Hostname: page-component-8448b6f56d-cfpbc Total loading time: 0 Render date: 2024-04-25T05:20:36.171Z Has data issue: false hasContentIssue false

Inorganic geochemistry of the type Caradoc series (Sandbian to middle Katian, Upper Ordovician), Onny valley, Shropshire, UK

Published online by Cambridge University Press:  07 February 2013

R. HANNIGAN
Affiliation:
Environmental Earth and Ocean Sciences, University of Massachusetts at Boston, 100 Morrissey Blvd, Boston, MA 02125, USA
M. E. BROOKFIELD*
Affiliation:
Environmental Earth and Ocean Sciences, University of Massachusetts at Boston, 100 Morrissey Blvd, Boston, MA 02125, USA
*
*Author for correspondence: mbrookfi@hotmail.com

Abstract

The geochemistry and petrology of the type section of the Caradoc Series in the Onny valley indicate that it was deposited on a marginal basin continental shelf similar to the western side of the present Sea of Japan. The lower beds form a transgressive–regressive sequence in which the rocks become less mature upwards. All the coarser sediments above the basal quartzites and conglomerates are greywackes in which the apparent muddy and ferrous matrix is due to the breakdown of unstable minerals and particles. Higher values of Na2O and Na/K ratios are found in the coarser shallow-water sandstones of the Horderley Sandstone Formation and decrease markedly in the succeeding beds, accompanied by an increase in K2O. Higher values of carbonate-corrected (and hence related) other major and minor elements like SiO2, CaO, P2O5, MnO and most trace elements correlate with the transgressive systems and maximum flooding surfaces of the three sequences recognized where they are related to condensation at those horizons. Chemical Indices of Alteration (CIA) suggest that the Horderley Sandstone Formation underwent greater predepositional physical weathering than lower and higher beds, which is compatible with the petrography, and were deposited during a cool phase within overall warm Sandbian–Katian times. Trace element ratios suggest an oxic to suboxic depositional environment.

Type
Original Articles
Copyright
Copyright © Cambridge University Press 2013 

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

Abanda, P. A. & Hannigan, R. 2006. Diagenetic controls of trace element partitioning in shales. Chemical Geology 230, 4259.Google Scholar
Algeo, T. J., Hannigan, R., Rowe, H., Brookfield, M. E., Baud, A., Krystyn, L. & Ellwood, B. B. 2007. Sequencing events across the Permian-Triassic boundary, Guryul Ravine (Kashmir, India). Palaeogeography, Palaeoclimatology, Palaeoecology 252, 328–46.Google Scholar
Algeo, T. J. & Lyons, T. W. 2006. Mo-total organic carbon covariation in modern anoxic marine environments: implications for analysis of paleoredox and paleohydrographic conditions. Paleoceanography 21, PA1016, doi:10.1029/2004PA001112, 23 pp.Google Scholar
Anand, R. R. & Gilkes, R. J. 1984. Weathering of ilmenite in a lateritic pallid zone. Clays and Clay Minerals 32, 363–74.Google Scholar
Armstrong, H. A., Baldini, J., Challands, T. J., Gröcke, D. R & Owen, A. W. 2009. Response of the Inter-tropical Convergence Zone to southern hemisphere cooling during the Upper Ordovician glaciation. Palaeogeography, Palaeoclimatology, Palaeoecology 284, 227–36.Google Scholar
Bancroft, B. B. 1933. Correlation Tables of the Costonian–Onnian in England and Wales. Gloucester (privately printed): Blakeney, 4 pp, 3 tables.Google Scholar
Bancroft, B. B. 1945. The brachiopod zonal indices of the stages Costonian to Onnian in Britain. Journal of Paleontology 19, 181252.Google Scholar
Bayon, G., German, C. R., Boella, R. M., Milton, J. A., Taylor, R. N. & Nesbitt, R. W. 2002. An improved method for extracting marine sediment fractions and its application to Sr and Nd isotopic analysis. Chemical Geology 187, 179–99.Google Scholar
Bergström, S. M., Chen, X., Gutiérrez-Marco, J. C. & Dronov, A. 2006. The new chronostratigraphic classification of the Ordovician System and its relations to major regional series and stages and to δ13C chemostratigraphy. Lethaia 41, 97107.Google Scholar
Bettley, R. M., Fortey, R. A. & Siveter, D. J. 2001. High-resolution correlation of Anglo-Welsh Middle to Upper Ordovician sequences and its relevance to international correlation. Journal of the Geological Society, London 158, 937–52.Google Scholar
Bevins, R. E. & Robinson, D. 1988. Low grade metamorphism of the Welsh Basin Lower Palaeozoic succession: an example of diastathermal metamorphism? Journal of the Geological Society, London 146, 885–90.Google Scholar
Bhatia, M. R. 1983. Plate tectonics and geochemical composition of sandstone. Journal of Geology 91, 611–27.Google Scholar
Bhatia, M. R. & Crook, K. A. W. 1986. Trace element characteristics of graywackes and tectonic setting discrimination of sedimentary basins. Contributions to Mineralogy and Petrology 92, 181–93.Google Scholar
