Home An evolutionary system of mineralogy, Part VII: The evolution of the igneous minerals (>2500 Ma)
Article
Licensed
Unlicensed Requires Authentication

An evolutionary system of mineralogy, Part VII: The evolution of the igneous minerals (>2500 Ma)

  • Robert M. Hazen ORCID logo , Shaunna M. Morrison , Anirudh Prabhu ORCID logo , Michael J. Walter and Jason R. Williams
Published/Copyright: August 31, 2023
Become an author with De Gruyter Brill

Abstract

Part VII of the evolutionary system of mineralogy catalogs, analyzes, and visualizes relationships among 919 natural kinds of primary igneous minerals, corresponding to 1665 mineral species approved by the International Mineralogical Association—minerals that are associated with the wide range of igneous rock types through 4.566 billion years of Earth history. A systematic survey of the mineral modes of 1850 varied igneous rocks from around the world reveals that 115 of these mineral kinds are frequent major and/or accessory phases. Of these most common primary igneous minerals, 69 are silicates, 19 are oxides, 13 are carbonates, and 6 are sulfides. Collectively, these 115 minerals incorporate at least 33 different essential chemical elements.

Patterns of coexistence among these minerals, revealed by network, Louvain community detection, and agglomerative hierarchical clustering analyses, point to four major communities of igneous primary phases, corresponding in large part to different compositional regimes: (1) silica-saturated, quartz- and/or alkali feldspar-dominant rocks, including rare-element granite pegmatites; (2) mafic/ultramafic rock series with major calcic plagioclase and/or mafic minerals; (3) silica-undersaturated rocks with major feldspathoids and/or analcime, including agpaitic rocks and their distinctive rare-element pegmatites; and (4) carbonatites and related carbonate-bearing rocks.

Igneous rocks display characteristics of an evolving chemical system, with significant increases in their minerals’ diversity and chemical complexity over the first two billion years of Earth history. Earth’s earliest igneous rocks (>4.56 Ga) were ultramafic in composition with 122 diferent minerals, followed closely by mafic rocks that were generated in large measure by decompression melting of those ultramafic lithologies (4.56 Ga). Quartz-normative granitic rocks and their extrusive equivalents (>4.4 Ga), formed primarily by partial melting of wet basalt, were added to the mineral inventory, which reached 246 diferent mineral kinds. Subsequently, four groups of igneous rocks with diagnostic concentrations of rare element minerals—layered igneous intrusions, complex granite pegmatites, alkaline igneous complexes, and carbonatites—all first appeared ~3 billion years ago. These more recent varied kinds of igneous rocks hold more than 700 diferent minerals, 500 of which are unique to these lithologies.

Network representations and heatmaps of primary igneous minerals illustrate Bowen’s reaction series of igneous mineral evolution, as well as his concepts of mineral associations and antipathies. Furthermore, phase relationships and reaction series associated with the minerals of a dozen major elements (H, Na, K, Mg, Ca, Fe, Al, Si, Ti, C, O, and S), as well as minor elements (notably Li, Be, Sr, Ba, Mn, B, Cr, Y, REE, Ti, Zr, Nb, Ta, P, and F), are embedded in these multi-dimensional visualizations.

Acknowledgments and funding

We are especially grateful to Jay Ague, Robert Downs, George Harlow, and Michael Wong for valuable discussions and reviews of an early version of this contribution. We also thank Associate Editor Steven Simon, and reviewers Mark Ghiorso and Ross Mitchell for their thorough, thoughtful, and constructive reviews.

Studies of mineral evolution and mineral ecology have been supported by the Alfred P. Sloan Foundation, the W.M. Keck Foundation, 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.

References cited

Ague, J.J. and Brimhall, G.H. (1988) Magmatic arc asymmetry and distribution of anomalous plutonic belts in the batholiths of California: Effects of assimilation, crustal thickness, and depth of crystallization. Geological Society of America Bulletin, 100, 912–927, https://doi.org/10.1130/0016-7606(1988)100<0912:MAAADO>2.3.CO;2.Search in Google Scholar

Almende, B.V. and Contributors, B. Thieurmel and T. Robert. (2021) visNetwork: Network Visualization using ‘vis.js’ Library. R package version 2.1.0. https://CRAN.R-project.org/package=visNetwork.Search in Google Scholar

Anenburg, M., Broom-Fendley, S., and Chen, W. (2021) Formation of rare earth deposits in carbonatites. Elements, 17, 327–332, https://doi.org/10.2138/gselements.17.5.327.Search in Google Scholar

Annen, C., Blundy, J.D., and Sparks, R.S.J. (2006) The genesis of intermediate and silicic magmas in deep crustal hot zones. Journal of Petrology, 47, 505–539, https://doi.org/10.1093/petrology/egi084.Search in Google Scholar

Anthony, J.W., Bideaux, R.A., Bladh, K.W., and Nichols, M.C. (1990–2003) Handbook of Mineralogy, 6 volumes. Mineralogical Society of America, Chantilly, Virginia 20151-1110, U. S.A. http://www.handbookofmineralogy.org/.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

Atencio, D., Andrade, M.B., Christy, A.G., Gieré, R., and Kartashov, P.M. (2010) The pyrochlore supergroup of minerals: Nomenclature. Canadian Mineralogist, 48, 673–698, https://doi.org/10.3749/canmin.48.3.673.Search in Google Scholar

Badro, J. and Walter, M., Eds. (2015) The early Earth: Accretion and differentiation. American Geophysical Union, Geophysical Monograph, vol. 212.Search in Google Scholar

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

Berdnikov, N., Kepezhinskas, P., Konovalova, N., and Kepezhinskas, N. (2022) Formation of gold alloys during differentiation of convergent zone magmas: Constraints from an Au-rich websterite in the Stanovoy suture zone (Russian Far East). Geosciences, 12, 126, https://doi.org/10.3390/geosciences12030126.Search in Google Scholar

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

Bizzarro, M., Simonetti, A., Stevenson, R.K., and David, J. (2002) Hf isotope evidence for a hidden mantle reservoir. Geology, 30, 771–774, https://doi.org/10.1130/0091-7613(2002)030<0771:HIEFAH>2.0.CO;2.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, P10008, https://doi.org/10.1088/1742-5468/2008/10/P10008.Search in Google Scholar

Bonnefoi, C.C., Provost, A., and Albarede, F. (1995) The “Daly gap” as a magmatic catastrophe. Nature, 378, 270–272, https://doi.org/10.1038/378270a0.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

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. (1912) The order of crystallization in igneous rocks. Journal of Geology, 20, 457–468.Search in Google Scholar

Bowen, N.L. (1913) The order of crystallization in igneous rocks. Journal of Geology, 21, 399–401.Search in Google Scholar

Bowen, N.L. (1915) Evolution of igneous rocks. Carnegie Institution of Washington Year Book, 15, 144–145.Search in Google Scholar

Bowen, N.L. (1922) The reaction principle in petrogenesis. The Journal of Geology, 30, 177–198, https://doi.org/10.1086/622871.Search in Google Scholar

Bowen, N.L. (1927) The origin of ultrabasic and related rocks. American Journal of Science, 14, 89–108, https://doi.org/10.2475/ajs.s5-14.80.89.Search in Google Scholar

Bowen, N.L. (1928) The Evolution of the Igneous Rocks, 602 p. Princeton University Press.Search in Google Scholar

Bowles, J.F.W., Howie, R.A., Vaughan, D.J., and Zussman, J. (2011) Rock-Forming Minerals, vol. 5A, 2nd ed.. Non-Silicates: Oxides, Hydroxides and Sulfides, 418 p. The Geological Society of London.Search in Google Scholar

Boyd, R. (2010) Homeostasis, higher taxa, and monophyly. Philosophy of Science, 77, 686–701, https://doi.org/10.1086/656551.Search in Google Scholar

Bradley, D.C. (2011) Secular trends in the geologic record and the supercontinent cycle. Earth-Science Reviews, 108, 16–33, https://doi.org/10.1016/j.earscirev.2011.05.003.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

Burnham, A.D. and Berry, A.J. (2017) Formation of Hadean granites by melting of igneous crust. Nature Geoscience, 10, 457–461, https://doi.org/10.1038/ngeo2942.Search in Google Scholar

Carmichael, I.S.E., Turner, F.J., and Verhoogen, J. (1973) Igneous Petrology, 739 p. McGraw-Hill Book Company.Search in Google Scholar

Castle, R.O. and Lindsley, D.H. (1993) An exsolution silica-pump model for the origin of myrmekite. Contributions to Mineralogy and Petrology, 115, 58–65, https://doi.org/10.1007/BF00712978.Search in Google Scholar

Chakhmouradian, A.R. (2006) High-field-strength elements in carbonatitic rocks: Geochemistry, crystal chemistry, and significance for constraining the sources of carbonatites. Chemical Geology, 235, 138–160, https://doi.org/10.1016/j.chemgeo.2006.06.008.Search in Google Scholar

Chang, L.L.Y., Howie, R.A., and Zussman, J. (1996) Rock-Forming Minerals, vol. 5B, 2nd ed., Sulphates, Carbonates, Phosphates and Halides, 383 p. Longman Group.Search in Google Scholar

Chappell, B.W. and White, A.J.R. (1974) Two contrasting granite types. Pacific Geology, 8, 173–174.Search in Google Scholar

Chappell, B.W. and White, A.J.R. (2001) Two contrasting granite types: 25 years later. Australian Journal of Earth Sciences, 48, 489–499, https://doi.org/10.1046/j.1440-0952.2001.00882.x.Search in Google Scholar

Chayes, F. (1963) Relative abundance of intermediate members of the oceanic basalt-trachyte association. Journal of Geophysical Research, 68, 1519–1534, https://doi.org/10.1029/JZ068i005p01519.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., Morrison, S.M., and Hazen, R.M. (2022) ESMD-Garnet dataset. Open Data Repository, https://doi.org/10.48484/camh-xy98.Search in Google Scholar

Christy, A.G. and Atencio, D. (2013) Clarification of status of species in the pyrochlore supergroup. Mineralogical Magazine, 77, 13–20, https://doi.org/10.1180/minmag.2013.077.1.02.Search in Google Scholar

Christy, A.G., Pekov, I.V., and Krivovichev, S.V. (2021) The distinctive mineralogy of carbonatites. Elements, 17, 333–338, https://doi.org/10.2138/gsele-ments.17.5.333.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

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

Dasgupta, R., Hirschmann, M.M., and Withers, A.C. (2004) Deep global cycling of carbon constrained by the solidus of anhydrous carbonated eclogite under upper mantle conditions. Earth and Planetary Science Letters, 227, 73–85, https://doi.org/10.1016/j.epsl.2004.08.004.Search in Google Scholar

Deer, W.A., Howie, R.A., and Zussman, J. (1982) Rock-Forming Minerals, vol. 1A, 2nd ed., Orthosilicates, 333 p. Longman.Search in Google Scholar

Deer, W.A., Howie, R.A., and Zussman, J. (1986) Rock-Forming Minerals, vol. 1B, 2nd ed., Disilicates and Ring Silicates, 638 p. John Wiley.Search in Google Scholar

Deer, W.A., Howie, R.A., and Zussman, J. (1997a) Rock-Forming Minerals, vol. 2A, 2nd ed., Single-Chain Silicates, 680 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, 2nd ed., Double-Chain Silicates, 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, 2nd ed., Framework Silicates: Feldspars, 972 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, 2nd ed., Framework Silicates: Silica Minerals, Feldspathoids and the Zeolites, 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, 2nd ed., Layered Silicates Excluding Micas and Clay Minerals, 320 p. The Geological Society of London.Search in Google Scholar

