Home An evolutionary system of mineralogy, Part VIII: The evolution of metamorphic minerals
Article
Licensed
Unlicensed Requires Authentication

An evolutionary system of mineralogy, Part VIII: The evolution of metamorphic minerals

  • Shaunna M. Morrison ORCID logo , Anirudh Prabhu ORCID logo and Robert M. Hazen ORCID logo EMAIL logo
Published/Copyright: September 24, 2024
Become an author with De Gruyter Brill

Abstract

Part VIII of the evolutionary system of mineralogy focuses on 1220 metamorphic mineral species, which correspond to 755 root mineral kinds associated with varied metamorphic rock types, most of which likely formed prior to the Phanerozoic Eon. A catalog of the mineral modes of 2785 metamorphic rocks from around the world reveals that 94 mineral kinds often occur as major phases. Of these common metamorphic minerals, 66 are silicates, 14 are oxides or hydroxides, 8 are carbonates or phosphates, 4 are sulfides, and 2 are polymorphs of carbon. Collectively, these 94 minerals incorporate 23 different essential chemical elements.

Patterns of coexistence among these 94 minerals, as revealed by network analysis and Louvain community detection, point to six major communities of metamorphic phases, three of which correspond to different pressure-temperature (P-T) regimes of metamorphosed siliceous igneous and sedimentary rocks, while three represent thermally altered carbonate and calc-silicate lithologies.

Metamorphic rocks display characteristics of an evolving chemical system, with significant increases in mineral diversity and chemical complexity through billions of years of Earth history. Earth’s first metamorphic minerals formed in thermally altered xenoliths and contact zones (hornfels and sanidinite facies) associated with early Hadean igneous activity (>4.5 Ga). The appearance of new Hadean lithologies, including clay-rich sediments, arkosic sandstones, and carbonates, provided additional protoliths for thermal metamorphism prior to 4.0 Ga. Orogenesis and erosion exposed extensive regional metamorphic terrains, with lithologies corresponding to the Barrovian sequence of index mineral metamorphic zones appearing by the Mesoarchean Era (>2.8 Ga). More recently, rapid subduction and rebound of crustal wedges, coupled with a shallowing geothermal gradient, has produced distinctive suites of blueschist, eclogite, and ultrahigh-pressure metamorphic suites (<1.0 Ga). The evolution of metamorphic minerals thus exemplifies changes in physical and chemical processes in Earth’s crust and upper mantle.

Funding statement: Studies of mineral evolution have been supported by the Deep-time Digital Earth (DDE) program, the John Templeton Foundation, the NASA Astrobiology Institute ENIGMA team, a private foundation, and the Carnegie Institution for Science. Any opinions, findings, or recommendations expressed herein are those of the authors and do not necessarily reflect the views of the National Aeronautics and Space Administration.

Acknowledgments

This manuscript benefited from multiple extensive and detailed reviews by John Ferry, who initially served as one of two reviewers of the submitted manuscript. His efforts fully warranted co-authorship, though he declined to be so recognized. Nevertheless, his informed and thoughtful contributions are reflected in every aspect of this contribution. We are also grateful to John Brady, Michael Brown, Douglas Rumble, Michael Walter, and Michael Wong for valuable discussions and reviews of an early version of this contribution. We also thank Associate Editor Simon Redfern and reviewer Jay Ague for their thorough, thoughtful, and constructive reviews.

References cited

Ackermand, D. and Raase, P. (1973) Coexisting zoisite and clinozoisite in biotite schists from the Hohe Tasuren, Austria. Contributions to Mineralogy and Petrology, 42, 333–341, https://doi.org/10.1007/BF00372611.Search in Google Scholar

Agrell, S.O. and Langley, J.M. (1958) The dolerite plug at Tievebulliagh, near Cushendall, Co. Antrim. Part (I): The thermal metamorphism. Proceedings of the Royal Irish Academy, 59B, 93–127.Search in Google Scholar

Alfors, J.T., Stinson, M.C., Matthews, R.A., and Pabst, A. (1965) Seven new barium minerals from eastern Fresno County, California. American Mineralogist, 50, 314–340.Search in Google Scholar

Almende, B.V. and Contributors, Benoit Thieurmel and Titouan Robert (2021) vis-Network: Network Visualization using ‘vis.js’ Library. R package version 2.1.0. https://CRAN.R-project.org/package=visNetwork. Accessed 11 September 2023.Search in Google Scholar

Anthony, J.W., Bideaux, R.A., Bladh, K.W., and Nichols, M.C. (1990–2003) Handbook of Mineralogy, 6 volumes. Mineral Data Publishing.Search in Google Scholar

Anzolini, C., Wang, F., Harris, G.A., Locock, A.J., Zhang, D., Nestola, F., Peruzzo, L., Jacobsen, S.D., and Pearson, D.G. (2019) Nixonite, Na2Ti6O13, a new mineral from a metasomatized mantle garnet pyroxenite from the western Rae Craton, Darby kimberlite field, Canada. American Mineralogist, 104, 1336–1344, https://doi.org/10.2138/am-2019-7023.Search in Google Scholar

Armbruster, T., Bonazzi, P., Akasaka, M., Bermanec, V., Chopin, C., Gieré, R., Heuss-Assbichler, S., Liebscher, A., Menchetti, S., Pan, Y., and others. (2006) Recommended nomenclature of the epidote-group minerals. European Journal of Mineralogy, 18, 551–567, https://doi.org/10.1127/0935-1221/2006/0018-0551.Search in Google Scholar

Atherton, M.P. (1964) The garnet isograd in pelitic rocks and its relation to metamorphic rocks. American Mineralogist, 49, 1331–1349.Search in Google Scholar

Augustithis, S.S. (1985) Atlas of the Textural Patterns of Metamorphosed (Transformed and Deformed) Rocks and their Genetic Significance, 401 p. Theophrastus Publications.Search in Google Scholar

Bai, W.J., Robinson, P., Fang, Q.S., Yang, J., Yan, B., Zhang, Z.M., Hu, X.F., Zhou, M.F., and Malpas, J. (2000) The PGE and base-metal alloys in the podiform chromitites of the Luobusa ophiolite, southern Tibet. Canadian Mineralogist, 38, 585–598, https://doi.org/10.2113/gscanmin.38.3.585.Search in Google Scholar

Bai, W.J., Zhou, M.F., and Robinson, P.T. (2011) Possibly diamond-bearing mantle peridotites and chromites in the Luobusa and Dongqiao ophiolites, Tibet. Canadian Journal of Earth Sciences, 30, 1650–1659, https://doi.org/10.1139/e93-143.Search in Google Scholar

Ballhaus, C., Wirth, R., Fonseca, R.O.C., Blanchard, H., Pröll, W., Bragagni, A., Nagel, T., Schreiber, A., Dittrich, S., Thome, V., and others. (2017) Ultra-high pressure and ultra-reduced minerals in ophiolites may form by lightning strikes. Geochemical Perspectives Letters, 5, 42–46.Search in Google Scholar

Banno, S. and Matsui, Y. (1965) Ecologite types and partition of Mg, Fe and Mn between clinopyroxene and garnet. Proceedings of the Japan Academy, 41, 716–721.Search in Google Scholar

Barnes, J. and Hut, P. (1986) A hierarchical O(N log N) force-calculation algorithm. Nature, 324, 446–449, https://doi.org/10.1038/324446a0.Search in Google Scholar

Barrow, G. (1893) On an intrusion of muscovite-biotite gneiss in the Southeast Highlands of Scotland, and its accompanying metamorphism. Quarterly Journal of the Geological Society of London, 49, 330–358, https://doi.org/10.1144/GSL.JGS.1893.049.01-04.52.Search in Google Scholar

Basta, E.Z. and Shaalan, M.M.B. (1974) Distribution of opaque minerals in the Tertiary volcanic rocks of Yemen and Aden. Neues Jahrbuch für Mineralogie, Abhandlungen, 121, 85–102.Search in Google Scholar

Bilgrami, S.A. (1956) Manganese silicate minerals from Chikla, Bhandara district, India. Mineralogical Magazine and Journal of the Mineralogical Society, 31, 236–244, https://doi.org/10.1180/minmag.1956.031.234.04.Search in Google Scholar

Binns, R.A. (1967) Barroisite-bearing eclogite from Naustdal, Sogn og Fjordane, Norway. Journal of Petrology, 8, 349–371, https://doi.org/10.1093/petrology/8.3.349.Search in Google Scholar

Bird, A. and Tobin, E. (2015) Natural kinds. In E.N. Zalta, Ed., The Stanford Encyclopedia of Philosophy, Spring 2015 ed., https://plato.stanford.edu/entries/natural-kinds/.Search in Google Scholar

Black, P.M. (1989) High-temperature calc-silicate hornfels in Northland, New Zealand. Bulletin—Royal Society of New Zealand, 26, 215–224.Search in Google Scholar

Blondel, V.D., Guillaume, J.-L., Lambiotte, R., and Lefebvre, E. (2008) Fast unfolding of communities in large networks. Journal of Statistical Mechanics, 2008, P10008, https://doi.org/10.1088/1742-5468/2008/10/P10008.Search in Google Scholar

Bluth, G.J.S. and Kump, L.R. (1991) Phanerozoic paleogeology. American Journal of Science, 291, 284–308, https://doi.org/10.2475/ajs.291.3.284.Search in Google Scholar

Bohlen, S.R., Montana, A., and Kerrick, D.M. (1991) Precise determinations of the equilibria kyanite-reversible-sillimanite and kyanite-reversible-andalusite and a revised triple point for Al2SiO5 polymorphs. American Mineralogist, 76, 677–680.Search in Google Scholar

Bostock, M., Ogievetsky, V., and Heer, J. (2011) D3: Data-driven documents. IEEE Transactions on Visualization and Computer Graphics, 17, 2301–2309, https://doi.org/10.1109/TVCG.2011.185.Search in Google Scholar

Botha, B.J.V., Ed. (1983) Namaqualand Metamorphic Complex. The Geological Society of South Africa.Search in Google Scholar

Boujibar, A., Howell, S., Zhang, S., Hystad, G., Prabhu, A., Liu, N., Stephan, T., Narkar, S., Eleish, A., Morrison, S.M., and others. (2021) Cluster analysis of presolar silicon carbide grains: Evaluation of their classification and astrophysical implications. The Astrophysical Journal Letters, 907, L39, https://doi.org/10.3847/2041-8213/abd102.Search in Google Scholar

Bowen, N.L. (1928) The Evolution of the Igneous Rocks, 332 p. Princeton University Press.Search in Google Scholar

Bowen, N.L. (1940) Progressive metamorphism of siliceous limestone and dolomite. The Journal of Geology, 48, 225–274, https://doi.org/10.1086/624885.Search in Google Scholar

Bowles, J.F.W., Howie, R.A., Vaughan, D.J., and Zussman, J. (2011) Rock-Forming Minerals, Vol. 5A: Non-Silicates: Oxides, Hydroxides and Sulfides, Second Edition, 418 p. The Geological Society of London.Search in Google Scholar

Boyd, R. (1991) Realism, anti-foundationalism and the enthusiasm for natural kinds. Philosophical Studies, 61, 127–148, https://doi.org/10.1007/BF00385837.Search in Google Scholar

Boyd, R. (1999) Homeostasis, species, and higher taxa. In R. Wilson, Ed., Species: New Interdisciplinary Essays, 141–186. Cambridge University Press.Search in Google Scholar

Brown, M. (2006) Duality of thermal regimes is the distinctive characteristic of plate tectonics since the Neoarchean. Geology, 34, 961–964, https://doi.org/10.1130/G22853A.1.Search in Google Scholar

Brown, M. (2007) Metamorphic conditions in orogenic belts: A record of secular change. International Geology Review, 49, 193–234, https://doi.org/10.2747/0020-6814.49.3.193.Search in Google Scholar

Brown, M. and Johnson, T.E. (2019) Time’s arrow, time’s cycle: Granulite metamorphism and geodynamics. Mineralogical Magazine, 83, 323–338, https://doi.Org/10.1180/mgm.2019.19.Search in Google Scholar

Buddington, A.F. (1952) Chemical petrology of some metamorphosed Adirondack gabbroic, syenitic and quartz syenitic rocks. American Journal of Science, Bowen Volume, 37–84.Search in Google Scholar

Burke, E.A.J. (2006) The end of CNMMN and CCM—Long live the CNMNC! Elements, 2, 388.Search in Google Scholar

Buseck, P.R. and Huang, B.-J. (1985) Conversion of carbonaceous material to graphite during metamorphism. Geochimica et Cosmochimica Acta, 49, 2003–2016, https://doi.org/10.1016/0016-7037(85)90059-6.Search in Google Scholar

Bustin, R.M. and Mathews, W.H. (1982) In situ gasification of coal, a natural example: History, petrology, and mechanics of combustion. Canadian Journal of Earth Sciences, 19, 514–523, https://doi.org/10.1139/e82-042.Search in Google Scholar

Button, A. (1982) Sedimentary iron deposits, evaporates and phosphorites: State of the art report. In H.D. Holland and M. Schidlowski, Eds., Mineral Deposits and the Evolution of the Biosphere, 259–273. Springer-Verlag.Search in Google Scholar

Cairncross, B. and Beukes, N.J. (2013) The Kalahari Manganese Field, the Adventure Continues. Struik Nature Publishers.Search in Google Scholar

Carpenter, A.B. (1967) Mineralogy and petrology of the system CaO-MgO-CO2-H2O at Crestmore, California. American Mineralogist, 52, 1351–1363.Search in Google Scholar

Carswell, D.A., Ed. (1990) Ecologite Facies Rocks, 411 p. Springer.Search in Google Scholar

Carswell, D.A., and Compagnoni, R., Eds. (2003) Ultrahigh Pressure Metamorphism, Vol. 5, 508 p. Mineralogical Society of Great Britain and Ireland.Search in Google Scholar

Chang, L.L.Y., Howie, R.A., and Zussman, J. (1996) Rock-Forming Minerals, Vol. 5B: Sulphates, Carbonates, Phosphates and Halides, Second edition, 392 p. Longman.Search in Google Scholar

Chatterjee, N.D. (1970) Synthesis and upper stability of paragonite. Contributions to Mineralogy and Petrology, 27, 244–257, https://doi.org/10.1007/BF00385781.Search in Google Scholar

Chiama, K., Gabor, M., Lupini, I., Rutledge, R., Nord, J.A., Zhang, S., Boujibar, A., and Hazen, R.M. (2020) Garnet: A comprehensive standardized database of garnet geochemical analyses integrating provenance and paragenesis. Geological Society of America Annual Meeting 2020, 52.Search in Google Scholar

Chiama, K., Gabor, M., Lupini, I., Rutledge, R., Nord, J.A., Zhang, S., Boujibar, A., Bullock, E.S., Walter, M.J., Spear, F., and others. (2022) ESMD-Garnet dataset. Open Data Repository, https://doi.org/10.48484/camh-xy98.Search in Google Scholar

Chopin, C. (1981) Talc-phengite: A widespread assemblage in high-grade pelitic blueschists of the Western Alps: A first record and some consequences. Journal of Petrology, 22, 628–650, https://doi.org/10.1093/petrology/22.4.628.Search in Google Scholar

Chopin, C. (1984) Coesite and pure pyrope in high-grade blueschists of the western Alps: A first record and some consequences. Contributions to Mineralogy and Petrology, 86, 107–118, https://doi.org/10.1007/BF00381838.Search in Google Scholar

Cleland, C.E., Hazen, R.M., and Morrison, S.M. (2021) Historical natural kinds in mineralogy: Systematizing contingency in the context of necessity. Proceedings of the National Academy of Sciences, 108, e2015370118 (8 p.)Search in Google Scholar

Coleman, R.G. and Clark, J.R. (1968) Pyroxenes in the blueschist facies of California. American Journal of Science, 266, 43–59, https://doi.org/10.2475/ajs.266.1.43.Search in Google Scholar

Coleman, R.G. and Lee, D.E. (1963) Glaucophane-bearing metamorphic rock types of the Cadazero area, California. Journal of Petrology, 4, 260–301, https://doi.org/10.1093/petrology/4.2.260.Search in Google Scholar

Coleman, R.G., Lee, D.E., Beatty, L.B., and Brannock, W.W. (1965) Eclogites and eclogites: Their differences and similarities. Geological Society of America Bulletin, 76, 483–508, https://doi.org/10.1130/0016-7606(1965)76[483:EAETDA]2.0.CO;2.Search in Google Scholar