Bowdler-Hicks, A., Ingham, J. K. & Owen, A. W. 2002. The taxonomy and stratigraphical significance of the Anglo-Welsh Cryptolithinae (Trinucleidae, Trilobita). Palaeontology 45, 1075–105.Google Scholar
Brenchley, P. J. 1969 a. The relationship between Caradocian volcanicity and sedimentation in North Wales. In The Pre-Cambrian and Lower Palaeozoic Rocks of Wales (ed. Wood, A.), pp. 181202. Cardiff: University of Wales Press.Google Scholar
Brenchley, P. J. 1969 b. Origin of matrix in Ordovician greywackes, Berwyn Hills, North Wales. Journal of Sedimentary Petrology 39, 1297–301.Google Scholar
Brenchley, P. J. & Newall, G. 1982. Storm-influenced inner-shelf sand lobes in the Caradoc (Ordovician) of Shropshire, England. Journal of Sedimentary Petrology 52, 1257–69.Google Scholar
Brenchley, P. J. & Pickerill, R. 1980. Shallow subtidal sediments of Soudleyan (Caradoc) age in the Berwyn Hills, North Wales and their palaoegeographic context. Proceedings of the Geologists’ Association 91, 177–94.Google Scholar
Brenchley, P. J., Rushton, A. W. A., Howells, A. & Cave, R. 2006. Cambrian and Ordovician: the early Palaeozoic tectonostratigraphic evolution of the Welsh Basin, Midland and Monian Terranes of Eastern Avalonia. In The Geology of England and Wales (eds Brenchley, P. J. & Rawson, P. F.), pp. 2574. London: The Geological Society.Google Scholar
Brewer, J. A., Allmendinger, R. W., Brown, L. D., Oliver, J. E. & Kaufman, S. 1982. COCORP profiling across the Rocky Mountain Front in southern Wyoming, Part 1: Laramide structure. Geological Society of America Bulletin 93, 1242–52.Google Scholar
Bridge, D. McC., Carney, J. N., Lawley, R. S. & Rushton, A. W. A. 1998. Geology of the Country around Coventry and Nuneaton. British Geological Survey Memoir for 1:50,000 Geological Sheet 169. London: The Stationery Office, 185 pp.Google Scholar
Brookfield, M. E., Shellnutt, J. G., Qi, L., Hannigan, R., Bhat, G. M. & Wignall, P. B. 2010. Platinum element group variations at the Permo-Triassic boundary in Kashmir and British Columbia and their significance. Chemical Geology 272, 1219.Google Scholar
Bullock, P., Fedoroff, N., Jongerius, A., Stoops, G., Tursina, T. & Bable, U. 1985. Handbook for Soil Thin Section Description. Wolverhampton: Waine Research Publications, 152 pp.Google Scholar
Cai, G., Guo, F., Liu, X., Sui, S., Li, C. & Zhao, L. 2008. Geochemistry of Neogene sedimentary rocks from the Jiyan basin, North China Block: the roles of grain size and clay minerals. Geochemical Journal 42, 381402.Google Scholar
Calvert, S. E. & Pedersen, T. 1993. Geochemistry of recent oxic and anoxic marine sediments: implications for the geological records. Marine Geology 11, 6788.Google Scholar
Calvert, S. E. & Pederson, T. F. 1996. Sedimentary geochemistry of manganese: implications for the environment of formation of manganiferous black shales. Economic Geology 91, 3650.Google Scholar
Calvert, S. E. & Pedersen, T. F. 2007. Element proxies for palaeoclimatic and palaeoceanographic variability. Developments in Marine Geology 1, 567644.Google Scholar
Carrano, C. J., Schellenberg, S., Amin, S.A., Green, D. H. & Küpper, F. C. 2009. Boron and marine life: a new look at an enigmatic bioelement. Marine Biotechnology 11, 431–40.Google Scholar
Castillo, S., Moreno, T., Querol, X., Alastuey, A., Cuevas, E., Herrmann, L., Mounkaila, M. & Gibbons, W. 2008. Trace element variation in size-fractionated African desert dusts. Journal of Arid Environments 72, 1034–45.Google Scholar
Catuneanu, O. 2006. Principles of Sequence Stratigraphy. Amsterdam: Elsevier, 375 pp.Google Scholar
Cauwet, G. 1987. Influence of sedimentological features on the distribution of trace metals in marine sediments. Marine Chemistry 22, 221–34.Google Scholar
Cave, R., Dean, W. T. & Hains, B. A. 1988. Age of the Ordovician andesitic conglomerate of Castle Hill, Montgomery, Powys, Wales. Geological Journal 23, 205–10.Google Scholar
Cha, H.-J., Choi, M. S., Lee, C.-B. & Shin, D.-H. 2007. Geochemistry of surface sediments in the southwestern East/Japan Sea. Journal of Asian Earth Sciences 29, 685–97.Google Scholar
Challands, T. J., Armstrong, H. A., Maloney, D. P., Davies, J. R., Wilson, D. & Owen, A. W. 2009. Organic-carbon deposition and coastal upwelling at mid-latitude during the Upper Ordovician (Late Katian): a case study from the Welsh basin, UK. Palaeogeography, Palaeoclimatology, Palaeoecology 273, 395410.Google Scholar
Chen, X., Rong, J., Fan, J., Zhan, R., Mitchell, C. E., Harper, D. A. T., Melchin, M. J., Peng, P., Finney, S. C. & Wang, X. 2006. The global boundary stratotype section and point (GSSP) for the base of the Hirnantian Stage (the uppermost of the Ordovician System). Episodes 29, 183–96.Google Scholar
