Hostname: page-component-848d4c4894-75dct Total loading time: 0 Render date: 2024-05-06T07:22:09.527Z Has data issue: false hasContentIssue false

The Layos Granite, Hercynian Complex of Toledo (Spain): an example of parautochthonous restite-rich granite in a granulitic area

Published online by Cambridge University Press:  03 November 2011

L. Barbero
Affiliation:
L. Barbero and C. Villaseca, Dpto. Petrología y Geoquímica, Fac. C.C. Geológicas,Universidad Complutense, 28040 Madrid, Spain

Abstract

The Layos Granite forms elongated massifs within the Toledo Complex of central Spain. It is late-tectonic with respect to the F2 regional phase and simultaneous with the metamorphic peak of the region, which reached a maximum temperature of 800–850°C and pressures of 400–600 MPa. Field studies indicate that this intrusion belongs to the “regional migmatite terrane granite” type. This granite is typically interlayered with sill-like veins and elongated bodies of cordierite/garnet-bearing leucogranites. Enclaves are widespread and comprise restitic types (quartz lumps, biotite, cordierite and sillimanite-rich enclaves) and refractory metamorphic country-rocks including orthogneisses, amphibolites, quartzites, conglomerates and calc-silicate rocks.

These granites vary from quartz-rich tonalites to melamonzogranites and define a S-type trend on a QAP plot. Cordierite and biotite are the mafic phases of the rocks. The particularly high percentage of cordierite (10%–30%) varies inversely with the silica content. Sillimanite is a common accessory mineral, always included in cordierite, suggesting a restitic origin. The mineral chemistry of the Layos Granite is similar to that of the leucogranites and country-rock peraluminous granulites (kinzigites), indicating a close approach to equilibrium. The uniform composition of plagioclase (An25), the high albitic content of the K-feldspar, the continuous variation in the Fe/Mg ratios of the mafic minerals, and the high Ti content of the biotites (2.5–6.5%) suggest a genetic relationship.

Geochemically, the Layos Granite is strongly peraluminous. Normative corundum lies between 4% and 10% and varies inversely with increase in SiO2. The CaO content is typically low (<1.25%) and shows little variation; similarly the LILE show a limited range. On many variation diagrams, linear trends from peraluminous granulites to the Layos Granite and associated leucogranite can be observed. The chemical characteristics argue against an igneous fractionation or fusion mechanism for the diversification of the Layos Granite. A restite unmixing model between a granulitic pole (represented by the granulites of the Toledo Complex) and a minimum melt (leucogranites) could explain the main chemical variation of the Layos Granite. Melting of a pelitic protolith under anhydrous conditions (biotite dehydration melting) could lead to minimum-temperature melt compositions and a strongly peraluminous residuum.

For the most mafic granites (61–63% SiO2), it is estimated that the trapped restite component must have been around 65%. This high proportion of restite is close to the estimated rheological critical melt fraction, but field evidence suggests that this critical value has been exceeded. This high restite component implies high viscosity of the melt which, together with the anhydrous assemblage of the Layos Granite and the associated leucogranites, indicates H2O-undersaturated melting conditions. Under such conditions, the high viscosity magma (crystal-liquid mush) had a restricted movement capacity, leading to the development of parautochthonous plutonic bodies.

