.983Z Has data issue: false hasContentIssue false

Status of thermobarometry in granitic batholiths

Published online by Cambridge University Press:  03 November 2011

J. Lawford Anderson
Affiliation:
J. Lawford Anderson, Department of Earth Sciences, University of Southern California, Los Angeles, CA 90089-0740, U.S.A.

Abstract:

Most granitic batholiths contain plutons which are composed of low-variance mineral assemblages amenable to quantification of the P conditions that characterise emplacement. Some mineral thermometers, such as those based on two feldspars or two Fe–Ti oxides, commonly undergo subsolidus re-equilibration. Others are more robust, including hornblende–plagioclase, hornblende–clinopyroxene, pyroxene–ilmenite, pyroxene–biotite, garnet–hornblende, muscovite-biotite and garnet–biotite. The quality of their calibration is variable and a major challenge resides in the large range of liquidus to solidus crystallisation temperatures that are incompletely preserved in mineral profiles. Further, the addition of components that affect Kd relations between non-ideal solutions remains inadequately understood. Estimation of solidus and near-solidus conditions derived from exchange thermometry often yield results >700°C and above that expected for crystallisation in the presence of an H2O-rich volatile phase. These results suggest that the assumption of crystallisation on an H2O-saturated solidus may not be an accurate characterisation of some granitic rocks.

Vapour undersaturation and volatile phase composition dramatically affect solidus temperatures. Equilibria including hypersthene–biotite–sanidine–quartz, fayalite–sanidine–biotite, and annite–sanidine–magnetite (ASM) allow estimation of Estimates by the latter assemblage, however, are highly dependent on . Oxygen fugacity varies widely (from two or more log units below the QFM buffer to a few log units below the HM buffer) and can have a strong affect on mafic phase composition. Ilmenite–magnetite, quartz–ulvospinel–ilmenite–fayalite (QUILF), annite–sanidine–magnetite, biotite–almandine–muscovite–magnetite (BAMM), and titanite–magnetite–quartz (TMQ) are equilibria providing a basis for the calculation of .

Granite barometry plays a critical part in constraining tectonic history. Metaluminous granites offer a range of barometers including ferrosilite–fayalite–quartz, garnet–plagioclase–hornblende–quartz and Al-in-hornblende. The latter barometer remains at the developmental stage, but has potential when the effects of temperature are considered. Likewise, peraluminous granites often contain mineral assemblages that enable pressure determinations, including garnet–biotite–muscovite–plagioclase and muscovite–biotite–alkali–feldspar–quartz. Limiting pressures can be obtained from the presence of magmatic epidote and, for low-Ca pegmatites or aplites, the presence of subsolvus versus hypersolvus alkali feldspars.

As with all barometers, the influence of temperature, , and choice of activity model are critical factors. Foremost is the fact that batholiths are not static features. Mineral compositions imperfectly record conditions acquired during ascent and over a range of temperature and pressure and great care must be taken in properly quantifying intensive parameters.

Type
Research Article
Copyright
Copyright © Royal Society of Edinburgh 1996

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

Ague, J. J.&Brimhall, G. H. 1988. Magmatic arc asymmetry and distribution of anomalous plutonic belts in the batholiths of California: effects of assimilation, crustal thickness, and depth of crystallization. BULL GEOL SOC AM 100, 912–27.2.3.CO;2>CrossRefGoogle Scholar
Andersen, D.J., Bishop, F. C.&Lindsley, D. H. 1991. Internally consistent solution models for Fe–Mg–Mn–Ti oxides: Fe–Mg–Ti oxides and olivine. AM MINERAL 76, 427–44.Google Scholar
Andersen, D.J., Lindsley, D. H.&Davidson, P. M. 1993. QUILF: a Pascal program to assess equilibria among Fe–Mg–Mn–Ti oxides, pyroxenes, olivine, and quartz. COMPUT GEOSCI 19, 1333–50.CrossRefGoogle Scholar