Christie-Blick, N. & Driscoll, N. W. 1995. Sequence stratigraphy. Annual Review of Earth and Planetary Sciences 95, 451–78.Google Scholar
Clifton, H. E. 2003. Supply, segregation, successions, and significance of shallow marine conglomeratic deposits. Bulletin of Canadian Petroleum Geology 51, 370–88.Google Scholar
Coelho, M. R. & Vidal-Torrado, P. 2003. Caracterização e Gênese de Perfis Plínticos Desenvolvidos de Arenito do Grupo Baura, II Mineralogia. Revista Brasileira de Ciência do Solo 27, 495507.Google Scholar
Compston, W. 2000. Interpretation of SHRIMP and isotope dilution zircon ages for the Palaeozoic time-scale: II. Silurian to Devonian. Mineralogical Magazine 64, 1127–46.Google Scholar
Compston, W. & Williams, I. S. 1992. Ion probe ages for the British and Silurian stratotypes. In Global Perspectives on Ordovician Geology (eds Webby, B. D. & Laurie, J. R.), pp. 5967. Rotterdam: Balkema.Google Scholar
Cox, R. & Lowe, D. R. 1996. Quantification of the effects of secondary matrix on the analysis of sandstone compositions, and a petrographic-chemical technique for retrieving original framework grain modes of altered sandstones. Journal of Sedimentary Research 66, 548–58.Google Scholar
Crusius, J., Calvert, S., Pedersen, T. & Sages, D. 1996. Rhenium and molybdenum enrichments in sediments as indicators of oxic, suboxic and sulfidic conditions of deposition. Earth and Planetary Science Letters 145, 6578.Google Scholar
Dean, W. T. 1958. The faunal succession in the Caradoc Series of south Shropshire. Bulletin of the British Museum of Natural History (Geology) 3, 191231.Google Scholar
Dean, W. T. 1960. The Ordovician rocks of the Chatwall district, Shropshire. Geological Magazine 97, 163–71.Google Scholar
Dean, W. T. 1964. The geology of the Ordovician and adjacent strata in the southern Caradoc district of Shropshire. Bulletin of the British Museum of Natural History (Geology) 9, 259–96.Google Scholar
Dennison, J. M. 1972. Statistical meaning in geologic field work. Geological Society of America Special Paper 146, 2538.Google Scholar
Deutsch, C., Sarmiento, J. L., Sigman, D. M., Gruber, N. & Dunne, J. P. 2007. Spatial coupling of nitrogen inputs and losses in the ocean. Nature 445, 163–7.Google Scholar
Dewey, J. & Rosenbaum, M. 2008. Future avenues of research in the Welsh Borderland, with particular reference to terrane tectonics. Proceedings of the Shropshire Geological Society 13, 104–13.Google Scholar
Dickinson, W. R. 1985. Interpreting provenance relations from detrital modes of sandstone. In Provenance of Arenites (ed. Zuffa, G. G.), pp. 333–61. Dordrecht: Reidal Publishing Company.Google Scholar
Dickinson, W. R. & Suczek, C. A. 1979. Plate tectonics and sandstone compositions. American Association of Petroleum Geologists Bulletin 63, 2164–82.Google Scholar
Dinelli, E., Tateo, F. & Summa, V. 2007. Geochemical and mineralogical proxies for grain size in mudstones and siltstones from the Pleistocene and Holocene of the Po River alluvial plain, Italy. Geological Society of America Special Paper 420, 2536.Google Scholar
Dodd, J. R. 1967. Magnesium and strontium in calcareous skeletons: a review. Journal of Paleontology 41, 1313–29.Google Scholar
Dominik, J. & Stanley, D. J. 1993. Boron, beryllium and sulfur in Holocene sediments and peats of the Nile delta. Egypt: their use as indicators of salinity and climate. Chemical Geology 104, 203–16.Google Scholar
Dymond, J., Suess, E. & Lyle, M. 1992, Barium in deep-sea sediment: a geochemical proxy for paleoproductivity. Paleoceanography 7, 163–81.Google Scholar
Eynatten, H. von, Barceló-Vidal, C. & Pawlowsky-Glahn, V. 2003. Composition and discrimination of sandstones: a statistical evaluation of different analytical methods. Journal of Sedimentary Research 73, 4757.Google Scholar
Fedo, C. M., Nesbitt, H. W. & Young, G. M. 1995. Unraveling the effects of potassium metasomatism in sedimentary rocks and paleosols, with implications for paleoweathering conditions and provenance. Geology 23, 921–4.Google Scholar
Fleet, W. F. 1925. The chief heavy detrital minerals in the rocks of the English Midlands. Geological Magazine 62, 98128.Google Scholar
Floyd, P. A., Shail, R., Leveridge, B. E. & Franke, W. 1991. Geochemistry and provenance of Rhenohercynian synorogenic sandstones: implications for tectonic environment discrimination. In Developments in Sedimentary Provenance Studies (eds Morton, A. C., Todd, S. P. & Haughton, P. D. W.), pp. 173–88. Geological Society of London, Special Publication no. 57.Google Scholar
Flügel, E. 1982. Microfacies Analysis of Limestones. Berlin: Springer Verlag, 633 pp.Google Scholar
Fortey, R. A., Harper, D. A. T., Ingham, J. K., Owen, A. W., Parkes, M. A., Rushton, A. W. A. & Woodcock, N. H. 2000. A Revised Correlation of Ordovician Rocks in the British Isles. Geological Society of London Special Report no. 24, 93 pp.Google Scholar
