Hostname: page-component-8448b6f56d-gtxcr Total loading time: 0 Render date: 2024-04-18T19:21:27.270Z Has data issue: false hasContentIssue false

Role of pressure solution in the formation of bedding-parallel calcite veins in an immature shale (Cretaceous, southern UK)

Published online by Cambridge University Press:  15 May 2018

QINGFENG MENG*
Affiliation:
Department of Earth Sciences, University of Oxford, South Parks Road, Oxford, OX1 3AN, UK
JOHN HOOKER
Affiliation:
Department of Earth Sciences, University of Oxford, South Parks Road, Oxford, OX1 3AN, UK
JOE CARTWRIGHT
Affiliation:
Department of Earth Sciences, University of Oxford, South Parks Road, Oxford, OX1 3AN, UK
*
Author for correspondence: meng.qingfeng@hotmail.com

Abstract

Bedding-parallel fibrous calcite veins in black shales (Cretaceous, southern UK) were investigated using a combined field, stable isotopic geochemistry, petrographic and crystallographic method to examine their formation mechanism. Calcite veins occur in all shale beds and are most abundant in the bituminous shales of the Chief Beef Beds. The calcite fibres in these veins exhibit either an antitaxial fibre growth with curvy stylolites as the median zone, or a predominantly syntaxial, upwards growth. The calcite veins range from –0.49 to 1.78‰ of δ13C values, and –6.53 to –0.03‰ of δ18O values, which are both similar to those of their host shales. Our petrographic observations demonstrate that subhorizontal and interconnecting microstylolite networks commonly occur within the calcite veins. Equant calcite grains in the median zones exhibit indenting, truncating and also interpenetrating grain contacts. It is interpreted that the fibrous calcite veins were sourced by neomorphic calcite from their host shales, with evidence from the δ13C signatures, pressure-solution features (stylolites, microstylolites and grain contact styles) and embedded fossil ghosts within the veins. The diagenetic fluids, from which calcite was precipitated, were a mixing of the original seawaters and 18O-depleted meteoric waters. Development of bedding-parallel calcite veins is considered to have been enhanced by pressure solution as a positive feedback mechanism, which was facilitated by the overburden pressure as the maximum principal stress. Calcite fibres, with a predominant subvertical c-axis orientation, exhibit a displacive growth in porous shales and a replacive growth at vein-limestone contacts. This study highlights the critical role of pressure solution in the formation of bedding-parallel calcite veins during burial and diagenesis of immature black shales.

Type
Original Article
Copyright
Copyright © Cambridge University Press 2018 

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

Al-Aasm, I., Coniglio, M. & Desrochers, A. 1995. Formation of complex fibrous calcite veins in Upper Triassic strata of Wrangellia Terrain, British Columbia, Canada. Sedimentary Geology 100, 8395.Google Scholar
Andrews, J. E. & Walton, W. 1990. Depositional environments within Middle Jurassic oyster-dominated lagoons: an integrated litho-, bio- and palynofacies study of the Duntulm Formation (Great Estuarine Group, Inner Hebrides). Earth and Environmental Science Transactions of The Royal Society of Edinburgh 81, 122.Google Scholar
Baron, M. & Parnell, J. 2007. Relationships between stylolites and cementation in sandstone reservoirs: Examples from the North Sea, UK and East Greenland. Sedimentary Geology 194, 1735.Google Scholar