Type
Research Article
Copyright
Copyright © Royal Society of Edinburgh 1992

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Andonaegui, P. 1990. Geoquímica y geocronología de los granitoides del sur de Toledo. Tesis doctoral, Universidad Complutense de Madrid, Spain.Google Scholar
Arzi, A. A. 1978. Critical phenomena in the rheology of partially melted rocks. TECTONOPHYSICS 44, 173–89.CrossRefGoogle Scholar
Barbero, L. & Villaseca, C. 1988. Gabros coroníticos en el macizo cristalino de Toledo. GEOGACETA 5, 67–9.Google Scholar
Barbero, L., Villaseca, C. & Andonaegui, P. 1990. On the origin of the gabbro-tonalite-monzogranite association from Toledo area (Hercynian Iberian belt). SCHWEIZ MINERAL PETROGR MITT 7, 209–21.Google Scholar
Barbey, P., Bertrand, J. M., Angoua, S. & Dautei, D. 1989. Petrology and U/Pb chronology of the Telohat migmatites (Aleksod, Central Hoggar). CONTRIB MINERAL PETROL 101, 207–19.Google Scholar
Barbey, P., Macaudiere, J. & Uzenti, J. P. 1990. High-pressure dehydration melting of metapelites: evidence from migmatites of Yaoundí (Cameroon). J PETROL 31, 401–28.Google Scholar
Bohlen, J. R. & Liotta, J. J. (1986). A barometer for garnet amphibolites and garnet granulites. J PETROL 27, 1024–34.Google Scholar
Bowden, P., Batchelor, R. A., Chappell, B. W., Didier, J. & Lameyre, J. 1984. Petrological, geochemical and source criteria for the classification of granitic rocks: a discussion. PHYS EARTH PLANET INTERIORS 35, 111.CrossRefGoogle Scholar
Brandebourger, E. 1984. Les granitoides hercyniens tardifs de la Sierra de Guadarrama (Système Central, Espagne). Pétrographie et geochimie, These 3eme Cycle, CNRS, France.Google Scholar
Brändle, J. L. & Cerqueira, I. 1972. Determinación de elementos menores en rocas silicatadas por fluorescencia de rayos X. ESTUD GEOL 28, 445–51.Google Scholar
Casillas, R. 1989. Las asociaciones plutónicas tardihercínicas del sector occidental de la Sierra de Guadarrama—Sistema Central Español, (Las Navas del Marqués–San Martín de Valdeiglesias). Petrología, geoquimíca, génesis y evolución. Tesis Doctoral, Universidad Complutense de Madrid, Spain.Google Scholar
Casquet, C., Fuster, J. M., Casado, J. M.González, Peinado, M. & Villaseca, C. 1988. Extensional tectonics and granite emplacement in the Spanish Central System. A discussion. EUROPEAN SCI FOUND (SPEC VOL), PROC 5th WORK, 6576.Google Scholar
Chappell, B. W., White, A. J. R. & Wyborn, D. 1987. The importance of residual source material (restite) in granite petrogenesis. J PETROL 28, 1111–38.CrossRefGoogle Scholar
Clarke, R. G. & Lyons, J. B. 1986. Petrogenesis of the Kingsman intrusive suite: peraluminous granitoids of Western New Hampshire. J PETROL 27, 1365–93.CrossRefGoogle Scholar
Debon, F. & Fort, P.Le 1983. A chemical-mineralogical classification of common plutonic rock and associations. TRANS R SOC EDINBURGH: EARTH SCI 73, 135–49.Google Scholar
Moro, A.Del 1987. Sistematica Rb/Sr di alcune magmatiti tardo-erciniche dell'area italiana. RIC SCI EDUC PERM 52, 107–32.Google Scholar
Dymek, F. R. 1983. Titanium, aluminium and interlayered cation substitution in biotite from high-grade gneisses, west Greenland. AM MINERAL 68, 880–99.Google Scholar
Ebadi, A. & Johannes, W. 1991. Beginning of melting and composition of first melts in the system Qz–Ab–Or–H2O–CO2. CONTRIB MINERAL PETROL 106, 286–95.Google Scholar
Ferry, J. M. & Spear, F. S. 1978. Experimental calibration of the partitioning of Fe and Mg between garnet and biotite. CONTRIB MINERAL PETROL 66, 113–7.Google Scholar
Flood, R. H. & Shaw, S. E. 1975. A cordierite-bearing granite suite from the New England batholith. N.S.W. Australia. CONTRIB MINERAL PETROL 52, 157–64.Google Scholar
Frost, B. R. & Chacko, T. 1989. The granulite uncertainty principle: limitations on thermobarometry in granulites. J GEOL 97, 435–50.Google Scholar