Coombs, D.S. (1960) Lower grade mineral facies in New Zealand. Report of the 21st International Geological Congress, 13, 339–351.Search in Google Scholar

Coombs, D.S. (1993) Dehydration veins in diagenetic very-low-grade metamorphic rocks: Features of the crustal seismogenic zone and their significance to mineral facies. Journal of Metamorphic Geology, 11, 389–399, https://doi.org/10.1111/j.1525-1314.1993.tb00156.x.Search in Google Scholar

Coombs, D.S., Kawachi, Y., Houghton, B.F., Hyden, G., Pringle, I.J., and Williams, J.G. (1977) Andradite and andradite-grossular solid solutions in very low-grade regionally metamorphosed rocks in southern New Zealand. Contributions to Mineralogy and Petrology, 63, 229–246, https://doi.org/10.1007/BF00375574.Search in Google Scholar

Csardi, G. and Nepusz, T. (2006) The igraph software package for complex network research. InterJournal. Complex Systems. 1695, https://igraph.orgSearch in Google Scholar

Daly, R.A. (1925) The geology of Ascension Island. Proceedings of the American Academy of Arts and Sciences, 60, 3–80, https://doi.org/10.2307/25130043.Search in Google Scholar

Davies, G.R., Nixon, P.H., Pearson, D.G., and Obata, M. (1993) Tectonic implications of graphitized diamonds from the Ronda peridotite massif, southern Spain. Geology, 21, 471–474, https://doi.org/10.1130/0091-7613(1993)021<0471:TIOGDF>2.3.CO;2.Search in Google Scholar

Davis, S.R. and Ferry, J.M. (1993) Fluid infiltration during contact metamorphism of interbedded marble and calc-silicate hornfels, Twin Lakes area, central Sierra Nevada, California. Journal of Metamorphic Geology, 11, 71–88, https://doi.org/10.1111/j.1525-1314.1993.tb00132.x.Search in Google Scholar

De Roever, W.P. (1956) Some differences between post-Paleozoic and older regional metamorphism. Geologie en Mijnbouw. New Series, 18e, 123–127.Search in Google Scholar

Deer, W.A., Howie, R.A., and Zussman, J. (1982) Rock-Forming Minerals, Vol. 1A: Orthosilicates, Second ed., 919 p. Longman.Search in Google Scholar

Deer, W.A., Howie, R.A., and Zussman, J. (1986) Rock-Forming Minerals, Vol. 1B: Disilicates and Ring Silicates, 629 p. Second ed. Wiley.Search in Google Scholar

Deer, W.A., Howie, R.A., and Zussman, J. (1997a) Rock-Forming Minerals, Vol. 2A: Single-Chain Silicates, Second ed. 668 p. The Geological Society of London.Search in Google Scholar

Deer, W.A., Howie, R.A., and Zussman, J. (1997b) Rock-Forming Minerals, Vol. 2B: Double-Chain Silicates, Second ed., 764 p. The Geological Society of London.Search in Google Scholar

Deer, W.A., Howie, R.A., and Zussman, J. (2001) Rock-Forming Minerals, Vol. 4A: Framework Silicates: Feldspars, Second ed., 984 p. The Geological Society of London.Search in Google Scholar

Deer, W.A., Howie, R.A., Wise, W.S., and Zussman, J. (2004) Rock-Forming Minerals, Vol. 4B: Framework Silicates: Silica Minerals, Feldspathoids and the Zeolites, Second ed., 982 p. The Geological Society of London.Search in Google Scholar

Deer, W.A., Howie, R.A., and Zussman, J. (2009) Rock-Forming Minerals, Vol. 3B: Layered Silicates Excluding Micas and Clay Minerals, Second ed., 328 p. The Geological Society of London.Search in Google Scholar

Deer, W.A., Howie, R.A., and Zussman, J. (1982–2013) Rock-Forming Minerals. Second ed., 11 volumes. Longman, Wiley, and The Geological Society of London.Search in Google Scholar

Diessel, C.F.K., Brothers, R.N., and Black, P.M. (1978) Coalification and graphitization in high-pressure schists in New Caledonia. Contributions to Mineralogy and Petrology, 68, 63–78, https://doi.org/10.1007/BF00375447.Search in Google Scholar

Dilek, Y. (2003) Ophiolite pulses, mantle plumes, and orogeny. Special Publication, Geological Society of London, 218, 9–19, https://doi.org/10.1144/GSL.SP.2003.218.01.02.Search in Google Scholar

Dobrzhinetskaya, L.F., Eide, E.A., Larsen, R.B., Sturt, B.A., Trønnes, R.G., Smith, D.C., Taylor, W.R., and Posukhova, T.V. (1995) Microdiamond in high-grade metamorphic rocks of the Western Gneiss region, Norway. Geology, 23, 597–600, https://doi.org/10.1130/0091-7613(1995)023<0597:MIHGMR>2.3.CO;2.Search in Google Scholar

Dobrzhinetskaya, L.F., Wirth, R., Yang, J., Hutcheon, I.D., Weber, P.K., and Green, H.W. II (2009) High-pressure highly reduced nitrides and oxides from chromitite of a Tibetan ophiolite. Proceedings of the National Academy of Sciences, 106, 19233–19238, https://doi.org/10.1073/pnas.0905514106.Search in Google Scholar

Dobrzhinetskaya, L.F., O’Bannon, E.F. III, and Sumino, H. (2022) Non-cratonic diamonds from UHP metamorphic terranes, ophiolites and volcanic sources. Reviews in Mineralogy and Geochemistry, 88, 191–256.Search in Google Scholar

Einaudi, M.T. and Burt, D.M. (1982) Terminology, classification and composition of skarn deposits. Economic Geology and the Bulletin of the Society of Economic Geologists, 77, 745–754, https://doi.org/10.2113/gsecongeo.77.4.745.Search in Google Scholar

Einaudi, M.T., Meinert, L.D., and Newberry, R.J. (1981) Skarn deposits. Economic Geology, 75th Anniversary Volume, 317–391.Search in Google Scholar

Ereshefsky, M. (2014) Species, historicity, and path dependency. Philosophy of Science, 81, 714–726, https://doi.org/10.1086/677202.Search in Google Scholar

Ernst, W.G. (1972) Occurrence and mineralogical evolution of blueschist belts with time. American Journal of Science, 272, 657–668, https://doi.org/10.2475/ajs.272.7.657.Search in Google Scholar

Eskola, P. (1920) The mineral facies of rocks. Norsk Geologisk Tidsskrift, 6, 143–194.Search in Google Scholar

Eskola, P. (1952) On the granulites of Lapland. American Journal of Science Part 1, 133–171.Search in Google Scholar

Evans, B.W. (1964) Coexisting albite and oligoclase in some schists from New Zealand. American Mineralogist, 49, 173–179.Search in Google Scholar

Falkowski, P., Scholes, R.J., Boyle, E., Canadell, J., Canfield, D., Elser, J., Gruber, N., Hibbard, K., Hogberg, P., Linder, S., and others. (2000) The global carbon cycle: A test of our knowledge of earth as a system. Science, 290, 291–296, https://doi.org/10.1126/science.290.5490.291.Search in Google Scholar

Ferré, E. (1989) Les gneiss à cordiérite-grenat-orthoamphibole de Topiti: Témoin possible d’un soile métamorphique du Protérozoique en Corse occidentale. Compte Rendus Academie des Science Paris, Series 2, 309, 893–898.Search in Google Scholar

Ferry, J.M. (1976) P, T, fCO2 and fH2O during metamorphism of calcareous sediments in the Waterville-Vassalboro area, south-central Maine. Contributions to Mineralogy and Petrology, 57, 119–143, https://doi.org/10.1007/BF00405221.Search in Google Scholar

Ferry, J.M. (1984) A biotite isograd in South-Central Maine, U.S.A.: Mineral reactions, fluid transfer, and heat transfer. Journal of Petrology, 25, 871–893, https://doi.org/10.1093/petrology/25.4.871.Search in Google Scholar

Ferry, J.M. (1988) Infiltration-driven metamorphism in Northern New England, USA. Journal of Petrology, 29, 1121–1159, https://doi.org/10.1093/petrology/29.6.1121.Search in Google Scholar

Ferry, J.M. (1989) Contact metamorphism of roof pendants at Hope Valley, Alpine County, California, USA. Contributions to Mineralogy and Petrology, 101, 402–417, https://doi.org/10.1007/BF00372214.Search in Google Scholar

Ferry, J.M. (1992) Regional metamorphism of the Waits River Formation, eastern Vermont: Delineation of a new type of giant metamorphic hydrothermal system. Journal of Petrology, 33, 45–94, https://doi.org/10.1093/petrology/33.1.45.Search in Google Scholar

Ferry, J.M. (1994) Overview of the petrologic record of fluid flow during regional metamorphism in northern New England. American Journal of Science, 294, 905–988, https://doi.org/10.2475/ajs.294.8.905.Search in Google Scholar

Ferry, J.M. (1995) Fluid flow during contact metamorphism of ophicarbonate rocks in the Bergell Aureole, Val Malenco, Italian Alps. Journal of Petrology, 36, 1039–1053, https://doi.org/10.1093/petrology/36.4.1039.Search in Google Scholar

Ferry, J.M. (1996) Prograde and retrograde fluid flow during contact metamorphism of siliceous carbonate rocks from the Ballachulish aureole, Scotland. Contributions to Mineralogy and Petrology, 124, 235–254, https://doi.org/10.1007/s004100050189.Search in Google Scholar

Ferry, J.M. (2007) The role of volatile transport by diffusion and dispersion in driving biotite-forming reactions during regional metamorphism of the Gile Mountain Formation, Vermont. American Mineralogist, 92, 1288–1302, https://doi.org/10.2138/am.2007.2429.Search in Google Scholar

Ferry, J.M. and Rumble, D. III (1997) Formation and destruction of periclase by fluid flow in two contact aureoles. Contributions to Mineralogy and Petrology, 128, 313–334, https://doi.org/10.1007/s004100050312.Search in Google Scholar

Ferry, J.M., Mutti, L.J., and Zuccala, G.J. (1987) Contact metamorphism/hydrothermal alteration of Tertiary basalts from the Isle of Skye, northwest Scotland. Contributions to Mineralogy and Petrology, 95, 166–181, https://doi.org/10.1007/BF00381266.Search in Google Scholar

Ferry, J.M., Wing, B.A., and Rumble, D. III (2001) Formation of wollastonite by chemically reactive fluid flow during contact metamorphism, Mt. Morrison Pendant, Sierra Nevada, California, USA. Journal of Petrology, 42, 1705–1728, https://doi.org/10.1093/petrology/42.9.1705.Search in Google Scholar

Ferry, J.M., Wing, B.A., Penniston-Dorland, S.C., and Rumble, D. (2002) Direction of fluid flow during contact metamorphism of siliceous carbonate rocks: New data for the Monzoni and Predazzo aureoles, northern Italy, and a global review. Contributions to Mineralogy and Petrology, 142, 679–699, https://doi.org/10.1007/s00410-001-0316-7.Search in Google Scholar

Ferry, J.M., Rumble, D. III, Wing, B.A., and Penniston-Dorland, S.C. (2005) A new interpretation of centimeter-scale variations in the progress of infiltration-driven metamorphic reactions: Case study of carbonated metaperidotite, Val d’Efra, Central Alps, Switzerland. Journal of Petrology, 46, 1725–1746, https://doi.org/10.1093/petrology/egi034.Search in Google Scholar

Ferry, J.M., Stubbs, J.E., Xu, H., Guan, Y., and Eiler, J.M. (2015) ankerite grains with dolomite cores: A diffusion chronometer for low- to medium-grade regionally metamorphosed clastic sediments. American Mineralogist, 100, 2443–2457, https://doi.org/10.2138/am-2015-5209.Search in Google Scholar

Fleet, M.E. (2003) Rock-Forming Minerals, Vol. 3A: Sheet Silicates: Micas, Second ed., 758 p. The Geological Society of London.Search in Google Scholar

Floran, R.J. and Papike, J.J. (1978) Mineralogy and petrology of the Gunflint Iron Formation, Minnesota–Ontario: Correlation of compositional and assemblage variations at low to moderate grade. Journal of Petrology, 19, 215–288, https://doi.org/ 10.1093/petrology/19.2.215.Search in Google Scholar

Fortunato, S. (2010) Community detection in graphs. Physics Reports, 486, 75–174, https://doi.org/10.1016/j.physrep.2009.11.002.Search in Google Scholar

Frondel, C. (1990) Historical overview of the development of mineralogical science at Franklin and Sterling Hill, Sussex County, New Jersey. In A.S. Wilkerson, Ed., Character and Origin of the Franklin-Sterling Hill Orebodies. Franklin, p. 3–13, Ogdensburg Mineralogical Society.Search in Google Scholar

Fulignati, P., Marianelli, P., Santacroce, R., and Sbrana, A. (2000) The skarn shell of the 1944 Vesuvius magma chamber. Genesis and P-T-X conditions from melt and fluid inclusion data. European Journal of Mineralogy, 12, 1025–1039, https://doi.org/10.1127/0935-1221/2000/0012-1025.Search in Google Scholar

Furnes, H., de Wit, M., Staudigel, H., Rosing, M., and Muehlenbachs, K. (2007) A vestige of Earth’s oldest ophiolite. Science, 315, 1704–1707, https://doi.org/10.1126/science.1139170.Search in Google Scholar

Gaines, R.V., Skinner, C., Foord, E.E., Mason, B., and Rosenzweig, A. (1997) Dana’s New Mineralogy, 1872 p. Wiley.Search in Google Scholar

Ganade, C.E., Rubatto, D., Lanari, P., Hermann, J., Tesser, L.R., and Caby, R. (2023) Fast exhumation of Earth’s earliest ultrahigh-pressure rocks in the West Gondwana orogen, Mali. Geology, 51, 647–651.Search in Google Scholar

Gates, A.E. and Speer, J.A. (1991) Allochemical retrograde metamorphism in shear zones: And example in metapelites, Virginia, USA. Journal of Metamorphic Geology, 9, 581–604, https://doi.org/10.1111/j.1525-1314.1991.tb00550.x.Search in Google Scholar

Girvan, M. and Newman, M.E.J. (2002) Community structure in social and biological networks. Proceedings of the National Academy of Sciences of the United States of America, 99, 7821–7826, https://doi.org/10.1073/pnas.122653799.Search in Google Scholar

Godman, M. (2019) Scientific realism with historical essences: The case of species. Synthese, 198, 3041–3057, https://doi.org/10.1007/s11229-018-02034-3.Search in Google Scholar

Goldschmidt, V.M. (1911) Die Kontakmetamorphose im Kristtianiagebiet. Norske Videnskabers Selskabs Skrifter I, Mat.-Naturv. Klasse, no. 1.Search in Google Scholar

Grapes, R. (2006) Pyrometamorphism, 2nd ed., 276 p. Springer.Search in Google Scholar

Gregory, D.D., Cracknell, M.J., Large, R.R., McGoldrick, P., Kuhn, S., Maslennikov, V.V., Baker, M.J., Fox, N., Belousov, I., Figueroa, M.C., and others. (2019) Distinguishing ore deposit type and barren sedimentary pyrite using laser ablation-inductively coupled plasma-mass spectrometry trace element data and statistical analysis of large data sets. Economic Geology and the Bulletin of the Society of Economic Geologists, 114, 771–786, https://doi.org/10.5382/econgeo.4654.Search in Google Scholar

Guidotti, C.V. (1984) Micas in metamorphic rocks. Reviews in Mineralogy, 13, 357–467.Search in Google Scholar

Guidotti, C.V. and Sassi, F.P. (1998) Petrogenetic significance of Na-K white mica mineralogy: Recent advances for metamorphic rocks. European Journal of Mineralogy, 10, 815–854, https://doi.org/10.1127/ejm/10/5/0815.Search in Google Scholar

Guidotti, C.V., Sassi, F.P., Blencoe, J.G., and Selverstone, J. (1994) The paragonitemuscovite solvus. I. P-T-X limits derived from the Na-K compositions of natural, quasibinary paragonite-muscovite pairs. Geochimica et Cosmochimica Acta, 58, 2269–2275, https://doi.org/10.1016/0016-7037(94)90009-4.Search in Google Scholar

Guitard, G. (1965) Les types de métamorphisme régionale à andalousite, cordiérite et almandine, et à andalousite, cordiérite, almandine et staurotide, dans la zone axiale des Pyrénees-Orientales; contributions à l’étude des types de métamorphisme de basse pression. Compte Rendus Academie des Science, Paris, 269D, 1159–1162.Search in Google Scholar