Bjorkum, P. A. 1996. How important is pressure in causing dissolution of quartz in sandstones? Journal of Sedimentary Research 66, 147–54.Google Scholar
Boggs, S. 2009. Petrology of Sedimentary Rocks. Cambridge: Cambridge University Press.Google Scholar
Bons, P. D. 2000. The formation of veins and their microstructures. Journal of the Virtual Explorer 2, doi: 10.3809/jvirtex.2000.00007.Google Scholar
Bons, P. D., Elburg, M. A. & Gomez-Rivas, E. 2012. A review of the formation of tectonic veins and their microstructures. Journal of Structural Geology 43, 3362.Google Scholar
Bons, P. D. & Montenari, M. 2005. The formation of antitaxial calcite veins with well-developed fibres, Oppaminda Creek, South Australia. Journal of Structural Geology 27, 231–48.Google Scholar
Braithwaite, C. 1989. Stylolites as open fluid conduits. Marine and Petroleum Geology 6, 93–6.Google Scholar
Bray, R. J., Duddy, I. R. & Green, P. F. 1998. Multiple heating episodes in the Wessex Basin: implications for geological evolution and hydrocarbon generation. In The Development, Evolution and Petroleum Geology of the Wessex Basin (eds Underhill, J. R.), pp. 199213. Geological Society, London, Special Publication no. 133.Google Scholar
British Geological Survey. 2000. 1:63,360/1:50,000 geological map series, New Series, Sheet number 342 (East) and 343, Swanage. British Geological Survey, Keyworth.Google Scholar
Buxton, T. M. & Sibley, D. F. 1981. Pressure solution features in a shallow burled limestone. Journal of Sedimentary Research 51, 1926.Google Scholar
Carozzi, A. V. & Von Bergen, D. 1987. Stylolitic porosity in carbonates: a critical factor for deep hydrocarbon production. Journal of Petroleum Geology 10, 267–82.Google Scholar
Clements, R. 1993. Type-section of the Purbeck Limestone Group, Durlston Bay, Swanage, Dorset. Proceedings of the Dorset Natural History and Archaeological Society 114, 181206.Google Scholar
Cobbold, P. R. & Rodrigues, N. 2007. Seepage forces, important factors in the formation of horizontal hydraulic fractures and bedding-parallel fibrous veins (‘beef’ and ‘cone-in-cone’). Geofluids 7, 313–22.Google Scholar
Cobbold, P. R., Zanella, A., Rodrigues, N. & Loseth, H. 2013. Bedding-parallel fibrous veins (beef and cone-in-cone): Worldwide occurrence and possible significance in terms of fluid overpressure, hydrocarbon generation and mineralization. Marine and Petroleum Geology 43, 120.Google Scholar
Conybeare, D. & Shaw, H. 2000. Fracturing, overpressure release and carbonate cementation in the Everest Complex, North Sea. Clay Minerals 35, 135–49.Google Scholar
Cosgrove, J. W. 2001. Hydraulic fracturing during the formation and deformation of a basin: a factor in the dewatering of low-permeability sediments. AAPG Bulletin 85, 737–48.Google Scholar
Dewers, T. & Ortoleva, P. 1990. A coupled reaction/transport/mechanical model for intergranular pressure solution, stylolites, and differential compaction and cementation in clean sandstones. Geochimica et Cosmochimica Acta 54, 1609–25.Google Scholar
El-Shahat, A. & West, I. 1983. Early and late lithification of aragonitic bivalve beds in the Purbeck Formation (Upper Jurassic – Lower Cretaceous) of southern England. Sedimentary Geology 35, 1541.Google Scholar
El-Tabakh, B. & Warren, J. K. 1998. Origin of fibrous gypsum in the Newark rift basin, eastern North America. Journal of Sedimentary Research 68, 8899.Google Scholar