Fortey, R. A., Harper, D. A. T., Ingham, J. K., Owen, A. W. & Rushton, A. W. A. 1995. A revision of Ordovician series and stages from the historical type area. Geological Magazine 132, 1530.Google Scholar
Fowler, S. W. 1977. Trace elements in zooplankton particulate products. Nature 269, 51–3.Google Scholar
Furst, M. J. 1981. Boron in siliceous materials as a paleosalinity indicator. Geochimica et Cosmochimica Acta 45, 113.Google Scholar
Gallego-Torres, D., Martinez-Ruis, F., De Lange, G. J., Jiminez-Espejo, F. J. & Ortega-Huertas, M. 2010. Trace-elemental derived paleoceanographic and paleoclimatic conditions for Pleistocene Mediterranean sapropels. Palaeogeography, Palaeoclimatology, Palaeoecology 293, 7689.Google Scholar
Gao, S., Luo, T.-C., Zhang, B.-R., Zhang, H.-F., Han, Y.-W., Zhao, Z.-D. & Hu, Y.-K. 1998. Chemical composition of the continental crust as revealed by studies in East China. Geochimica et Cosmochimica Acta 62, 1959–75.Google Scholar
Gibbons, W. & Young, T. P. 1999. Mid-Caradoc magmatism in central Llyn, rhyolite petrogenesis, and the evolution of the Snowdonia volcanic corridor in NW Wales. Journal of the Geological Society, London 156, 301–16.Google Scholar
Govin, A., Holzwarth, U., Heslop, D., Keeling, L. F., Zabel, M., Mulitza, S., Collins, J. A. & Chiessi, C. M. 2012. Distribution of major elements in Atlantic surface sediments (36°N–49°S): imprint of terrigenous input and continental weathering. Geochemistry, Geophysics, Geosystems 13, doi: 10.1029/2011GC003785, 23 pp.Google Scholar
Govindaraju, K. 1994. 1994 compilation of working values and sample descriptions for 383 geostandards. Geostandards Newsletter 18 (Supplement S1), 1158.Google Scholar
Greig, D. C., Wright, J. E., Hains, B. A. & Mitchell, G. H. 1968. Geology of the Country around Church Stretton, Craven Arms, Wenlock Edge and Brown Clee. Memoir of the Geological Survey of Great Britain. London: Her Majesty's Stationery Office, 379 pp.Google Scholar
Haase, K. M., Stroncik, N. A., Hékinian, R. & Stoffers, P. 2005. Nd-depleted andesites from the Pacific-Antarctic rise as analogs for early continental crust. Geology 33, 921–4.Google Scholar
Hannigan, R. E. & Basu, A. R. 1998. Late diagenetic trace element remobilization in organic-rich black shales of the Taconic foreland basin of Quebec, Ontario and New York. In Sahel and Mudstone: volume II, Petrography, Petrophysics, Geochemistry, and Economic Geology (eds Schieber, J. W. Z. & Sethi, P. S.), pp. 209–33. Stuttgart: E. Schweizerbart'sche Verlagsbuchhandlung.Google Scholar
Hannigan, R. E & Sholkovitz, E. R. 2001. The development of middle rare earth element enrichments in freshwaters: weathering of phosphate minerals. Chemical Geology 175, 495508.Google Scholar
Harper, D. A. T. 1978. The occurrence of the Ordovician brachiopod Heterorthis alternata (J. de C. Sowerby) in the topmost Onnian of the type Caradoc area. Geological Magazine 115, 301–4.CrossRefGoogle Scholar
Hassler, D. R., Peucker-Ehrenbrink, B. & Ravizza, G. E. 2000. Rapid determination of Os isotopic composition by sparging OsO4 into a magnetic-sector ICP-MS. Chemical Geology 166, 114.Google Scholar
Heiken, G. & Wohletz, K. 1985. Volcanic Ash. Berkeley: University of California Press, 246 pp.Google Scholar
Henderson, G. M. 2002. New oceanic proxies for paleoclimate. Earth and Planetary Science Letters 203, 113.Google Scholar
Hobday, D. K. & Reading, H. G. 1972. Fair weather versus storm processes in shallow marine sand bar sequences in the late Precambrian of Finnmark, north Norway. Journal of Sedimentary Research 42, 318–24.Google Scholar
Hoppie, B. & Garrison, R. E. 2002. The Cuyama strike-slip basin, California, U.S.A.: an exemplar of contrasting syntectonic and post-tectonic strata. Journal of Sedimentary Research 72, 268–87.Google Scholar
Howells, M. F., Reedman, A. J. & Campbell, S. D. G. 1991. Ordovician (Caradoc) Marginal Basin Volcanism in Snowdonia (North-West Wales). London: Her Majesty's Stationery Office, British Geological Survey, 191 ppGoogle Scholar
Hughes, R. A. 1989. Llandeilo and Caradoc Graptolites of the Builth and Shelve Inliers. London: Monograph of the Palaeontological Society, 89 pp., 5 plates.Google Scholar
Hurst, J. M. 1979 a. The environment of deposition of the Caradoc Alternata Limestone and contiguous deposits. Geological Journal 14, 1540.Google Scholar
Hurst, J. M. 1979 b. Evolution, succession and replacement in the type upper Caradoc (Ordovician) benthic faunas of England. Palaeogeography, Palaeoclimatology, Palaeoecology 27, 189246.Google Scholar
Ingersoll, R. V., Kretchemer, A. G. & Valles, P. K. 1993. The effect of sampling scale on actualistic sandstone petrofacies. Sedimentology 40, 937–53.Google Scholar