Ganguly, J. & Saxena, S. K. 1984. Mixing properties of aluminosilicate garnets: constrains for natural and experimental data and application to geothermo-barometry. AM MINERAL 69, 8897.Google Scholar
Ghent, E. D. & Stout, M. Z. 1984. TiO2 activity in metamorphosed pelitic and basic rocks: principles and applications to metamorphism in southeastern Canadian Cordillera. CONTRIB MINERAL PETROL 86, 248–55.CrossRefGoogle Scholar
Grapes, R. H. 1985. Melting and thermal reconstruction of pelitic xenolith, Wehr volcano, east Eifel, west Germany. J PETROL 27, 343–96.Google Scholar
Green, D. & Ringwood, A. 1967. An experimental investigation of the gabbro to eclogite transformation and its petrological applications. GEOCHIM COSMOCHIM ACTA 31, 767833.CrossRefGoogle Scholar
Gromet, L. P., Dymek, R. F., Haskin, L. A. & Korotev, R. L. 1984. The “North American Shale composite”: its compilation, major and trace element characteristics. GEOCHIM COSMOCHIM ACTA 48, 2469–82.Google Scholar
Hodges, K. V. & Spear, F. S. 1982. Geothermometry, geobarometry and the A12O3 triple point at Mt Moosilauke, New Hampshire. AM MINERAL 67, 1118–34.Google Scholar
Holdaway, M. J. & Lee, S. N. 1977. Fe-Mg cordierite stability in high-grade pelitic rocks based on experimental, theoretical and natural observations. CONTRIB MINERAL PETROL 63, 175–98.CrossRefGoogle Scholar
Ibarguchi, J. I. G. & Martínez, F. J. 1982. Petrology of garnet-cordierite-sillimanite gneisses from the E1 Tormes thermal Dome, Iberian Hercynian foldbelt (W Spain). CONTRIB MINERAL PETROL 80, 1424.Google Scholar
Jones, K. A. & Brown, M. 1990. High-temperature ‘clockwise’ P-T paths and melting in the development of regional migmatites: an example from southern Brittany, France. J MET GEOL 8, 551–78.Google Scholar
Julivert, M., Fontboté, J. M., Ribeiro, A. & Conde, L. E. 1974. Memoria explicativa del mapa tectónico de la Península Ibérica y Baleares. Escala 1:100,000. I.G.M.E., Madrid, Spain.Google Scholar
Lavrenteva, E. V. & Perchuck, L. L. 1981. Cordierite-garnet thermometer. A collection of theses. ACAD SCI USSR 259, 607700.Google Scholar
Breton, N.Le & Thompson, A. B. 1988. Fluid-absent (dehydration) melting of biotite in metapelites in the early stages of crustal anatexis. CONTRIB MINERAL PETROL 99, 226–37.Google Scholar
Escorza, C.Martín & Martínez, J.López 1978. Análisis mesoestructural en la Unidad Migmatítica de Toledo. ESTUD GEOL 34, 3443.Google Scholar
Masuda, A., Nakamura, N. & Tanaka, T. 1973. Fine structures of mutual normalized rare earth patterns of chondrites. GEOCHIM COSMOCHIM ACTA 37, 239–48.Google Scholar
McRae, N. D. & Nesbitt, H. W. 1980. Partial melting of common metasedimentary rocks: a mass balance approach. CONTRIB MINERAL PETROL 75, 21–6.Google Scholar
Miller, C. F. 1985. Are strongly peraluminous magmas derived from pelitic sedimentary sources? J GEOL 93, 673–89.Google Scholar
Miller, C. F., Watson, E. B. & Harrison, T. M. 1988. Perspectives on the source, segregation and transport of granitoid magmas. TRANS R SOC EDINBURGH EARTH SCI 79, 135–56.Google Scholar
Minster, F. J. & Allègre, C. J. 1977. Systematic use of trace elements in igneous processes. Part I: Fractional crystallization processes in volcanic suites. CONTRIB MINERAL PETROL 60, 5775.Google Scholar
Moller, P. & Muecke, G. K. 1984. Significance of europium anomalies in silicate melts and crystal melt equilibria: a re-evaluation. CONTRIB MINERAL PETROL 87, 242–50.Google Scholar
Patiño, A. & Johnston, A. 1991. Phase equilibria and melt productivity in the pelitic system: implications for the origin of peraluminous granitoids and aluminous granulites. CONTRIB MINERAL PETROL 107, 202–18.Google Scholar
Rottura, A., Bargossi, G. M., Caironi, V., D'Amico, C. & Maccarrone, E. 1989. Petrology and geochemistry of late-Hercynian granites from the Western Central System of the Iberian Massif. EUR J MINER 1, 667–83.Google Scholar