Andersen, D. J.&Lindsley, D. H. 1988. Internally consistent solution models for Fe–Mg–Mn–Ti oxides: Fe–Ti oxides. AM MINERAL 73, 714–26.Google Scholar
Anderson, J. L. 1980. Mineral equilibria, and crystallization conditions in the late Precambrian Wolf River rapakivj massif, Wisconsin. AM J SCIENCE 280, 289332.CrossRefGoogle Scholar
Anderson, J. L. 1983. Proterozoic anorogenic granite plutonism of North America. MEM GEOL SOC AM 161, 133–52.Google Scholar
Anderson, J. L. 1988. Core complexes of the Mojave-Sonoran Desert: conditions of plutonism, mylonitization, and decompression. In Ernst, W. G. (ed.) Metamorphism and crustal evolution of the western United States, Rubey Volume VII, 503–25. Englewood Cliffs: Prentice Hall.Google Scholar
Anderson, J. L. 1992. Compositional variation within the high-Mg, tonalitic Mount Stuart batholith, north Cascades, Washington. GEOL SOC AM ABSTR PROGRAM 24, 3.Google Scholar
Anderson, J. L.&Bender, E. E. 1989. Nature and origin of Proterozoic A-type granitic magmatism in the southwestern United States. LITHOS 23, 1952.CrossRefGoogle Scholar
Anderson, J. L.&Morrison, J. 1992. The role of anorogenic granites in the Proterozoic crustal development of North America. In Condie, K. C. (ed.) Proterozoic crustal evolution, 263–99. New York: Elsevier.CrossRefGoogle Scholar
Anderson, J. L.&Paterson, S. R. 1991. Emplacement of the Cretaceous Mt. Stuart Batholith, central Cascades, Washington. GEOL SOC AM ABSTR PROGRAM 23, A387.Google Scholar
Anderson, J. L.&Rowley, M. C. 1981. Synkinematic intrusion of two-mica and associated metaluminous granitoids. Whipple Mountains, California. CAN MINERAL 19, 83101.Google Scholar
Anderson, J. L.&Smith, D. R. 1995. The effect of temperature and oxygen fugacity on Al-in-hornblende barometry. AM MINERAL 80, 549–59.CrossRefGoogle Scholar
Anderson, J. L.&Thomas, W. M. 1985. Proterozoic anorogenic two-mica granites: Silver Plume and St. Vrain batholiths of Colorado. GEOLOGY 13, 177–80.2.0.CO;2>CrossRefGoogle Scholar
Anderson, J.L., Barth, A. P.&Young, E. D. 1988. Mid-crustal roots of Cordilleran metamorphic core complexes. GEOLOGY 16, 366–9.2.3.CO;2>CrossRefGoogle Scholar
Anderson, J.L., Barth, A. P.. Young, E. D., Davis, M. J., Farber, D., Hayes, E. M.&Johnson, K. A. 1992. Plutonism across the Tujunga–North American terrane boundary: a middle to upper crustal view of two juxtaposed arcs. In Bartholomew, M. J., Hyndman, D. W., Mogk, D. W.&Mason, R., (eds) Characterization and comparison of ancient and Mesozoic continental margins, 205–30. Dordrecht: Kluwer Academic.Google Scholar
Anovitz, L. M.&Essene, E. J. 1987. Compatibility of geobarometers in the system CaO–FeO–Al2O3–SiO2–TiO2 (CFAST); implications for garnet mixing models. J GEOL 95, 635–45.CrossRefGoogle Scholar
Barth, A. P. 1989. Mesozoic rock units in the upper plate of the Vincent thrust fault, San Gabriel Mountains, southern California. Ph.D. Thesis. University of Southern California.Google Scholar
Barth, A. P. 1990. Mid-crustal emplacement of Mesozoic plutons, San Gabriel Mountains. California, and implications for the geologic history of the San Gabriel Terrane. In Anderson, J. L. (ed.) The nature and origin of cordilleran magmatism. GEOL SOC AM MEM. 174, 3345.Google Scholar
Barth, A. P.. Wooden, J. L., Tosdal, R. M.&Morrison, J. 1995. Crustal contamination in the petrogenesis of a calc-alkalic rock series: Josephine Mountain intrusion, California. BULL GEOL SOC AM 107, 201–12.2.3.CO;2>CrossRefGoogle Scholar
Beckerman, G. M., Robinson, J. P.&Anderson, J. L. 1982. The Teutonia batholith: a large intrusive complex of Jurassic and Cretaceous age in the eastern Mojave Desert, California. In Frost, E. G.&Martin, D. L. (eds) Mesozoic–Cenozoic tectonic evolution of the Colorado River region, California, Arizona, and Nevada, 205221. San Diego: Cordilleran Publishers.Google Scholar