Deer, W.A., Howie, R.A., and Zussman, J. (1982–2013) Rock-Forming Minerals. 2nd ed. 11 volumes. Longman, Wiley, and The Geological Society of London.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

Elliott, H.A.L., Wall, F., Chakhmouradian, A.R., Siegfried, P.R., Dahlgren, S., Weatherley, S., Finch, A.A., Marks, M.A.W., Dowman, E., and Deady, E. (2018) Fenites associated with carbonatite complexes: A review. Ore Geology Reviews, 93, 38–59, https://doi.org/10.1016/j.oregeorev.2017.12.003.Search in Google Scholar

Ewing, R.C. (1976) A numerical approach toward the classification of complex, orthorhombic, rare-earth, AB2O6-type Nb-Ta-Ti oxides. Canadian Mineralogist, 14, 111–119.Search in Google Scholar

Fleet, M.E. (2003) Rock-Forming Minerals, vol. 3A, 2nd ed., Sheet Silicates: Micas, 780 p. The Geological Society of London.Search in Google Scholar

Foley, S.F., Yaxley, G.M., Rosenthal, A., Buhre, S., Kiseeva, E.S., Rapp, R.P., and Jacob, D.E. (2009) The composition of near-solidus melts of peridotite in the presence of CO2 and H2O between 40 and 60 kbar. Lithos, 112, (Supplement 1), 274–283, https://doi.org/10.1016/j.lithos.2009.03.020.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

Gaines, R.V., Skinner, C., Foord, E.E., Mason, B., and Rosenzweig, A. (1997) Dana’s New Mineralogy, 1819 p. Wiley.Search in Google Scholar

Galili, T., O’Callaghan, A., Sidi, J., and Sievert, C. (2018) heatmaply: An R package for creating interactive cluster heatmaps for online publishing. Bioinformatics (Oxford, England), 34, 1600–1602, https://doi.org/10.1093/bioinformatics/btx657.Search in Google Scholar

Ghiorso, M.S. and Sack, R.O. (1995) Chemical mass transfer in magmatic processes IV. A revised and internally consistent thermodynamic model for the interpolation and extrapolation of liquid-solid equilibria in magmatic systems at elevated temperatures and pressures. Contributions to Mineralogy and Petrology, 119, 197–212, https://doi.org/10.1007/BF00307281.Search in Google Scholar

Ghiorso, M.S., Hirschmann, M.M., Reiners, P.W., and Kress, V.C. III. (2002) The pMELTS: A revision of MELTS for improved calculation of phase relations and major element partitioning related to partial melting of the mantle to 3 GPa. Geochemistry, Geophysics, Geosystems, 3, 1–35, https://doi.org/10.1029/2001GC000217.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, 99, 7821–7826, https://doi.org/10.1073/pnas.122653799.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

Grew, E.S. and Hazen, R.M. (2014) Beryllium mineral evolution. American Mineralogist, 99, 999–1021, https://doi.org/10.2138/am.2014.4675.Search in Google Scholar

Griffin, W.L., Huang, J.-X., Thomassot, E., Gain, S.E.M., Toledo, V., and O’Reilly, S.Y. (2018) Super-reducing conditions in ancient and modern volcanic systems: Sources and behaviours of carbon-rich fluids in the lithospheric mantle. Mineralogy and Petrology, 112, (S1), S101–S114, https://doi.org/10.1007/s00710-018-0575-x.Search in Google Scholar

Guilbert, J.M. and Park, C.F. Jr. (2007) The Geology of Ore Deposits, 985 p. Waveland Press.Search in Google Scholar

Harker, A. (1964) Harker’s Petrology for Students, Eighth Edition Revised by C.R. Tilley, S.R. Nockolds, M. Black, 283 p. Cambridge University Press.Search in Google Scholar

Harrison, T.M. (2009) The Hadean crust: Evidence from > 4 Ga zircons. Annual Review of Earth and Planetary Sciences, 37, 479–505, https://doi.org/10.1146/annurev.earth.031208.100151.Search in Google Scholar

Harrison, T.M., Bell, E.A., and Boehnke, P. (2017) Hadean zircon petrochronology. Reviews in Mineralogy and Geochemistry, 83, 329–363, https://doi.org/10.2138/rmg.2017.83.11.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., Oberti, R., Harlow, G.E., Maresch, W.V., Martin, R.F., Schumacher, J.C., and Welch, M.D. (2011) Nomenclature of the amphibole supergroup. American Mineralogist, 97, 2031–2048, https://doi.org/10.2138/am.2012.4276.Search in Google Scholar

Hazen, R.M. (2014) Data-driven abductive discovery in mineralogy. American Mineralogist, 99, 2165–2170, https://doi.org/10.2138/am-2014-4895.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. 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. and Morrison, S.M. (2021) An evolutionary system of mineralogy, Part V: Aqueous and thermal alteration of planetesimals (~4565 to 4550 Ma). American Mineralogist, 106, 1388–1419, https://doi.org/10.2138/am-2021-7760.Search in Google Scholar

Hazen, R.M. and Morrison, S.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., Liu, X.-M., Downs, R.T., Golden, J.J., Pires, A.J., Grew, E.S., Hystad, G., Estrada, C., and Sverjensky, D.A. (2014) Mineral evolution: Episodic metallogenesis, the supercontinent cycle, and the coevolving geosphere and biosphere. Society of Economic Geologists Special Publication, 18, 1–15, https://doi.org/10.5382/SP.18.01.Search in Google Scholar

Hazen, R.M., Downs, R.T., Eleish, A., Fox, P., Gagné, O., Golden, J.J., Grew, E.S., Hummer, D.R., Hystad, G., Krivovichev, S.V., and others. (2019) Data-driven discovery in mineralogy: Recent advances in data resources, analysis, and visualization. Engineering (Beijing), 5, 397–405, https://doi.org/10.1016/j. eng.2019.03.006.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

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

Jagoutz, O. and Klein, B. (2018) On the importance of crystallization-differentiation for the generation of SiO2-rich melts and the compositional build-up of ARC (and continental) crust. American Journal of Science, 318, 29–63, https://doi.org/10.2475/01.2018.03.Search in Google Scholar

Johannsen, A. (1932) A Descriptive Petrography of the Igneous Rocks, vol. II, The Quartz-Bearing Rocks, 429 p. The University of Chicago Press.Search in Google Scholar

Johannsen, A. (1937) A Descriptive Petrography of the Igneous Rocks, vol. III, The Intermediate Rocks, 369 p. The University of Chicago Press.Search in Google Scholar

Johannsen, A. (1938) A Descriptive Petrography of the Igneous Rocks, vol. IV. Part I: The Feldspathoid Rocks, Part II: The Peridotites and Perknites, 523 p. The University of Chicago Press.Search in Google Scholar

Johnsen, O. and Grice, J.D. (1999) The crystal chemistry of the eudialyte group. Canadian Mineralogist, 37, 865–891.Search in Google Scholar

Johnson, T.E., Brown, M., Gardiner, N.J., Kirkland, C.L., and Smithies, R.H. (2017) Earth’s first stable continents did not form by subduction. Nature, 543, 239–242, https://doi.org/10.1038/nature21383.Search in Google Scholar

Jones, A.P., Genge, M., and Carmody, L. (2013) Carbonate melts and carbonatites. Reviews in Mineralogy and Geochemistry, 75, 289–322, https://doi.org/10.2138/rmg.2013.75.10.Search in Google Scholar

Kamenetsky, V.S., Doroshkevich, A.G., Elliott, H.A.L., and Zaitsev, A.N. (2021) Carbonatites: Contrasting, complex, and controversial. Elements, 17, 307–314, https://doi.org/10.2138/gselements.17.5.307.Search in Google Scholar

Kapustin, Y.L. (2010) Features of fenitization around carbonatite bodies. International Geology Review, 25, 1393–1404.Search in Google Scholar

Kjarsgaard, B.A., Hamilton, D.L., and Peterson, T.D. (1995) Peralkaline nepheline/carbonatite liquid immiscibility: Comparison of phase compositions in experiments and natural lavas from Oldoinyo Lengai. In K. Bell and J. Keller, Eds., Carbonatite volcanism: Oldoinyo Lengai and the Petrogenesis of Natrocarbonatites, pp. 163–190. Springer-Verlag.Search in Google Scholar

Kogarko, L.N., Kononova, V.A., Orlova, M.P., and Woolley, A.R. (1995) Alkaline Rocks and Carbonatites of the World. Part Two. Chapman & Hall.Search in Google Scholar

Korenaga, J. (2021) Hadean geodynamics and the nature of early continental crust. Precambrian Research, 359, 106178, https://doi.org/10.1016/j.precamres.2021.106178.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

Kutyrev, A., Zelenski, M., Nekrylov, N., Savelyev, D., Kontonikas-Charos, A., and Kamenetsky, V.S. (2021) Noble metals in arc basaltic magmas worldwide: A case study of modern and pre-historic lavas of the Tolbachik volcano, Kamchatka. Frontiers in Earth Science (Lausanne), 9, 791465, https://doi.org/10.3389/feart.2021.791465.Search in Google Scholar

Laporte, J. (2004) Natural Kinds and Conceptual Change, 232 p. Cambridge University Press.Search in Google Scholar

Le Bas, M.J. (2008) Fenites associated with carbonatites. Canadian Mineralogist, 46, 915–932, https://doi.org/10.3749/canmin.46.4.915.Search in Google Scholar

Lester, G.W., Clark, A.H., Kyser, T.K., and Naslund, H.R. (2013) Experiments on liquid immiscibility in silicate melts with H2O, P, S, F and Cl: Implications for natural magmas. Contributions to Mineralogy and Petrology, 166, 329–349, https://doi.org/10.1007/s00410-013-0878-1.Search in Google Scholar

Levin, E.M., Robbins, C., and Reser, M.K. (1964) Phase Diagrams for Ceramists, 395 p. The American Ceramics Society.Search in Google Scholar

London, D. (1986) The magmatic-hydrothermal transition in the Tanco rare-element pegmatite: Evidence from fluid inclusions and phase equilibrium experiments. American Mineralogist, 71, 376–395.Search in Google Scholar

London, D. (2008) Pegmatites. Mineralogical Association of Canada.Search in Google Scholar

Maas, R., Kinny, P.D., Williams, I.S., Froude, D.O., and Compston, W. (1992) The Earth’s oldest known crust: A geochronological and geochemical study of 3900–4200 Ma old detrital zircons from Mt. Narryer and Jack Hills Western Australia. Geochimica et Cosmochimica Acta, 56, 1281–1300, https://doi.org/10.1016/0016-7037(92)90062-N.Search in Google Scholar

Macdonald, R., Bagiński, B., Belkin, H.E., and Stachowicz, M. (2019) Composition, paragenesis, and alteration of the chevkinite group of minerals. American Mineralogist, 104, 348–369, https://doi.org/10.2138/am-2019-6772.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

Maier, W.D. (2005) Platinum-group element (PGE) deposits and occurrences: Mineralization styles, genetic concepts, and exploration criteria. Journal of African Earth Sciences, 41, 165–191, https://doi.org/10.1016/j.jafrearsci.2005.03.004.Search in Google Scholar