Hacker, B.R. (2006) Pressures and temperatures of ultrahigh-pressure metamorphism: Implications for UHP tectonics and H2O in subducting slabs. International Geology Review, 48, 1053–1066, https://doi.org/10.2747/0020-6814.48.12.1053.Search in Google Scholar

Halferdahl, L.B. (1961) Chloritoid: Its composition, X-ray and optical properties, stability and occurrence. Journal of Petrology, 2, 49–135, https://doi.org/10.1093/petrology/2.1.49.Search in Google Scholar

Hall, A.J., Boyce, A.J., and Fallick, A.E. (1987) Iron sulfides in metasediments; support for a retrogressive pyrrhotite to pyrite reaction. Chemical Geology, 65, 305–310.Search in Google Scholar

Harker, A. (1950) Metamorphism: A Study of the Transformations of Rock-Masses, 362 p. E.P. Dutton & Co.Search in Google Scholar

Harley, S.L. (2008) Refining the P-T records of UHT crustal metamorphism. Journal of Metamorphic Geology, 26, 125–154, https://doi.org/10.1111/j.1525-1314.2008.00765.x.Search in Google Scholar

Ferry, J.M. (2021) UHT metamorphism. In D. Alderton and S.A. Elias, Encyclopedia of Geology, 2nd ed., 522–552. Elsevier.Search in Google Scholar

Hatert, F., Mills, S.J., Hawthorne, F.C., and Rumsey, M.S. (2021) A comment on “An evolutionary system of mineralogy: Proposal for a classification of planetary materials based on natural kind clustering.” American Mineralogist, 106, 150–153, https://doi.org/10.2138/am-2021-7590.Search in Google Scholar

Hawley, K. and Bird, A. (2011) What are natural kinds? Philosophical Perspectives, 25, 205–221, https://doi.org/10.1111/j.1520-8583.2011.00212.x.Search in Google Scholar

Hawthorne, F.C., Griep, J.L., and Curtis, L. (1980) A three amphibole assemblage from the Talon Lake sill, Peterborough County, Ontario. Canadian Mineralogist, 18, 275–284.Search in Google Scholar

Hawthorne, F.C., Oberti, R., Harlow, G.E., Maresch, W.V., Martin, R.F., Schumacher, J.C., and Welch, M.D. (2012) Nomenclature of the amphibole supergroup. American Mineralogist, 97, 2031–2048, https://doi.org/10.2138/am.2012.4276.Search in Google Scholar

Hawthorne, F.C., Mills, S.J., Hatert, F., and Rumsey, M.S. (2021) Ontology, archetypes and the definition of “mineral species.” Mineralogical Magazine, 85, 125–131, https://doi.org/10.1180/mgm.2021.21.Search in Google Scholar

Hazen, R.M. (2019) An evolutionary system of mineralogy: Proposal for a classification of planetary materials based on natural kind clustering. American Mineralogist, 104, 810–816, https://doi.org/10.2138/am-2019-6709CCBYNCND.Search in Google Scholar

Hazen, R.M. (2021) Reply to “A comment on ‘An evolutionary system of mineralogy: Proposal for a classification of planetary materials based on natural kind clustering’.” American Mineralogist, 106, 154–156, https://doi.org/10.2138/am-2021-7773.Search in Google Scholar

Hazen, R.M. and Ausubel, J.H. (2016) On the nature and significance of rarity in mineralogy. American Mineralogist, 101, 1245–1251, https://doi.org/10.2138/am-2016-5601CCBY.Search in Google Scholar

Hazen, R.M. and Morrison, S.M. (2020) An evolutionary system of mineralogy. Part I: Stellar mineralogy (>13 to 4.6 Ga). American Mineralogist, 105, 627–651, https://doi.org/10.2138/am-2020-7173.Search in Google Scholar

Hazen, R.M. (2021) An evolutionary system of mineralogy, Part V: Aqueous and thermal alteration of planetesimals (~4.565 to 4.550 Ma). American Mineralogist, 106, 1388–1419, https://doi.org/10.2138/am-2021-7760.Search in Google Scholar

Hazen, R.M. (2022) On the paragenetic modes of minerals: A mineral evolution perspective. American Mineralogist, 107, 1262–1287, https://doi.org/10.2138/am-2022-8099.Search in Google Scholar

Hazen, R.M., Papineau, D., Bleeker, W., Downs, R.T., Ferry, J.M., McCoy, T.L., Sverjensky, D.A., and Yang, H. (2008) Mineral evolution. American Mineralogist, 93, 1693–1720, https://doi.org/10.2138/am.2008.2955.Search in Google Scholar

Hazen, R.M., Morrison, S.M., and Prabhu, A. (2021) An evolutionary system of mineralogy. Part III: Primary chondrule mineralogy (4566 to 4561 Ma). American Mineralogist, 106, 325–350, https://doi.org/10.2138/am-2020-7564.Search in Google Scholar

Hazen, R.M., Morrison, S.M., Krivovichev, S.L., and Downs, R.T. (2022) Lumping and splitting: Toward a classification of mineral natural kinds. American Mineralogist, 107, 1288–1301, https://doi.org/10.2138/am-2022-8105.Search in Google Scholar

Hazen, R.M., Morrison, S.M., Prabhu, A., Walter, M.J., and Williams, J.R. (2023) An evolutionary system of mineralogy, Part VII: The evolution of the igneous minerals (>2500 Ma). American Mineralogist, 108, 1620–1641, https://doi.org/10.2138/am-2022-8539.Search in Google Scholar

Heaney, P.J. (2016) Time’s arrow in the trees of life and minerals. American Mineralogist, 101, 1027–1035, https://doi.org/10.2138/am-2016-5419.Search in Google Scholar

Henry, D.J. and Dutrow, B.L. (2012) Tourmaline at diagenetic to low-grade metamorphic conditions: Its petrologic applicability. Lithos, 154, 16–32, https://doi.org/10.1016/j.lithos.2012.08.013.Search in Google Scholar

Henry, D.J., Novak, M., Hawthorne, F.C., Ertl, A., Dutrow, B.L., Uher, P., and Pezzotta, F. (2011) Nomenclature of the tourmaline-supeigroup minerals. American Mineralogist, 96, 895–913, https://doi.org/10.2138/am.2011.3636.Search in Google Scholar

Hodges, K.V. and Spear, F.S. (1982) Geothermometry, geobarometry and the Al2SiO5 triple point at Mt. Moosilauke, New-Hampshire. American Mineralogist, 67, 1118–1134.Search in Google Scholar

Holder, R.M., Viete, D.R., Brown, M., and Johnson, T.E. (2019) Metamorphism and the evolution of plate tectonics. Nature, 572, 378–381, https://doi.org/10.1038/s41586-019-1462-2.Search in Google Scholar

Hori, F. (1962) On the load metamorphic formation of rhodonite, tephroite and manganosite. Scientific Papers of the College of General Education, University of Tokyo, 12, 117–142.Search in Google Scholar

Hoschek, G. (1984) Alpine metamorphism of calcareous sediments in the Western Hohe Tauren, Tyrol: Mineral equilibria in COHS fluids. Contributions to Mineralogy and Petrology, 87, 129–137, https://doi.org/10.1007/BF00376219.Search in Google Scholar

Hystad, G., Boujibar, A., Liu, N., Nittler, L.R., and Hazen, R.M. (2021) Evaluation of the classification of presolar silicon carbide grains using consensus clustering with resampling methods: An assessment of the confidence of grain assignments. Monthly Notices of the Royal Astronomical Society, 510, 334–350, https://doi.org/10.1093/mnras/stab3478.Search in Google Scholar

Iwasaki, M. (1960) Barroisitic amphibole from Bizan in eastern Sikoku, Japan. Chishitsugaku Zasshi, 66, 625–630, https://doi.org/10.5575/geosoc.66.625.Search in Google Scholar

Jacob, D.E. and Mikhail, S. (2022) Polycrystalline diamonds from kimberlites: Snapshots of rapid and episodic diamond formation in the lithospheric mantle. Reviews in Mineralogy and Geochemistry, 88, 167–189, https://doi.org/10.2138/rmg.2022.88.03.Search in Google Scholar

Jahn, B.-M., Caby, R., and Monie, P. (2001) The oldest UHP eclogites of the World: Age of UHP metamorphism, nature of protoliths and tectonic implications. Chemical Geology, 178, 143–158, https://doi.org/10.1016/S0009-2541(01)00264-9.Search in Google Scholar

Jan, M.Q. and Symes, R.F. (1977) Piemontite schists from Upper Swat, north-west Pakistan. Mineralogical Magazine, 41, 537–540, https://doi.org/10.1180/minmag.1977.041.320.22.Search in Google Scholar

Joplin, G.A. (1968) A Petrography of Australian Metamorphic Rocks, 262 p. Elsevier.Search in Google Scholar

Kappler, A., Pasquero, C., Konhauser, K.O., and Newman, D.K. (2005) Deposition of banded iron formations by anoxygenic phototrophic Fe(II)-oxidizing bacteria. Geology, 33, 865–868, https://doi.org/10.1130/G21658.1.Search in Google Scholar

Khalidi, M.A. (2013) Natural Categories and Human Kinds: Classification in the Natural and Social Sciences, 264 p. Cambridge University Press.Search in Google Scholar

Kimball, K.L. and Spear, F.S. (1984) Metamorphic petrology of the Jackson County Iron Formation, Wisconsin. Canadian Mineralogist, 22, 605–619.Search in Google Scholar

Kjarsgaard, B.A., de Wit, M., Heaman, L.M., Pearson, D.G., Stiefenhofer, J., Janusczcak, N., and Shirey, S.B. (2022) A review of the geology of global diamond mines and deposits. Reviews in Mineralogy and Geochemistry, 88, 1–117, https://doi.org/10.2138/rmg.2022.88.01.Search in Google Scholar

Klein, C. (1966) Mineralogy and petrology of the metamorphosed Wabush iron formation, southwestern Labrador. Journal of Petrology, 7, 246–305, https://doi.org/10.1093/petrology/7.2.246.Search in Google Scholar

Klein, C. (2005) Some Precambrian banded iron-formations (BIFs) from around the world: Their age, geologic setting, mineralogy, metamorphism, geochemistry, and origins. American Mineralogist, 90, 1473–1499, https://doi.org/10.2138/am.2005.1871.Search in Google Scholar

Kranck, S.H. (1961) A study of phase equilibria in a metamorphic iron formation. Journal of Petrology, 2, 137–184, https://doi.org/10.1093/petrology/2.2.137.Search in Google Scholar

Krivovichev, S.V., Krivovichev, V.G., and Hazen, R.M. (2018) Structural and chemical complexity of minerals: Correlations and time evolution. European Journal of Mineralogy, 30, 231–236, https://doi.org/10.1127/ejm/2018/0030-2694.Search in Google Scholar

Kusky, T.M., Ed. (2004) Precambrian Ophiolites and Related Rocks, 772 p. Elsevier.Search in Google Scholar

Kusky, T.M., Li, J.-H., and Tucker, R.D. (2001) The Archean Dongwanzi ophiolite complex, North China craton: 2.505-billion-year-old oceanic crust and mantle. Science, 292, 1142–1145, https://doi.org/10.1126/science.1059426.Search in Google Scholar

Lal, R.K. (1969) Retrogression of cordierite to kyanite and andalusite at Fishtail Lake, Ontario, Canada. Mineralogical Magazine, 37, 466–471, https://doi.org/10.1180/minmag.1969.037.288.06.Search in Google Scholar

Landis, C.A. (1971) Graphitization of dispersed carbonaceous material in metamorphic rocks. Contributions to Mineralogy and Petrology, 30, 34–45, https://doi.org/10.1007/BF00373366.Search in Google Scholar

Larsen, E.S. and Foshag, W.F. (1921) Merwinite, a new calcium magnesium orthosilicate from Crestmore, California. American Mineralogist, 6, 143–148.Search in Google Scholar

Leach, D.L., Sangster, D.F., Kelley, K.D., Large, R.R., Garven, G., Allen, C.R., Gutzmer, J., and Walters, S. (2005) Sediment-hosted lead-zinc deposits: A global perspective. Economic Geology, 100th Anniversary Volume, 561–607.Search in Google Scholar

Leech, M. and Ernst, G. (1998) Graphite pseudomorphs after diamond? A carbon isotope and spectroscopic study of graphite cuboids from the Maksyutov Complex, south Ural Mountains, Russia. Geochimica et Cosmochimica Acta, 62, 2143–2154, https://doi.org/10.1016/S0016-7037(98)00142-2.Search in Google Scholar

Léger, A. and Ferry, J.M. (1993) Fluid infiltration and regional metamorphism of the Waits River Formation, north-east Vermont, USA. Journal of Metamorphic Geology, 11, 3–29, https://doi.org/10.1111/j.1525-1314.1993.tb00128.x.Search in Google Scholar

Lipp, A.G., Shorttle, O., Sperling, E.A., Brocks, J.J., Cole, D.B., Crockford, P.W., Del Mouro, L., Dewing, K., Dornbos, S.Q., Emmings, J.F., and others. (2021) The composition and weathering of the continents over geologic time. Geochemical Perspectives Letters, 17, 21–26, https://doi.org/10.7185/geochemlet.2109.Search in Google Scholar

Litasov, K.D., Kagi, H., and Bekker, T.B. (2019) Enigmatic super-reduced phases in corundum from natural rocks: Possible contamination from artificial abrasive materials or metallurgical slags. Lithos, 340-341, 181–190, https://doi.org/10.1016/j.lithos.2019.05.013.Search in Google Scholar

Luth, R.W. (2003) Mantle volatiles-distribution and consequences. In R.W. Carlson, Ed., The Mantle and Core, p. 319–361. Elsevier-Pergamon.Search in Google Scholar

Magnus, P.D. (2012) Scientific Enquiry and Natural Kinds: From Mallards to Planets, 210 p. Palgrave MacMillan.Search in Google Scholar

Manning, C.E. and Frezzotti, M.L. (2020) Subduction-zone fluids. Elements, 16, 395–400, https://doi.org/10.2138/gselements.16.6.395.Search in Google Scholar

Marincea, S. and Dumitras, D.-G. (2019) Contrasting types of boron-bearing deposits in magnesian skarns from Romania. Ore Geology Reviews, 112, 102952.Search in Google Scholar

Mills, S.J., Hatert, F., Nickel, E.H., and Ferraris, G. (2009) The standardization of mineral group hierarchies: Application to recent nomenclature proposals. European Journal of Mineralogy, 21, 1073–1080, https://doi.org/10.1127/0935-1221/2009/0021-1994.Search in Google Scholar

Miyashiro, A. (1961) Evolution of metamorphic belts. Journal of Petrology, 2, 277–311, https://doi.org/10.1093/petrology/2.3.277.Search in Google Scholar

Moore, A.C. (1973) Carbonatite and kimberlites in Australia: A review of the evidence. Minerals Science and Engineering, 5, 81–91.Search in Google Scholar

Moores, E.M. (2002) Pre–1 Ga (pre-Rodinian) ophiolites: Their tectonic and environmental implications. Geological Society of America Bulletin, 114, 80–95, https://doi.org/10.1130/0016-7606(2002)114<0080:PGPROT>2.0.CO;2.Search in Google Scholar

Monchoux, P. (1972) Roches à sapphirine aux contact des lherzolites pyrénéennes. Contributions to Mineralogy and Petrology, 37, 47–64.Search in Google Scholar

Morrison, S.M. and Hazen, R.M. (2020) An evolutionary system of mineralogy. Part II: Interstellar and solar nebula primary condensation mineralogy (>4.565 Ga). American Mineralogist, 105, 1508–1535.Search in Google Scholar

Morrison, S.M. and Hazen, R.M. (2021) An evolutionary system of mineralogy, part IV: Planetesimal differentiation and impact mineralization (4.566 to 4.560 Ga). American Mineralogist, 106, 730–761, https://doi.org/10.2138/am-2021-7632.Search in Google Scholar

Morrison, S.M., Liu, C., Eleish, A., Prabhu, A., Li, C., Ralph, J., Downs, R.T., Golden, J.J., Fox, P., Hummer, D.R., and others. (2017) Network analysis of mineralogical systems. American Mineralogist, 102, 1588–1596, https://doi.org/10.2138/am-2017-6104CCBYNCND.Search in Google Scholar