Berman, R. G. 1988. Internally consistent thermodynamic data for minerals in the system Na2O–K2O–CaO–MgO–FeO–Fe2O3–Al2O3–SiO2–TiO2–H2O–CO2. J PETROL 29, 445522.CrossRefGoogle Scholar
Berman, R. G. 1990. Mixing properties of Ca–Mg–Fe–Mn garnets. AM MINERAL 75, 328–44.Google Scholar
Bishop, F. C. 1980. The distribution of Fe2+ and Mg between coexisting ilmenite and pyroxene with applications to geothermometry. AM J SCI 280, 4677.CrossRefGoogle Scholar
Blundy, J. D.&Holland, T. J. B. 1990. Calcic amphibole equilibria and a new amphibole–plagioclase geothermometer. CONTRIB MINERAL PETROL 104, 208–24.CrossRefGoogle Scholar
Bohlen, S. R.&Boettcher, A. L. 1981. Experimental investigation and geological application of orthopyroxene barometry. AM MINERAL 66, 951–64.Google Scholar
Bohlen, S. R.&Lindsley, D. H. 1987. Thermometry and barometry of igneous and metamorphic rocks. ANNU REV EARTH PLANET SCI 15, 397420.CrossRefGoogle Scholar
Bohlen, S. R., Essene, E. J.&Boettcher, A. L. 1980. Reinvestigation and application of olivine–quartz–orthopyroxene barometry. EARTH PLANET SCI LETT 47, 110.CrossRefGoogle Scholar
Bohlen, S. R., Boettcher, A. L., Wall, V. J.&Clemens, J. D. 1983. Stability of phlogopite–quartz and sanidine–quartz: a model for melting in the lower crust. CONTRIB MINERAL PETROL 83, 270–7.CrossRefGoogle Scholar
Brown, W. L.&Parsons, I. 1981. Towards a more practical two-feldspar thermometer. AM MINERAL 70, 356–61.Google Scholar
Buddington, A. L.&Lindsley, D. H. 1964. Iron–titanium oxide minerals and synthetic equivalents. J PETROL 5, 310–57.CrossRefGoogle Scholar
Carmichael, I. S. E. 1991. The redox states of basic and silicic magmas: a reflection of their source regions? CONTRIB MINERAL PETROL 106, 129–41.CrossRefGoogle Scholar
Chatterjee, N. D.&Johannes, W. 1974. Thermal stability and standard thermodynamic properties of synthetic 2Ml muscovite. KAl2AlSi3O10(OH)2. CONTRIB MINERAL PETROL 49, 89114.CrossRefGoogle Scholar
Chipera, S. J.&Perkins, D. 1988. Evaluation of biotite–garnet geothermometers: application to the English River subprovince. Ontario. CONTRIB MINERAL PETROL 98, 40–8.CrossRefGoogle Scholar
Chou, I.-M. 1978. Calibration of oxygen buffers at elevated P and T using the hydrogen fugacity sensor. AM MINERAL 63, 690703.Google Scholar
Clemens, J. D. 1995. Phlogopite stability in the silica-saturated portion of the system KAlO2-MgO-SiO2-H2O: new data and a reappraisal of phase relations to 1·5 GPa. AM MINERAL 80, 982–97.CrossRefGoogle Scholar
Cosca, M. A., Essene, E. J.&Bowman, J. R. 1991. Complete chemical analyses of metamorphic hornblendes: implications for normalizations, calculated H2O activities, and thermobarometry. CONTRIB MINERAL PETROL 108, 472–84.CrossRefGoogle Scholar
Cotkin, S. J.&Medaris, L. G. 1993. Evaluation of the crystallization conditions for the calc-alkaline Russian Peak intrusive complex. Klamath Mountains, Northern California. J PETROL 34, 543–71.CrossRefGoogle Scholar
Cullers, R. L., Griffin, T., Bickford, M. E.&Anderson, J. L. 1992. Origin and chemical evolution of the 1360 Ma-old San Isabel batholith, Wet Mountains, Colorado U.S.A.: A mid-crustal granite of anorogenic affinities. BULL GEOL SOC AMER 104, 316–28.2.3.CO;2>CrossRefGoogle Scholar
Cullers, R. L., Stone, J., Anderson, J. L., Sassarini, N.&Bickford, M. E. 1993. Petrogenesis of Mesoproterozoic Oak Creek and West McCoy Gulch plutons, Colorado: an example of cumulate unmixing of mid-crustal, two mica granite of anorogenic affinity. PRECAMBRIAN RES 62, 139–69.CrossRefGoogle Scholar
Cygan, G. L., Chou, I.-Ming&Sherman, D. M. 1991. Reassessment of the annite breakdown reaction using new hydrothermal experimental techniques. EOS, TRANS AM GEOPHYS UNION 72, 313.Google Scholar
Czamanske, G. K., Ishihara, S.&Atkin, S. A. 1981. Chemistry of rock-forming minerals of the Cretaceous–Paleocene batholith in southwestern Japan and implications for magma genesis. J GEOPHYS RES 86, 10431–69.CrossRefGoogle Scholar