Marks, M.A.W. and Markl, G. (2017) A global review on agpaitic rocks. Earth-Science Reviews, 173, 229–258, https://doi.org/10.1016/j.earscirev.2017.06.002.Search in Google Scholar

Merlino, S. and Perchiazzi, N. (1988) Modular mineralogy in the cuspidine group of minerals. Canadian Mineralogist, 26, 933–943.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

Mitchell, R.H., Ed. (1996a) Undersaturated Alkaline Rocks: Mineralogy, Petrogenesis, and Economic Potential, 312 p. Mineralogical Association of Canada Short Course, 24.Search in Google Scholar

Mitchell, R.H., Ed. (1996b) Classification of undersaturated and related alkaline rocks. Mineralogical Association of Canada Short Course, 24, 1–22.Search in Google Scholar

Mitchell, R.N., Spencer, C.J., Kirscher, U., and Wilde, S.A. (2022) Plate tectonic-like cycles since the Hadean: Initiated or inherited? Geology, 50, 827–831, https://doi.org/10.1130/G49939.1.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., Buongiorno, J., Downs, R. T., Eleish, A., Fox, P., Giovannelli, D., Golden, J.J., Hummer, D.R., Hystad, G., Kellogg, L.H., and others. (2020) Exploring carbon mineral systems: Recent advances in C mineral evolution, mineral ecology, and network analysis. Frontiers in Earth Science (Lausanne), 8, 208, https://doi.org/10.3389/feart.2020.00208.Search in Google Scholar

Morrison, S.M., Hazen, R.M., and Prabhu, A. (2023) An evolutionary system of mineralogy, part VI: Earth’s earliest Hadean crust (>4370 Ma). American Mineralogist, 108(1), 42–58.Search in Google Scholar

Moynier, F., Yin, Q.-Z., Irisawa, K., Boyet, M., Jacobsen, B., and Rosing, M.T. (2010) Coupled 182W-142Nd constraint for early Earth differentiation. Proceedings of the National Academy of Sciences, 107, 10810–10814, https://doi.org/10.1073/pnas.0913605107.Search in Google Scholar

Mungall, J.E. and Naldrett, A.J. (2008) Ore deposits of the platinum-group elements. Elements, 4, 253–258, https://doi.org/10.2113/GSELEMENTS.4.4.253.Search in Google Scholar

Nance, R.D., Murphy, J.B., and Santosh, M. (2014) The supercontinent cycle: A retrospective essay. Gondwana Research, 25, 4–29, https://doi.org/10.1016/j.gr.2012.12.026.Search in Google Scholar

Newman, M.E.J. (2010) Networks: An Introduction, 772 p. Oxford University Press.Search in Google Scholar

O’Driscoll, B. and Van Tongeren, J.A. (2017) Layered intrusions: From petrological paradigms to precious metal repositories. Elements, 13, 383–389, https://doi.org/10.2138/gselements.13.6.383.Search in Google Scholar

Ouzegane, K., Fourcade, S., Kienast, J.-R., and Javoy, M. (1988) New carbonatite complexes in the Archaean In’Ouzzal nucleus (Ahaggar, Algeria): Mineralogical and geochemical data. Contributions to Mineralogy and Petrology, 98, 277–292, https://doi.org/10.1007/BF00375179.Search in Google Scholar

Philpotts, A.R. and Ague, J.J. (2009) Principles of Igneous and Metamorphic Petrology, 2nd ed., 720 p. Cambridge University Press.Search in Google Scholar

Rastsvetaeva, R.K. and Chukanov, N.V. (2012) Classification of eudialyte-group minerals. Geology of Ore Deposits, 54, 487–497, https://doi.org/10.1134/S1075701512070069.Search in Google Scholar

Rastsvetaeva, R.K., Chukanov, N.V., and Aksenov, S.M. (2016) The crystal chemistry of lamprophyllite-related minerals: A review. European Journal of Mineralogy, 28, 915–930, https://doi.org/10.1127/ejm/2016/0028-2560.Search in Google Scholar

Rokarch, L. and Maimon, O. (2006) Clustering methods. In O. Maimon and L. Rokarch, Eds., Data Mining and Knowledge Discovery Handbook, 321–352. Springer.Search in Google Scholar

Rollinson, H. (2007) Early Earth Systems: A Geochemical Approach, 296 p. Blackwell Publishing.Search in Google Scholar

Roycroft, P.D. (1991) Magmatically zoned muscovite from the peraluminous two-mica granites of the Leinster batholith, southeast Ireland. Geology, 19, 437–440, https://doi.org/10.1130/0091-7613(1991)019<0437:MZMFTP>2.3.CO;2.Search in Google Scholar

Rudnick, R.L. and Gao, S. (2003) Composition of the continental crust. Treatise on Geochemistry, 3, 1–64.Search in Google Scholar

Salvioli-Mariani, E., Boschetti, T., Toscani, L., Montanini, A., Petriglieri, J.R., and Bersani, D. (2020) Multi-stage rodingitization of ophiolite bodies from Northern Apennines (Italy): Constraints from petrography, geochemistry and thermodynamic modeling. Geoscience Frontiers, 11, 2103–2125, https://doi.org/10.1016/j.gsf.2020.04.017.Search in Google Scholar

Savelyev, D.P., Kamenetsky, V.S., Danyushevsky, L.V., Botcharnikov, R.E., Kamenetsky, M.B., Park, J.-W., Portnyagin, M.V., Olin, P., Krasheninnikov, S.P., Hauff, F., and others. (2018) Immiscible sulfide melts in primitive oceanic magmas: Evidence and implications from picrite lavas (Eastern Kamchatka, Russia). American Mineralogist, 103, 886–898, https://doi.org/10.2138/am-2018-6352.Search in Google Scholar

Schertl, H.-P., Mills, S.J., and Maresch, W.V. (2018) A Compendium of IMA-Approved Mineral Nomenclature, 353 p. International Mineralogical Association.Search in Google Scholar

Schrenk, M.O., Brazelton, W.J., and Lang, S.Q. (2013) Serpentinization, carbon, and deep life. Reviews in Mineralogy and Geochemistry, 75, 575–606, https://doi.org/10.2138/rmg.2013.75.18.Search in Google Scholar

Shand, J.S. (1943) Eruptive Rocks: Their Genesis, Composition, Classification, with a Chapter on Meteorites. 2nd ed., 444 p. John Wiley and Sons.Search in Google Scholar

Simandl, G.J. and Paradis, S. (2018) Carbonatites: Related ore deposits, resources, footprint, and exploration methods. Applied Earth Sciences, 117, 123–152.Search in Google Scholar

Sizova, E., Gerya, T., Brown, M., and Perchuk, L.L. (2010) Subduction styles in the Precambrian: Insight from numerical experiments. Lithos, 116, 209–229, https://doi.org/10.1016/j.lithos.2009.05.028.Search in Google Scholar

Sizova, E., Gerya, T., and Brown, M. (2014) Contrasting styles of Phanerozoic and Precambrian continental collision. Gondwana Research, 25, 522–545, https://doi.org/10.1016/j.gr.2012.12.011.Search in Google Scholar

Sjöqvist, A.S.L., Cornell, D.H., Andersen, T., Christensson, U.I., and Berg, J.T. (2017) Magmatic age of rare-earth element and zirconium mineralisation at the Norra Kärr alkaline complex, southern Sweden, determined by U-Pb and Lu-Hf isotope analyses of metasomatic zircon and eudialyte. Lithos, 294–295, 73–86, https://doi.org/10.1016/j.lithos.2017.09.023.Search in Google Scholar

Sokolova, E. and Cámara, F. (2017) The seidozerite supergroup of TS-block minerals: nomenclature and classification, with change of the following names: Rinkite to rinkite-(Ce), mosandrite to mosandrite-(Ce), hainite to hainite-(Y) and innelite-1T to innelite- 1A. Mineralogical Magazine, 81, 1457–1484, https://doi.org/10.1180/minmag.2017.081.010.Search in Google Scholar

Speer, J.A. (1984) Micas in igneous rocks. Reviews in Mineralogy, 13, 299–356.Search in Google Scholar

Tassara, S., Ague, J.J., and Valencia, V. (2021) The deep magmatic cumulate roots of the Acadian orogen, eastern North America. Geology, 49, 168–173, https://doi.org/10.1130/G47887.1.Search in Google Scholar

Teertstra, D.K. and Černý, P. (1995) First natural occurrences of end-member pollucite: A product of low-temperature reequilibrium. European Journal of Mineralogy, 7, 1137–1148, https://doi.org/10.1127/ejm/7/5/1137.Search in Google Scholar

Tilley, C.E. (1921) Reviews: Die Eruptivgesteine des Kristianiagebietes. IV. Das Fengebiet in Telemerk, Norwegen. By W.C. Brøgger. Vid. Selsk. Skrifter I.M.N.Kl. No. 9, pp. 1–408. 1920. Geological Magazine, 58, 549–554, https://doi.org/10.1017/S0016756800105217.Search in Google Scholar

Tkachev, A.V. (2011) Evolution of metallogeny of granitic pegmatites associated with orogens throughout geological time. Special Publication—Geological Society of London, 350, 7–23, https://doi.org/10.1144/SP350.2.Search in Google Scholar

Trønnes, R.G., Baron, M.A., Eigenmann, K.R., Guren, M.G., Heyn, B.H., Løken, A., and Mohn, C.E. (2019) Core formation, mantle differentiation and core-mantle interaction within Earth and the terrestrial planets. Tectonophysics, 760, 165–198, https://doi.org/10.1016/j.tecto.2018.10.021.Search in Google Scholar

Valley, J.W., Cavosie, A.J., Ushikubo, T., Reinhard, D.A., Lawrence, D.F., Larson, D.J., Clifton, P.H., Kelly, T.F., Wilde, S.A., Moser, D.E., and others. (2014) Hadean age for a post-magma-ocean zircon confirmed by atom-probe tomography. Nature Geoscience, 7, 219–223, https://doi.org/10.1038/ngeo2075.Search in Google Scholar

Van Kranendonk, M.J., Smithies, R.H., and Bennett, V.C., Eds. (2007) Earth’s Oldest Rocks. Developments in Precambrian Geology, 15, 1330 p. Elsevier.Search in Google Scholar

Wan, B., Yang, X., Tian, X., Yuan, H., Kirscher, U., and Mitchell, R.N. (2020) Seismological evidence for the earliest global subduction network at 2 Ga. Science Advances, 6, eabc5491 (9 p.).Search in Google Scholar

Weidendorfer, D., Schmidt, M.W., and Mattsson, H.B. (2016) Fractional crystallization of Si-undersaturated alkaline magmas leading to unmixing of carbonatites on Brava Island (Cape Verde) and a general model of carbonatite genesis in alkaline magma suites. Contributions to Mineralogy and Petrology, 171, 43, https://doi.org/10.1007/s00410-016-1249-5.Search in Google Scholar

Whalen, J.B. (1985) Geochemistry of an island-arc plutonic suite: The Uasilau-Yau Yau intrusive complex, New Britain, P.N.G. Journal of Petrology, 26, 603–632, https://doi.org/10.1093/petrology/26.3.603.Search in Google Scholar

Wilde, S.A., Valley, J.W., Peck, W.H., and Graham, C.M. (2001) Evidence from detrital zircons for the existence of continental crust and oceans on the Earth 4.4 Gyr ago. Nature, 409, 175–178, https://doi.org/10.1038/35051550.Search in Google Scholar