Evans, M. A. 1995. Fluid inclusions in veins from the Middle Devonian shales: A record of deformation conditions and fluid evolution in the Appalachian Plateau. Geological Society of America Bulletin 107, 327–39.Google Scholar
Finkel, E. A. & Wilkinson, B. H. 1990. Stylolitization as source of cement in Mississippian Salem Limestone, west-central Indiana. AAPG Bulletin 74, 174–86.Google Scholar
Fitches, W., Cave, R., Craig, J. & Maltman, A. 1986. Early veins as evidence of detachment in the Lower Palaeozoic rocks of the Welsh Basin. Journal of Structural Geology 8, 607–20.Google Scholar
Fletcher, R. C. & Pollard, D. D. 1981. Anticrack model for pressure solution surfaces. Geology 9, 419–24.Google Scholar
Fossen, H. 2016. Structural Geology. Cambridge: Cambridge University Press.Google Scholar
Fowler, T. 1996. Flexural-slip generated bedding-parallel veins from central Victoria, Australia. Journal of Structural Geology 18, 1399–415.Google Scholar
Fowler, T. & Winsor, C. 1997. Characteristics and occurrence of bedding-parallel slip surfaces and laminated veins in chevron folds from the Bendigo-Castlemaine goldfields: implications for flexural-slip folding. Journal of Structural Geology 19, 799815.Google Scholar
Goulty, N., Ramdham, A. & Jones, S. 2012. Chemical compaction of mudrocks in the presence of overpressure. Petroleum Geoscience 18, 471–9.Google Scholar
Greenhalgh, E. 2016. The Jurassic Shales of the Wessex Area: Geology and Shale Oil and Shale Gas Resource Estimation. London: British Geological Survey for the Oil and Gas Authority, 82 pp.Google Scholar
Gustavson, T. C., Hovorka, S. D. & Dutton, A. R. 1994. Origin of satin spar veins in evaporite basins. Journal of Sedimentary Research 64, 8894.Google Scholar
Guzzetta, G. 1984. Kinematics of stylolite formation and physics of the pressure-solution process. Tectonophysics 101, 383–94.Google Scholar
Hara, H. & Hisada, K. I. 2007. Tectono-metamorphic evolution of the Cretaceous Shimanto accretionary complex, central Japan: Constraints from a fluid inclusion analysis of syn-tectonic veins. Island Arc 16, 5768.Google Scholar
Heap, M. J., Baud, P., Reuschle, T. & Meredith, P. G. 2014. Stylolites in limestones: Barriers to fluid flow? Geology 42, 51–4.Google Scholar
Hendry, J. P., Ditchfield, P. W. & Marshall, J. D. 1995. Two-stage neomorphism of Jurassic aragonitic bivalves: implications for early diagenesis. Journal of Sedimentary Research 65, 214–24.Google Scholar
Heydari, E. 2000. Porosity loss, fluid flow, and mass transfer in limestone reservoirs: application to the Upper Jurassic Smackover Formation, Mississippi. AAPG Bulletin 84, 100–18.Google Scholar
Hilgers, C. & Urai, J. L. 2005. On the arrangement of solid inclusions in fibrous veins and the role of the crack-seal mechanism. Journal of Structural Geology 27, 481–94.Google Scholar
Hillier, R. & Cosgrove, J. 2002. Core and seismic observations of overpressure-related deformation within Eocene sediments of the Outer Moray Firth, UKCS. Petroleum Geoscience 8, 141–9.Google Scholar
Hooker, J. N., Cartwright, J., Stepehnson, B., Silver, C. R., Dickson, A. J. & Hsieh, Y.-T. 2017. Fluid evolution in fracturing black shales, Appalachian Basin. AAPG Bulletin 101, 1203–38.Google Scholar
Hopson, P. M., Wilkinson, I. P. & Woods, M. A. 2008. A stratigraphical framework for the Lower Cretaceous of England. Research Report RR/08/03, British Geological Survey, Keyworth.Google Scholar
Horton, A. 1995. Geology of the Country Around Thame. Champaign: Balogh Scientific Books.Google Scholar