Jarvis, I., Murphy, A. M. & Gale, A. S. 2001. Geochemistry of pelagic and hemipelagic carbonates: criteria for identifying systems tracts and sea-level changes. Journal of the Geological Society, London 158, 685–96.Google Scholar
Johnson, H. D. 1977. Shallow marine sand bar sequences: an example from the late Precambrian of North Norway. Sedimentology 24, 245–70.Google Scholar
Johnsson, M. J. 1993. The system controlling the composition of clastic sediments. Geological Society of America Special Paper 285, 119.Google Scholar
Jones, B. & Manning, D. A. C. 1994. Comparison of geochemical indices used for the interpretation of palaeoredox conditions in ancient mudstones. Chemical Geology 111, 111–29.Google Scholar
Knoll, M. A. & James, W. C. 1987. Effect of the advent and diversification of vascular land plants on mineral weathering through geological time. Geology 15, 1099–102.Google Scholar
Kokelaar, P. 1988. Tectonic controls of Ordovician arc and marginal basin volcanism in Wales. Journal of the Geological Society, London 145, 759–75.Google Scholar
Kumar, N., Anderson, R. F., Mortlock, R. A., Froehlich, P. N., Kubik, P., Dittrich-Hannen, B. & Suter, M. 1995. Increased biological productivity and export production in the glacial Southern Ocean. Nature 378, 675–80.Google Scholar
Landing, W. M. & Bruland, K. W. 1987. The contrasting biogeochemistry of iron and manganese in the Pacific Ocean. Geochimica et Cosmochimica Acta 51, 2943.Google Scholar
LaPorte, D. F., Holmden, C., Patterson, W. P., Loxton, J. D., Melchin, M. J., Mitchell, C. E., Finney, S. C. & Sheets, H. D. 2009. Local and global perspectives on carbon and nitrogen cycling during the Hirnantian glaciation. Palaeogeography, Palaeoclimatology, Palaeoecology 276, 182–95.Google Scholar
Leat, P. T. & Thorpe, R. S. 1989. Snowdon basalts and the cessation of Caledonian subduction by the Longvillian. Journal of the Geological Society, London 146, 965–70.Google Scholar
Lev, S. M., McLennan, S. M. & Hanson, G. N. 1999. Mineralogic controls on REE mobility during black-shale diagenesis. Journal of Sedimentary Research 69, 1071–82.Google Scholar
Li, Y.-H. 2000. A Compendium of Geochemistry from Solar Nebula to the Human Brain. Princeton: Princeton University Press, 475 pp.Google Scholar
Lynas, B. D. 1988. Evidence for dextral oblique-slip faulting in the Shelve Ordovician Inlier, Welsh Borderland: implications for the South British Caledonides. Geological Journal 23, 3957.Google Scholar
Lyons, T. W., Werne, J. P., Hollander, D. J. & Murray, R. W. 2003. Contrasting sulfur geochemistry and Fe/Al and Mo/Al ratios across the last oxic-to-anoxic transition in the Cariacou Basin, Venezuela. Chemical Geology 195, 131–57.Google Scholar
Mack, G. H. 1984. Exceptions to the relationship between plate tectonics and sandstone composition. Journal of Sedimentary Petrology 54, 212–20.Google Scholar
Masuzawa, T., Noriki, S., Kurosaki, T., Tsunagai, S. & Koyama, M. 1989. Compositional change of settling particles with water depth in the Japan Sea. Marine Chemistry 27, 6178.Google Scholar
Mathieu, R., Pagel, M., Clauer, N., De Windt, L., Cabrera, J. & Boisson, J. Y. 2000. Paleocirculation in shales: a mineralogical and geochemical study of calcite veins from the Tournemire tunnel site (Aveyron, France). European Journal of Mineralogy 12, 377–90.Google Scholar
Maynard, J. B., Valloni, R. & Yu, H.-S. 1982. Composition of modern deep-sea sands from arc-related basins. In Trench-Forearc Geology: Sedimentation and Tectonics on Modern and Ancient Active Plate Margins (ed. Leggett, J. K.), pp. 551–61. Geological Society of London, Special Publication no. 10.Google Scholar
McLennan, S. M., Hemmings, S., McDaniel, D. K. & Hanson, G. N. 1993. Geochemical approaches to sedimentation, provenance and tectonics. Geological Society of America Special Paper 285, 2140.Google Scholar
McLennan, S. M. & Taylor, S. R. 1980. Geochemical standards for sedimentary rocks: trace-element data for U.S.G.S. standards SCo-1, MAG-1 and SGR-1. Chemical Geology 29, 333343.Google Scholar
Merriman, R. J. 2006. Clay mineral assemblages in British Lower Palaeozoic mudrocks. Clay Minerals 41, 473512.Google Scholar
Milne, A. R. & Fitzpatrick, R. W. 1977. Titanium and zirconium minerals. In Minerals in Soil Environments (eds Dixon, J. B. & Weed, S. B.), pp. 1131–205. Madison, WI: Soil Science Society of America.Google Scholar
Minai, Y., Matsumoto, R. & Tominaga, T. 1977. Geochemistry of deep sea sediments from the Nankai Trough, the Japan Trench, and adjacent regions. In Initial Reports of the Deep Sea Drilling Project vol. 87 (eds Kagami, H., Karig, D. E., Colbourn, W. T., et al.), pp. 643–57. College Station, Texas.Google Scholar
Moreno, T., Querol, X., Castillo, S., Alastuey, A., Cuevas, E., Herrmann, L., Mounkaila, M., Elvira, J. & Gibbons, W. 2006. Geochemical variations in Aeolian mineral particles from the Sahara-Sahel dust corridor. Chemosphere 65, 261–70.Google Scholar