Morrison, S.M., Prabhu, A., and Hazen, R.M. (2023) An evolutionary system of mineralogy, part VI: Earth’s earliest Hadean crust (>4370 Ma). American Mineralogist, 108, 42–58, https://doi.org/10.2138/am-2022-8329.Search in Google Scholar

Murdoch, J. (1951) Perovskite. American Mineralogist, 36, 573–580.Search in Google Scholar

Myer, G.H. (1966) New data on zoisite and epidote. American Journal of Science, 264, 364–385, https://doi.org/10.2475/ajs.264.5.364.Search in Google Scholar

Nakajima, Y., Uchida, E., Imai, H., and Ohno, H. (1992) Brucite-bearing white rock and the genetically related basalt dyke in the Nabeyama carbonate formation of the Kuzuu district, Tochigi Prefecture, Japan. Japanese Journal of Mineralogy and Petrology, 87, 445–459 (in Japanese).Search in Google Scholar

Newman, M.E.J. (2010) Networks: An Introduction, 784 p. Oxford University Press.Search in Google Scholar

Nutman, A.P. and Friend, C.R.L. (2007) Comment on “A vestige of Earth’s oldest ophiolite.” Science, 318, 746, https://doi.org/10.1126/science.1144148.Search in Google Scholar

Ohnmacht, W. (1974) Petrogenesis of carbonate-orthopyroxenites (sagvandites) and related rocks from Troms, northern Norway. Journal of Petrology, 15, 303–324, https://doi.org/10.1093/petrology/15.2.303.Search in Google Scholar

O’Reilly, S.Y. and Griffin, W.L. (2012) Mantle metasomatism. In D.E. Harlov and H. Austrheim, Eds., Metasomatism and the Chemical Transformation of Rock, 471–533. Springer.Search in Google Scholar

Palin, R.M. and White, R.W. (2016) Emergence of blueschists on Earth linked to secular changes in oceanic crust composition. Nature Geoscience, 9, 60–64, https://doi.org/10.1038/ngeo2605.Search in Google Scholar

Passchier, C.W. and Trouw, R.A.J. (2005) Microtectonics, 2nd ed. Springer.Search in Google Scholar

Pattison, D.R.M. (2001) Instability of Al2SiO5 “triple point” assemblages in muscovite+biotite+quartz-bearing metapelites, with implications. American Mineralogist, 86, 1414–1422, https://doi.org/10.2138/am-2001-11-1210.Search in Google Scholar

Pearson, D.G., Davies, G.R., Nixon, P.H., and Milledge, H.J. (1989) Graphitized diamonds from a peridotite massif in Morocco and implications for anomalous diamond occurrences. Nature, 338, 60–62, https://doi.org/10.1038/338060a0.Search in Google Scholar

Penniston-Dorland, S.C. and Ferry, J.M. (2006) Development of spatial variations in reaction progress during regional metamorphism of micaceous carbonate rocks, northern New England. American Journal of Science, 306, 475–524, https://doi.org/10.2475/07.2006.01.Search in Google Scholar

Peters, S.E. and Husson, J.M. (2017) Sediment cycling on continental and oceanic crust. Geology, 45, 323–326, https://doi.org/10.1130/G38861.1.Search in Google Scholar

Peters, T.A., Koestler, R.J., Peters, J.J., and Grube, C.H. (1983) Minerals of the Buckwheat dolomite, Franklin, New Jersey. The Mineralogical Record, 14, 183–194.Search in Google Scholar

Peters, S.E., Husson, J.M., and Czaplewski, J. (2018) Macrostrat: A platform for geological data integration and deep-time Earth crust research. Geochemistry, Geophysics, Geosystems, 19, 1393–1409, https://doi.org/10.1029/2018GC007467.Search in Google Scholar

Philpotts, A.R. and Ague, J.J. (2009) Principles of Igneous and Metamorphic Petrology, 2nd ed. Cambridge University Press.Search in Google Scholar

Pinger, A.W. (1950) Geology of the Franklin-Sterling area, Sussex county, New Jersey. International Geological Congress, 18, Part VII, 77–87.Search in Google Scholar

Pinsent, R.H. and Hirst, D.M. (1977) The metamorphism of the Blue River ultramafic body, Cassiar, British Columbia, Canada. Journal of Petrology, 18, 567–594, https://doi.org/10.1093/petrology/18.4.567.Search in Google Scholar

Post, J.E. (1999) Manganese oxide minerals: Crystal structures and economic and environmental significance. Proceedings of the National Academy of Sciences of the United States of America, 96, 3447–3454, https://doi.org/10.1073/pnas.96.7.3447.Search in Google Scholar

Raith, M. (1976) The Al-Fe(III) epidote miscibility gap in metamorphic profile through the Penninic series of the Tauren Window, Austria. Contributions to Mineralogy and Petrology, 57, 99–117, https://doi.org/10.1007/BF00392855.Search in Google Scholar

Rakovan, J. and Waychunas, G. (1996) Luminescence in minerals. The Mineralogical Record, 27, 7–19.Search in Google Scholar

Ramberg, H. (1952) The Origin of Metamorphic and Metasomatic Rocks, 335 p. University of Chicago Press.Search in Google Scholar

Read, H.H. (1923) Geology of the country around Banff, Huntly, and Turiff (Lower Banfffshire and North-West Aberdeenshire). Memoirs of the Geological Survey of Great Britain, 240 p. H.M. Stationery Office.Search in Google Scholar

Reverdatto, V.V. (1970) Pyrometamorphism of limestone and the temperature of basaltic magmas. Lithos, 3, 135–143, https://doi.org/10.1016/0024-4937(70)90069-1.Search in Google Scholar

Reverdatto, V.V. (1973) The Facies of Contact Metamorphism, 263 p. Australian National University (D.A. Brown, trans.).Search in Google Scholar

Ross, J.V., Bustin, R.M., and Rouzaud, J.N. (1991) Graphitization of high rank coals—the role of shear stain: Experimental considerations. Organic Geochemistry, 17, 585–596, https://doi.org/10.1016/0146-6380(91)90002-2.Search in Google Scholar

Roy, S. (1965) Comparative study of the metamorphosed manganese protores of the world: The problem of the nomenclature of the gondites and kodurites. Economic Geology and the Bulletin of the Society of Economic Geologists, 60, 1238–1260, https://doi.org/10.2113/gsecongeo.60.6.1238.Search in Google Scholar

Schertl, H.-P., Schreyer, W., and Chopin, C. (1991) The pyrope-coesite rocks and their country rocks at Parigi, Dora Maira Massif, Western Alps: Detailed petrography, mineral chemistry and P-T path. Contributions to Mineralogy and Petrology, 108, 1–21, https://doi.org/10.1007/BF00307322.Search in Google Scholar

Schertl, H.-P., Mills, S.J., and Maresch, W.V. (2018) A Compendium of IMA-Approved Mineral Nomenclature. International Mineralogical Association.Search in Google Scholar

Schreyer, W. (1977) Whiteschists: Their composition and pressure-temperature regime based on experimental, field, and petrographic evidence. Tectonophysics, 43, 127–144, https://doi.org/10.1016/0040-1951(77)90009-9.Search in Google Scholar

Schreyer, W., Ohnmacht, W., and Mannchen, J. (1972) Carbonate-orthopyroxenites (sagvandites) from Troms, northern Norway. Lithos, 5, 345–364, https://doi.org/10.1016/0024-4937(72)90089-8.Search in Google Scholar

Shedlock, R.J. and Essene, E.J. (1979) Mineralogy and petrology of a tactite near Helena, Montana. Journal of Petrology, 20, 71–97, https://doi.org/10.1093/petrology/20.1.71.Search in Google Scholar

Shimazaki, H. (1977) Grossular-spessartine-almandine garnets from some Japanese scheelite skarns. Canadian Mineralogist, 15, 74–80.Search in Google Scholar

Shirey, S.B., Cartigny, P., Frost, D.J., Keshav, S., Nestola, F., Nimis, P., Pearson, D.G., Sobolev, N.V., and Walter, M.J. (2013) Diamonds and the geology of mantle carbon. Reviews in Mineralogy and Geochemistry, 75, 355–421, https://doi.org/10.2138/rmg.2013.75.12.Search in Google Scholar

Simmons, E.C., Lindsley, D.H., and Papike, J.J. (1974) Phase relations and crystallization sequence in a contact-metamorphosed rock from the Gunflint Iron Formation, Minnesota. Journal of Petrology, 15, 539–565, https://doi.org/10.1093/petrology/15.3.539.Search in Google Scholar

Smith, D.G.W. (1969) A reinvestigation of the pseudobrookite from Havredal (Bamble), Norway. Norsk Geologisk Tidsskrift, 49, 285–288.Search in Google Scholar

Spear, F.S. (1982) Phase equilibria of amphibolites from the Post Pond Volcanics, Mt Cube Quadrangle, Vermont. Journal of Petrology, 23, 383–426, https://doi.org/10.1093/petrology/23.3.383.Search in Google Scholar

Spooner, E.T.C. and Fyfe, W.S. (1973) Sub-sea-floor metamorphism, heat and mass transfer. Contributions to Mineralogy and Petrology, 42, 287–304, https://doi.org/10.1007/BF00372607.Search in Google Scholar

Springer, R.K. (1974) Contact metamorphosed ultramafic rock in the Western Sierra Nevada foothills, California. Journal of Petrology, 15, 160–195, https://doi.org/10.1093/petrology/15.1.160.Search in Google Scholar

Spry, P.G., Plimer, I.R., and Teale, G.S. (2008) Did the giant Broken Hill (Australia) Zn- Pb-Ag deposit melt? Ore Geology Reviews, 34, 223–241, https://doi.org/10.1016/j.oregeorev.2007.11.001.Search in Google Scholar

Stachel, T., Aulbach, S., and Harris, J.W. (2022) Mineral inclusions in lithospheric diamonds. Reviews in Mineralogy and Geochemistry, 88, 307–391, https://doi.org/10.2138/rmg.2022.88.06.Search in Google Scholar

Stern, R.J. (2018) The evolution of plate tectonics. Philosophical Transactions of the Royal Society A: Mathematical, Physical, and Engineering Sciences, 376, 20170406, https://doi.org/10.1098/rsta.2017.0406.Search in Google Scholar

Stewart, F.H. (1942) Chemical data on a silica-poor argillaceous hornfels and its constituent minerals. Mineralogical Magazine and Journal of the Mineralogical Society, 26, 260–266, https://doi.org/10.1180/minmag.1942.026.178.04.Search in Google Scholar

Suzuki, K. (1977) Local equilibrium during the contact metamorphism of siliceous dolomites in Kasuga-Mura, Gifu-ken, Japan. Contributions to Mineralogy and Petrology, 61, 79–89, https://doi.org/10.1007/BF00375946.Search in Google Scholar

Sylvester, G.C. and Anderson, G.M. (1976) The Davis nepheline pegmatite and associated nepheline gneisses near Bancroft, Ontario. Canadian Journal of Earth Sciences, 13, 249–265, https://doi.org/10.1139/e76-027.Search in Google Scholar

Thompson, J.B. Jr. and Thompson, A.B. (1976) A model system for mineral facies in pelitic schists. Contributions to Mineralogy and Petrology, 58, 243–277, https://doi.org/10.1007/BF00402355.Search in Google Scholar

Tilley, C.E. (1926) On garnet in pelitic contact zones. Mineralogical Magazine and Journal of the Mineralogical Society, 21, 47–50, https://doi.org/10.1180/minmag.1926.021.113.03.Search in Google Scholar

Tilley, C.E. (1927) Vesuvianite and grossular as products of regional metamorphism. Geological Magazine, 64, 372–376, https://doi.org/10.1017/S0016756800103619.Search in Google Scholar

Tilley, C.E. (1948) Earlier stages in the metamorphism of siliceous dolomites. Mineralogical Magazine and Journal of the Mineralogical Society, 28, 272–276, https://doi.org/10.1180/minmag.1948.028.200.03.Search in Google Scholar

Tilley, C.E. (1951) The zoned contact skarns of the Broadford area, Skye: A study of boron-fluorine metasomatism in dolomites. Mineralogical Magazine and Journal of the Mineralogical Society, 29, 621–666, https://doi.org/10.1180/minmag.1951.029.214.01.Search in Google Scholar

Tilley, C.E. and Vincent, H.C.G. (1948) The occurrence of an orthorhombic high-temperature form of Ca2SiO4 (bredigite) in the Scawt Hill contact-zone and as a constituent of slags. Mineralogical Magazine and Journal of the Mineralogical Society, 28, 255–271, https://doi.org/10.1180/minmag.1948.028.200.02.Search in Google Scholar

Tilley, C.E., Nockolds, S.R., and Black, M. (1964) Harker’s Petrology for Students, 8th ed., 444 p. Cambridge University Press.Search in Google Scholar

Trouw, R.A.J., Passchier, C.W., and Wiersma, D.J. (2009) Atlas of Mylonites and Related Microstructures, 322 p. Springer.Search in Google Scholar

Tulloch, A.J. (1979) Secondary Ca-Al silicates as low-grade alteration products of granitoid biotite. Contributions to Mineralogy and Petrology, 69, 105–117, https://doi.org/10.1007/BF00371854.Search in Google Scholar

Turner, F.J. (1967) Thermodynamic appraisal of Nuckols steps in progressive metamorphism of siliceous dolomitic limestones. Neues Jahrbuch fur Mineralogie, 1967, 1–22.Search in Google Scholar

Tuttle, O.F. and Harker, R.I. (1957) Synthesis of spurrite and the reaction wollastonite + calcite → spurrite + carbon dioxide. American Journal of Science, 255, 226–234, https://doi.org/10.2475/ajs.255.3.226.Search in Google Scholar

Van Valkenburg, A. (1961) Synthesis of the humites nMg2SiO4·Mg(F, OH)2. Journal of Research of the National Bureau of Standards A: Physics and Chemistry, 65A, 415–428, https://doi.org/10.6028/jres.065A.043.Search in Google Scholar

Vernon, R.H. (2008) Principles of Metamorphic Petrology, 478 p. Cambridge University Press.Search in Google Scholar

Warner, R.D. and Luth, W.C. (1973) Two-phase data for the join monticellite (CaMgSiC4)–forsterite (Mg2SiO4): Experimental results and numerical analysis. American Mineralogist, 58, 998–1008.Search in Google Scholar

Watters, W.A. (1958) Some zoned skarns from granite-marble contacts near Puyvalador, in the Querigut area, eastern Pyrenees, and their petrogenesis. Mineralogical Magazine and Journal of the Mineralogical Society, 31,703–725, https://doi.org/10-1180/minmag.1958.031.240.01.Search in Google Scholar

Weeks, W.F. (1956) A thermochemical study of equilibrium relations during metamorphism of siliceous carbonate rocks. The Journal of Geology, 64, 245–270, https://doi.org/10.1086/626347.Search in Google Scholar

White, A.J.R. (1959) Scapolite-bearing marbles and calc-silicate rocks from Tungkillo and Milendelia, South Australia. Geological Magazine, 96, 285–306, https://doi.org/10.1017/S0016756800060763.Search in Google Scholar

Whitney, D.L. (2002) Coexisting andalusite, kyanite, and sillimanite: Sequential formation of three Al2SiC5 polymorphs during progressive metamorphism near the triple point, Sivrihisar, Turkey. American Mineralogist, 87, 405–416, https://doi.org/10.2138/am-2002-0404.Search in Google Scholar

Wilson, M.J. (2013) Rock-Forming Minerals, Vol. 3C: Sheet Silicates: Clay Minerals, Second ed., 724 p. The Geological Society of London.Search in Google Scholar

Wing, B.A., Ferry, J.M., and Harrison, T.M. (2003) Prograde destruction and formation of monazite and allanite during contact and regional metamorphism of pelites: Petrology and geochronology. Contributions to Mineralogy and Petrology, 145, 228–250, https://doi.org/10.1007/s00410-003-0446-1.Search in Google Scholar

Woodland, A.W. (1939) The petrography and petrology of the Lower Cambrian manganese ore of west Merionethshire. Quarterly Journal of the Geological Society of London, 95, 1–36, https://doi.org/10.1144/GSL.JGS.1939.065.01-04.03.Search in Google Scholar

Zedgenizov, D.A., Shatskiy, A., Ragozin, A.L., Kagi, H., and Shatsky, V.S. (2014) Merwinite in diamond from São Luiz, Brazil: A new mineral of the Ca-rich mantle environment. American Mineralogist, 99, 547–550, https://doi.org/10.2138/am.2014.4767.Search in Google Scholar

Zen, E. (1969) The stability relations of the polymorphs of aluminum silicate: A survey and some comments. American Journal of Science, 267, 297–309.Search in Google Scholar

Zhai, M., Zhao, G., and Zhang, Q. (2002) Is the Dongwanzi complex an Archean ophiolite? Science, 295, 923, https://doi.org/10.1126/science.295.5557.923a.Search in Google Scholar

Zharikov, V.A. and Shmulovich, K.L. (1969) High temperature mineral equilibria in the system CaO-SiO2-CO2. Geochemistry International, 6, 853–869.Search in Google Scholar

Zheng, Y.-F. and Chen, R.-X. (2017) Regional metamorphism at extreme conditions: Implications for orogeny at convergent plate margins. Journal of Asian Earth Sciences, 145, 46–73, https://doi.org/10.1016/j.jseaes.2017.03.009.Search in Google Scholar

Appendix I. Systematic mineralogy of metamorphic minerals

Appendix I presents a systematic mineralogy of the 94 most frequently encountered metamorphic minerals. Online Materials[1] Table S1 provides a list of 1220 metamorphic mineral species, the corresponding 755 metamorphic mineral kinds, and the distribution of these phases among 8 major groups of metamorphic rocks. This conversion of 1220 metamorphic minerals into 755 natural kinds requires several modifications to the IMA list, as detailed in Hazen et al. (2022). In 568 instances, the IMA species name (e.g., augite) is identical to the natural kind name (augite). Note that in this contribution we italicize the names of mineral natural kinds to distinguish them from IMA-CNMNC-approved mineral species. In the case of 652 IMA-approved species, we lump groups of two or more IMA species into single natural kinds. These numerous examples, resulting in a reduction from 652 species to 187 root natural kinds, are detailed in Online Materials[1] Table S2 (see also Online Materials[1] Read-Me File 2). For example, pumpellyite combines 9 IMA-approved species of the pumpellyite group. In 18 instances, the name assigned to the natural kind is a group name that is not itself an approved IMA species name. Thus, hornblende lumps 26 IMA-approved species of calcic amphiboles, while biotite encompasses 6 species of Fe-bearing trioctahedral micas.