Houseknecht, D. W. 1984. Influence of grain size and temperature on intergranular pressure solution, quartz cementation, and porosity in a quartzose sandstone. Journal of Sedimentary Research 54, 348–61.Google Scholar
Houseknecht, D. W. 1988. Intergranular pressure solution in four quartzose sandstones. Journal of Sedimentary Research 58, 228–46.Google Scholar
Hudson, J. 1975. Carbon isotopes and limestone cement. Geology 3, 1922.Google Scholar
Hudson, J. 1978. Concretions, isotopes, and the diagenetic history of the Oxford Clay (Jurassic) of central England. Sedimentology 25, 339–70.Google Scholar
Humphreys, F. J. 2004. Characterisation of fine-scale microstructures by electron backscatter diffraction (EBSD). Scripta Materialia 51, 771–6.Google Scholar
Jamison, W. 2013. Bed-parallel expansion seams and shear surfaces in shales. Geoconvention 2013 (abstract), Calgary Canada, 6–12. http://www.searchanddiscovery.com/documents/2014/51053jamison/ndx_jamison.pdf.Google Scholar
Jessell, M., Willman, C. & Gray, D. 1994. Bedding parallel veins and their relationship to folding. Journal of Structural Geology 16, 753–67.Google Scholar
Jowett, E. C. 1987. Formation of sulfide-calcite veinlets in the Kupferschiefer Cu-Ag deposits in Poland by natural hydrofracturing during basin subsidence. The Journal of Geology 95, 513–26.Google Scholar
Kamb, W. B. 1959. Theory of preferred crystal orientation developed by crystallization under stress. The Journal of Geology 67, 153–70.Google Scholar
Kelka, U., Veveakis, M., Koehn, D. & Beaudoin, N. 2017. Zebra rocks: compaction waves create ore deposits. Scientific Reports 7, article no. 14260.Google Scholar
Koehn, D. & Passchier, C. W. 2000. Shear sense indicators in striped bedding-veins. Journal of Structural Geology 22, 1141–51.Google Scholar
Koehn, D., Rood, M., Beaudoin, N., Chung, P., Bons, P. & Gomez-Rivas, E. 2016. A new stylolite classification scheme to estimate compaction and local permeability variations. Sedimentary Geology 346, 6071.Google Scholar
Lacombe, O. 2010. Calcite twins, a tool for tectonic studies in thrust belts and stable orogenic forelands. Oil & Gas Science and Technology 65, 809–38.Google Scholar
Lehner, F. K. 1995. A model for intergranular pressure solution in open systems. Tectonophysics 245, 153–70.Google Scholar
Li, R., Dong, S., Lehrmann, D. & Duan, L. 2013. Tectonically driven organic fluid migration in the Dabashan Foreland Belt: Evidenced by geochemistry and geothermometry of vein-filling fibrous calcite with organic inclusions. Journal of Asian Earth Sciences 75, 202–12.Google Scholar
Lindgreen, H. 1985. Diagenesis and primary migration in Upper Jurassic claystone source rocks in North Sea. AAPG Bulletin 69, 525–36.Google Scholar
Machel, H. G. 1985. Fibrous gypsum and fibrous anhydrite in veins. Sedimentology 32, 443–54.Google Scholar
Maher, H. D., Ogata, K. & Braathen, A. 2016. Cone-in-cone and beef mineralization associated with Triassic growth basin faulting and shallow shale diagenesis, Edgeøya, Svalbard. Geological Magazine 154, 201–16.Google Scholar
Maliva, R. G. & Dickson, J. 1992. The mechanism of skeletal aragonite neomorphism: evidence from neomorphosed mollusks from the upper Purbeck Formation (Late Jurassic - Early Cretaceous), southern England. Sedimentary Geology 76, 221–32.Google Scholar
Mandl, G. 2005. Rock Joints. Berlin: Springer.Google Scholar