Dachs, E. 1994. Annite stability revised. 1. Hydrogen-sensor data for the reaction annite = sanidine + magnetite + H2. CONTRIB MINERAL PETROL 117, 229–40.CrossRefGoogle Scholar
Dallmeyer, R. D. 1974. The role of crystal structure in controlling the partitioning of Mg and Fe2+ between coexisting garnet and biotite. AM MINERAL 59, 201–3.Google Scholar
Dasgupta, S., Sengupta, P., Guha, D.&Fukuoka, M. 1991. A refined garnet–biotite Fe–Mg exchange thermometer and its application in amphibolites and granulites. CONTRIB MINERAL PETROL 109, 130–7.CrossRefGoogle Scholar
Davidson, P. M.&Lindsley, D. H. 1985. Thermodynamic analysis of quadrilateral pyroxene. Part II. Model calibration from experiments and applications to geothermometry. CONTRIB MINERAL PETROL 91, 390404.CrossRefGoogle Scholar
Davidson, P. M.&Lindsley, D. H. 1989. Thermodynamic analysis of pyroxene–olivine–quartz equilibria in the system CaO–MgO–FeO–SiO2. AM MINERAL 74, 1830.Google Scholar
Davidson, P. M.&Mukhopadhyay, D. K. 1984. Ca–Fe–Mg olivines: phase relations and a solution model. CONTRIB MINERAL PETROL 86, 256–63.CrossRefGoogle Scholar
Dawes, R. L.&Evans, B. W. 1991. Mineralogy and geothermobarometry of magmatic epidote-bearing dikes, Front Range, Colorado. BULL GEOL SOC AM 103, 1017–31.2.3.CO;2>CrossRefGoogle Scholar
Elkins, L. T.&Grove, T. L. 1990. Ternary feldspar experiments and thermodynamic models. AM MINERAL 75, 544–59.Google Scholar
Ellis, D. J.&Green, D. H. 1979. An experimental study of the effect of Ca upon garnet–clinopyroxene Fe–Mg exchange equilibria. CONTRIB MINERAL PETROL 71, 1322.CrossRefGoogle Scholar
Emslie, R. F.&Stirling, J. A. R. 1993. Rapakivi and related granitoids of the Nain plutonic suite: geochemistry, mineral assemblages, and fluid equilibria. CAN MINERAL 31, 821–47.Google Scholar
Essene, E. J. 1982. Geologic thermometry and barometry. In Ferry, J. M. (ed.) Characterization of metamorphism through mineral equilibria. MINERAL SOC AM REV MINERAL 10, 153206.Google Scholar
Essene, E. J. 1989. The current status of thermobarometry in metamorphic rocks. In Daly, J. S., Cliff, R. A.&Yardley, B. W. D. (eds) Evolution of metamorphic belts. SPEC PUBL GEOL SOC LONDON 43, 144.CrossRefGoogle Scholar
Ewart, A., Hildreth, W.&Carmichael, I. S. E. 1975. Quaternary acid magmas in New Zealand. CONTRIB MINERAL PETROL 51, 127.CrossRefGoogle Scholar
Ferry, J. M.&Spear, F. S. 1978. Experimental calibration of the partitioning of Fe and Mg between biotite and garnet. CONTRIB MINERAL PETROL 66, 113–7.CrossRefGoogle Scholar
Fonarev, V. I.&Konilov, A. N. 1986. Experimental study of Fe–Mg distribution between biotite and orthopyroxene at P = 490MPa. CONTRIB MINERAL PETROL 93, 227–35.CrossRefGoogle Scholar
Frost, B. R. 1991. Introduction to oxygen fugacity and its petrologic importance. In Lindsley, D. H. (ed.) Oxide minerals. MINERAL SOC AM REV MINERAL 25, 19.Google Scholar
Frost, B. R.&Lindsley, D. H. 1992. Equilibria among Fe–Ti oxides, pyroxenes, olivine, and quartz: part II. Application. AM MINERAL 77, 1004–20.Google Scholar
Frost, B. R., Lindsley, D. H.&Andersen, D. J. 1988. Fe–Ti oxidesilicate equilibria: assemblages with fayalitic olivine. AM MINERAL 73, 727–40.Google Scholar
Fuhrman, M. L.&Lindsley, D. H. 1988. Ternary feldspar modeling and themometry. AM MINERAL 73, 201–15.Google Scholar
Fuhrman, M. L., Frost, B. R.&Lindsley, D. H. 1988. Crystallization conditions of the Sybille monzosyenite, Laramie anorthosite complex, Wyoming. J PETROL 29, 699729.CrossRefGoogle Scholar
Ganguly, J.&Saxena, S. 1984. Mixing properties of aluminosilicate garnets: constraints from natural and experimental data, and applications to geothermo-barometry. AM MINERAL 69, 8897.Google Scholar
Ghent, E. D.&Stout, M. Z. 1981. Geobarometry and geothermometry of plagioclase–biotite–garnet–muscovite assemblages. CONTRIB MINERAL PETROL 76, 92–7.CrossRefGoogle Scholar
Ghiorso, M. S. 1990. Thermodynamic properties of hematite–ilmenite–geikielite solid solutions. CONTRIB MINERAL PETROL 104, 645–67.CrossRefGoogle Scholar
