Vous êtes sur la page 1sur 6

Introduction

The Karoo basin is a retroarc foreland basin (Catuneanu et al. 1998) with a present area of approximately 600 000km2 (Cole, 1992). Sedimentation in the Karoo basin began in the Late Carboniferous at around 300 Ma. Sedimentation progressed up until the breakup of Gondwana during the Middle Jurassic (Catuneanu et al. 2005). The thickness of the basin is ~12km in the southeastern part towards the eastern end of the Karoo Trough (Johnson). The Karoo stratigraphic sequence can be divided into 5 groups namely the Dwyka, Ecca, Beaufort and Strormberg Groups. (Catuneanu et al. 2005).

Tectonic Setting
The Karoo Basin subsidence framework is characteristic of accumulation convergent tectonics related to first order cycle supercontinent assembly and break-up. Throughout the Karoo interval accommodation was generated via a series of compressional and accretionary events which took place along the southern margin of Gondwana (Catuneanu et al, 1998; Catuneanu, 2004; Cole; 1992). According to Catuneanu et al (2005), the southern (convergent) margin occasioned shallowangle subduction of the palaeo-Pacific plate beneath the supercontinent (Lock, 1980). This resulted in compression associated with collision and terrane accretion which spawned the 6000km long PanGondwanian fold-thrust belt. The Cape Fold Belt exists as a preserved remnant of this ancient belt and also signifies the supracrustal (tectonic) load which controlled the creation of the Karoo retroarc foreland system including the main Karoo Basin. Flexural tectonics together with dynamic subsidence acted as the main controls on accommodation inside the main Karoo Basin. Orogenic loading from the southern super continental margin imposed flexural tectonics which resulted in the initial subsidence of the Karoo foredeep. The continuation of tectonic loading has led to the portioning of the entire foreland system into foredeep, forebuldge, and backbuldge flexural provinces (Catuneanu, 2004). Flexural tectonics was later enhanced by dynamic subsidence which created additional accommodation across the entire foreland system (Pysklywec and Mitrovica, 1999). Dynamic subsidence occurs via sub-lithospheric loading of the overriding plate caused by the drag force generated by viscous mantle corner flow joined to the subduction plate. This occurs specifically where subduction is rapid and/or taking place at a shallow angle beneath the retroarc foreland system (Mitrovica et al., 1989; Gurnis, 1992; Holt and Stern, 1994; Burgess et al., 1997). Dynamic loading lags behind in time after the initiation of subduction and tectonic loading. This is as a result of the time taken for the subducting slab to reach far enough beneath the overriding plate to generate a viscous corner flow. The early stages of retroarc foreland system evolution are dominated by flexural tectonics when the foredeep is underfilled and the forebuldge is elevated above the base level which is then subject to erosion. The interaction of base level changes and sediment supply controls the degree to which available accommodation is filled by sedimentation. This may then be defined as the underfilled, filled, and overfilled stages in the development of the foreland system. In relation to this are depositional processes dominated by deep marine, shallow marine and fluvial sedimentation styles respectively (Sinclair and Allen, 1992; Catuneanu, 2004). Changes from underfilled to overfilled stages are best perceived in the foredeep, because the forebuldge may be subject to erosion in the