Marshall, J. D. 1982. Isotopic composition of displacive fibrous calcite veins: reversals in pore-water composition trends during burial diagenesis. Journal of Sedimentary Research 52, 615–30.Google Scholar
McLane, M. 1995. Sedimentology. Oxford: Oxford University Press.Google Scholar
Means, W. & Li, T. 2001. A laboratory simulation of fibrous veins: some first observations. Journal of Structural Geology 23, 857–63.Google Scholar
Meng, Q., Hooker, J. & Cartwright, J. 2017a. Early overpressuring in organic-rich shales during burial: evidence from fibrous calcite veins in the Lower Jurassic Shales-with-Beef Member in the Wessex Basin, UK. Journal of the Geological Society 174: 869–82.Google Scholar
Meng, Q., Hooker, J. & Cartwright, J. 2017b. Genesis of natural hydraulic fractures as an indicator of basin inversion. Journal of Structural Geology 102, 120.Google Scholar
Meng, Q., Hooker, J. & Cartwright, J. 2018. Displacive widening of fibrous calcite veins: insights into force of crystallization. Journal of Sedimentary Research 88, 327–43.Google Scholar
O'brien, D. K., Manghnani, M. H., Tribble, J. S. & Wenk, H.-R. 1993. Preferred orientation and velocity anisotropy in marine clay-bearing calcareous sediments. In Carbonate Microfabrics (eds Rezak, R. & Lavoie, D. L.), pp. 149–61. Berlin: Springer.Google Scholar
Oldershaw, A. & Scoffin, T. 1967. The source of ferroan and non-ferroan calcite cements in the Halkin and Wenlock Limestones. Geological Journal 5, 309–20.Google Scholar
Park, W. C. & Schot, E. H. 1968. Stylolites: their nature and origin. Journal of Sedimentary Research 38, 175–91.Google Scholar
Parnell, J., Ansong, G. & Veale, C. 1994. Petrology of the bitumen (manjak) deposits of Barbados: hydrocarbon migration in an accretionary prism. Marine and Petroleum Geology 11, 743–55.Google Scholar
Parnell, J. & Carey, P. F. 1995. Emplacement of bitumen (asphaltite) veins in the Neuquén Basin, Argentina. AAPG Bulletin 79, 1798–815.Google Scholar
Parnell, J., Carey, P. & Monson, B. 1996. Fluid inclusion constraints on temperatures of petroleum migration from authigenic quartz in bitumen veins. Chemical Geology 129, 217–26.Google Scholar
Parnell, J., Honghan, C., Middleton, D., Haggan, T. & Carey, P. 2000. Significance of fibrous mineral veins in hydrocarbon migration: fluid inclusion studies. Journal of Geochemical Exploration 69, 623–7.Google Scholar
Paxton, S., Szabo, J., Ajdukiewicz, J. & Klimentidis, R. 2002. Construction of an intergranular volume compaction curve for evaluating and predicting compaction and porosity loss in rigid-grain sandstone reservoirs. AAPG Bulletin 86, 2047–67.Google Scholar
Pittman, E. D. & Larese, R. E. 1991. Compaction of lithic sands: experimental results and applications. AAPG Bulletin 75, 1279–99.Google Scholar
Radley, J. D. 2009. Archaic-style shell concentrations in brackish-water settings: Lower Cretaceous (Wealden) examples from southern England. Cretaceous Research 30, 710–6.Google Scholar
Radley, J. D., Barker, M. J. & Munt, M. C. 1998. Bivalve trace fossils (Lockeia) from the Barnes High Sandstone (Wealden Group, Lower Cretaceous) of the Wessex Sub-basin, southern England. Cretaceous Research 19, 505–9.Google Scholar
Railsback, L. B. 2002. An atlas of pressure solution features. http://www.gly.uga.edu/railsback/PDFintro1.html.Google Scholar
Ramsay, J. 1980. The crack-seal mechanism of rock deformation. Nature 284, 135–9.Google Scholar