We include five phases that do not correspond to an IMA-CNMNC-approved species in Online Materials[1] Table S1. In the instance of plagioclase, we recognize intermediate compositions of the albite (NaAlSi3C8)–anorthite (CaAl2Si2C8) solid solution with 0.15 < Ca/(Ca+Na) < 0.85 as a separate natural kind. Similarly, olivine refers to intermediate compositions of the forsterite (Mg2SiO4)–fayalite (Fe2SiO4) solid solution for which 0.3 < Fe/(Fe+Mg) < 0.7. Fe-Dolomite refers to intermediate compositions of the dolomite–ankerite solid solution, where 0.15 < Fe/(Fe+Mg) < 0.50, many examples of which are incorrectly described as ankerite (Ferry et al. 2015). Phengite is a fine-grained variety of muscovite [K(Al,Mg,Fe)2–3(AlSi3)C10(CH)2], typically with excess Si, that is common in high-pressure metamorphic deposits. We also introduce silicate glass, an important amorphous phase in some pyrometamorphic lithologies, as a natural kind.

Of the 755 mineral kinds recorded in Online Materials[1] Tables S1 and S2, 94 phases are relatively common based on their occurrence in at least 10 rocks in 2785 metamorphic rocks (Table 3; Online Materials[1] Table S3). Online Materials[1] Table S3 also records a literature reference for each rock mode, the rock’s locality, and information on the type of metamorphism, the metamorphic facies, and its protolith. Here we present brief descriptions of these 94 mineral kinds, arranged according to the New Dana Classification (Gaines et al. 1997).

Native elements

Two polymorphs of carbon (C), graphite and diamond, with 105 and 10 occurrences in Online Materials[1] Table S3, respectively, are the only native element minerals that occur in any significant abundance in metamorphic rocks, though more than two dozen rare native elements and metal alloys are listed in Online Materials[1] Table S1. Graphite most frequently occurs in carbon-rich metapelites (Landis 1971; Diessel et al. 1978; Buseck and Huang 1985). However, graphite is also reported to occur as a product of retrograde metamorphism of diamond, at times as euhedral pseudomorphs (Pearson et al. 1989; Ferry 1992; Davies et al. 1993; Leech and Ernst 1998). In addition, Ross et al. (1991) reported an occurrence of graphite formed by the shearing of coal—an example of Phanerozoic metamorphism of a biotic protolith.

Most diamond formation occurs in the mantle by precipitation from carbon-rich fluids (Jacob and Mikhail 2022; Kjarsgaard et al. 2022). However, in some instances, micrometer-scale diamond forms during subduction, ultra-deep metamorphism, and subsequent rebound of carbon-bearing crustal wedges (Dobrzhinetskaya et al. 1995, 2022).

Sulfides

More than 50 sulfide and other chalcogenide minerals have been reported from metamorphic rocks (Online Materials[1] Table S1). However, only four iron-bearing sulfides, pyrite (FeS2), pyrrhotite (Fe7S8), chalcopyrite (CuFeS2), and pentlandite [(Ni,Fe)9S8], are relatively common. Pyrrhotite is the most commonly observed sulfide in our survey, occurring in 192 metamorphic rocks. It frequently occurs with graphite in higher-grade regional metamorphic rocks (Hoschek 1984; Ferry 1992), and it may form by solid-state transformation of pyrite coupled with sulfur loss (Bowles et al. 2011). It is also a common sulfide mineral associated with skarn deposits (Einaudi and Burt 1982).

Pyrite (FeS2) is the next most frequently reported metamorphic sulfide, with 120 occurrences in Online Materials[1] Table S3. Pyrite is found in all metamorphic grades, including high-pressure regimes, and may form through solid-state transformation of pyrrhotite (Hall et al. 1987). Pyrite also occurs via contact metamorphism associated with skarn formation (Einaudi et al. 1981) and in shear zones (e.g., Harker 1950).

Chalcopyrite, the only copper-bearing mineral among the most common metamorphic phases, is reported from 99 rocks in our survey. Most occurrences are as minor grains in regionally metamorphosed igneous and sedimentary lithologies (Ferry 1984, 1992, 1994; Ferry et al. 2001). It is likely that chalcopyrite, as an opaque and often micrometer-scale phase, is underrepresented in our study.

The 21 occurrences of pentlandite, the only common nickel-bearing metamorphic minerals, are recorded exclusively from ultramafic lithologies (Ferry 1995; Ferry et al. 2005).

In addition, it should be noted that sphalerite (ZnS) occurs in 9 of the rocks surveyed—a number that likely significantly underestimates the frequency of this opaque and typically minute zinc-bearing phase.

Oxides

Oxide minerals occur in the entire range of metamorphic rocks. More than 50 species, most of them relatively rare, have been reported (Online Materials[1] Table S1). We list 13 mineral kinds among the most frequently encountered metamorphic minerals.

The simple oxide rutile (TiO2) is reported in 297 of the metamorphic rocks from our study (Online Materials[1] Table S3), most often in regional metamorphic rocks by transformation of prior Ti-bearing phases, including titanite, ilmenite, and titaniferous micas and magnetite (Bowles et al. 2011), but also in a range of pyrometamorphic (Grapes 2006), contact metamorphic (Reverdatto 1973), and shear zone (Harker 1950) rocks.

Additionally, among the more common simple oxides is corundum (Al2O3; 85 occurrences), observed most frequently in silica-poor lithologies subjected to high temperature, notably in contact metamorphic zones. Al-rich lithologies with prominent corundum, often in association with spinel, mullite, cordierite, and/or sanidine, are known as “emery.” Corundum also occurs in pyrometamorphosed limestone xenoliths (Joplin 1968; Grapes 2006), as well as in high-grade regional metamorphic rocks (Harker 1950; Augustithis 1985).

Periclase (MgO) is a common mineral in calcite/dolomite marbles, often in association with brucite (Carpenter 1967; Bowles et al. 2011). As early as 1940, Bowen described periclase as part of an evolutionary thermal metamorphic sequence (Bowen 1940). We record 36 occurrences of periclase, primarily in limestone xenoliths and contact metamorphic environments (Augustithis 1985; Ferry and Rumble 1997; Grapes 2006).

Hematite (Fe2O3), represented by 24 occurrences in our tabulation, forms from Fe-rich protoliths in oxidized environments. Contexts of metamorphic hematite include xenoliths, contact environments, iron formations, and regional metamorphism (Bowles et al. 2011).

We record 18 occurrences of the zirconium oxide baddeleyite (ZrO2) in both contact and regional metamorphic contexts (Ferry 2007). That number that may underrepresent the frequency of baddeleyite because of its typical low modal abundance and small grain size.

The most frequently encountered metamorphic oxides are members of the spinel group, which occur in 672 (24%) of the 2785 metamorphic rocks we tabulated. With the general formula [(Mg,Fe2+)(Al,Fe3+,Cr3+,Ti)2O4], the spinel group encompasses a complex range of solid solutions (Bowles et al. 2011), of which four end-members are among the most commonly reported metamorphic minerals: magnetite (Fe2+Fe23+O4), spinel (MgAl2O4; although “spinel” may also refer to the mineral group rather than the species in some modes), chromite (Fe2+Cr23+O4), and hercynite (Fe2+Al2O4). In addition, several reports cite “pleonaste,” which refers to Fe2+-bearing intermediate compositions of the spinel-hercynite solid solution-examples that we include with spinel.

Magnetite is by far the most frequently reported metamorphic oxide, occurring in 429 (15%) of the 2785 metamorphic rocks we compiled, and in a wide range of metamorphic contexts (Joplin 1968; Reverdatto 1973; Grapes 2006). Magnetite often forms by thermal alteration of ferric iron oxide/hydroxides, as well as through the introduction of Fe-rich fluids—a situation in which the distinction between metamorphism and metasomatism may be blurred.

The Mg-Al oxide spinel, with 263 occurrences in our list, is found in numerous contact and regional metamorphic contexts, principally as a high-temperature mineral in metacarbonates and Al-rich protoliths (Harker 1950; Botha 1983; Carswell 1990; Grapes 2006).

Hercynite, which forms in high-grade metamorphic environments, often in association with corundum, mullite, and/or sillimanite, is represented by 34 examples in Online Materials[1] Table S3. Metamorphic contexts include xenolith, contact, and regional environments (Harker 1950; Grapes 2006; Bowles et al. 2011).

We list 12 occurrences of metamorphic chromite—a number that likely underestimates this most common chromium mineral because it is opaque and easily mistaken for other Fe-bearing oxides. Chromite is most often associated with ultramafic lithologies, particularly ophiolites.

Four titanium-bearing double oxides are listed among the 94 most common metamorphic phases. We record ilmenite (FeTiO3) from 181 xenoliths, contact metamorphic rocks, or regional metamorphic formations, notably forming via alteration of mafic and Fe-bearing lithologies (Reverdatto 1973; Carswell 1990; Grapes 2006). Geikielite (MgTiO3), the magnesium analog of ilmenite, occurs in 17 contact and regional metasedimentary rocks in Online Materials[1] Table S3. Geikeilite protoliths include carbonates, calc silicates, pelites, and sandstones. Perovskite (CaTiO3) is reported from 12 rocks, primarily in xenoliths and contact metamorphic environments with impure limestone (Murdoch 1951; Fulignati et al. 2000; Grapes 2006). Pseudobrookite (Fe23+TiO5), which we record in 15 primarily xenolith and contact metamorphic rocks, forms most often by the high-temperature oxidation of ilmenite (Agrell and Langley 1958; Smith 1969; Basta and Shaalan 1974).

Brucite [Mg(OH)2], the only relatively common hydroxide mineral in metamorphic rocks, occurred in 31 rocks of our survey. It is typically the consequence of hydration of periclase in altered Ca-Mg carbonates (Nakajima et al. 1992; Ferry and Rumble 1997; Ferry et al. 2002; Bowles et al. 2011).

Carbonates

At least 70 carbonate minerals have been reported from metamorphic rocks (Online Materials[1] Table S1), though only seven species occur with any significant frequency. By far the most abundant carbonates are calcite (CaCO3), dolomite [CaMg(CO3)2], and the intermediate composition Fe-dolomite [dolomite with a significant ankerite CaFe(CO3)2 component], with 645, 152, and 116 occurrences in our compilation, respectively. Calcite and dolomite most often occur in metamorphosed limestones and other carbonate-bearing protoliths (Harker 1950; Reverdatto 1973; Chang et al. 1996; Grapes 2006). In many instances, these phases form through recrystallization of prior carbonates to form a marble (Chang et al. 1996; Philpotts and Ague 2009). Fe-dolomite (often reported as “ankerite” in the petrologic literature) is found primarily in regionally metamorphosed sediments (Ferry 1992, 1994, 2007; Ferry et al. 2015) where it formed through carbonation reactions during metamorphism/ metasomatism (Spooner and Fyfe 1973).

Metamorphic magnesite (MgCO3) with 16 occurrences is characteristic of altered ophiolites, where it occurs in association with talc and serpentine. It typically forms via the carbonation of Mg-bearing oxides and silicates (Chang et al. 1996).

Aragonite, a high-pressure form of CaCO3, was recorded in 15 rocks of eclogite, blueschist, and ultrahigh-pressure facies carbonate rocks (Carswell 1990; Carswell and Compagnoni 2003; Philpotts and Ague 2009).

Two silicate carbonates, spurrite [Ca5(SiO4)2(CO3)] with 51 examples and tilleyite [Ca5Si2O7(CO3)2] with 16 examples, arise when calcite and wollastonite react at high temperature (Tuttle and Harker 1957; Zharikov and Shmulovich 1969).

Phosphates

Of the more than 50 metamorphic phosphates recorded in Online Materials[1] Table S1, only the calcium phosphate apatite [Ca5(PO4)3(F,OH)] is widely reported, with 155 occurrences in Online Materials[1] Table S3. We lump two common species, fluorapatite and hydroxylapatite, which are rarely differentiated in reports of metamorphic mineral modes. The majority of these apatite occurrences are in high-grade metamorphosed mafic igneous rocks, including granulites and eclogites (Harker 1950; Joplin 1968; Carswell 1990). In addition, apatite has been reported from contact metamorphic and shear environments (Harker 1950; Joplin 1968).

Silicates

Silicates constitute the majority of metamorphic minerals, both volumetrically and in terms of diversity. In Online Materials[1] Tables S1 and S2 we record 746 silicate mineral species, corresponding to 418 root natural kinds, of which 66 are frequently encountered metamorphic phases. In the following sections we review these more common silicates.

Nesosilicates or orthosilicates

Orthosilicates, with silicon exclusively in insular SiO44 structural groups, are characteristic minerals in environments with relatively low Si, notably those associated with carbonate, calc-silicate, or aluminous protoliths (Deer et al. 1982). We detail 21 orthosilicates in addition to the orthosilicate-carbonate mineral spurrite, described above. Of these 22 phases, 16 contain essential Ca and/or Al.

Olivine Group: We recognize four members of the olivine group [(Mg,Fe,Ca)2SiO4] as important metamorphic minerals (Deer et al. 1982). The Mg olivine forsterite (ideally Mg2SiO4) is reported in 173 of the rocks we surveyed, notably via contact, regional, or high-pressure metamorphism of silica-poor ultramafic (Springer 1974; Pinsent and Hirst 1977) or carbonate-bearing (Weeks 1956; Schreyer et al. 1972; Suzuki 1977) protoliths. We also distinguish olivine as intermediate Mg-Fe compositions with 0.3 < Fe/(Fe+Mg) < 0.7, which are common in metamorphosed mafic and ultramafic rocks (53 occurrences; Reverdatto 1973; Ferry et al. 1987).

Fayalite (ideally Fe2SiO4) is much less common, occurring in 10 rocks with Fe-rich protoliths, including mafic rocks and iron formations (Joplin 1968; Simmons et al. 1974; Floran and Papike 1978; Carswell 1990). Note that metamorphic Fe-dominant olivines with intermediate compositions are less common than examples close to either end-member (Deer et al. 1982).

Monticellite (CaMgSiO4), with 77 occurrences, is commonly found in contact metamorphic environments with siliceous carbonate lithologies, often forming with increasing temperature at the expense of diopside, forsterite, and/or wollastonite (Bowen 1940; Turner 1967; Deer et al. 1982). Monticellite frequently co-occurs with forsterite, as the solid solution between these two olivine group minerals is limited (Warner and Luth 1973).