Woolley, A.R. (1987) Alkaline Rocks and Carbonatites of the World. Part 1: North and South America, 216 p. British Museum of Natural History.Search in Google Scholar

Woolley, A.R. (2001) Alkaline Rocks and Carbonatites of the World. Part 3: Africa, 372 p. Geological Society of London.Search in Google Scholar

Woolley, A.R. (2019) Alkaline Rocks and Carbonatites of the World. Part 4: Antarctica, Asia and Europe (Excluding the former USSR), Australasia and Oceanic Islands, 562 p. Geological Society of London.Search in Google Scholar

Woolley, A.R. and Kjarsgaard, B.A. (2008) Paragenetic types of carbonatite as indicated by the diversity and relative abundances of associated silicate rocks: Evidence from a global database. Canadian Mineralogist, 46, 741–752, https://doi.org/10.3749/canmin.46.4.741.Search in Google Scholar

Xiong, Q., Griffin, W.L., Huang, J.-X., Gain, S.E.M., Toledo, V., Pearson, N.J., and O’Reilly, S. Y. (2017) Super-reduced mineral assemblages on “ophiolitic” chromitites and peridotites: The view from Mount Carmel. European Journal of Mineralogy, 29, 557–570, https://doi.org/10.1127/ejm/2017/0029-2646.Search in Google Scholar

Xiong, F., Xu, X., Mugnaioli, E., Gemmi, M., Wirth, R., Grew, E.S., Robinson, P.T., and Yang, J. (2020) Two new minerals, badengzhuite, TiP, and zhiqinite, TiSi2, from the Cr-11 chromitite orebody, Luobusa ophiolite, Tibet, China: Is this evidence for super-reduced mantle-derived fluids? European Journal of Mineralogy, 32, 557–574, https://doi.org/10.5194/ejm-32-557-2020.Search in Google Scholar

Yaxley, G.M. and Brey, G.P. (2004) Phase relations of carbonate-bearing eclogite assemblages from 2.5 to 3.5 GPa: Implications for petrogenesis of carbonatites. Contributions to Mineralogy and Petrology, 146, 606–619, https://doi.org/10.1007/s00410-003-0517-3.Search in Google Scholar

Yaxley, G.M., Kjarsgaard, B.A., and Jaques, A.L. (2021) Evolution of carbonatite magmas in the upper mantle and crust. Elements, 17, 315–320, https://doi.org/10.2138/gselements.17.5.315.Search in Google Scholar

Yoder, H.S. Jr. (1976) Generation of Basaltic Magma, 281 p. National Academy of Sciences Press.Search in Google Scholar

Yoder, H.S. Jr. (1979) The Evolution of the Igneous Rocks: Fiftieth Anniversary Perspectives, 600 p. Princeton University Press.Search in Google Scholar

Yoder, H.S. Jr. (1992) Norman L. Bowen (1886–1956): MIT class of 1912. First predoctoral fellow of the Geophysical Laboratory. Earth Sciences History, 11, 45–55, https://doi.org/10.17704/eshi.11.1.u8w2610560328526.Search in Google Scholar

Yoder, H.S. Jr. (1998) Norman L. Bowen: The experimental approach to petrology. GSA Today, 10–11.Search in Google Scholar

Zaitsev, A.N. and Keller, J. (2006) Mineralogical and chemical transformation of Oldoinyo Lengai natrocarbonates, Tanzania. Lithos, 91, 191–207, https://doi.org/10.1016/j.lithos.2006.03.018.Search in Google Scholar

Zaitsev, A.N., Wenzel, T., Vennemann, T., and Markl, G. (2013) Tindewret volcano, Kenya: An altered natrocarbonatite locality? Mineralogical Magazine, 77, 213–226, https://doi.org/10.1180/minmag.2013.077.3.01.Search in Google Scholar

Appendix I Systematic mineralogy of primary igneous minerals

Appendix I presents a systematic mineralogy of the most frequently encountered primary igneous minerals. A complete list of primary igneous minerals appears in Online Materials[1] Table OM2, which lists several attributes of 919 igneous mineral kinds, including their chemical formulas (column O), their relative abundances (column B), and the distribution of these phases among eight major groups of igneous rocks (columns F to M). Relative abundances are indicated with a qualitative scale, based on modal data in Online Materials[1] Table OM3 (see also Online Materials[1] Read-Me File 3): 51 of the most abundant igneous minerals, each with at least 10 occurrences as a major mineral (>5 vol%) in Online Materials[1] Table OM3, are designated with a “4” in column B, while 64 additional minerals with fewer than 10 occurrences as a major mineral, but in most cases 5 or more occurrences as an accessory phase (<5 vol%) are denoted with a “3.” Online Materials[1] Table OM2 also lists 182 uncommon phases designated “2” and 622 rare minerals (known as trace minerals in 5 or fewer igneous rocks) designated “1.” Online Materials[1] Table OM2 records the total number of occurrences of each mineral (Column C), as well as the number of those occurrences that are >5 vol% (Column D) and <5 vol% (Column E), based on igneous rock mode data in Online Materials[1] Table OM3.

Mineral species vs. natural kinds: The conversion of 1665 IMA-CNMNC-approved mineral species into 919 natural kinds requires several modifications to the IMA list, as detailed in Hazen et al. (2022). In 651 instances, the IMA species name (e.g., bertrandite) is identical to the natural kind name (bertrandite). 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 1001 IMA-approved species, we lump groups of two or more IMA species into single natural kinds. These numerous examples, resulting in a reduction from 1001 species to 251 natural kinds, are detailed in Online Materials[1] Table OM6 (see also Online Materials[1] Read-Me File 6). For example, “eudialyte” combines 30 IMA-approved species of the eudialyte group, while “labuntsovite” groups 27 species of the labuntsovite group. In 12 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.

In three instances, we recognize intermediate compositions of solid solutions between IMA-CNMNC-approved mineral species as separate natural kinds. For example, we include “perthite” (intermediate compositions of alkali feldspar with distinctive exsolution of Na- and K-rich feldspar), in addition to the Narich end-member albite, and the K-rich end-members sanidine, microcline, and orthoclase. Similarly, we define “plagioclase” as intermediate compositions of the albite-anorthite series with between 30 and 70 mol% anorthite component (i.e., corresponding to the obsolete feldspar types “andesine” and “labradorite”), along with albite (which includes albite and the obsolete “oligoclase”) and anorthite (“bytownite” and anorthite). Similarly, we define “magnesiowüstite” as intermediate compositions of the solid solution between periclase (MgO) and wüstite (FeO). We also introduce two kinds of Si-bearing glass as natural kinds: the SiO2-rich volcanic glass, “obsidian,” with SiO2 >>70 wt%, as well as a broad range of compositions of “silicate glass,” including glass of basaltic composition.

The core data of this study are found in Online Materials[1] Table OM3 (see also Online Materials[1] Read-Me File 3), which is a spreadsheet that details the distribution of 115 of the most common primary igneous minerals (columns E to DO) among 1850 igneous rock modes (lines 4 to 1853). Each major mineral phase (>5 vol%) is designated by “2” in the appropriate matrix element, whereas minor/accessory phases are designated “1.” In addition, columns A, B, and C record a literature reference for the rock mode, the rock’s locality, and the cited (often obsolete) rock name, respectively, while column D lists one of the 8 different 3-letter abbreviations (see above) for each rock.

Note that one consequence of this data consolidation exercise was a modest modification of the distribution of primary igneous minerals among 8 igneous paragenetic modes, compared to those recorded in two earlier studies [Hazen and Morrison (2022), their Online Materials[1] Table OM2; and Hazen et al. (2022), their Online Materials[1] Table OM1]. Consequently, in Online Materials[1] Table OM2 we have modified the distributions of 74 of the most common primary igneous minerals among the 8 broad categories compared to these earlier studies, in particular, adding 148 new mineral/mode combinations that we found in our compilations, while removing 19 combinations that did not occur in any of the 1850 tabulated rocks.

In constructing this data resource, we relied heavily on the monumental Descriptive Petrography of the Igneous Rocks of Albert Johannsen (1932, 1937, 1938), who documented qualitative or quantitative mineral modes for more than one thousand varied igneous rocks. From this source, we tabulated 983 modes, including 271 rocks from Volume II (“The Quartz-Bearing Rocks”), 278 rocks from Volume III (“The Intermediate Rocks”), and 434 rocks from Volume IV (“The Feldspathoid Rocks” and “The Peridotites and Perknites”).

To these analyses we added 795 igneous rock modes from a wide range of alkaline rocks and carbonatites, which were transcribed from entries in the exhaustive four-volume work, Alkaline Rocks and Carbonatites of the World by Alan R. Woolley and colleagues (Woolley 1987, 2001, 2019; Kogarko et al. 1995). These data were especially important because Johannsen questioned the occurrence of carbonates as primary igneous minerals [in spite of earlier findings by Brøgger, Tilley, and others, as summarized by Tilley (1921)], while the identification of many rare accessory minerals in agpaitic and miaskitic rocks postdated Johannsen’s work.

In addition, we compiled qualitative modal data on 72 rocks described in Rock-Forming Minerals (Deer et al. 1982–2013) and/or on https://mindat.org (accessed 20 January 2022), primarily complex granite pegmatites, which were not systematically surveyed by Johannsen or Woolley. The resulting table of coexisting phases among 115 of the most common primary mineral kinds in 1850 igneous rocks provides the raw data for much of the analysis in this contribution (Online Materials[1] Table OM3 and associated Online Materials[1] Read-Me File 3).

Primary vs. secondary igneous minerals: A recurrent difficulty in any treatment of igneous minerals relates to the definition of a primary phase. We define “primary” igneous minerals as those that formed during the initial cooling of a magma body. In the majority of rocks cited in Online Materials[1] Table OM3, igneous minerals crystallized directly from magma and display equilibrium eugranitic textures such as well defined triple junctions (e.g., Harker 1964; Philpotts and Ague 2009). The resulting suites of minerals are primary. As some igneous rocks cool and their late-stage fluids become increasingly aqueous and enriched in incompatible elements, depositional environments may become hydrothermal or even pneumatolytic (e.g., some occurrences of tourmaline, topaz, cassiterite, and/or fluorite in complex granite pegmatites; Anthony et al. 1990–2003; London 2008). We include all of these minerals as primary igneous phases, because they represent a continuous sequence of magma cooling and crystallization.

We also include as primary minerals those phases that arise through solid-state transformation on gradual cooling. Examples include the Al-Si ordering transitions from sanidine to orthoclase, as well as from orthoclase to microcline, in the alkali feldspar system [(K,Na)AlSi3O8]; the exsolution of pigeonite from augite in the clinopyroxene system [(Ca,Mg,Fe)2Si2O6]; and the transformation of high-temperature forms of SiO2 (e.g., tridymite or cristobalite) to quartz.