Ghiorso, M. S.&Sack, R. O. 1991. Fe–Ti oxide thermometry: thermodynamic formulation and estimation of intensive variables in silicic magmas. CONTRIB MINERAL PETROL 108, 485510.CrossRefGoogle Scholar
Gordon, T. M. 1992. Generalized thermobarometry; solution of the inverse chemical equilibrium problem using data for individual species. GEOCHIM COSMOCHIM ACTA 56, 1793–800.CrossRefGoogle Scholar
Graham, C. M.&Powell, R. 1984. A garnet–hornblende geothermometer: calibration, testing, and application to the Pelona schist, southern California: J METAMORPH GEOL 2, 1331.CrossRefGoogle Scholar
Green, N. L.&Usdansky, S. I. 1986. Ternary-feldspar mixing relations and thermobarometry. AM MINERAL 71, 1100–8.Google Scholar
Green, T. H.&Adam, J. 1991. Assessment of the garnet–clinopyroxene Fe–Mg exchange thermometer using new experimental data. J METAMORPH PETROL 9, 341–7.CrossRefGoogle Scholar
Hammarstrom, J. M.&Zen, E-an. 1986. Aluminum in hornblende, an empirical igneous geobarometer. AM MINERAL 71, 1297–313.Google Scholar
Harley, S. L. 1984. Comparison of the garnet-orthopyroxene geobarometer with recent experimental studies and applications to natural assemblages. J PETROL 25, 697712.CrossRefGoogle Scholar
Harrison, T. M.&Watson, E. B. 1984. The behavior of apatite during crustal anatexis: equilibrium and kinetic considerations. GEOCHIM COSMOCHIM ACTA 48, 1467–77.CrossRefGoogle Scholar
Haselton, H. T., Hovis, G. L., Hemingway, B. S.&Robie, R. A. 1983. Calorimetric investigation of the excess entropy of mixing in analbite–sanidine solid solutions: lack of evidence for Na, K short-range order and implications for two-feldspar thermometry. AM MINERAL 68, 398413.Google Scholar
Hayes, E. M. 1992. Petrology of Jurassic plutons and older crystalline units, the Cargo Muchacho Mountains, southeastern California. M.S. Thesis, University of Southern California.Google Scholar
Helgeson, H. C, Delaney, J. M., Nesbitt, H. W.&Bird, D. K. 1978. Summary and critique of the thermodynamic properties of rockforming minerals. AM J SCI 278A, 1229.Google Scholar
Hodges, K. V.&Crowley, P. D. 1985. Error estimation and empirical geothermobarometry for pelitic systems. AM MINERAL 70, 702–9.Google Scholar
Hodges, K. V.&McKenna, L. W. 1987. Realistic propagation of uncertainties in geologic thermobarometry. AM MINERAL 72, 671–80.Google Scholar
Hodges, K. V.&Spear, F. S. 1982. Geothermometry, geobarometry, and the Al2SiO5 triple point at Mt. Moosilauke, New Hampshire. AM MINERAL 67, 1118–34.Google Scholar
Hoisch, T. D. 1991. Equilibria with the mineral assemblage quartz = muscovite + biotite + garnet + plagioclase. CONTRIB MINERAL PETROL 108, 4354.CrossRefGoogle Scholar
Holland, T.&Blundy, J. 1994. Non-ideal interactions in calcic amphiboles and their bearing on amphibole–plagioclase thermometry. CONTRIB MINERAL PETROL 116, 433–47.CrossRefGoogle Scholar
Hollister, L. S., Grissom, G. C, Peters, E. K., Stowell, H. H.&Sisson, V. B. 1987. Confirmation of the empirical correlation of Al in hornblende with pressure of solidification of calc-alkaline plutons. AM MINERAL 72, 231–9.Google Scholar
Indares, A.&Martignole, J. 1985. Biotite–garnet geothermometry in the granulite facies: the influence of Ti and Al in biotite. AM MINERAL 70, 272–8.Google Scholar
Ishihara, S. 1977. The magnetite-series and ilmenite-series granitic rocks. MIN GEOL 27, 293305.Google Scholar
Johnson, M. C.&Rutherford, M. J. 1989. Experimental calibration of an aluminum-in-hornblende geobarometer with application to Long Valley caldera (California) volcanic rocks. GEOLOGY 17, 837–41.2.3.CO;2>CrossRefGoogle Scholar
Kohn, M. J.&Spear, F. S. 1989. Empirical calibration of geobarometers for the assemblage garnet + hornblende + plagioclase + quartz. AM MINEAL 74, 7784.Google Scholar
Kohn, M. J.&Spear, F. S. 1990. Two new geobarometers for garnet amphibolites, with applications to southeastern Vermont. AM MINERAL 75, 8996.Google Scholar