Garnet Group: The garnet group is represented by four relatively common metamorphic phases, occurring in 745 (27%) of 2785 rocks in our survey. Garnets collectively display a significant compositional range, typically with solid solutions among two or three end-members (Deer et al. 1982; Chiama et al. 2020, 2022). Ideal end-members of these minerals are almandine [Fe32+Al2(SiO4)3], andradite [Ca3Fe32+(SiO4)3], grossular [Ca3Al2(SiO4)3], and pyrope [Mg3Al2(SiO4)3], often with a significant spessartine [Mn32+Al2(SiO4)3] component, as well, though true Mn-dominant spessartine is recorded in only 9 occurrences in our compilation (Woodland 1939; Roy 1965; Jan and Symes 1977).

In some instances, such as pyrope-almandine-spessartine (“pyralspite”) from ecologites and other high-grade metamorphic rocks, grossular-andradite (“grandite”) from the contact metamorphism of carbonate-bearing sediments, and contact metamorphic garnets in the grossular-spessartine-almandine field (Shimazaki 1977), the compositional ranges among end-members may be continuous, thus warranting lumping of species into a single metamorphic mineral kind. However, until cluster analysis (Gregory et al. 2019; Boujibar et al. 2021; Hystad et al. 2021) can be performed on a wide range of garnet compositions from known paragenetic environments, we will treat these five types of metamorphic garnet separately.

Almandine, with 367 occurrences in Online Materials[1] Table S3, is the commonest garnet in metamorphic rocks (Harker 1950; Joplin 1968; Reverdatto 1973; Botha 1983; Augustithis 1985). Most almandine forms in a regional metamorphic context, derived from mafic or pelitic protoliths (Atherton 1964; Deer et al. 1982), including high-pressure examples from blueschist (Coleman and Lee 1963; Banno and Matsui 1965), eclogite (Coleman et al. 1965), and granulite (Buddington 1952; Eskola 1952) facies. In addition, almandine from contact metamorphism of pelites is not uncommon (Tilley 1926; Stewart 1942), while it also occurs in some metamorphosed iron formations (Klein 1966).

Pyrope’s 192 entries are overwhelmingly from high-pressure metamorphic environments, in many instances from ecologite-grade rocks with mafic precursors, most commonly in association with omphacite (Carswell 1990). Metamorphic pyrope typically has a significant almandine component (Deer et al. 1982).

The great majority of 162 grossular occurrences in Online Materials[1] Table S3 arise from contact metamorphism of calcareous rocks, often in association with diopside and/or wollastonite (Watters 1958; Reverdatto 1973), or in regional metamorphic formations, also with carbonate-bearing protoliths (Tilley 1927; Sylvester and Anderson 1976). In addition, 28 occurrences of the Ca-Fe3+ garnet andradite arise predominantly from contact and regional metamorphism of calc-silicate rocks (Harker 1950; White 1959; Shedlock and Essene 1979). In several instances, contact metamorphic garnets have so-called “grandite” compositions intermediate between grossular and andradite (Coombs et al. 1977; Tulloch 1979).

Three additional Ca-Mg orthosilicates, bredigite [Ca7Mg(SiO4)4] with 6 occurrences (Tilley and Vincent 1948; Grapes 2006), larnite (Ca2SiO4) with 24 occurrences (Deer et al. 1986), and merwinite [Ca3Mg(SiO4)2] with 38 occurrences (Larsen and Foshag 1921; Reverdatto 1973), are frequently found in contact metamorphosed calc-silicate protoliths, often in association with the calc-silicates melilite, rankinite, and spurrite (Joplin 1968; Deer et al. 1986; Grapes 2006). Merwinite is also reported as an ultrahigh-pressure mantle phase (Zedgenizov et al. 2014).

We lump three compositionally similar IMA species—humite, clinohumite, and hydroxylclinohumite—into humite [Mg7–9(SiO4)4(F,OH)2]. Members of the humite group differ in the ratios of two structural modules, one of forsterite composition [Mg2(SiO4)] and the other of brucite composition [Mg(OH,F)2]. Reports of metamorphic mineral modes seldom distinguish between humite [Mg9(SiO4)4(OH,F)2] and clinohumite [Mg7(SiO4)4(OH,F)2], nor between the OH- and F-dominant species. We record 21 occurrences of humite, all of which are characteristic of the contact metamorphism of dolomite-bearing sediments (Tilley 1951; Joplin 1968; Deer et al. 1982). Note that two other members of the humite group (Van Valkenburg 1961), norbergite [Mg3(SiO4)4(OH,F)2] and chondrodite [Mg5(SiO4)4(OH,F)2], are also contact metamorphic minerals, but did not appear as common phases in our tabulations of metamorphic rock modes.

Three aluminosilicate (Al2SiO5) polymorphs, andalusite, kyanite, and sillimanite, are abundant constituents of many metapelites, with 146, 102, and 235 occurrences in Online Materials[1] Table S3, respectively. These phases, which can coexist at their invariant triple point (~500 °C and 0.4 GPa; Hodges and Spear 1982; Bohlen et al. 1991; Pattison 2001), have received special attention for their ability to document the pressure-temperature regimes of their host rocks (Barrow 1893; Zen 1969; Deer et al. 1982; Whitney 2002; Philpotts and Ague 2009). Sillimanite, the high-temperature, low-pressure polymorph, is a common phase in various metapelites subjected to hornblende hornfels, granulite, and pyrometamorphic (sanidinite) conditions (Reverdatto 1973; Botha 1983; Grapes 2006). Mullite [Al4+2xSi2–2xO10–x (x ≈ 0.4)] is also a high-temperature, low-pressure orthosilicate that we record from 62 pyrometamorphic rocks, often in association with sillimanite (Grapes 2006).

Andalusite forms in pelitic protoliths at low pressure and moderate temperature (<770 °C), notably from albite-epidote hornfels and hornblende hornfels facies (Read 1923; Guitard 1965; Reverdatto 1973). Kyanite, the highest-pressure crustal polymorph of Al2SiO5, is frequently encountered in regional and high-pressure metamorphic rocks with aluminous precursors (Carswell 1990; Carswell and Compagnoni 2003). The aluminosilicates may record either prograde or retrograde metamorphism. For example, Lal (1969) described andalusite and kyanite formed via retrograde metamorphism from cordierite-bearing rocks, and Gates and Speer (1991) record retrograde kyanite after sillimanite in metapelite shear zones.

Chloritoid [(Fe2+,Mg,Mn2+)Al2O(SiO4)(OH)2], with 46 occurrences in our tabulation, is most commonly formed by regional or high-pressure metamorphism of pelitic rocks (Joplin 1968; Carswell 1990). We lump three IMA-CNMNC-approved species, chloritoid [Fe2+Al2O(SiO4)(OH)2], magnesiochloritoid [MgAl2O(SiO4)(OH)2], and ottrelite [Mn2+Al2O(SiO4)(OH)2], because they form a continuous solid solution and they are rarely differentiated in reports of metamorphic rock modes. The broad pressure-temperature stability field of chloritoid leads to a wide range of assemblages, from low-grade, clay-mineral- and phengite-bearing facies to high-grade rocks with kyanite, pyrope-almandine, and/or staurolite (Halferdahl 1961).

Staurolite [(Fe2+,Mg)2Al9Si4O23(OH)] is another common phase derived by regional or high-pressure metamorphism of pelitic sediments. Its 52 occurrences in Online Materials[1] Table S3 reflect a range of P-T conditions of formation, from low-grade assemblages with chloritoid and quartz, medium-grade assemblages with almandine and kyanite, and high-grade assemblages with sillimanite and plagioclase (Deer et al. 1982; Augustithis 1985; Carswell 1990). Staurolite is also observed in the contact metamorphism of pelites (Reverdatto 1973; Grapes 2006).

Titanite (CaTiSiO5; commonly reported as “sphene”) occurs as a minor phase in 186 metamorphic rocks in our survey in various contexts (Harker 1950; Joplin 1968; Reverdatto 1973; Carswell 1990). Zircon (ZrSiO4), another volumetrically minor phase, is listed in 66 of 2785 rocks in our tabulation, including a wide range of contact and regional metamorphic lithologies (Joplin 1968; Carswell 1990; Grapes 2006). Titanite and zircon are particularly durable accessory minerals that are widespread in igneous and sedimentary formations; therefore, their occurrence in metamorphic rocks sometimes derives from protolith minerals that have been little altered.

Though not among the more common metamorphic orthosilicates, willemite (Zn2SiO4) is an important mineral in some metamorphosed Pb-Zn deposits, such as the sillimanite-grade deposits at Franklin, New Jersey (Pinger 1950; Frondel 1990). In this case, which may be relevant to minerals in many metamorphic environments, willemite occurs both as a major phase in the ore and as a secondary phase in thin hydrothermal veins. These two generations of willemite, furthermore, have distinct properties: both forms are fluorescent, but only the secondary willemite has persistent luminescence as a consequence of its greater arsenic content (Rakovan and Waychunas 1996). With their distinct modes of formation and attributes, these coexisting forms of willemite represent two different mineral kinds in our evolutionary system.

Sorosilicates or disilicates

Sorosilicates incorporate the double-tetrahedron pyrosilicate group (Si2O76). We find 8 sorosilicate root natural kinds, corresponding to at least 35 IMA-CNMNC-approved species, among the 94 most frequently encountered metamorphic mineral kinds. All of these phases, in addition to the disilicate-carbonate mineral tilleyite described earlier, are calcium-bearing minerals that occur most frequently in the contact metamorphic zones of limestone and dolomite.

The most common metamorphic sorosilicates are from the diverse epidote group (Deer et al. 1986; Armbruster et al. 2006). We lump 4 monoclinic species (including Fe3+-bearing clinozoisite) into epidote [Ca2(Al2Fe3+)[Si2O7][SiO4]O(OH)] with 149 occurrences; 5 rare-earth element-bearing epidote group minerals into allanite [(CaCe)(AlAlFe2+)O[Si2O7][SiO4](OH)] with 16 occurrences (though certainly under-reported); and orthorhombic zoisite [Ca2Al3[Si2O7][SiO4]O(OH)] with 156 occurrences. In addition, Mn-bearing piemontite [Ca2Al2Mn3+(Si2O7)(SiO4)O(OH)] is an important phase in metamorphosed manganese deposits, though we record only 7 occurrences. Metamorphic epidote is found most commonly via contact metamorphism of carbonate-bearing sediments and mafic igneous rocks (Joplin 1968; Reverdatto 1973), but also in regional (Harker 1950), high-pressure (Carswell 1990), and xenolith (Grapes 2006) contexts. Note, however, that it may be difficult to distinguish metamorphic epidote and zoisite (see below) from examples formed by metasomatism (Joplin 1968).

Zoisite generally forms at lower metamorphic grades than epidote, though it can coexist with epidote in medium-grade regional metamorphic rocks derived from calcareous sediments or mafic igneous rocks (Myer 1966; Ackermand and Raase 1973; Raith 1976). Zoisite is also common in kyanite-bearing ecologite (Carswell and Compagnoni 2003), where it may form by a prograde reaction from lawsonite (Deer et al. 1986).

We lump two species, lawsonite and itoigawaite, into the root kind lawsonite [(Ca,Sr)Al2(Si2O7)(OH)2·H2O], with 11 occurrences in Online Materials[1] Table S3. Lawsonite forms exclusively in high-pressure blueschist or eclogite facies rocks (Philpotts and Ague 2009), commonly in association with glaucophane and an epidote group mineral, either zoisite or epidote (Coleman et al. 1965; Carswell 1990).

Rankinite (Ca3Si2O7), with 20 occurrences, is found exclusively in contact metamorphic rocks, commonly in association with larnite, melilite, spurrite, and/or wollastonite (Reverdatto 1973; Grapes 2006).

The melilite group includes the solid solution between åkermanite [Ca2(Al2SiO7)] and gehlenite [Ca2(MgSi2O7)], as well as alumoåkermanite [(Ca,Na)2(Al,Mg,Fe2+) (Si2O7)]—minerals that we lump into the root mineral kind melilite. With 108 occurrences in Online Materials[1] Table S3, melilite is a common mineral in pyrometamorphosed siliceous limestone and dolomite, particularly at pyroxene hornfels and sanidinite facies (Reverdatto 1973; Grapes 2006), often forming at the expense of diopside or anorthite (Bowen 1940; Reverdatto 1970).

Pumpellyite [Ca2(Al,Fe2+,Fe3+)3(Si2O7)(SiO4)(OH,O)2·H2O] lumps 9 closely related species of Ca-Al-Fe sorosilicates that are found most frequently in the low-grade zeolite and pumpellyite-prehnite facies of regional metamorphism. We list 10 occurrences, all of which occur in low-grade metapelites (Joplin 1968; Botha 1983; Augustithis 1985). However, Deer et al. (1986) note that Al-rich pumpellyite also occurs in blueschist facies, and Fe- and Mn-rich pumpellyite may occur in mineralized skarn zones.

We lump 10 IMA-CNMNC-approved species, most of which are rare compositional variants, into vesuvianite [(Ca,Na)19(Al,Mg,Fe)13(SiO4)10(Si2O7)4(OH,F,O)10]. We record 15 occurrences in contact metamorphosed limestone, in which it is a characteristic skarn mineral, commonly in association with diopside, grossular, and/or wollastonite (Harker 1950). Vesuvianite occurs less commonly in regional metamorphosed limestones (Tilley 1927; Deer et al. 1982). Note that, as with the example of epidote, it may be difficult to distinguish vesuvianite formed by metamorphism vs. metasomatism (Joplin 1968).

Cyclosilicates

Members of the cordierite, tourmaline, and osulmilite groups are relatively common metamorphic cyclosilicates. We lump the species cordierite (with Mg) and sekaninaite (with Fe2+) into the root mineral kind cordierite [(Mg,Fe2+)2Al4Si5O18], which, with 395 occurrences in Online Materials[1] Table S3, is among the most common minerals in contact and regionally metamorphosed pelites (Joplin 1968; Reverdatto 1973; Botha 1983; Grapes 2006). Deer et al. (1986) detail a wide range of cordierite parageneses, including pyrometamorphosed xenoliths, contact metamorphosed argillaceous sediments, and a range of regional metamorphic facies, including low-pressure, high-temperature assemblages with andalusite; moderate-pressure assemblages with sillimanite and garnet; and high-pressure assemblages with kyanite.

Tourmaline [(Na,Ca,□)(Mg,Al,Fe2+,Fe3+,Ti4+)3(Al,Fe3+,Mg)6(Si6O18)(BO3)3 (OH)3(O,F)] is the only common boron-bearing mineral in metamorphic rocks (Deer et al. 1986; Henry et al. 2011; Henry and Dutrow 2012). We lump 18 IMA-CNMNC-approved species of the tourmaline group (Online Materials[1] Table S1), all of which have been reported from regional metamorphic environments. The 47 tourmaline occurrences listed in Online Materials[1] Table S3, including several examples of tourmaline-dominant tourmalinites, are from metapelites (Harker 1950; Joplin 1968). Joplin (1968) suggests that metamorphic tourmaline occurs in 3 distinct ways: as a remnant mineral of the protolith, through metamorphism of a borate-containing lithology, or as the result of boron metasomatism.

Osumilite [(K,Na)(Fe2+,Mg)2(Al,Fe3+)3(Si,Al)12O30] was reported in 13 of our mineral modes, typically from ultrahigh-temperature metamorphosed pelites, in which it commonly occurs with cordierite, orthopyroxene, sanidine, and/or sillimanite (Harley 2021).

Inosilicates

Among the 94 relatively common metamorphic minerals listed in Table 2, 16 are chain silicates, including several members of the pyroxene (7 kinds) and amphibole (7 kinds) groups (Deer et al. 1997a, 1997b). Inosilicates are one of the most common classes of metamorphic minerals, occurring in 1387 (50%) of the 2785 metamorphic rocks in Online Materials[1] Table S3.

Pyroxene Group: We consider 7 root kinds of pyroxene group single-chain silicates, most of which lie in or near the [(Ca,Mg,Fe)2Si2O6] quadrilateral (Deer et al. 1997a).