However, ambiguity may arise as a consequence of auto-metasomatism— the self-alteration of an igneous body by interaction with its own hot fluids during cooling. Potential examples include processes such as fenitization [alkali metasomatism associated with the intrusion of alkaline rocks and carbonatites (Le Bas 2008; Kapustin 2010; Elliott et al. 2018)]; myrmekitization [the metasomatic alteration of K-feldspar to quartz plus Na-rich plagioclase Castle and Lindsley (1993)]; rodingitization [auto-metasomatism of ultramafic rocks (Salvioli-Mariani et al. 2020)]; saussuritization [the in situ alteration of calcic plagioclase to zoisite, chlorite, and other phases (Deer et al. 1986)]; and serpentinization [alteration/oxidation of mafic silicates to serpentine, brucite, magnetite, and other phases (Schrenk et al. 2013)]. We do not recognize these alteration products as primary igneous phases, even if the fluids involved in the alteration were derived exclusively from the magma. In addition, we do not list such oft-reported secondary phases as “iddingsite” (clay minerals and hematite after forsterite), “liebenerite” (muscovite after nepheline), “martite” (hematite after magnetite), “palagonite” (clay minerals after silicate glass), “pinite” (sericite and clay minerals after nepheline), “pseudoleucite” (nepheline and K-feldspar after leucite), “smaragdite” (actinolite after diopside), and “uralite” (hornblende after augite). However, we do include the precursor primary igneous phases from which these alteration products derive; i.e., when “martite” is reported in a modal analysis, we list magnetite as a primary phase.

Systematic mineralogy: Of the 919 primary mineral kinds recorded in Online Materials[1] Table OM1, 115 phases are relatively common based on their representation in 1850 igneous rocks (Table 2; Online Materials[1] Tables OM2 and OM3). Here we present brief descriptions of these mineral kinds, arranged according to the New Dana Classification (Gaines et al. 1997).

Native elements

Graphite (C), which has been documented as a primary igneous mineral in a few alkaline rocks and carbonatites, is the only native element that occurs in any significant abundance.

More than 20 platinum and platinum group element alloys (e.g., Ru, Rh, Pd, Os, Ir, and Pt), as well as gold, silver, and their alloys, notably with copper, are important economic resources in some layered igneous intrusions and ultramafic/mafic lithologies (Kutyrev et al. 2021; Berdnikov et al. 2022). However, these phases, though at times economically important, are volumetrically minor. Several native metals (e.g., Al, Fe, Ti, Zn) representing extremely reduced environments also occur as a trace phase in some ultramafic lithologies, though their paragenesis is a matter of some controversy (Dobrzhinetskaya et al. 2009; Xiong et al. 2017, 2020; Griffin et al. 2018).

Sulfides

Numerous sulfide minerals are associated with igneous rocks, notably in the context of economically valuable late-stage hydrothermal deposits (Guilbert and Park 2007). Sulfide and silicate melts tend to be immiscible (Lester et al. 2013; Savelyev et al. 2018), so crystallization of O- vs. S-bearing minerals often follows parallel pathways, at times with S- and O-rich melts separated by density.

Pyrite (FeS2), the most common igneous sulfide, is reported as an accessory phase in 288 (i.e., 16%) of the rocks in Online Materials[1] Table OM3. Pyrite is thus by far the most frequent S-bearing primary phase in igneous rocks. The host lithologies span the range from granites and their pegmatites to intermediate, mafic, and ultramafic rocks, as well as varied alkaline rocks and carbonatites. This unusual versatility parallels the observation that pyrite displays the largest number of different paragenetic modes of any mineral (Hazen and Morrison 2022).

We record pyrrhotite (Fe7S8) from 64 igneous rocks, spanning a range of lithologies similar to that of pyrite. In addition, we record chalcopyrite (CuFeS2; 38 occurrences), galena (PbS; 37), molybenite (MoS2; 11), and sphalerite (ZnS; 30) as suspected primary phases from varied igneous rocks, including granites and their pegmatites, alkaline rocks, and carbonatites.

At least 40 other sulfides, along with dozens of volumetrically minor arsenides, antimonides, and sulfosalts, appear to be primary igneous phases (Online Materials[1] Table OM2). However, in many instances it is uncertain whether these minerals associated with igneous rocks are primary igneous phases, as opposed to arising from later hydrothermal and/or alteration processes.

Oxides

Oxides are ubiquitous accessory phases, as well as occasional major phases, in the entire range of igneous rocks. More than 100 species, most of them volumetrically trivial, have been reported (Online Materials[1] Table OM1). The most common simple oxide is rutile (TiO2), which is reported in 100 of the igneous rocks from our study (Online Materials[1] Table OM3), most often in granites, syenites, and carbonatites. The TiO2 polymorphs, anatase (20 occurrences) and brookite (6), are less common, occurring primarily in granitic rocks and carbonatites (Bowles et al. 2011). An added complexity with the TiO2 polymorphs is recognizing primary vs. secondary occurrences. For example, brookite may occur as an alteration product of titanite, whereas brookite may alter to rutile.

Additionally, among the more common simple oxides are baddeleyite (ZrO2), which we list from 37 carbonatites and alkaline rocks; corundum (Al2O3; 20 occurrences), observed most commonly in Al-rich syenites; and cassiterite (SnO2; 21 occurrences) and uraninite (UO2; 9 occurrences), usually from granites and their complex pegmatites.

By far the most frequently encountered oxides are members of the spinel group, which occur in more than half of the igneous rocks we tabulated. This binary oxide group, with the general formula [(Mg,Fe2+)(Al,Fe3+,Cr3+,Ti)2O4], encompasses a complex range of solid solutions, spanning eight major end-members: chromite F e 2 + C r 2 3 + O 4 , hercynite (Fe2+Al2O4), magnesiochromite (MgAl2O4), magnesioferrite M g F e 2 3 + O 4 magnetite F e 2 + F e 2 3 + O 4 , qandilite (Mg2Ti4+O4), spinel (MgAl2O4), and ulvöspinel F e 2 2 + T i 4 + O 4 , with some specimens incorporating significant amounts of other cations, including Ca, Mn2+, Zn, V3+, Ti3+, and/or Si (Bowles et al. 2011). Because most of these examples are opaque in thin section, petrographers often do not differentiate the several spinel species. Alternatively, generic names such as “pleonaste”, “chrome-spinel”, and “picotite” are employed (Table 2).

In this study, we recognize four natural kinds of oxide spinels. Magnetite is by far the most frequently reported oxide, occurring in 1161 of the 1850 igneous rocks we compiled (63%), and appearing as a major phase (>5 vol%) in 292 of those rocks. Included in our tabulations of primary magnetite are minerals identified as “iron ores,” “martite” (hematite pseudomorphs after magnetite), and “Ti-magnetite” or “titanomagnetite.” In addition, when explicitly reported, we distinguish chromite (31 occurrences; including “chrome-spinel”) and hercynite (8; also cited as “picotite”) from mostly mafic/ultramafic rocks, and spinel (53; often reported as “pleonaste”) from mafic/ultramafic rocks and carbonatites.

Ilmenite (FeTiO3) is another common binary oxide, which occurs in 330 (18%) of the rocks we compiled. Like magnetite, ilmenite is an opaque Fe-rich phase that occurs across the spectrum of igneous rocks.

Perovskite (CaTiO3) was reported from 137 basic and alkaline rocks and carbonatites; it is a major phase in 27 of those lithologies.

Complex oxides are important phases in many highly differentiated rocks, notably rare-element granitic and agpaitic pegmatites. We lump six IMA-CNMNC-approved species of the columbite-tantalite series [(Fe,Mg)(Ta,Nb)2O6] into columbite, which appears as a minor phase in 62 rare-element pegmatites. Similarly, 7 species of the euxenite group [(Y,Ca,Ce,U,Th)(Nb,Ta,Ti)2O6], reported from 8 highly differentiated rocks, are lumped into euxenite; 7 species from the fergusonite group [(Y,REE)(Nb,Ta)O4] from 7 rocks are lumped into fergusonite; 4 species from the zirconolite group [(Ca,Y)Zr(Ti,Mg,Al)2O7] from 9 agpaitic rocks are lumped into zirconolite; and 8 species from the aeschynite group [(Y,REE,Ca,Th) (Ti,Nb,Ta)2(O,OH)6] from 10 rocks are lumped into aeschynite.

The pyrochlore oxide supergroup [(Ca,Na,REE,U,o)2(Nb,Ta,Sb,Ti,W)2O6 (O,OH,F,H2O)] presents significant challenges in species identification for the petrographer. The IMA-CNMNC has recognized five mineral groups (pyrochlore, microlite, roméite, betafite, and elsmoreite, with Nb, Ta, Sb, Ti, and W dominant, respectively), which are further divided into at least 38 species (Atencio et al. 2010; Christy and Atencio 2013). The petrographic literature only records the supergroup or group name and does not distinguish among these numerous species. Accordingly, in Online Materials[1] Tables OM2, OM3, and OM6 we lump 7 species into the Nb-dominant pyrochlore, which is recorded from 172 igneous rocks (mostly complex pegmatites, carbonatites, and alkaline rocks), and 10 species into Tadominant microlite, reported from 27 rocks of similar lithologies.

Halides

At least 20 halide minerals have been reported as primary phases in igneous rocks (Online Materials[1] Table 2). However, fluorite (CaF2) is the only common igneous halide (Chang et al. 1996). We record 213 fluorite-bearing igneous rocks, mostly granites, alkaline rocks, and carbonatites.

Carbonates

Over the past century, carbonate minerals have been increasingly recognized as primary igneous minerals (Tilley 1921; Chang et al. 1996), with more than 60 recorded species (Online Materials[1] Table OM1), most notably in carbonatites (Woolley 1987, 2001, 2019; Kogarko et al. 1995). The majority of occurrences of igneous carbonate minerals feature divalent cations with a general formula [(Ca,Mg,Fe)2(CO3)2]— a compositional span that encompasses five igneous species approved by the IMA-CNMNC: calcite (CaCO3), magnesite (MgCO3), siderite (FeCO3), dolomite [CaMg(CO3)2], and ankerite [CaFe(CO3)2]. An important mineralogical distinction occurs between carbonates in which all divalent cation sites have the same average composition (calcite, magnesite, and siderite), vs. those that display cation ordering in alternating sites (dolomite, ankerite).

Petrographers employ these same five carbonate mineral names; however, the nomenclature in the petrographic literature is at times confused. Calcite is correctly named in 347 igneous rocks, the great majority of which are Ca-dominant carbonatites known as “sovites” or “alvikites.” Dolomite, reported from 99 rocks in our compilation, in most instances occurs as the major mineral in Ca-Mg-bearing carbonatites, sometimes recorded as “beforsites” or “rauhaugites.” However, in these instances “dolomite” may represent either the ordered Ca-Mg carbonate or disordered Mg-bearing calcite.

The two iron-bearing carbonatites reported in the carbonatite literature, ankerite (50 occurrences) and siderite (11 occurrences), often do not strictly conform to the IMA species definitions. In many instances, “ankerite” in the petrographic literature denotes an Fe-bearing dolomite with significant Fe substituting for Mg, though Mg > Fe in most examples. Similarly, “siderite” may correspond to calcite with significant Fe substituting for Ca, even if Ca > Fe. In Online Materials[1] Table OM3 we record the mineral name included in the original literature (in particular as recorded by Woolley 1987, 2001, 2019; Kogarko et al. 1995).

In addition to these most common carbonate minerals, a few rocks incorporate other divalent-cation carbonates that occur as primary igneous phases: magnesite (MgCO3; 3 occurrences), rhodochrosite (MnCO3; 2), strontianite (SrCO3; 26), norsethite [BaMg(CO3)2; 2], barytocalcite [BaCa(CO3)2; 2], and alstonite [BaCa(CO3)2; 2].