absence of dynamic loading or it may accommodate (at best) shallow marine to fluvial environments even when the foredeep is underfilled (Catuneanu et al., 2002). The first-order Cape fold belt (CFB) cycle is further analysed by looking at the second-order tectonic pulses. The latter is determined using radiometric dating on CFB metamorphic rocks (Catuneanu et al, 1998) using Ar39/Ar40 isotope system (Cole, 1992) and also interpretation of the foreland stratigraphy. The orogenic pulses are followed by orogenic quiescence and each of these stages within the stratigraphic record of the Karoo is marked by distinct patterns of deposition as well as basin geometry changes. According to Catuneanu et al (1998) there were eight tectonic paroxysms (P1 P8) throughout the sequence and each is followed by sometimes discrete quiescence (Q1 Q8). The authors suggest that there were two distinct phases of hinge line movement, the first phase is that of the forward movement of the orogenic front causing progradation of the hinge line; and the second stage is that of retrogradation of the hinge line. The first phase occurs between ~ 300 246 Ma and is characterised by the deposition of the Dwyka Group, Ecca Group and early Adelaide Subgroup. This phase had four tectonic loading pulse stages (P1 P4) and after each pulse followed by a stage of quiescence (Q1 Q4), this resulted in the movement of the foreland away from the CFB and hence the migration of the sedimentary basin hinge line (Catuneanu et al, 1998). During the quiescence stage second-order unloading occurs as a result of erosion of the orogenic front and is then deposited in the basin forming the earlier mentioned sequences. During this phase, there was an out-of-phase deposition between the sediments found in the proximal versus those in the distal areas of the same group/subgroup/formation. Hence, the migration of the stratigraphic hinge line towards the north during the pulses (Catuneanu et al, 1998). The second phase occurs between ~242 184 Ma, it is characterised by two other sub-phases one of rapid, and of slow retrogradation of the foreland system towards the CFB respectively. With the first sub-phase, advancement of the orogenic front ceased at the end of the Permian, however, orogenic loading continued which is thought to be as a result of piggyback thrusting during pulses P5 and P6. This led to the retrogradation of the centre of weight in the orogenic belt and as a consequence the basin experienced high rates of retrogradation of the hinge-line. During this subphase (Early Middle Traissic) the rest of the Beaufort Group was laid down and an age gap between the top of the Tarkstad Group and the bottom of the Stormberg Group (Catuneanu et al, 1998). The second sub-phase (Late Triassic-Middle Jurassic) is characterised by slow rate of the retrogradation of the foreland system due to the gradual change of the centre of weight in the orogenic belt which was a result of second-oredr unloading due to erosion. This sub-phase is considered to have experienced prolonged periods of unloading (Q6, Q7 and Q8) with minor orogenic pulses (P7 and P8) which led to the uplift and erosion of the forebulge leading the unconformities found in the Molteno Formation. The end of the deposition of the Karoo sequence (Middle Jurassic) caused by the extrusion of the Drakensberg Group dolerites which was an indicator of first-order extension (Catuneanu et al, 1998).

Sedimentation Dwyka Group


The Dwyka Group was deposited during the Late Carboniferous to Early Permian (Catuneanu et al. 2005). It lies on glaciated Precambrian bedrock in the northern part of basin and unconformably in the south over the Cape Supergroup (Johnson). It was deposited in a predominantly glacio-marine environment (Catuneanu et al. 2005). The Dwyka Group is diachrounous in age being younger in the distal region and older in the proximal region. Lateral changes of this group depend on

the variability of the topography and indicate a transition from marine facies in the foredeep to continental facies to the north. The foredeep contains silt-dominated marine diamictites which extends up to 800m and also contains dropstones that originated from floating ice. The forebulge contains predominantly tillite derived from continental ice sheets (Catuneanu et al. 2005). The tillite cyclically grades upwards into fine-grained clastic rocks. In the proximal region, nine upward-fining cycles occur and this is indicated by a transition from terrestrial or subaqueous moraines at the base to glaciolacustrine shales at the top. It continues laterally and uniformly. The lateral continuity of the distal region is difficult to correlate due to irregular thickness and complex facies relationships. The direction of ice movement was to the south indicating that the Dwyka at that time dipped to the south. Gradual deglaciation of the continental area occurred resulting in coal-bearing fluviodeltaic sequences (Catuneanu et al. 1998).

Ecca Group
The Ecca Group was deposited in the Permian (Johnson). It is diachronous, young in the distal and older in the proximal (Catuneanu et al. 1998). Its aerial extent is equivalent to that of the Karoo basin and its thickness is 3000m in the proximal region (Catuneanu et al. 2005). The Ecca Group can be divided into the Prince Albert, Whitehill, Ripon, Collingham and Fort Brown formations in the proximal region and the Pietermaritzburg, Vryheid and Volkrust formations in the distal region (Catuneanu et al. 1998). Formation Prince Albert Whitehill Ripon Collingham Fort Brown Pietermaritzburg Vryheid Volkrust Facies Greenish-grey shale and graded silty layers Carbonaceous shale, chert bands and lenses Greywacke siltstone and shale (Bouma sequences) Alternating siltstone and shale with yellow tuff and volcanic ash Greenish- grey shale and sandstone Shale Sandstone, shale and coal beds Dark shale intercalated with fine grained sandstone Environment Deep water Deep water, pelagic and reducing Deep water and submarine fan Submarine fan, pelagic and aeolian Regressive shallow marine Deep marine Fluviodeltaic Deep to shallow marine

Beaufort Group
The Beaufort Group was fluvially deposited during the Permeo-Triassic (Catuneanu et al. 2005). The Beaufort Group can be subdivided into two subgroups: Adelaide and Tarkastad Subgroup. The Adelaide Subgroup represents the lower Beaufort Group and its proximal facies include Koonap, Middleton and Balfour formations with the Normandien Formation containing the distal facies. The base of the Adelaide Subgroup is diachronous but conformable in the proximal and distal regions. It also displays a facies contact indicating a transition from marine to non-marine environment. The Koonap Formation consists of greenish silty mudstones and sandstones that occurs in fining-upward cycles. The depositional environment changes from a high energy braided river systems grading upwards to a low energy meandering systems. The Middleton Formation consists of upward fining maroon and greenish-grey mudstones interbedded with sandstones which were deposited in a low energy meandering and lacustrine environment. There is an unconformable contact between the Middleton and Balfour Formations indicating a sudden transition from low energy meandering facies to high energy braided facies. The Balfour Formation consists of yellowish and bluish-greenish-grey sandstones which are interbedded with dark mudstones in an upward fining sequence. The environment they were deposited in changed from braided rivers to a meandering system. The