Morford, J. L. & Emerson, S. E. 1999. The geochemistry of redox sensitive trace metals in sediments. Geochimica et Cosmochimica Acta 63, 1715–50.Google Scholar
Murchison, R. I. 1834. On the structure and classification of the Transition rocks of Shropshire, Herefordshire, and part of Wales, and on the lines of disturbance which have affected that series of deposits, including the Valley of Elevation at Woolhope. Proceedings of the Geological Society of London 2: 1318.Google Scholar
Murray, R.W. & Leinen, I. 1993. Chemical transport to the seafloor of the equatorial Pacific across a latitudinal transect at 135°W: tracking sedimentary major, trace, and rare earth element fluxes at the Equator and the ITCZ. Geochimica et Cosmochimica Acta 57, 4141–63.Google Scholar
Nadeau, P. H. 2010. Earth's energy “Golden Zone”: a synthesis from mineralogical research. Clay Minerals 46, 124.Google Scholar
Nameroff, T. J., Balistrieri, L. S. & Murray, J. W. 2002. Suboxic trace metal geochemistry in the Eastern Tropical North Pacific. Geochimica et Cosmochimica Acta 7, 1139–58.Google Scholar
Nesbitt, H. W. & Young, G. M. 1982. Early Proterozoic climates and plate motions inferred from major element chemistry of lutites. Nature 199, 715–17.Google Scholar
Ogg, J. G., Ogg, G. & Gradstein, F. M. 2008. The Concise Geologic Time Scale. Cambridge: Cambridge University Press, 177 pp.Google Scholar
Orton, G. J. 1991. Emergence of subaqueous depositional environments in advance of a major ignimbrite eruption, Capel Curig Volcanic Formation, Ordovician, North Wales – an example of regional volcanotectonic uplift? Sedimentary Geology 74, 251–86.Google Scholar
Owen, A. W. & Ingham, J. K. 1988. The stratigraphical distribution and taxonomy of the trilobite Onnia in the type Onnian stage of the uppermost Caradoc. Palaeontology 31, 829–55.Google Scholar
Pain, C. F. 1971. Micromorphology of soils developed from volcanic ash and river alluvium in the Kokoda valley, Northern District, Papua. Journal of Soil Science 22, 275–80.Google Scholar
Pancost, R. D., Freeman, K. H., Patzkowsky, M. E., Wavrek, D. A. & Collister, J. W. 1998. Molecular indicators of redox and marine photoautotroph composition in the late Middle Ordovician of Iowa, U.S.A. Organic Geochemistry 29, 1649–62.Google Scholar
Parnell, J. 1987. Secondary porosity in hydrocarbon-bearing transgressive sandstones on an unstable Lower Palaeozoic continental shelf. In Diagenesis of Sedimentary Sequences (ed. Marshall, J. D.), pp. 297312. Geological Society of London, Special Publication no. 36.Google Scholar
Perry, A. A. Jr. 1972. Diagenesis and the validity of the boron paleosalinity technique. American Journal of Science 272, 150–60.Google Scholar
Pharaoh, T. C., Webb, P. C., Thorpe, R. S. & Beckinsale, R. D. 1987. Geochemical evidence for the tectonic setting of Late Proterozoic volcanic suites in central England. In Geochemistry and Mineralization of Proterozoic Volcanic Suites (eds Pharaoh, T. C., Beckinsale, R. D. & Rickard, D.), pp. 541–52. Geological Society of London, Special Publication no. 33.Google Scholar
Pharaoh, T. C., Winchester, J. A., Verniers, J., Lassen, A. & Seghedi, A. 2006. The western accretionary margin of the East European craton: an overview. Geological Society of London Memoir 32, 291311.Google Scholar
Pickerill, R. K. & Brenchley, P. J. 1979. Caradoc marine benthic communities in the south Berwyn hills, North Wales. Palaeontology 22, 229–64.Google Scholar
Piper, D. Z. & Perkins, R. B. 2004. A modern vs. Permian black shale – the hydrography, primary productivity, and water-column chemistry of deposition, Chemical Geology 206, 177–97.Google Scholar
Potts, P. J., Tindle, A. G. & Webb, P. C. 1992. Geochemical Reference Material Compositions. Dunbeath: Whittles Publishing, 313 pp.Google Scholar
Ravizza, G. & Pyle, D. 1997. PGE and Os isotopic analysis of single sample aliquots with NiS fire assay preconcentration. Chemical Geology 141, 251–68.Google Scholar
Reitz, A., Pfeifer, K., De Lange, G. J. & Klump, J. 2004. Biogenic barium and the detrital Ba/Al ratio: a comparison of their direct and indirect determination. Marine Geology 204, 289300.Google Scholar
Rimmer, S. M. 2004. Geochemical paleoredox indicators in Devonian-Mississippian black shales, Central Appalachian Basin (USA). Chemical Geology 206, 373–91.Google Scholar
Riquier, L., Tribovillard, N., Averbuch, O., Devleeschouwer, X. & Riboulleau, A. 2006. Late Frasnian Kellwasser horizons of the Harz Mountains (Germany): two oxygen-deficient periods resulting from different mechanisms. Chemical Geology 233, 137–55.Google Scholar
Rosenau, N. A., Herrmann, A. D. & Leslie, S. A. 2012. Conodont apatite δ18O values from a platform margin setting, Oklahoma, USA: implications for initiation of Late Ordovician icehouse conditions. Palaeogeography, Palaeoclimatology, Palaeoecology 315–316, 172–80.Google Scholar