Several rare alkali carbonates, including burbankite [(Na,Ca)3(Sr,Ba,Ce)3 (CO3)5, fairchildite [K2Ca(CO3)2], gregoryite [Na2(CO3)], and nyererite [Na2Ca(CO3)2], occur as abundant primary phases in rare alkali carbonatites, such as those erupted at Oldoinyo Lengai in Tanzania. These minerals may be relatively common in young carbonatites, but they are ephemeral and easily lost in the geological record (Zaitsev and Keller 2006). As a result, alkali carbonatites may be significantly under-represented in the modern sample (Zaitsev et al. 2013). Consequently, our catalog records fewer than five primary igneous localities for each alkali carbonate mineral. Nevertheless, we include gregoryite and nyererite in our list of the 115 most common primary igneous minerals, because they are the two dominant carbonate minerals in alkali carbonatites.

Of special note are four groups of rare-earth element carbonates that occur as accessory minerals in some carbonatites. We lump two species of the ancylite group [(La,Ce)Sr(CO3)2(OH)·H2O] into ancylite, which is recorded from 17 carbonatites and complex pegmatites. Similarly, we lump 7 species of the bastnaesite group into bastnaesite [(Y,REE)(CO3)(F,OH); 45 occurrences]; 2 species of the parisite group into parisite [Ca(REE)2(CO3)3F2; 22 occurrences]; and 4 species of the synchesite group into synchesite [(Ca,Ba)(Y,REE)(CO3)2F; 20 occurrences].

Sulfates

A few sulfate minerals with divalent cations Ca, Sr, and Ba have been reported as primary igneous minerals. Baryte (BaSO4; often reported as “barite”) is the most common example, occurring in 79 in igneous rocks, primarily carbonatites.

Phosphates

Primary igneous phosphates are common accessory phases, with more than 100 recorded species, most of which occur in trace amounts (Online Materials[1] Table OM1). “Apatite” [Ca5(PO4)3(F,OH,Cl)] is the most common primary igneous mineral, reported in 1234 of 1850 (67%) of rocks in our compilation (Online Materials[1] Table OM3). The great majority of these occurrences correspond to the F-dominant species, fluorapatite, which is common across the full range of igneous rocks as the most common P- and F-bearing phase. However, a small fraction of apatite occurrences may be hydroxylapatite or chlorapatite, which are occasionally encountered as igneous minerals (Chang et al. 1996).

Two types of rare-earth element phosphates, the monazite and xenotime groups, occur frequently as accessory minerals in granites and their pegmatites, carbonatites, and other igneous lithologies (Chang et al. 1996). We lump 4 species of the monazite group [(REE)PO4] into monazite, which we record from 94 rocks. Three species of the xenotime group [(Y,REE)(P,As)O4] are combined into xenotime, with 23 occurrences.

We also lump three species of Li-bearing phosphates into amblygonite [Li(Al,Fe3+)PO4(F,OH)], which we record from 16 rare-element granite pegmatites, typically in association with beryl, lepidolite, and spodumene.

Silicates

Silicates are by far the most common primary igneous phases, both volumetrically and in terms of diversity. In Online Materials[1] Table OM2 we record 375 different silicate mineral kinds, of which 69 are frequently encountered major phases and/or common accessory minerals. In the following sections we review these more common silicates.

Nesosilicates or Orthosilicates

Three members of the olivine group [(Mg,Fe,Ca)2SiO4] are important primary igneous minerals (Deer et al. 1982). Olivine with Mg/Fe >1 (i.e., forsterite) is reported in 412 of the rocks we surveyed, notably in mafic and ultramafic rocks, as well as alkaline rocks. Fayalite (with Mg/Fe < 1; ideally Fe2SiO4) is much less common, occurring in 30 rocks, mostly granites, syenites, and their extrusive equivalents. Monticellite (CaMgSiO4), though more familiar as a skarn mineral in the context of contact metamorphism, also occurs occasionally in carbonatites, alkaline rocks, and mafic lithologies (Deer et al. 1982). Here, we record monticellite as a primary igneous mineral in 24 rocks.

Four members of the garnet group—almandine F e 3 2 + A l 2 S i O 4 3 , andradite C a 3 F e 2 3 + S i O 4 3 , pyrope [Mg3Al2(SiO4)3], and spessartine M n 3 2 + A l 2 S i O 4 3 - are reported as primary minerals in igneous rocks (Deer et al. 1982). Note, however, that almost all igneous garnets are solid solutions of three or more end-members (Chiama et al. 2020, 2022). Andradite, also cited as “melanite” or “schorlomite,” is the only common igneous garnet, recorded in 199 of the rocks in our compilation, including granitic, intermediate, basic, and alkaline rocks. In addition, Almandine is reported from 7 granite pegmatites, whereas spessartine is found in 7 granite or syenite pegmatites.

Clinohumite [Mg9(SiO4)4(F,OH)2], an orthosilicate most commonly associated with metamorphosed carbonate sediments, is recorded as a likely primary phase from 6 carbonatites and ultramafic rocks in our inventory.

Among the most common silicate accessory igneous minerals are the orthosilicates zircon (ZrSiO4), listed in 400 of 1850 rocks in our tabulation (22%); its isomorph thorite (ThSiO4) identified in 30 rocks; and titanite (CaTiSiO5; commonly reported as “sphene”), which occurs in 518 rocks (28%). Both zircon and sphene occur in a wide range of igneous lithologies, whereas thorite is found primarily in alkaline rocks, carbonatites, and complex granite pegmatites.

Several orthosilicates are most frequently found in complex granite pegmatites. Topaz (Al2SiO4F2), phenakite (Be2SiO4), and eucryptite (LiAlSiO4), recorded from 26, 9, and 5 rocks in Online Materials1 Table OM3, respectively, are commonly associated with beryl, lepidolite, muscovite, and tourmaline. Note that topaz is usually a pneumatolytic phase, and thus one of the last primary igneous minerals to crystallize (Deer et al. 1982). Johannsen (1932, p. 21) notes that “It is sometimes rather difficult to draw the line between contact metamorphosed and true igneous quartz-topaz rocks. In each case the minerals are pneumatolytic; but in the former a pre-existing rock was saturated and changed by the mineralizers; while in the latter the mineralizer-rich solutions crystallized as quartz and topaz.”

The britholite group [(Y,REE,Ca)5(SiO4)3(OH,F)], also known as “beckelite,” encompasses several silicate apatites. We lump 5 species of this group into britholite, which we record in 14 carbonatites, granites, and alkaline rocks.

Sorosilicates or disilicates

Melilite is the group name for the solid solution between åkermanite [Ca2(Al2SiO7)] and gehlenite [Ca2(MgSi2O7)] (Deer et al. 1986). Melilite is a common major mineral in basic alkaline magmas, being represented in 83 rocks of our compilation, with 63 examples containing more than 5 modal percent.

Several sorosilicates occur as occasional accessory phases in igneous rocks. For example, the beryllium silicate bertrandite [Be4Si2O7(OH)2] is recorded from 6 complex granite pegmatites in Online Materials1 Table OM3.

Five other important examples of accessory igneous sorosilicates come from mineral groups that incorporate rare elements. Most frequently encountered are species of the allanite subgroup (also called “orthite”), which are Y- and REE-rich members of the diverse epidote group (Armbruster et al. 2006) with a range of compositions: [(Ca,Mn)(Y,REE)(Al,Fe3+,Fe2+)(Si2O7)(SiO4)O(OH)]. We lump 9 species of this subgroup into allanite, which appears in 95 of the igneous rocks in our compilation, most of which are in felsic lithologies from granite to syenite. Allanite is present in major amounts in only one example, a carbonatite from Ahaggar, Algeria (Ouzegane et al. 1988). Note that several other epidote group minerals, including zoisite, clinozoisite, epidote, and piemontite (all hydrous Ca-bearing aluminosilicates), are sometimes reported from igneous rocks, but we ascribe these phases to secondary processes, such as metamorphism or hydrous alteration (e.g., saussuritization; Deer et al. 1982, 2001).

Members of the chevkinite group ( R E E ) 4 T i 4 + , F e 2 + , F e 3 + , Z r , M n , C r , W , 5 O 8 (Si2O7)2], including 6 species that we lump into chevkinite, occur as scarce accessory minerals in alkaline rocks (Macdonald et al. 2019). We list 14 occurrences in Online Materials[1] Table OM3.

The lamprophyllite group (Rastsvetaeva et al. 2016), with 8 species lumped into lamprophyllite [corresponding to (Sr,Ba)(Na,K)Ti@@@4+Na3Ti4+(Si2O7)2O2(OH)2], has a similar paragenesis, with 13 occurrences in alkaline rocks.

The seidozerite supergroup includes a suite of complex and diverse minerals (Sokolova and Cámara 2017), with at least 48 approved species (https://rruff.info/ima, accessed 20 January 2022). We lump 12 of these species from the rinkite group [(Na,Ca,Mn,Y,REE)4(H2O,□)2(Ti4+,Zr,Nb)(Si2O7)2(OH,F)4–x(H2O)x] into rinkite. Our tabulations record 39 occurrences of rinkite as an accessory mineral in nepheline syenites and their pegmatites.

We lump five monoclinic members of the wöhlerite group (Merlino and Perchiazzi 1988) with compositions corresponding to [(Na,Ca)(Ca,Mn2+,Fe2+)2 (Ti4+,Nb5+,Zr4+)(Si2O7)(O,F)2] into wöhlerite. The 35 occurrences recorded in Online Materials[1] Table OM3 are all from alkaline rocks.

Cyclosilicates

Members of the beryl group [(Na,Cs,□)Be2(Be,Li,B)(Al,Mg,Fe3+)2Si6O18] are by far the most abundant Be-bearing minerals (Deer et al. 1986). We lump 6 species into beryl, and record occurrences in 41 igneous rocks, most notably in granitic pegmatites, where it often occurs as a major phase in association with albite, elbaite, lepidolite, and various scarce lithium and beryllium minerals.

The diverse minerals of the tourmaline supergroup, including more than 30 IMA-approved species, are the most common boron-bearing phases in igneous rocks (Deer et al. 1986; Henry et al. 2011). We divide the igneous tourmalines into two kinds. The more common tourmaline, embracing 20 different IMA-CNMNC-approved species and sometimes reported in the petrographic literature as “dravite” (with Mg > Fe2+) or “schorl” (with Fe2+ > Mg), has a broad compositional range of [(Na,Ca,□)(Mg,Al,Fe2+,Fe3+,Ti4+)3(Al,Fe3+,Mg)6(Si6O18)(BO3)3(OH)3(O,F)]. Tourmaline occurs widely in igneous rocks, with 54 occurrences in our tabulation. We also distinguish 5 Li-bearing tourmaline species as elbaite (Henry et al. 2011), which we record from 24 complex granite pegmatites.

We lump the two species of the solid solution between catapleiite [Na2Zr(Si3O9)·2H2O] and calciocatapleiite [CaZr(Si3O9)·2H2O] into catapleiite. We report 13 occurrences in alkaline rocks, in only one of which is catapleiite a major phase—an agpaitic nepheline syenite from the Norra Kärr alkaline complex, Sweden (Sjöqvist et al. 2017).

The eudialyte group boasts at least 30 approved species corresponding to end-member compositions in the general formula [(Na,Ca,REE)(Fe,Mn)(Zr,Ti) (Si3O9)2(OH,F)·nH2O] (Johnsen and Grice 1999; Rastsvetaeva and Chukanov 2012). While petrographers rarely identify the exact species, they often distinguish between Na-dominant “eudialyte” and more calcic “eucolite.” Here we lump all 30 species of the group into eudialyte, as detailed in Hazen et al. (2022). We report 69 alkaline rocks with eudialyte, 14 of which incorporate >5 mode percent in eudialyte-bearing agpaitic pegmatites.