Kohn, M. J.&Spear, F. S. 1991a. Error propagation for barometers: 1. Accuracy and precision of experimentally located end-member reactions. AM MINERAL 76, 128–37.Google Scholar
Kohn, M. J.&Spear, F. S. 1991b. Error propagation for barometers: 2. Application to rocks. AM MINERAL 76, 138–47.Google Scholar
Koziol, A. M. 1990. Activity-composition relationships of binary Ca–Fe and Ca–Mn garnets determined by reversed, displaced equilibrium experiments. AM MINERAL 75, 319–27.Google Scholar
Koziol, A. M.&Newton, R. C. 1988. Redetermination of the anorthite breakdown reaction and improvement of the plagioclase–garnet–Al2SiO5–quartz geobarometer. AM MINERAL 73, 216–23.Google Scholar
Kretz, R. 1982. Transfer and exchange equilibria in a portion of the pyroxene quadrilateral as deduced from natural and experimental data. GEOCHIM COSMOCHIM. ACTA 46, 411–22.CrossRefGoogle Scholar
Kretz, R.&Jen, L. S. 1978. Effect of temperature on the distribution of Mg and Fe2+ between calcic pyroxene and hornblende. CAN MINERAL 16, 533–7.Google Scholar
Kullerud, K. 1995. Chlorine, titanium, and barium rich biotites: factors controlling biotite composition and implications for garnet–biotite geothermometry. CONTRIB MINERAL PETROL 120, 4259.CrossRefGoogle Scholar
Lee, H.&Ganguly, J. 1987. Equilibrium compositions of coexisting garnet and orthopyroxene: experimental determinations in the system FeO–MgO–Al2O3–SiO2 and applications. J PETROL 29, 93114.CrossRefGoogle Scholar
Lindsley, D. H. 1983. Pyroxene thermometry. AM MINERAL 68, 477–93.Google Scholar
Lindsley, D. H.&Frost, B. R. 1992. Equilibria among Fe–Ti oxides, pyroxenes, olivine, and quartz: part I. Theory. AM MINERAL 77, 9871003.Google Scholar
Lindsley, D. H.&Spencer, K. J. 1982. Fe–Ti oxide geothermometry: reducing analyses of coexisting Ti–magnetite (Mt) and ilmenite (Ilm). EOS, TRANS AM GEOPHYS UNION 63, 471.Google Scholar
Lindsley, D. H., Frost, B. R., Andersen, D. J.&Davidson, P. J. 1990. Fe–Ti oxide equilibria: assemblages with orthopyroxene. In Spencer, R. J.&Chou, I.-M (eds) Fluid-mineral interactions. SPEC PUBL GEOCHEM SOC 2, 103–19.Google Scholar
Luhr, J. F., Carmichael, I. S. E.&Varekamp, J. C. 1984. The 1982 eruptions of El Chichon volcano, Chiapas, Mexico: mineralogy and petrology of the anhydrite-bearing pumices. J VOLCANOL GEOTHERM RES 23, 69108.CrossRefGoogle Scholar
Ludington, S. 1978. The biotite–apatite geothermometer revisited. AM MINERAL 63, 551–3.Google Scholar
Massonne, H. J.&Schreyer, W. 1987. Phengite geobarometry based on the limiting assemblage with K–feldspar, phlogopite, and quartz. CONTRIB MINERAL PETROL 96, 212–24.CrossRefGoogle Scholar
Mayo, D. P. 1994. Estimating crystallization conditions in metaluminous granodiorite using microtextures and equilibria involving mafic silicates and oxides. GEOL SOC AM ABSTR PROGRAM 26, 71.Google Scholar
McKenna, L. W.&Hodges, K. V. 1988. Accuracy versus precision in locating reaction boundaries: implications for the garnet–plagioclase–aluminum silicate–quartz geobarometer. AM MINERAL 73, 1205–8.Google Scholar
Miller, C. F., Stoddard, E. F., Bradfish, L. J.&Dollase, W. A. 1981. Composition of plutonic muscovite: genetic implications. CAN MINERAL 19, 2534.Google Scholar
Munoz, J. L.&Ludington, S. D. 1974. Fluoride–hydroxyl exchange in biotite. AM MINERAL 274, 396413.Google Scholar
Munoz, J. L.&Swenson, A. 1981. Chloride–hydroxyl exchange in biotite and estimation of relative HCL/HF activities in hydrothermal fluids. ECON GEOL 76, 2212–21.CrossRefGoogle Scholar
Myers, J.&Eugster, H. P. 1983. The system Fe–Si–O: oxygen buffer calibrations to 1500 °K. CONTRIB MINERAL PETROL 82, 7590.CrossRefGoogle Scholar
Nabelek, C. R.&Lindsley, D. H. 1985. Tetrahedral Al in amphibole: a potential thermometer for some mafic rocks. GEOL SOC AM ABSTR PROGRAM 17, 673.Google Scholar
Newton, R. C.&Haselton, H. T. 1981. Thermodynamics of the garnet–plagioclase–Al2SiO5–quartz geobarometer. In Newton, R. C. (ed.) Thermodynamics of minerals and melts, 131–47. New York: Springer Verlag.CrossRefGoogle Scholar