Roser, B. P., Cooper, R. A., Nathan, S. & Tulloch, A. J. 1996. Reconnaissance sandstone geochemistry, provenance, and tectonic setting of the lower Paleozoic terranes of the West Coast and Nelson, New Zealand. New Zealand Journal of Geology and Geophysics 39, 116 Google Scholar
Ross, R. J. Jr., Naeser, C. W., Izett, G. A., Obradovich, J. D., Bassett, M. G., Hughes, C. P., Cocks, L. R. M., Dean, W. T., Ingham, J. K., Jenkins, C. J., Rickards, R. B., Sheldon, P. R., Toghill, P., Whittington, H. B. & Zalasiewicz, J. 1982. Fission-track dating of British Ordovician and Silurian stratotypes. Geological Magazine 119, 135–53.Google Scholar
Rushton, A. W. A., Owen, A. W., Owens, R. M. & Prigmore, J. K. 1999. British Cambrian to Ordovician Stratigraphy. Peterborough: Joint Nature Conservation Committee, 435 pp.Google Scholar
Sadler, P. M., Cooper, R. A. & Melchin, M. A. 2009. High-resolution early Paleozoic (Ordovician-Silurian) time scales. Geological Society of America Bulletin 121, 887906.Google Scholar
Savage, K. M. & Potter, P. E. 1991. Petrology of modern sands of the rios Guaviare and Inirida, southern Colombia: tropical climate and sand composition. Journal of Geology 99, 289–98.Google Scholar
Savage, N. M. & Bassett, M. G. 1985. Caradoc-Ashgill conodont faunas from Wales and the Welsh borderland. Palaeontology 28, 679713.Google Scholar
Schofield, D. 2009. What's in the Welsh Basin? Insights into the evolution of Central Wales and the Welsh Borderlands during the Lower Palaeozoic. Proceedings of the Shropshire Geological Society 14, 117.Google Scholar
Shankar, R., Subbarao, K. V. & Kolla, V. 1987. Geochemistry of surface sediments from the Arabina Sea. Marine Geology 76, 253–79.Google Scholar
Shimmield, G. B. & Mowbray, S. 1991. The inorganic geochemical record of the northwest Arabian Sea: a history of productivity variation over the last 400 ky from Sites 722 and 724. Proceedings of the Ocean Drilling Program, Scientific Results, vol. 117 (eds Prell, W. L., Niitsuma, N. et al.), pp. 409–29. College Station: Texas.Google Scholar
Shimmield, G. B., Mowbray, S. R. & Weedon, G. P. 1990. A 350 ka history of the Indian Southwest Monsoon—evidence from deep-sea cores, northwest Arabian Sea. Transactions of the Royal Society of Edinburgh 81, 289–99.Google Scholar
Strachan, I., Temple, J. & Williams, A. 1948. The age of the Neptunian dyke at Hazler hill. Geological Magazine 85, 276–8.Google Scholar
Suttner, L. J., Basu, A. & Mack, G. H. 1981. Climate and the origin of quartz arenites. Journal of Sedimentary Petrology 51, 1235–46.Google Scholar
Taylor, S. R. & McLennan, S. M. 1985. The Continental Crust: Its Composition and Evolution. Oxford: Blackwell Scientific, 328 pp.Google Scholar
Thirlwall, M. F., Smith, T. E., Graham, A. M., Theodorou, N., Hollings, P., Davidson, J. P. & Arculus, R. J. 1994. High field strength element anomalies in arc lavas: source or process? Journal of Petrology 35, 819–38.Google Scholar
Thorpe, R. S., Leat, P. T., Mann, A. C., Howells, M. F., Reedman, A. J. & Campbell, S. D. G. 1993. Magmatic evolution of the Ordovician Snowdon volcanic centre, North Wales, (UK). Journal of Petrology 34, 711–41.Google Scholar
Toghill, P. 1992 a. Onny Valley, Shropshire: Geology Teaching Trail. Geologists’ Association Field Guide 45, 21 pp.Google Scholar
Toghill, P. 1992 b. The Shelveian event, a late Ordovician tectonic episode in Southern Britain (Eastern Avalonia). Proceedings of the Geologists’ Association 103, 31–5.Google Scholar
Totten, M. W., Hanan, M. A. & Weaver, B. L. 2000. Beyond whole-rock geochemistry of shales: the importance of assessing mineralogic controls for revealing tectonic discriminants of multiple sediment sources for the Ouachita flysch deposits. Geological Society of America Bulletin 112, 1012–22.Google Scholar
Trench, A., Torsvik, T. H., Smethurst, M. A., Woodcock, N. H. & Metcalfe, R. 1991. A palaeomagnetic study of the Builth Wells-Llandrindod Wells Ordovician inlier, Wales: palaeogeographic and structural implications. Geophysical Journal International 105, 477–89.Google Scholar
Tribovillard, N., Algeo, T. J., Lyons, T. & Riboulleau, A. 2006. Trace metals as paleoredox and paleoproductivity proxies: an update. Chemical Geology 232, 1232.Google Scholar
Tucker, M. 2001. Sedimentary Petrology. Oxford: Blackwell, 262 pp.Google Scholar
Tucker, R. D. & McKerrow, W. S. 1995. Early Paleozoic chronology: a review in light of new U-Pb zircon ages from Newfoundland and Britain. Canadian Journal of Earth Sciences 32, 368–79.Google Scholar
Turgeon, S. & Brumsack, H.-J. 2006. Anoxic vs dysoxic events reflected in sediment geochemistry during the Cenomanian–Turonian Boundary Event (Cretaceous) in the Umbria–Marche Basin of central Italy. Chemical Geology 234, 321–39.Google Scholar