Inosilicates

Pyroxene Group: Chain silicates, notably pyroxenes and amphiboles, are found in the majority of igneous rocks (Deer et al. 1997a, 1997b); they occur in 79% of the modes recorded in Online Materials[1] Table OM3. We recognize six different kinds of clinopyroxenes based on their major element compositional ranges in the system [(Na,Li,Mg,Fe2+,Ca,Al,Fe3+)2Si2O6], noting that continuous solid solutions occur among several of these pyroxenes. Aegirine includes Na- and Fe3+-bearing clinopyroxenes that range from the IMA-approved species aegirine (NaFe3+Si2O6) to aegirine-augite [(Ca,Na)(Fe3+,Mg,Fe2+)Si2O6]. Aegirine is an abundant phase that we record from 521 rocks in Online Materials[1] Table OM3 (28%), usually in alkaline lithologies.

Augite, one of the few IMA-approved species that is not defined by a single end-member composition, represents clinopyroxenes in the [(Ca,Mg,Fe2+)2Si2O6] system with Ca occupying between ~25 and ~45 at% of the Ca-Mg-Fe sites. We adopt the IMA definition for augite, which we report from 504 rocks that span a range from granitic to ultramafic lithologies. Pigeonite, similarly, is defined for clinopyroxenes with ~5 to ~15 at% Ca in the Ca-Mg-Fe sites. We recognize pigeonite as a relatively uncommon primary igneous phase in terrestrial rocks. We record only 16 instances in 1850 rocks compiled—a sharp contrast to meteoritic clinopyroxenes that often display pigeonite lamellae exsolved from augite (Deer et al. 1997a; Morrison and Hazen 2021).

We recognize two kinds of Ca-rich clinopyroxene, diopside (ideally CaMgSi2O6) and hedenbergite (ideally CaFe2+Si2O6). Both of these minerals display compositional plasticity, with significant Mg-Fe2+ solid solution in both end-members. Additionally, diopside often incorporates significant amounts of Al, Fe3+, Ti, and Cr in variants sometimes named “chrome-diopside,” “diallage,” “fassaite,” “malacolite,” or “salite.” We record 228 occurrences of diopside in a wide variety of igneous rocks. The Fe2+-dominant calcic pyroxene, hedenbergite, is much less common, restricted to 15 occurrences of granitic or alkaline rocks in our compilation.

The clinopyroxene spodumene (LiAlSi2O6), an important lithium ore mineral, is restricted to 24 complex granite pegmatites in our study.

We document orthorhombic pyroxenes, with end-members enstatite (MgSiO3) and ferrosilite (Fe2+SiO3), in 135 rocks, primarily in intermediate and mafic igneous rocks. Many citations refer to intermediate Mg-Fe2+-bearing compositions “bronzite” or “hypersthene;” we call all of these occurrences orthoenstatite—a name that underscores the Mg-dominant composition of most examples, and that distinguishes the orthorhombic pyroxene from clinoenstatite.

Pyroxenoid Group: Wollastonite (CaSiO3), most commonly associated with metamorphosed limestone, occurs as a primary phase in a few igneous rocks, as well. We list 15 occurrences in alkaline rocks.

Amphibole Group: With at least 110 IMA-CNMNC-approved species, the amphibole supergroup of double-chain silicates is the most diverse of all mineral structure types (Hawthorne et al. 2011). Amphiboles occur in 775 (42%) of the 1850 rocks in Online Materials[1] Table OM3. Identifying natural kinds of amphiboles through the construction of comprehensive data resources and application of cluster analysis (e.g., Ewing 1976; Gregory et al. 2019; Boujibar et al. 2021; Hystad et al. 2021) remains a challenging ambition. Here we provisionally lump these varied amphibole species into 6 groups of primary igneous minerals.

The hornblende group of calcic amphiboles, with a general formula [(o,Na,K) Ca2(Mg,Fe2+,Al,Fe3+)5(Si,Al)8O22(OH,F,Cl)2], is the most common igneous double-chain silicate (Deer et al. 1997b; Hawthorne et al. 2011). A wide range of solid solutions, including Na-K in alkali sites, Mg-Fe-Al in octahedral sites, Al-Si in tetrahedral sites, and OH-F-Cl (see Online Materials[1] Table OM6 for a list of end-member compositions) have been documented (Deer et al. 1997b). We lump 26 species into hornblende, which we record from 487 granitic, intermediate, and mafic rocks (Online Materials[1] Table OM3).

Kaersutite [NaCa2(Mg3Fe3+Ti4+)(Si6Al2)O22O2], encompassing 2 IMA-approved species, is an anhydrous Na-bearing calcic amphibole that occurs primarily in alkalic, mafic, and ultramafic rocks. We record 19 examples in Online Materials[1] Table OM3.

Richterite encompasses 15 species of calcic alkali amphiboles with the general formula [(Na,K)(Na,Ca)2(Mg,Fe2+,Al,Fe3+)5(Si6Al2)O22(OH)2], which likely represents a continuous range of solid solutions (Deer et al. 1997b). We list 47 examples from various alkaline rocks.

We lump 10 species of alkali amphiboles spanning the compositional range [(Na,K)Na2(Mg,Fe2+,Al,Fe3+)5Si7(Si,Al)O22(OH,F)2] into arfvedsonite. We list 229 occurrences, primarily from alkaline rocks. We distinguish riebeckite, a group of 3 alkali amphibole species with vacancies, corresponding to [□Na2(Mg,Fe2+)3 Fe@@@@3+Si8O22(OH,F)2]. Online Materials[1] Table OM3 includes 68 occurrences, primarily in alkali granites and other alkaline rocks.

Actinolite, a calcic amphibole with vacancies [□Ca2(Mg,Fe2+)5Si8O22(OH,F)2], is an intermediate member of the tremolite-actinolite (Mg-Fe2+) series. We lump all 10 occurrences into actinolite, which is most often associated with metamorphic rocks and only occurs rarely as a primary igneous mineral in carbonatites and other Ca-rich lithologies.

The pectolite group of single-chain silicates, with 4 species lumped into pectolite [(Na,Li)(Ca,Mn2+)2Si3O8(OH)], is a late-stage mineral that forms near the transition from magmatic to hydrothermal crystallization in some mafic and alkaline rocks (Deer et al. 1997a). We list 12 occurrences of pectolite, in most of which it is associated with nepheline and aegirine.

Two additional inosilicate groups, both branched-chain silicates, are associated with alkaline rocks and their pegmatites. We lump 12 members of the astrophyllite group [(K,Na,Li,Cs,□,Ca)3(Fe2+,Mn2+,Mg,Na)7(Ti4+,Zr,Nb)2(Si4O12)2O2 (OH)4(OH,O,F)(H2O)n] into astrophyllite, which occurs in 37 agpaitic rocks listed in Online Materials1 Table OM3. We also include aenigmatite N a 4 F e 10 2 + T i 2 O 4 S i 12 O 36 , recorded in 53 varied alkaline rocks.

Phyllosilicates

Mica Group: Mica group minerals occur abundantly as primary igneous minerals (Speer 1984; Fleet 2003), being found in 1107 of 1850 rocks (60%) compiled in this survey (Online Materials[1] Table OM3). In 646 instances (35%), one or more kinds of mica is a major phase. At least 34 mica species have been approved by the IMA-CNMNC; however, from a petrographic perspective we recognize only 4 mica groups as primary igneous phases. By far the most common mica group is biotite K F e 2 2 + F e 2 + , M g , M n 2 + S i , A l , F e 3 + 2 S i 2 O 10 ( O H , F , C l ) 2 which encompasses a range of Fe-bearing, dark-colored trioctahedral micas that have been reported in 866 of 1850 rock modes (47%), and as a major phase in 466 of those rocks (25%). Biotite lumps 6 species of micas [Hazen et al. (2022); Online Materials[1] Table OM3; see Online Materials[1] Table OM6 for a list of end-member compositions].

Phlogopite includes various Mg-dominant, light-colored trioctahedral micas [KMg2(Mg,Fe2+,Mn2+,Fe3+,Ti4+)(Si,Al,Fe3+)2Si2O10(OH,F)2], representing 6 IMA-CNMNC-approved species (see Online Materials[1] Table OM6 for a list of end-member compositions). It was reported in 176 rocks, primarily carbonatites and alkaline rocks, in which it is a major phase (>5 vol%) in 117 examples.

The dioctahedral aluminous mica, muscovite [KAl2(Si3Al)O10(OH)2], occurs in 91 rocks, in 22 of which it is a major phase. Muscovite is a common primary mineral in granitic rocks and their pegmatites. However, it can also be a secondary mineral in igneous rocks—occurrences that can usually be distinguished by their textures (Speer 1984; Roycroft 1991; Fleet 2003).

In addition, at least 11 species of lithium-bearing micas have been approved; we lump all of these into lepidolite, which is recorded in 42 complex pegmatites in our survey.

An important task for future research is a compilation of data on numerous igneous mica compositions and their distinctive lithologies, with the objective of defining mica natural kinds through cluster analysis. It is very likely that such widely occurring igneous minerals as biotite and muscovite should be split into multiple natural kinds.

In addition to the mica group minerals, petalite (LiAlSi4O10) is a relatively uncommon sheet silicate that we record in 19 complex granite pegmatites (Deer et al. 2009). It is almost always associated with beryl, lepidolite, and tourmaline and/or elbaite.

Tectosilicates

Silica Group: Quartz, tridymite, and cristobalite, all polymorphs of SiO2, occur as primary igneous phases (Deer et al. 2004). However, only quartz is frequently encountered as a major igneous mineral. We record quartz from 531 rocks, most of which are varieties of granite and their pegmatites.

Tridymite is a high-temperature form of SiO2 that must cool rapidly; therefore it occurs most frequently in rhyolites and other acidic extrusive rocks. We list 5 instances in our compilation.

The volcanic glass, obsidian, though not an approved mineral, is an important silica-rich phase in rhyolites. Obsidian, which typically contains > >70 wt% SiO2, and usually with ~90 wt% (SiO2 + Al2O3), may represent >99 vol% of rapidly cooled, silica-rich extrusives. We record 5 examples in Online Materials[1] Table OM3, though that number underestimates the frequency of obsidian occurrence because the modes of extrusive rocks are rarely reported.

Feldspar Group: Minerals of the feldspar group are the most common phases in igneous rocks, appearing in 1375 of 1850 diverse rocks in our study (74%), while forming an estimated 50 vol% of Earth’s crust (Deer et al. 2001; Rudnick and Gao 2003). The only major rock-forming feldspars occur with compositions close to two binary systems: alkali feldspars (NaAlSi3O8 to KAlSi3O8) and plagioclase feldspars (NaAlSi3O8 to CaAl2Si2O8). Each of these series has complexities related to petrographic nomenclature.

The IMA-CNMNC recognizes the two end-members of the plagioclase series—albite and anorthite—as valid species. Petrographers, on the other hand, have traditionally split plagioclase into six compositional types: albite, oligoclase, andesine, bytownite, labradorite, and anorthite (with compositional boundaries of 0 to 10, 10 to 30, 30 to 50, 50 to 70, 70 to 90, and 90 to 100 at% anorthite content, respectively). Complexities arise both because many igneous plagioclase crystals are zoned across two or three of these types, and because specimens often display fine-scaled exsolution, termed peristerite, Bøggild intergrowths, and Huttenlocher intergrowths (e.g., Deer et al. 2001). In our survey we adopt a compromise nomenclature, with three compositional ranges. Na-rich albite and oligoclase are lumped into albite (543 occurrences; Online Materials[1] Table OM3); intermediate andesine and bytownite are lumped into plagioclase (210 occurrences), and Ca-rich labradorite and anorthite are lumped into anorthite (245 occurrences).