Orthopyroxene lumps 281 occurrences of orthorhombic pyroxenes, most often described as enstatite (the Mg end-member) or “hypersthene” (with Mg~Fe2+) but sometimes “bronzite” (with Mg > Fe2+) or ferrosilite (the Fe2+ end-member), always lying close to the MgSiO3–Fe2+SiO3 binary. Orthopyroxene most often occurs in granulite, ecologite, and UHT facies of metamorphosed ultramafic and mafic igneous rocks, in which it is often the most abundant mafic phase (Joplin 1968; Augustithis 1985; Carswell and Compagnoni 2003). It also occurs in lower-grade regional metamorphic rocks (Joplin 1968; Botha 1983), with iron-rich examples in metamorphosed iron formations (Kranck 1961; Simmons et al. 1974). Orthopyroxene is not uncommon in contact metamorphic environments, including pyrometamorphosed xenoliths (Reverdatto 1973; Grapes 2006). In some instances, Mg-rich orthopyroxene is associated with carbonate minerals (Schreyer et al. 1972; Ohnmacht 1974).

Three Ca-bearing clinopyroxenes, diopside [Ca(Mg,Fe2+)Si2O6; 409 occurrences in Online Materials[1] Table S3), hedenbergite (CaFe2+Si2O6; 16 occurrences), and the ternary solid solution augite [(Ca,Mg,Fe2+)2Si2O6, typically with 0.5 < Ca/(Mg+Fe) < 0.9; 150 occurrences], are common in a wide variety of metamorphic rocks (Deer et al. 1997a). Pyroxenes close to the continuous CaMg–CaFe2+ solid solution between diopside and hedenbergite are most typical of thermally metamorphosed carbonate and calc-silicate rocks, occurring in xenoliths and contact metamorphic contexts (Grapes 2006). We define diopside broadly to include most intermediate Mg-Fe compositions (e.g., “salite” and “ferrosalite”), as well as Al-bearing “fassaite.” More than 80% of occurrences of diopside, the most abundant pyroxene in our survey, arise from contact metamorphism. Diopside also occurs in regional and high-pressure metamorphic rocks, with several examples from amphibolite facies (Harker 1950; Augustithis 1985) and eclogite facies (Carswell 1990; Carswell and Compagnoni 2003).

Hedenbergite displays much the same parageneses as diopside, but with Fe-rich protoliths (Joplin 1968; Augustithis 1985). We debated whether to lump these two end-members, but hedenbergite appears to form a discrete cluster of metamorphic clinopyroxenes with low Mg. Cluster analysis of igneous and metamorphic clinopyroxenes represents an important future research goal.

The IMA-CNMNC-approved species augite, including clinopyroxenes in the [(Ca,Mg,Fe2+)2Si2O6] system with Ca occupying between ~25 and ~45 atom percent of the Ca-Mg-Fe sites, is equivalent to our root kind augite. The 150 occurrences in Online Materials[1] Table S3, while mostly from contact metamorphism or pyrometamorphism of ultramafic/mafic lithologies (Reverdatto 1973; Grapes 2006), also include representatives derived by regional metamorphism of mafic and intermediate igneous protoliths (Harker 1950; Joplin 1968; Ferry et al. 1987).

Three Na-bearing kinds of clinopyroxene, aegirine [(Ca,Na)(Fe3+,Mg,Fe2+) Si2O6; 16 occurrences], jadeite (NaAlSi2O6; 34 occurrences), and omphacite [(Ca,Na) (Mg,Fe,Al)Si2O6; 151 occurrences] are especially characteristic of high-pressure metamorphic environments. Aegirine, in which we lump two IMA-CNMNC-approved species aegirine (NaFe3+Si2O6) and aegirine-augite [(Ca,Na)(Fe3+,Mg,Fe2+)Si2O6], is the least common of these phases in metamorphic rocks, being found primarily in the context of mafic and intermediate igneous rocks subjected to ultrahigh pressure and ecologite facies (Carswell 1990), though aegirine also occurs via contact metamorphism of alkaline rocks (Reverdatto 1973). A complication is the formation of aegirine through sodium metasomatism of prior pyroxenes (Moore 1973; Deer et al. 1997a).

Jadeite is a diagnostic phase found exclusively in high-pressure metamorphic environments, including blueschist facies, eclogite facies, and ultrahigh-pressure metamorphic rocks (Carswell 1990). It often forms through the iconic reaction albitejadeite + quartz (Deer et al. 1997a, and references therein). Jadeite commonly incorporates up to 15 mol% of an aegirine/omphacite component; however, a significant compositional gap separates jadeite from these phases, which often coexist in high-pressure assemblages (Coleman and Clark 1968).

Omphacite, which represents a solid solution among aegirine, diopside, and jadeite, is a relatively common phase in high-pressure metamorphic rocks. All 151 occurrences in our tabulation were reported from blueschist, eclogite, or ultrahigh-pressure environments, most often with ultramafic or mafic protoliths and often in association with glaucophane, pyrope/almandine, and quartz/coesite (Carswell 1990; Carswell and Compagnoni 2003).

Pyroxenoid Group: Two members of the inosilicate pyroxenoid group, wollastonite (CaSiO3; 158 occurrences) and the Mn-bearing rhodonite [CaMn3Mn(Si5O15); 8 occurrences, not listed among the top 94 phases], are most commonly associated with skarn zones. Almost all wollastonite reports are from carbonate or calc-silicate protoliths subjected to pyroxene hornfels or sanidinite facies metamorphism (Reverdatto 1973; Grapes 2006).

Rhodonite, which lumps 4 closely related species of Mn pyroxenoids, is most frequently encountered in the high-pressure metamorphic environments of Mn-rich protoliths (Carswell 1990), notably by reaction of rhodochrosite (MnCO3; Hori 1962), though it can also form through Mn metasomatism (Bilgrami 1956).

Amphibole Group: The amphibole group of double-chain silicates, which boasts more than 110 IMA-CNMNC-approved species (https://rruff.info/ima; accessed 13 January 2023), is likely the most diverse of all mineral structure types (Deer et al. 1997b; Hawthorne et al. 2012). Here, we provisionally lump 55 amphibole species known to occur in metamorphic rocks into 7 root mineral kinds. It should be noted, however, that the variety of amphibole parageneses, coupled with the extensive and complex solid solutions and miscibility gaps among many species, render any suggestion of amphibole mineral kinds tentative, at best. In the context of metamorphism, we have yet to determine if different facies, different protoliths, effects of metasomatism, prograde vs. retrograde formation, and other factors may yield numerous distinct combinations of paragenesis and attributes. We require data resources with analyses of tens of thousands of well-characterized amphibole specimens, coupled with advanced methods of cluster analysis (Boujibar et al. 2021; Hystad et al. 2021). Such an epic endeavor could represent a lifetime of fruitful study for an ambitious young mineralogist.

Anthophyllite [□(Mg,Fe2+)2(Mg,Fe2+,Fe3+,Al)5(Si,Al)8O22(OH)2; with 52 occurrences in Online Materials[1] Table S3], lumps 6 species of the complex anthophyllite/ferro-anthophyllite/gedrite/ferro-gedrite solid solution of orthorhombic amphiboles (Ferré 1989). An unresolved question regards the possible presence of a miscibility gap in this system between Al-rich (at times with Na) and Al-poor orthoamphiboles (Hawthorne et al. 1980; Spear 1982). If so, then at least two root mineral kinds would be warranted. Most of the examples in our compilation arise from hornblende-hornfels or pyroxene-hornfels facies contact metamorphism of ultramafic/mafic igneous rocks or pelitic sediments, often in association with biotite, cordierite, and quartz (Reverdatto 1973). We also record several instances of anthophyllite in regional metamorphic settings, including amphibolite and granulite facies metamorphism of pelites and ultramafic rocks (Joplin 1968).

The closely related clinopyroxenes cummingtonite [□Mg2Mg5Si8O22(OH)2; with 12 occurrences] and grunerite [ Fe22+Fe52+Si8O22(OH)2; with 8 occurrences, hence not listed among the top 94] are the monoclinic polymorphs of anthophyllite and ferro-anthophyllite. All but one of the cummingtonite examples in our tabulation, with Mg/ (Mg+Fe) generally >0.4 (i.e., in some instances with Fe > Mg), are from metapelites subjected to hornblende-hornfels facies contact metamorphism (Reverdatto 1973). The more iron-rich grunerite examples, by contrast, are primarily from amphibolite or granulite facies regionally metamorphosed iron formations (Joplin 1968; Kimball and Spear 1984). Therefore, we provisionally distinguish these two closely related mineral kinds based on paragenetic mode, even though they may display continuous solid solution between the Mg and Fe2+ end-members. It should be noted that as a result of miscibility gaps, cummingtonite and grunerite often occur in assemblages with multiple amphiboles, including calcic hornblende, Al-bearing anthophyllite, and/or a sodic amphibole (Deer et al. 1997b).

Several calcic clinoamphiboles are common constituents of metamorphic rocks. Tremolite [ Ca2(Mg5.04.5Fe0.00.52+)Si8O22(OH,F)2; with 122 occurrences] is typically an almost pure Ca-Mg phase (i.e., low Fe2+) formed from ultramafic/mafic igneous or calc-silicate sediments in contact, regional, and high-pressure metamorphic environments. Most of the examples in Online Materials[1] Table S3 are from muscovite-hornfels, hornblende-hornfels, or pyroxene-hornfels contact metamorphic zones, in which tremolite is associated with diopside, dolomite, grossular, talc, and/or other Ca-Mg phases (Reverdatto 1973). We lump the OH- and F-bearing species, which display a complete solid solution and share the same paragenesis.

Actinolite [□Ca2(Mg,Fe2+)5Si8O22(OH,F)2; with 74 occurrences] lumps the species actinolite and ferro-actinolite, spanning a range 0.9 > Mg/(Mg+Fe2+) > 0 (Deer et al. 1997b). Though chemically and structurally similar to tremolite, actinolite is distinguished both by its greater Fe2+ content and by its common association in metapelites or metabasites with biotite, epidote or zoisite, and/or quartz in high-pressure, regional, or contact metamorphic settings (Harker 1950; Joplin 1968; Botha 1983; Carswell 1990).

We lump 26 IMA-CNMNC-approved metamorphic species of Ca-(±Na,K)-clinoamphiboles into hornblende [(Na,K)Ca2(Mg,Fe2+,Al,Fe3+)5(Si,Al)8O22 (OH,F,Cl)2; with 387 occurrences]. This complex group displays significant compositional plasticity, with solid solutions among Na, K, and vacancies in alkali sites; Mg-Fe2+-Fe3+-Al in octahedral sites; and Al-Fe3+-Si in tetrahedral sites, as well as among OH, F, and Cl (Deer et al. 1997b; Hawthorne et al. 2012; see Online Materials[1] Tables S1 and S2). Hornblende, a defining phase in hornblende-hornfels and amphibolite facies rocks, appears in numerous metamorphic environments, including contact metamorphism (albite-epidote to sanidinite facies; Reverdatto 1973; Grapes 2006), regional metamorphism (greenschist to granulite facies; Harker 1950; Joplin 1968), and high-pressure metamorphism (blueschist to eclogite facies; Reverdatto 1973; Carswell 1990). Hornblende protoliths, similarly, span a wide range of igneous, sedimentary, and metamorphic rocks. Hornblende in metamorphic rocks commonly coexists with other amphiboles, including anthophyllite, cummingtonite, and grunerite (Deer et al. 1997b). Given the complexity of this mineral group, cluster analyses of numerous hornblende samples based on composition, paragenesis, and mineral associations, would doubtless reveal many distinct kinds of hornblende.

We lump a wide range of Na-Ca clinoamphiboles into richterite, defined here as [(□,Na)(NaCa)(Mg,Fe2+,Al,Fe3+)5(Si,Al,Fe3+)8O22(OH)2]. Although we record only 12 occurrences in Online Materials[1] Table S3, most of which were originally described as barroisite [nominally (□NaCa)(Mg3Al2)(Si7Al)O22(OH)2] or winchite [(□NaCa) (Mg4Al)Si8O22(OH)2], we lump 16 IMA-CNMNC-approved species into the root natural kind richterite, while acknowledging that much more work is needed to fully characterize these minerals and their associated parageneses. Most of the richterite occurrences that we record are metamorphosed mafic rocks from eclogite facies, almost always in association with omphacite, pyrope, and rutile (Binns 1967; Carswell 1990), though it is reported from regionally metamorphosed basalt, as well (Iwasaki 1960).

The sodium amphibole glaucophane [□Na2(Mg,Fe2+)3Al2Si8O22(OH)2; with 79 occurrences] lumps 2 species, glaucophane and ferro-glaucophane, which form a Mg-Fe2+ solid solution. All of the examples in Online Materials[1] Table S3 are from high-pressure metamorphic rocks (most commonly blueschist or eclogite facies, but also ultrahigh-pressure facies) of mafic/intermediate igneous rocks or Mg-bearing sediments (Augustithis 1985; Carswell 1990).

One additional inosilicate group, sapphirine [(Mg,Fe2+,Al,Fe3+)8O2(Al,Si)6O18, 28 occurrences], is important as a key indicator of the temperature (>900 °C) of ultra high-temperature metamorphic rocks formed from ultramafic protoliths (Monchoux 1972; Deer et al. 1997a; Carswell and Compagnoni 2003; Harley 2021). It commonly occurs in association with orthopyroxene and sillimanite.

Phyllosilicates

Mica Group: With 1244 occurrences in Online Materials[1] Table S3 (45% of the 2795 rocks surveyed), the mica minerals are prominent constituents of many metamorphic lithologies (Guidotti 1984; Fleet 2003). We lump 15 IMA-CNMNC-approved species into five root mineral kinds of micas: biotite, phlogopite, phengite, muscovite, and paragonite. However, there undoubtedly exist many more kinds of metamorphic micas, the identification of which will require the construction of extensive mica databases and application of cluster analysis (Hazen 2019). In particular, widely occurring metamorphic biotite and muscovite will likely be split into multiple natural kinds.

We define the familiar group of dark-colored, Fe2+-bearing trioctahedral mica species as biotite [ KFe22+(Fe2+,Mg,Mn2+)(Si,Al,Fe3+)2Si2O10(OH,F,Cl)2; 822 occurrences]. We lump 6 IMA-CNMNC-approved species, which are themselves rarely identified in the petrographic literature. Few minerals occur in as a diverse array of metamorphic environments as biotite, which we record from low-pressure pyrometamorphic, contact, regional, and high-pressure metamorphism, typically of pelites, but also of ultramafic, mafic, intermediate, acidic, and (rarely) agpaitic igneous rocks, as well as Fe-bearing impure carbonate and calc-silicate protoliths (Harker 1950; Joplin 1968; Reverdatto 1973; Botha 1983; Carswell 1990; Fleet 2003; Grapes 2006).

We also lump 6 Mg-dominant species of trioctahedral micas, which are generally lighter in color than biotite, as phlogopite [KMg2(Mg,Fe2+,Mn2+,Fe3+,Ti4+) (Si,Al,Fe3+)2Si2O10(OH,F)2; with 78 occurrences]. Though often combined with biotite in some descriptions of mica (e.g., Fleet 2003), and consequently sometimes reported as biotite in the petrologic literature, phlogopite displays distinct mineral associations in metamorphosed high-Mg, low-Fe environments, including ultramafic and dolomitic carbonate protoliths (Joplin 1968; Reverdatto 1973). Owing to their diverse parageneses and compositional range, trioctahedral micas represent yet another mineral group that is ripe for investigation by cluster analysis.

The dioctahedral aluminous mica muscovite [K(Al,Fe3+,Cr)2(Si3Al)O10(OH)2; 515 occurrences], commonly reported with the varietal names illite, phengite, or sericite, is most characteristic of regionally and contact metamorphosed pelites (Reverdatto 1973; Philpotts and Ague 2009), which represent most of the occurrences recorded in Online Materials[1] Table S3. Muscovite, among the first phases to form during diagenesis of clay minerals (Fleet 2003), also occurs in a wide range of other contexts, including contact and regional metamorphosed arkosic, calc-silicate, and impure carbonate sediments (Philpotts and Ague 2009), as well as various igneous protoliths (Harker 1950; Carswell 1990).

We also include the fine-grained, Si-rich white mica phengite (110 occurrences) as a separate kind, even though it falls under IMA’s definition of muscovite. Phengite is most commonly associated with high-pressure metamorphism (Carswell 1990).