Vakhrameev, V. A. 1991. Jurassic and Cretaceous Floras and Climates of the Earth. Cambridge: Cambridge University Press, 340 pp.Google Scholar
Valloni, R. 1985. Reading provenance from modern marine sands. In Provenance of Arenites (ed. Zuffa, G. G.), pp. 309–32. Dordrecht: D. Reidel.Google Scholar
Van der Weijden, C. H. 2002. Pitfalls of normalization of marine geochemical data using a common divisor. Marine Geology 184, 167–87.Google Scholar
Vandenbroucke, T. R. A., Anciletta, A., Fortey, R. A. & Verniers, J. 2009. A modern assessment of Ordovician chitinozoans from the Shelve and Caradoc areas, Shropshire, and their significance for correlation. Geological Magazine 146, 216–36.Google Scholar
Vandenbroucke, T. R. A., Armstrong, H. A., Williams, M., Paris, F., Zalasiewicz, J. A., Sabbe, K., Nölvak, J., Challands, T. J., Verniers, J. & Servais, T. 2010. Polar front shift and atmospheric CO2 during the glacial maximum of the Early Paleozoic Icehouse. Proceedings of the National Academy of Sciences (USA) 107, 14983–6.Google Scholar
Veizer, J. & Mackenzie, F. T. 2004. Evolution of sedimentary rocks. In Treatise on Geochemistry, Vol. 7; Sediments, Diagenesis and Sedimentary Rocks (ed. Mackenzie, F. T.), pp. 369407. Amsterdam: Elsevier.Google Scholar
Webby, B. D., Cooper, R. A., Bergström, S. M. & Paris, F. 2004. Stratigraphic framework and time slices. In The Great Ordovician Biodiversification Event (eds Webby, B. D., Paris, F., Droser, M. L. & Percival, I. G.), pp. 41–7. New York: Columbia University Press.Google Scholar
Wei, G., Liu, Y., Li, X., Shao, L. & Liang, X. 2003. Climatic impact on Al, K, Sc and Ti in marine sediments: evidence from ODP Site 1144, South China Sea. Geochemical Journal 37, 593602.Google Scholar
Whittard, W. F. 1953. Report of Summer Field Meeting in South Shropshire. Proceedings of the Geologists’ Association 19, 173–83.Google Scholar
Whittard, W. F. 1979. An account of the Ordovician rocks of the Shelve Inlier in west Salop and part of north Powys. Bulletin of the British Museum (Natural History) Geology 33 (1), 1169.Google Scholar
Wickman, F. E. 1962. The amount of material necessary for a trace element analysis. Arkiv för Mineralogi Och Geologi 10, 131–9.Google Scholar
Wignall, P. B. 1994. Black Shales. Oxford: Oxford University Press, 130 pp.Google Scholar
Wignall, P. B. & Myers, K. J. 1988. Interpreting the benthic oxygen levels in mudrocks, a new approach. Geology 16, 452–5.Google Scholar
Wilde, P., Quinby-Hunt, M. S. & Erdtmann, B.-D. 1996. The whole-rock cerium anomaly: a potential indicator of eustatic sea-level changes in shales of the anoxic facies. Sedimentary Geology 101, 4353.Google Scholar
Wilde, P., Lyons, T.W. & Quinby-Hunt, M. S. 2004. Organic carbon proxies in black shales: molybdenum. Chemical Geology 206, 167–76.Google Scholar
Williams, J., Basu, A., Bhargava, O. N., Ahluwalia, A. D. & Hannigan, R. 2012. Resolving original signatures from a sea of overprint – the geochemistry of the Gungri Shale (Upper Permian, Spiti Valley, India). Chemical Geology 324–5, 5972.Google Scholar
Woodcock, N. H. 1984 a. The Pontesford Lineament, Welsh Borderland. Journal of the Geological Society, London 141,1001–14.Google Scholar
Woodcock, N. H. 1984 b. Early Palaeozoic sedimentation and tectonics in Wales. Proceedings of the Geologists’ Association 95, 323–35.Google Scholar
Yan, D., Chen, D., Wang, Q. & Wang, J. 2010. Large-scale climatic fluctuations in the latest Ordovician on the Yangtze block, south China. Geology 38, 599602.Google Scholar
Yarincik, K. M., Murray, R. W. & Peterson, L. C. 2000. Climatically sensitive eolian and hemipelagic deposition in the Cariaco basin, Venezuela, over the past 578,000 years: results from Al/Ti and K/Al. Paleoceanography 15, 210–28.Google Scholar
Young, G. M., Minter, W. E. L. & Theron, J. N. 2004. Geochemistry and palaeogeography of upper Ordovician glaciogenic sedimentary rocks in the Table Mountain Group, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 214, 323–45.Google Scholar
Yuan, C., Sun, M., Yang, J., Zhou, H. & Zhou, M.-F. 2004. Nb-depleted, continental rift related metavolcanic rocks (West Kunlun): implication for the rifting of the Tarim Craton from Gondwanaland. In Aspects of the Tectonic Evolution of China (eds Malpas, J., Fletcher, C. J. N., Ali, J. R. & Aitchison, J. C.), pp. 131–43. Geological Society of London, Special Publication no. 226.Google Scholar
Zhao, M., Zhang, B., Bian, L., Xiao, Z., Li, M., Peng, Y. & Qing, S. 2000. Features of the type III-like source rock and its generated natural gas. Chinese Science Bulletin 45, 857–61.Google Scholar
Supplementary material: File

Hannigan Supplementary Material

Appendix

Download Hannigan Supplementary Material(File)
File 27.1 KB