O'Neil, J. R. 1986. Theoretical and experimental aspects of isotopic fractionation. In Valley, J. W., Taylor, H. P.&O'Neil, J. R. (eds) Stable isotopes in high-temperature geologic processes. MINERAL SOC AM REV MINERAL 167, 140.Google Scholar
Paterson, S. R., Miller, R. B., Anderson, J. L., Lund, S., Bendixen, J., Taylor, N.&Fink, T. 1994. Emplacement and evolution of the Mt. Stuart batholith. In Swanson, D. A.&Haugerud, R. A. (ed.) Geologic field trips of the Pacific Northwest, Vol. 2. GEOL SOC AM GUIDE 2F1-47.Google Scholar
Pease, V., Foster, D., Wooden, J., O'Sullivan, P.&Argent, J. 1995. Tertiary plutonism and extension in the Sacramento Mountains, SE California U.S.A. EOS, TRANS AM GEOPHYS UNION 76, 639.Google Scholar
Perkins, D.&Vielzeuf, D. 1992. Experimental investigation of Fe–Mg distribution between olivine and clinopyroxene: implications for mixing properties of Fe–Mg in clinopyroxene and garnet–clinopyroxene thermometry. AM MINERAL 77, 774–83.Google Scholar
Perkins, D., Essene, E. J.&Wall, V. J. 1987. THERMO: a computer program for calculation of mixed volatile equilibria. AM MINERAL 72, 446–7.Google Scholar
Perchuck, L. L.&Lavrent'eva, I. V. 1983. Experimental investigation of exchange equilibria in the system cordierite–garnet–biotite. In Saxena, S. K. (ed.) Kinetics and equilibrium in mineral systems, 199239. New York: Springer Verlag.CrossRefGoogle Scholar
Perchuck, L. L., Aranovich, L. Y., Podlesskii, K. K., Lavrant'eva, I. V., Gerasimov, V. Y., Fed'kin, V. V., Kitsul, V. I., Karsakov, L. P.&Berdnikov, N. V. 1985. Precambrian granulites of the Aldan shield, eastern Siberia, USSR. J METAMORPH GEOL 3, 265310.CrossRefGoogle Scholar
Poli, S.&Schmidt, M. W. 1992. A comment on ‘calcic amphibole equilibria and a new amphibole-plagioclase geothermometer’. CONTRIB MINERAL PETROL 111, 273–82.CrossRefGoogle Scholar
Powell, R.&Evans, J. A. 1983. A new geobarometer for the assemblage biotite–muscovite–chlorite–quartz. J METAMORPH GEOL 1, 331–6.CrossRefGoogle Scholar
Powell, R. E.&Holland, T. J. B. 1988. An internally consistent dataset with uncertainties and correlations: 3. Applications to geobarometry, worked examples and a computer program. J METAMORPH GEOL 6, 173204.CrossRefGoogle Scholar
Rutherford, M. J.&Hill, P. M. 1993. Magma ascent rates from amphibole breakdown: an experimental study applied to the 1980–1986 Mount St. Helens eruption. J GEOPHYS RES 98, 19 667–85.CrossRefGoogle Scholar
Sack, R. O.&Ghiorso, M. S. 1991. An internally consistent model for the thermodynamic properties of Fe–Mg titanomagnetite–aluminate spinels. CONTRIB MINERAL PETROL 106, 474505.CrossRefGoogle Scholar
Schmidt, M. W. 1992. Amphibole composition in tonalite as a function of pressure: an experimental calibration of the Al-in-hornblende barometer. CONTRIB MINERAL PETROL 110, 304–10.CrossRefGoogle Scholar
Schmidt, M. W. 1993. Phase relations and compositions in tonalite as a function of pressure: an experimental study at 650°C. AM J SCI 293, 1011–60.CrossRefGoogle Scholar
Sengupta, P., Dasgupta, S., Bhattacharya, P. K.&Mukherjee, M. 1990. An orthopyroxene–biotite geothermometer and its application in crustal granulites and mantle-derived rocks. J METAMORPH GEOL 8, 191–7.CrossRefGoogle Scholar
Smith, J. V. 1974. Feldspar minerals. Heidelberg: Springer.Google Scholar
Spear, F. S. 1980. NaSi=CaAl exchange equilibrium between plagioclase and amphibole, an empirical model. CONTRIB MINERAL PETROL 72, 3341.CrossRefGoogle Scholar
Spear, F. S. 1981. Amphibole–plagioclase equilibria: an empirical model for the reaction albite + tremolite = edenite + 4 quartz. CONTRIB MINERAL PETROL 77, 355–64.CrossRefGoogle Scholar
Spear, F. S. 1993. Metamorphic phase equilibria and pressure–temperature time paths. MINERAL SOC MONOGR.Google Scholar
Spear, F. S.&Kimball, K. L. 1984. Recamp—a fortran IV program for estimating Fe3+ contents in amphiboles. COMPUT GEOSCI 10, 317–25.CrossRefGoogle Scholar