The sodium trioctahedral mica paragonite [NaAl2(Si3Al)O10(OH)2; 67 occurrences] is characteristic of high-pressure eclogite facies metamorphism of mafic igneous rocks, often in association with glaucophane, kyanite, omphacite, and/or pyrope (Carswell 1990). Paragonite, often intimately intermixed with phengite (with which it has limited solid solution; e.g., Thompson and Thompson 1976; Guidotti et al. 1994), also occurs in low- and medium-grade metapelites, in which it can form by both prograde and retrograde reactions (Chatterjee 1970; Guidotti 1984; Guidotti and Sassi 1998; Fleet 2003). Paragonite also frequently co-occurs with the so-called “brittle mica” margarite [CaAl2(Si2Al2)O10(OH)2] in a wide range of metamorphic grades of metasediments (Guidotti 1984; Fleet 2003); however, margarite is relatively rare in comparison to the micas described above, being reported from only one of the metamorphic rocks in our survey (Carswell and Compagnoni 2003, Table 2 therein).

Other phyllosilicates: More than 30 other IMA-CNMNC-approved layer silicates occur in metamorphic rocks (Online Materials[1] Table S1). Most of these minerals (e.g., apophyllite group, gillespite, pyrophyllite, stilpnomelane, zussmanite) occur only rarely in metamorphic rocks. Note that we do not list diagenetically formed clay minerals as metamorphic phases (Wilson 2013); they will be considered further in Part IX of this series. However, chlorite, prehnite, serpentine, and talc are included in our list of 94 relatively common metamorphic phases (Deer et al. 2009).

Chlorite [(Mg,Fe2+)5(Al,Fe3+)(Si3AlO10)(OH)8; 339 occurrences] encompasses 3 IMA-CNMNC-approved species of Mg-Fe2+-Al-(Fe3+) layer silicates: chamosite, clinochlore, and sudoite. Chlorite is common in pelites subjected to greenschist and amphibolite facies metamorphism (Harker 1950; Joplin 1968; Botha 1983), as well as from muscovite-, hornblende-, and pyroxene-hornfels facies contact metamorphism of pelites and basic igneous rocks (Joplin 1968; Reverdatto 1973), often in association with albite, biotite, muscovite, and/or quartz. In addition, Coleman et al. (1965) and Carswell (1990) record more than a dozen examples of chlorite in eclogite facies high-pressure metamorphism. Chlorite forms through various pathways, including prograde and retrograde metamorphism, both with and without external aqueous fluids. As such, chlorite represents a mineral whose often uncertain parageneses grade continuously from regional or “burial” metamorphism to metasomatism to hydrothermal alteration (Deer et al. 2009).

We lump 5 IMA-CNMNC-approved species, including three structural variants of Mg3Si2O5(OH)4 (antigorite, chrysotile, and lizardite), aluminous amesite, and Fe- bearing greenalite, into serpentine [(Mg,Fe2+,Al,Fe3+)3(Al,Si)Si(OH)4; 68 occurrences]. Most examples in Online Materials[1] Table S3 are from low- to moderate-grade regional metamorphism of ultramafic/mafic igneous or pelitic protoliths (Joplin 1968; Philpotts and Ague 2009), in which they form primarily by retrograde/hydrothermal reactions from olivine and Mg-rich pyroxene or by prograde metamorphism of serpentinite (Deer et al. 2009, and references therein). Given its varied modes of formation, coupled with multiple polymorphs, cluster analysis of metamorphic serpentine is warranted.

Talc [Mg3Si4O10(OH)2; 73 occurrences] is characteristic of Mg-rich protoliths over a wide range of pressure-temperature condition. Examples include thermal metamorphism of dolomite-bearing sediments (Tilley 1948; Reverdatto 1973; Augustithis 1985), high-pressure (blueschist and eclogite facies) metamorphism of basic igneous rocks (Chopin 1981; Carswell 1990), and greenschist to amphibolite grade regional metamorphism of ultramafic rocks (Harker 1950; Joplin 1968). Of special note are high-pressure to ultrahigh-pressure (>0.6 GPa) talc-kyanite-(quartz/coesite) assemblages known as “whiteschists” (Schreyer 1977), which form in the Mg-Al-Si-H (“MASH”) system, at times with PH2O approximately equal to the total pressure.

Prehnite [Ca2Al(Si3Al)O10(OH)2; 16 occurrences], though most familiar as a hydrothermal phase associated with zeolites in amygdaloidal basalt, is also common in the eponymous prehnite-pumpellyite facies of low-grade regional metamorphism (Coombs 1960; Philpotts and Ague 2009). Prehnite, often in association with chlorite and quartz, is also present occasionally in metamorphosed basic igneous and pelitic sedimentary rocks from zeolite to lower amphibolite grades (Harker 1950; Joplin 1968; Botha 1983), at times the result of retrograde reactions (Coombs 1993).

Tectosilicates

A wide range of framework silicates, including the silica group, feldspars, feldspathoids, and zeolites (Deer et al. 2001, 2004), occur in metamorphic rocks, with one or more examples reported in 1595 (67%) of the 2785 rocks surveyed in Online Materials[1] Table S3. We focus attention on 10 mineral kinds that occur most frequently.

Silica Group: Four silica group minerals—quartz, high-pressure coesite, and high-temperature cristobalite and tridymite—span the entire range of metamorphic environments, occurring in all but the most Si-deficient rocks (Deer et al. 2004).

Quartz (SiO2), with 1353 occurrences in our study (49% of rocks in Online Materials[1] Table S3), is the most common metamorphic mineral. It occurs in pyrometamorphosed xenoliths, contact metamorphic rocks, metamorphosed iron-manganese formations, high-pressure and regional metamorphic settings, metasomatized rocks, and shear zones (Online Materials[1] Table S3 and references therein), often by recrystallization of protolith quartz (Deer et al. 2004). Quartz is stable in all crust and upper mantle pressure-temperature regimes except above ~2.7 GPa, where it transforms to coesite, or at low pressure above ~850 °C, where cristobalite and tridymite are the stable silica phases.

Coesite (SiO2; 23 occurrences) is restricted to ultrahigh pressure (>2.7 GPa) metamorphic environments, where it is typically associated with kyanite, omphacite, and/or pyrope (Carswell and Compagnoni 2003). Several reports describe coesite as inclusions in upper mantle phases, including pyrope (Chopin 1984; Schertl et al. 1991) and diamond (Stachel et al. 2022, and references therein).

Cristobalite (10 occurrences) and tridymite (44 occurrences) are high-temperature, low-pressure polymorphs of SiO2 that occur almost exclusively in sedimentary rocks that have been thermally metamorphosed (pyroxene hornfels or sanidinite facies) by basic igneous rocks (Agrell and Langley 1958; Reverdatto 1973; Black 1989; Grapes 2006). These two minerals co-occur in 7 of the 10 reported rocks with cristobalite. Tridymite and cristobalite also are associated in some burning coal deposits with temperatures that may exceed 1100 °C (Bustin and Mathews 1982; Grapes 2006), and therefore are the consequence of Phanerozoic biological precursors (to be considered in Part XII).

Feldspar Group: Metamorphic feldspar group minerals display compositions close to two binary solid solutions (Deer et al. 2001): the Na-Ca plagioclase feldspars and the Na-K alkali feldspars. In both instances we suggest modifications of the nomenclature approved by the IMA-CNMNC.

In the case of the plagioclase series [(CaAl,NaSi)AlSi2O8], we identify albite as compositions close to NaAlSi3O8 [Na/(Na+Ca) > 0.85 and often with >10 mol% KAlSi3O8], anorthite as close to CaAl2Si2O8 [Ca/(Ca+Na) > 0.85], and plagioclase as having intermediate compositions between ~An20 and ~An70 as valid root mineral kinds. We justify this division based on the existence of the so-called peristerite and Huttenlocker miscibility gaps between ~An2–17 and An65–88 respectively. As a consequence, several authors have recorded coexisting albite-plagioclase and plagioclase-anorthite pairs (Evans 1964; Botha 1983).

Albite, with 177 occurrences in Online Materials[1] Table S3, is observed in a wide range of thermal, regional, and high-pressure metamorphic environments, with both igneous and sedimentary protoliths (Harker 1950; Joplin 1968; Reverdatto 1973; Philpotts and Ague 2009). Anorthite (78 occurrences) is more restricted in its occurrences, being found primarily as a contact metamorphic mineral derived from calc-silicate and carbonate-bearing sediments (Grapes 2006), though it is also found in regionally metamorphosed mafic and calc-silicate rocks from amphibolite to granulite facies (Joplin 1968). Plagioclase (809 occurrences), like quartz, albite, and kspar (see below), is not a particularly diagnostic phase in metamorphic rocks because it occurs across the full spectrum of thermal, regional, and high-pressure metamorphic environments, with an equally broad range of igneous, sedimentary, and metamorphic protoliths. For a given protolith, the anorthite content of plagioclase tends to increase with metamorphic grade (Deer et al. 2001). Note that most reports of “plagioclase” in the older metamorphic literature lack compositional information; thus, some of these occurrences may fit our definitions of albite or anorthite.

The alkali feldspars are complicated by the existence of three K-rich (KAlSi3O8) variants—the higher-temperature (>500 °C) monoclinic sanidine and two lower-temperature triclinic phases, microcline and orthoclase, which are often reported as “kspar” in the literature of metamorphic petrology. An additional consideration is that alkali feldspars of intermediate compositions often exsolve Na- and K-rich phases, typically reported as “perthite.” In our study, we adopt the name kspar for microcline and orthoclase, and record both albite and kspar for perthite.

Sanidine (102 occurrences) is most commonly found in thermally metamorphosed rocks of pyroxene hornfels or sanidinite grade (Grapes 2006), though it also has been reported from ultrahigh-pressure metamorphism of pelites (Carswell 1990; Carswell and Compagnoni 2003) and ultrahigh-temperature regimes (Harley 2021).

Kspar (411 occurrences) has been reported from thermal (zeolite to pyroxene hornfels facies), regional (amphibolite to granulite facies), and high-pressure (eclogite to ultrahigh-pressure facies) metamorphic environments. Protoliths for kspar include mafic, acidic, and agpaitic igneous rocks and arkosic, pelitic, and carbonate-bearing sedimentary rocks (Harker 1950; Joplin 1968; Reverdatto 1973; Carswell 1990; Deer et al. 2001).

At the temperatures of UHT metamorphism (>900 °C), an additional complication is the occurrence of Ca-Na-K ternary feldspars, which typically exsolve to a perthite with coexisting plagioclase and alkali feldspar lamellae (Harley 2008; Harley 2021, Fig. 20 therein).

Scapolite Group: Three species of the scapolite group are lumped as scapolite [(Na,Ca)4Al3(Al,Si)3Si6O24(CO3,SO4,Cl), with 33 occurrences]. Scapolite is not infrequently observed in medium- to high-grade contact (hornblende-hornfels and pyroxene hornfels facies) and regional (amphibolite and granulite facies) metamorphosed pelites and calc-silicate sediments (Joplin 1968; Reverdatto 1973), and has also been reported from the albite-epidote hornfels facies contact metamorphism of amygdaloidal basalt (Joplin 1968).

Zeolite Group: The zeolite facies is the lowest pressure-temperature regime of metamorphism, with temperatures <200 °C at pressures <0.3 GPa (Philpotts and Ague 2009). Most zeolite minerals form via low-temperature aqueous processes, including fluid interactions with cooling basalt and authigenesis (Deer et al. 2004). Nevertheless, some zeolite occurrences are attributed to metamorphism, sensu stricto. Though not sufficiently abundant to include among the most common metamorphic phases, analcime (2 occurrences), laumontite (1), and wairakite (2), as well as 6 undifferentiated reports of “zeolite,” were listed in modes of low-grade metamorphosed mafic igneous rocks and pelites (Joplin 1968).

Feldspathoid Group: None of the members of the sub-silicic feldspathoid framework silicate group is common in metamorphic rocks. Nevertheless, kalsilite, leucite, and nepheline (with 5, 2, and 4 occurrences, respectively) are representative of undersaturated pyrometamorphosed calc-silicates (Grapes 2006).

To these framework silicates, we add silicate glass [(Si,Al,Ca,Mg,Fe)O; SiO2 >70 wt%; with 68 occurrences] as an important yet often poorly characterized metamorphic phase in thermal metamorphic environments, particularly of arkosic sandstones and pelites (Reverdatto 1973; Grapes 2006). Two varietal names of metamorphic glass are “buchite,” which forms when a silica-rich pelitic rock is altered by igneous contact, and “porcellanite,” a glass derived from pyrometamorphosed clay, marl, shale, or bauxite (Grapes 2006). Melting and glass formation may occur as low as ~650 °C at 05. GPa in a granite protolith, or >1000 °C at low pressure and dry conditions. Grapes (2006) details how “Si-rich glass” in many pyrometamorphic zones typically contains significant Al2O3 and alkalis, a consequence of quartz-feldspar melting. Note that pure SiO2 melts at 1700 °C—a temperature only attainable by lightning strikes or bolide impacts.

Rarer Metamorphic Minerals: In addition to the 94 mineral kinds outlined above, 661 other mineral kinds occur rarely as trace phases in metamorphic rocks, as documented by reports in numerous primary sources and compilations, notably Anthony et al. (1990–2003) and references cited in https://mindat.org and https://rruff.info/ima (both accessed 20 January 2023). Most of these scarce minerals, which are listed in Online Materials[1] Table S1, were not recorded from any of the 2785 metamorphic rock modes in Online Materials[1] Table S3.

However, a few of the less common minerals in Online Materials[1] Table S1, were also noted in one or more metamorphic rock modes. Among these minor minerals, listed alphabetically, are alleghanyite (1 occurrence), analcime (2), anhydrite (2), ankerite (6), ardennite (1), arsenopyrite (1), axinite (4), bornite (1), braunite (7), bredigite (6), brownmillerite (5), bustamite (4), calzirtite (3), celestine (1), chondrodite (4), cuspidine (2), deerite (2), diaspore (1), fluorite (1), fluormayenite (5), friedelite (1), galaxite (1), giuseppeite (1), grunerite (8), hausmannite (1), hillbrandite (1), hügbomite (2), ilvaite (2), jacobsite (1), kalsilite (5), kornerupine (1), kutnohorite (1), laumontite (1), leucite (2), margarite (1), monazite (3), nepheline (4), norbergite (1), piemontite (7), pigeonite (2), pyrophanite (1), pyroxmangite (2), qandilite (3), rhodocrosite (2), rhodonite (8), riebeckite (4), scawtite (1), siderite (9), sonolite (1), sphalerite (9), spessartine (9), stilpnomelane (3), suenoite (5), tephroite (1), thompsonite (3), uvarovite (2), wairakite (2), and “zeolite” (6).

Received: 2023-03-19
Accepted: 2024-01-31
Published Online: 2024-09-24
Published in Print: 2024-10-28

© 2024 by Mineralogical Society of America

Articles in the same Issue

  1. Trace element fractionation in magnetite as a function of Fe depletion from ore fluids at the Baijian Fe-(Co) skarn deposit, eastern China: Implications for Co mineralization in Fe skarns
  2. Evidence for oceans pre-4300 Ma confirmed by preserved igneous compositions in Hadean zircon
  3. Experimental vs. natural fulgurite: A comparison and implications for the formation process
  4. Illitization of smectite influenced by chemical weathering and its potential control of anatase formation in altered volcanic ashes
  5. A modified genetic model for multiple pulsed mineralized processes at the giant Qulong porphyry Cu-Mo mineralization system
  6. Lead and noble gas isotopic constraints on the origin of Te-bearing adularia-sericite epithermal Au-Ag deposits in a calc-alkaline magmatic arc, NE China
  7. Wenlanzhangite–(Y) from the Yushui deposit, South China: A potential proxy for tracing the redox state of ore formation
  8. High-resolution SIMS U-Th-Pb geochronology of small-size (<5 μm) monazite: Constraints on the timing of Qiuling sediment-hosted gold deposit, South Qinling Orogen, central China
  9. An evolutionary system of mineralogy, Part VIII: The evolution of metamorphic minerals
  10. Gamma-enhancement of reflected light images as a routine method for assessment of compositional heterogeneity in common low-reflectance Fe-bearing minerals
  11. Polysomatic intergrowths between amphiboles and non-classical pyriboles in magnetite: Smallest-scale features recording a protracted geological history
  12. Mineralogical and geochemical facets of the massive deposition of stibnite-metastibnite at a seafloor hydrothermal field (Wakamiko Crater, Kagoshima Bay, Ryukyu Volcanic Arc)
  13. High-pressure phase transition in clinochlore: IIa polytype stabilization
  14. Memorial of Larry Wayne Finger (1940–2024)
Downloaded on 30.10.2025 from https://www.degruyterbrill.com/document/doi/10.2138/am-2023-9004/html
Scroll to top button