Renard, F., Dysthe, D., Feder, J., Bjorlykke, K. & Jamtveit, B. 2001. Enhanced pressure solution creep rates induced by clay particles: Experimental evidence in salt aggregates. Geophysical Research Letters 28, 1295–8.Google Scholar
Renard, F., Gratier, J. P. & Jamtveit, B. 2000. Kinetics of crack-sealing, intergranular pressure solution, and compaction around active faults. Journal of Structural Geology 22, 1395–407.Google Scholar
Renard, F., Ortoleva, P. & Gratier, J. P. 1997. Pressure solution in sandstones: influence of clays and dependence on temperature and stress. Tectonophysics 280, 257–66.Google Scholar
Rezaee, M. R. & Tingate, P. R. 1997. Origin of quartz cement in the Tirrawarra sandstone, southern Cooper basin, South Australia. Journal of Sedimentary Research 67, 168–77.Google Scholar
Richter, D. K. & Fuchtbauer, H. 1978. Ferroan calcite replacement indicates former magnesian calcite skeletons. Sedimentology 25, 843–60.Google Scholar
Robinson, S. A. & Hesselbo, S. P. 2004. Fossil-wood carbon-isotope stratigraphy of the non-marine Wealden Group (Lower Cretaceous, southern England). Journal of the Geological Society 161, 133–45.Google Scholar
Rodrigues, N., Cobbold, P. R., Loseth, H. & Ruffet, G. 2009. Widespread bedding-parallel veins of fibrous calcite (‘beef’) in a mature source rock (Vaca Muerta Fm, Neuquén Basin, Argentina): evidence for overpressure and horizontal compression. Journal of the Geological Society 166, 695709.Google Scholar
Rukin, N. 1990. The diagenesis of the Shales-with-beef of the Lower Lias, West Dorset. PhD Thesis, University of Liverpool. Published thesis.Google Scholar
Rutter, E. 1983. Pressure solution in nature, theory and experiment. Journal of the Geological Society 140, 725–40.Google Scholar
Rutter, E. & Elliott, D. 1976. The kinetics of rock deformation by pressure solution. Philosophical Transactions of the Royal Society of London A: Mathematical, Physical and Engineering Sciences 283, 203–19.Google Scholar
Schnyder, J., Ruffel, A., Deconinck, J.-F. & Baudin, F. 2006. Conjunctive use of spectral gamma-ray logs and clay mineralogy in defining late Jurassic–early Cretaceous palaeoclimate change (Dorset, UK). Palaeogeography, Palaeoclimatology, Palaeoecology 229, 303–20.Google Scholar
Scholle, P. A. 1977. Chalk diagenesis and its relation to petroleum exploration: oil from chalks, a modern miracle? AAPG Bulletin 61, 9821009.Google Scholar
Scotchman, I. 1987. Clay diagenesis in the Kimmeridge Clay Formation, onshore UK, and its relation to organic maturation. Mineralogical Magazine 51, 535–51.Google Scholar
Scotchman, I. 1989. Diagenesis of the Kimmeridge Clay formation, onshore UK. Journal of the Geological Society 146, 285303.Google Scholar
Selles-Martinez, J. 1994. New insights in the origin of cone-in-cone structures. Carbonates and Evaporites 9, 172–86.Google Scholar
Selles-Martinez, J. 1996. Concretion morphology, classification and genesis. Earth-Science Reviews 41, 177210.Google Scholar
Shearman, D., Mossop, G., Dunsmore, H. & Martin, M. 1972. Origin of gypsum veins by hydraulic fracture. Institution of Mining and Metallurgy, Transactions, Section B: Applied Earth Science 81, 149–55.Google Scholar
Sibley, D. F. & Blatt, H. 1976. Intergranular pressure solution and cementation of the Tuscarora orthoquartzite. Journal of Sedimentary Research 46, 881–96.Google Scholar
Stewart, D., Ruffell, A., Wach, G. & Goldring, R. 1991. Lagoonal sedimentation and fluctuating salinities in the Vectis Formation (Wealden Group, Lower Cretaceous) of the Isle of Wight, southern England. Sedimentary Geology 72, 117–34.Google Scholar