The alkali feldspars present their own complexities in nomenclature. In the petrographic literature, sodium-rich varieties include albite and K-bearing “anorthoclase,” as well as “oligoclase,” all of which we record as albite. Intermediate alkali feldspar compositions commonly display exsolution, either perthite with Na-rich exsolution lamellae in a K-rich feldspar, or antiperthite with K-rich exsolution lamellae in a Na-rich feldspar. We classify all of these alkali feldspars with prominent exsolution textures as perthite, which occurs in 304 rocks of Online Materials[1] Table OM3. Feldspars near the potassium-rich end-member occur in several structure types, identified as sanidine, orthoclase, and microcline depending primarily on the ordered state of Al and Si (which is in turn a function of cooling history). In the petrographic literature they are often referred to collectively as “K-spar” without further identification. Sanidine is the high-temperature form with disordered Al-Si; it typically occurs in rapidly cooled magmas from near-surface environments and is reported from 82 extrusive rocks in our survey. Orthoclase (447 occurrences) and microcline (310 occurrences) incorporate ordered Al-Si arrangements, indicative of more slowly cooled environments. In some reports, petrographers are able to distinguish microcline based primarily on its distinctive pattern of fine-scale cross-hatched twinning. Other potassic feldspars are identified as orthoclase or simply K-spar. With acknowledged uncertainties, we identify sanidine and microcline when those minerals are explicitly named; otherwise we lump orthoclase and K-spar into orthoclase.

Feldspathoid Group: Feldspathoid group minerals, defined as alkali- or Cabearing phases that are chemically similar to feldspar but incorporating less silica, are major primary igneous phases in a range of silica-undersaturated, alkaline rocks (Deer et al. 2004). They appear in 659 of 1850 rocks (35%) tabulated in Online Materials[1] Table OM3, though that number likely exaggerates the volumetric importance of feldspathoids because of our inclusion of 795 modes of alkaline rocks and carbonatites from Woolley’s comprehensive review (Woolley 1987, 2001, 2019; Kogarko et al. 1995). Nevertheless, in many subsilicic rocks feldspathoid minerals, notably nepheline, take the place of feldspar as the volumetrically dominant rock-forming phase.

Nepheline [Na6(K,Cao)2(Al8Si8O32)] is the most frequently encountered feldspathoid in our tabulations with 567 occurrences. It is the dominant mineral in nepheline syenites, in which it is often associated with alkali feldspars (Deer et al. 2004). Nepheline also occurs with plagioclase in various mafic lithologies, for example nepheline gabbros.

Other common feldspathoids include leucite [K(Si2Al)O6], recorded in 103 rocks, at times as “pseudoleucite,” in which nepheline, K-feldspar, and other phases replace primary leucite. Sodalite [Na4Si3Al3O12Cl], listed in the modes of 158 alkaline rocks, and the structurally related hauyne [Na3(Ca,Na,K)(Si3Al3) O12(SO4)·(H2O)], found in 63 rocks, commonly occur in association with nepheline or leucite. Kalsilite [(K,Na)AlSiO4], the K-dominant isomorph of nepheline, is much less abundant, reported from only 6 alkaline rocks in our compilation.

The cancrinite group, with 22 species lumped into cancrinite [(Na,K,Cao)8 (Al6Si6)O24(CO3,SO4,Cl,OH)2·nH2O], presents an added complication. Cancrinite is a relatively common feldspathoid, appearing in the modes of 119 alkaline igneous rocks. However, because cancrinite paragenesis requires a fluid rich in carbonate or sulfate ions, it can form both as a late-stage primary igneous phase and as a secondary mineral, for example through alteration of nepheline by reaction with carbonate-rich fluids. We have excluded cancrinite from Online Materials[1] Table OM3 when it is explicitly described as a secondary phase. However, in many other instances the mode of cancrinite formation is unknown. In those instances, we record cancrinite as a primary phase.

Scapolite, with 3 lumped species [(Na,Ca)4(Al,Si)12O24(CO3,SO4,Cl)], usually occurs as a metamorphic or metasomatic phase. In Online Materials1 Table OM3 we list 12 occurrences that may represent formation during initial magma cooling of alkaline or mafic lithologies, though uncertainties remain.

Zeolite Group: The great majority of zeolite group framework aluminosilicates (Deer et al. 2004), as well as closely associated minerals such as prehnite and apophyllite (Deer et al. 2009), occur as secondary/hydrothermal phases deposited by post-magmatic fluid interaction/alteration of vesicular basalt and other igneous rocks. However, we list 3 kinds of zeolite that also may occur as late-stage primary phases in igneous rocks.

Analcime [NaAlSi2O6·H2O] is the most abundant of these phases, with 151 occurrences in our list (in 88 of which it is reported as a major phase). Analcime is most commonly associated with nepheline in a wide range of intrusive and extrusive lithologies, typically alkaline and silica-undersaturated rocks (Johannsen 1938; Deer et al. 2004).

Natrolite [(Na,Ca)2(Si3Al2)O10·2H2O] is reported as a primary zeolite in the modes of 17 rocks in Online Materials[1] Table OM3, notably as a late-stage major phase in some agpaitic pegmatites, in which it may form from auto-metasomatism during cooling. Natrolite usually occurs with nepheline and aegirine in alkaline rocks.

Pollucite [Cs(Si2Al)O6·nH2O] is a common late-stage (T < 500 °C) zeolite phase (London 1986, 2008; Teertstra and Černý 1995), though it also occurs as a secondary zeolite (Deer et al. 2004). We report 22 occurrences, exclusively from complex rare-element granite pegmatites (Online Materials[1] Table OM3), in which it is almost always associated with beryl, lepidolite, and one or more tourmaline group minerals. In spite of the scarcity of Cs, representing <5 ppm of atoms in Earth’s crust (Rudnick and Gao 2005), pollucite can be present as a major phase in the terminal crystallization of a pegmatite, with crystal pods in the Tanco pegmatite, Manitoba, Canada, exceeding 10 m in length (London 2008).

Silicate glass: Primary igneous glass with a mixture of rock-forming elements [e.g., (Si,Al,Ca,Mg,Fe,O)] is an important primary extrusive igneous phase, most notably as glass of basaltic composition but also in a range of intermediate to mafic extrusive rocks. We provisionally employ silicate glass as a catch-all term for such amorphous igneous phases that are less Si-rich than obsidian (i.e., <70 wt% SiO2), as recorded in 48 modes in Online Materials[1] Table OM3. Future research on the tabulation of igneous glass compositions and cluster analysis of their distributions will be required to determine if silicate glass and obsidian represent a continuum, or if there are multiple natural kinds of Si-bearing glass in extrusive igneous rocks.

Rarer Primary Igneous Minerals: More than 800 other minerals (in addition to the 115 kinds outlined above) occur rarely as trace phases in igneous 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 2022). Most of these scarce minerals, which are listed in Online Materials[1] Table OM2, were not recorded from any of the 1850 igneous rock modes in Online Materials[1] Table OM3.

However, a few of the less common minerals that are designated “2” (i.e., uncommon accessory phases) in Online Materials[1] Table OM2, were also recorded in one or more igneous rock modes. Among these minor minerals, listed alphabetically, are agrellite (1 occurrence), armalcolite (2), baotite (2), calzirtite (2), canasite (1), carbocernaite (2), cerianite (1), cerite (3), charoite (1), chrysoberyl (3), chrichtonite group (3), dalyite group (2), daqingshanite (1), elpidite (3), fersmanite (1), florencite (2), gadolinite group (1), gittinsite (2), hellandite (1), helvine group (2), hogbomite group (2), hollandite group (3), isokite (1), joaquinite group (1), karnasurite (1), labuntsovite (3), lanthanite group (1), lazurite (1), lievrite (2), lorenzenite (2), lovozerite group (1), lueshite (1), murmanite (2), narsarsukite (2), natroniobite (1), neotocite (1), neptunite group (2), pachnolite (1), rhabdophane group (1), rhönite (4), sahamalite (1), samarskite group (1), scheelite (2), steenstrupine (2), thorianite (3), tinaksite (1), uraninite (1), vlasovite (1), weberite (1), wolframite (1), wulfenite (3), and zirkelite (4).

Received: 2022-03-24
Accepted: 2022-10-12
Published Online: 2023-08-31
Published in Print: 2023-09-26

© 2023 by Mineralogical Society of America

Articles in the same Issue

  1. Fluorine-rich mafic lower crust in the southern Rocky Mountains: The role of pre-enrichment in generating fluorine-rich silicic magmas and porphyry Mo deposits
  2. Apatite in brachinites: Insights into thermal history and halogen evolution
  3. A high-pressure structural transition of norsethite-type BaFe(CO3)2: Comparison with BaMg(CO3)2 and BaMn(CO3)2
  4. An evolutionary system of mineralogy, Part VII: The evolution of the igneous minerals (>2500 Ma)
  5. Oriented secondary magnetite micro-inclusions in plagioclase from oceanic gabbro
  6. A multi-methodological study of the bastnäsite-synchysite polysomatic series: Tips and tricks of polysome identification and the origin of syntactic intergrowths
  7. Petrogenesis of Chang’E-5 mare basalts: Clues from the trace elements in plagioclase
  8. Experimental investigation of trace element partitioning between amphibole and alkali basaltic melt: Toward a more general partitioning model with implications for amphibole fractionation at deep crustal levels
  9. Grain-scale zircon Hf isotope heterogeneity inherited from sediment-metasomatized mantle: Geochemical and Nd-Hf-Pb-O isotopic constraints on Early Cretaceous intrusions in central Lhasa Terrane, Tibetan Plateau
  10. Mechanism and kinetics of the pseudomorphic replacement of anhydrite by calcium phosphate phases at hydrothermal conditions
  11. Vacancy infilling during the crystallization of Fe-deficient hematite: An in situ synchrotron X-ray diffraction study of non-classical crystal growth
  12. Simulated diagenesis of the iron-silica precipitates in banded iron formations
  13. Wave vector and field vector orientation dependence of Fe K pre-edge X-ray absorption features in clinopyroxenes
  14. Structure and compressibility of Fe-bearing Al-phase D
  15. Synthesis of boehmite-type GaOOH: A new polymorph of Ga oxyhydroxide and geochemical implications
  16. Scheelite U-Pb geochronology and trace element geochemistry fingerprint W mineralization in the giant Zhuxi W deposit, South China
  17. A rare sekaninaite occurrence in the Nenana Coal Basin, Alaska Range, Alaska
  18. Slyudyankaite, Na28Ca4(Si24Al24O96)(SO4)6(S6)1/3(CO2)·2H2O, a new sodalite-group mineral from the Malo-Bystrinskoe lazurite deposit, Baikal Lake area, Russia
  19. Ruizhongite, (Ag2□)Pb3Ge2S8, a thiogermanate mineral from the Wusihe Pb-Zn deposit, Sichuan Province, Southwest China
Downloaded on 15.9.2025 from https://www.degruyterbrill.com/document/doi/10.2138/am-2022-8539/html
Scroll to top button