Spear, F. S., Selverstone, J., Hickmott, D., Crowley, P.&Hodges, K. 1984. P–T paths from garnet zoning: A new technique for deciphering tectonic processes in crystalline terranes. GEOLOGY 12, 8790.2.0.CO;2>CrossRefGoogle Scholar
Speer, J. A. 1987. Evolution of magmatic AFM mineral assemblages in granitoid rocks: the hornblende + melt = biotite reaction in the Liberty Hill pluton, South Carolina. AM MINERAL 72, 863–78.Google Scholar
Spencer, J. J.&Lindsley, D. H. 1981. A solution model for coexisting iron-titanium oxides. AM MINERAL 66, 1189–202.Google Scholar
Stormer, J. C. 1975. A practical two-feldspar thermometer. AM MINERAL 60, 667–74.Google Scholar
Stormer, J. C. 1983. The effects of recalculation on estimates of temperature and oxygen fugacity from analyses of multicomponent iron-titanium oxides. AM MINERAL 68, 586–94.Google Scholar
Thomas, W. M.&Ernst, W. G. 1990. The aluminum content of hornblende in calc-alkaline granitic rocks: a mineralogic barometer calibrated experimentally to 12 kbars. In Spencer, R. J.&Chou, I.-M. (ed) Fluid–mineral interactions: a tribute to H. P. Eugster. GEOCHEM SOC SPEC PUBL 2, 5963.Google Scholar
Valley, J. W.&Essene, E. J. 1980. Calc-silicate reactions in Adirondack marbles: the role of fluids and solid solution. BULL GEOL SOC AM 91, 114–7.2.0.CO;2>CrossRefGoogle Scholar
Velde, B. 1965. Phengitic micas: synthesis, stability, and natural occurrence. AM J SCI 263, 886913.CrossRefGoogle Scholar
Vielzeuf, D.&Clemens, J. D. 1992. The fluid absent melting of phlogopite + quartz: experiments and models. AM MINERAL 77, 1206–22.Google Scholar
Watson, E. B.&Harrison, T. M. 1983. Zircon saturation revisited: temperature and compositional effects in a variety of crustal magma types. EARTH PLANET SCI LETT 64, 295304.CrossRefGoogle Scholar
Wells, P. R. A. 1977. Pyroxene thermometry in simple and complex systems: CONTRIB MINERAL PETROL 62, 129139.CrossRefGoogle Scholar
Whitney, J. A.&Stormer, J. C. Jr 1977. The distribution of NaAlSi3O8 between coexisting microcline and plagioclase and its effect on geothermometric calculations. AM MINERAL 62, 687–91.Google Scholar
Williams, M. L.&Grambling, J. A. 1990. Manganese, ferric iron, and the equilibrium between garnet and biotite. AM MINERAL 75, 886908.Google Scholar
Wones, D. R. 1981. Mafic silicates as indicators of intensive parameters in granitic magmas. MIN GEOL 31, 191212.Google Scholar
Wones, D. R. 1989. Significance of the assemblage titanite + magnetite + quartz in granitic rocks. AM MINERAL 74, 744–9.Google Scholar
Wones, D. R.&Eugster, H. P. 1965. Stability of biotite: experiment, theory, and application. AM MINERAL 50, 1228–72.Google Scholar
Wood, B. J.&Banno, S. 1973. Garnet–orthopyroxene and orthopyroxene–clinopyroxene relationships in simple and complex systems. CONTRIB MINERAL PETROL 42, 109–27.CrossRefGoogle Scholar
Wyllie, P. J. 1984. Constraints imposed by experimental petrology on possible and impossible maga sources and products. PHIL TRANS R SOC LONDON A310, 439–56.Google Scholar
Young, E. D. 1990. Geothermobarometric and geochemical studies of two crystalline terrains of the eastern Mojave Desert, USA. Ph.D. Thesis, University of Southern California.Google Scholar
Zen, E-an. 1985. An oxygen buffer for some peraluminous granites and metamorphic rocks. AM MINERAL 70, 6573.Google Scholar
Zen, E-An. 1988. Phase relations of peraluminous granitic rocks and their petrogenetic implications. ANNU REV EARTH PLANET SCI 16, 2151.CrossRefGoogle Scholar
Zen, E-an&Hammarstrom, J. M. 1984a. Magmatic epidote and its petrologic significance. GEOLOGY 12, 515–8.2.0.CO;2>CrossRefGoogle Scholar
Zen, E-an&Hammarstrom, J. M. 1984b. Mineralogy and a petrogenetic model for the tonalite pluton at Bushy Point, Revillagigedo Island, Ketchikan 1° × 2° quadrangle, southeastern Alaska, U.S. GEOL SURVEY CIRC 939, 118–23.Google Scholar
zhu, C.&Sverjensky, D. A. 1991. Partitioning of F–C1–OH between minerals and hydrothermal fluids. GEOCHIM COSMOCHIM ACTA 55 1837–58CrossRefGoogle Scholar