Rottura, A., Bargossi, G. M., Caironi, V., Moro, A.Del, Maccarrone, E., Macera, P., Paglionico, A.Petrini, R.Piccarreta, G. & Poli, G. 1990. Petrogenesis of contrasting Hercynian granitoids from the Calabrian Arc, southern Italy. LITHOS 24, 97119.CrossRefGoogle Scholar
Sawyer, E. W. 1987. The role of partial melting and fractional crystallisation in determining discordant migmatite leucosome compositions. J PETROL 28, 445–73.CrossRefGoogle Scholar
Sengupta, P., Dasgupta, S. K., Bhattacharya, P. K. & Mukherjee, M. 1990. An orthopyroxene-biotite geothermometer and its application in crustal granulites and mantle derived rocks. J MET GEOL 8, 191–8.Google Scholar
Speer, J. A. 1981. Petrology of cordierite- and almandine-bearing granitoid plutons of the southern Appalachian piedmont U.S.A. CAN MINERAL 19, 3546.Google Scholar
Streckeisen, A. 1976. To each plutonic rock its proper name. EARTH SCI REV 12, 113.Google Scholar
Thompson, R. N., Morrison, M. A., Hendry, G. L. & Parry, S. J. 1984. An assessment on the relative roles of crust and mantle magma genesis: an elemental approach. PHILOS TRANS R SOC LONDON A310, 549–90.Google Scholar
Van der Molen, I. & Paterson, M. S. 1979. Experimental deformation of partially-melted granite. CONTRIB MINERAL PETROL 70, 299318.Google Scholar
Vialette, Y., Casquet, C., Fúster, J. M., Ibarrola, E., Navidad, M., Peinado, M. & Villaseca, C. 1987. Geochronological study of orthogneisses from the Sierra de Guadarrama (Spanish Central System). N JARB MINER MH JG H10, 465–79.Google Scholar
Vielzeuf, D. 1983. The spinel and quartz associations in high grade xenoliths from Tallante (SE Spain) and their potential use in geothermometry and barometry. CONTRIB MINERAL PETROL 82, 301–11.CrossRefGoogle Scholar
Vielzeuf, D., Clemens, J. D., Pin, C. & Moinet, E. 1990. Granites, granulites and crustal evolution. In Vielzeuf, D. & Vidal, Ph. (eds) Granulites and crustal evolution. NATO ASI SCI SER C 311, 5985.Google Scholar
Vielzeuf, D. & Holloway, J. R. 1988. Experimental determination of the fluid-absent melting relations in the pelitic system. Consequences for crustal differentiation. CONTRIB MINERAL PETROL 98, 257–76.Google Scholar
Villaseca, C. 1983. Evolución metamórfica del sector centroseptentrional de la Sierra de Guadarrama. Tesis Doctoral 216/84, Universidad Complutense de Madrid, Spain.Google Scholar
Villaseca, C. & Barbero, L. (in press). Los granates de rocas metapelíticas de la región central española: implicaciones en el origen de granates en granitoides. GEOGACETA (in press).Google Scholar
Vry, J. K., Brown, E. B. & Valley, J. W. 1990. Cordierite volatile content and the role of CO2 in high-grade metamorphism. AM MINERAL 75, 7188.Google Scholar
Watson, E. B. 1987. The role of accessory minerals in granitoid geochemistry. 1ST HUTTON SYMP ABSTR, 1921.Google Scholar
Wells, P. R. A. 1977. Pyroxene thermometry in simple and complex systems. CONTRIB MINERAL PETROL 62, 129–39.Google Scholar
Whalen, J. B. & Chappell, B. W. 1988. Opaque mineralogy and mafic mineral chemistry of I- and S-type granites of the Lachlan fold belt, southern Australia. AM MINERAL 73, 281–96.Google Scholar
White, A. J. R. & Chappell, B. W. 1987. Some supracrustal (S-type) granites of the Lachlan Fold Belt. TRANS R SOC EDINBURGH EARTH SCI 79, 169–81.Google Scholar
Wickham, S. M. 1987a. Crustal anatexis and granite petrogenesis during low pressure regional metamorphism in the Trois Seigneurs massif, Pyrenees, France. J PETROL 28, 127–69.Google Scholar
Wickham, S. M. 1987b. The segregation and emplacement of granitic magmas. J GEOL SOC LONDON 144, 281–97.CrossRefGoogle Scholar
Wildberg, H. G. H., Bischoff, L. & Baumann, A. 1989. U-Pb ages of zircons from meta-igneous and metasedimentary rocks of the Sierra de Guadarrama: implications for the Central Iberian crustal evolution. CONTRIB MINERAL PETROL 103, 253–62.CrossRefGoogle Scholar