Stoneley, R. 1983. Fibrous calcite veins, overpressures, and primary oil migration. AAPG Bulletin 67, 1427–8.Google Scholar
Suchy, V., Dobes, P., Filip, J., Stejskal, M. & Zeman, A. 2002. Conditions for veining in the Barrandian Basin (Lower Palaeozoic), Czech Republic: evidence from fluid inclusion and apatite fission track analysis. Tectonophysics 348, 2550.Google Scholar
Taber, S. 1918. The origin of veinlets in the Silurian and Devonian strata of central New York. The Journal of Geology 26, 5673.Google Scholar
Tada, R., Maliva, R. & Siever, R. 1987. A new mechanism for pressure solution in porous quartzose sandstone. Geochimica et Cosmochimica Acta 51, 2295–301.Google Scholar
Tada, R. & Siever, R. 1989. Pressure solution during diagenesis. Annual Review of Earth and Planetary Sciences 17, 89118.Google Scholar
Thomson, A. 1959. Pressure solution and porosity. In Silica in Sediments (eds Ireland, H. A.), pp. 92111. Society of Economic Paleontologists and Mineralogists, Special Publication no. 7.Google Scholar
Underhill, J. R. & Stoneley, R. 1998. Introduction to the development, evolution and petroleum geology of the Wessex Basin. In: The Development, Evolution and Petroleum Geology of the Wessex Basin (eds Underhill, J. R.), pp. 118. Geological Society, London, Special Publication no. 133.Google Scholar
Vannucchi, P. 2001. Monitoring paleo-fluid pressure through vein microstructures. Journal of Geodynamics 32, 567–81.Google Scholar
Watts, N. 1978. Displacive calcite: evidence from recent and ancient calcretes. Geology 6, 699703.Google Scholar
Westhead, R. & Mather, A. 1996. An updated lithostratigraphy for the Purbeck Limestone Group in the Dorset type-area. Proceedings of the Geologists’ Association 107, 117–28.Google Scholar
Weyl, P. K. 1959. Pressure solution and the force of crystallization: a phenomenological theory. Journal of Geophysical Research 64, 2001–25.Google Scholar
Wiltschko, D. V. & Morse, J. W. 2001. Crystallization pressure versus “crack seal” as the mechanism for banded veins. Geology 29, 7982.Google Scholar
Wolff, G. A., Rukin, N. & Marshall, J. D. 1992. Geochemistry of an early diagenetic concretion from the Birchi Bed (L. Lias, W. Dorset, UK). Organic Geochemistry 19, 431–44.Google Scholar
Worden, R. H., Oxtoby, N. H. & Smalley, P. C. 1998. Can oil emplacement prevent quartz cementation in sandstones? Petroleum Geoscience 4, 129–37.Google Scholar
Yasuhara, H., Elsworth, D. & Polak, A. 2003. A mechanistic model for compaction of granular aggregates moderated by pressure solution. Journal of Geophysical Research: Solid Earth 108, doi: 10.1029/2003JB002536.Google Scholar
Zanella, A., Cobbold, P. R. & Boassen, T. 2015. Natural hydraulic fractures in the Wessex Basin, SW England: widespread distribution, composition and history. Marine and Petroleum Geology 68, 438–48.Google Scholar
Zanella, A., Cobbold, P. R. & De Veslud, C. L. C. 2014. Physical modelling of chemical compaction, overpressure development, hydraulic fracturing and thrust detachments in organic-rich source rock. Marine and Petroleum Geology 55, 262–74.Google Scholar
Zanella, A., Cobbold, P. R., Ruffet, G. & Leanza, H. A. 2015. Geological evidence for fluid overpressure, hydraulic fracturing and strong heating during maturation and migration of hydrocarbons in Mesozoic rocks of the northern Neuquén Basin, Mendoza Province, Argentina. Journal of South American Earth Sciences 62, 229–42.Google Scholar