Normandien Formation consists of interbedded sandstones and mudstones which were deposited by meandering streams and channels surrounded by semiarid floodplains (Catuneanu et al. 1998). The Tarkastad Subgroup hosts the Katberg and Burgersdorp Formations in the proximal region and the Verkykerskop and Driekoppen Formations in the distal region. The Katberg Formation consists of thick, laterally extensive, light olive grey, coarse-grained sandstones which are made up of marcoforms. They were deposited in a shallow braided environment. The Burgersdorp Formation conformably overlies the Katberg Formation consists of thick, laterally inextensive, olive grey, fine to medium grained sandstones overlain by red maroon siltstones and mudstones. This represents a fining upward sequence that indicates a mixed-load meandering river and floodplain depositional environment. The Verkykerkop Formation consists of thin, laterally extensive, medium to finegrained sandstones made up of macroforms. The Driekoppen Formation consists of fine grained sandstones overlain by thick, massive to laminated siltstones and mudstones. These deposits indicate a suspended-load-dominated meandering river depostional system. There is a gradual transition between the conformable proximal and distal regions of the subgroup (Catuneanu et al. 1998).

Stormberg Group
The Lower Stormberg assemblage was deposited in the Late Triassic and the Upper Stormberg assemblage in the Early Jurassic (Johnson). The Stormberg Group can be divided into the Molteno, Elliot and Clarens Formations (Catuneanu et al. 1998).The Molteno Formation consists of tabular sheets of medium to coarse grained sandstones composed of macroforms. The sheets extend laterally and indicate deposition by a braided stream on a braidplain. Siltstones, mudstones and coal deposits can occur and the depositional environment include fills of abandoned channel tracks within ponded water on the braidplain. The Elliot Formation consists predominantly of red floodplain mudstones with lesser sandstones. These indicated deposition within mixed-load meandering systems. Wind-blown sediments also occur which was introduced to the system resulting in a loessic dust component. These wind-blown deposits indicate a transitions towards a desert dominated environment influencing the deposition of the Clarens Formation. The Clarens Formation consists of yellow finegrained sandstones, sandy siltstones and mudstones with lesser coarse-grained sandstones. Deposition of these sediments occurred in a desert environment influencing the formation of wind-blown dunes and shallow playa lakes in the colder surrounding environment. Climate became less harsh with time and wet desert processes became dominant. The gradational transition between the fluvial-dominated Elliot Formation and the Aeolian-dominated Clarens Formation can be seen as an one upwardcoarsening sequence (Catuneanu et al. 1998).

Conclusion

References
Burgess, P.M., Gurnis, M., Moresi, L. (1997). Formation of sequences in the cratonic interior of North America by interaction between mantle, eustatic, and stratigraphic processes. GSA Bulletin, 108 (12), 15151535. Catuneanu, O. (2004). Retroarc foreland systemsevolution through time. Journal of African Earth Sciences, 38, 225242. Catuneanu, O. and Hancoxt. P.J. (1998). Reciprocal flexural behavior and contrasting stratigraphies: a new basin development model for the Karoo retroarc foreland system, South Africa. Basin Research, 10, 417-439 Catuneanu, O., Hancox, P.J., Cairncross, B., Rubidge, B.S. (2002). Foredeep submarine fans and forebulge deltas: orogenic off-loading in the underfilled Karoo Basin. Journal of African Earth Sciences, 35, 489502. Catuneanu, O., Wopfner, H., Eriksson, P.G., Cairncross, B., Rubidge, B.S., Smith, R.M.H., Hancox, P.J. (2005). The Karoo Basins of south-central Africa. Journal of African Earth Sciences, 43, 211253. Cloetingh, S., Lankreijer, A., de Wit, M.J. and Martinez, I. (1992). Subsidence history analysis and forward modeling of the Cape and Karoo Supergroups. . In: de Wit, M.J., Ransome, I.G.D. (Eds.), Inversion Tectonics of the Cape Fold Belt, Karoo and Cretaceous Basins of Southern Africa. Balkema, Rotterdam, pp. 239-247. Cole, D.I. (1992). Evolution and development of the Karoo Basin. In: de Wit, M.J., Ransome, I.G.D. (Eds.), Inversion Tectonics of the Cape Fold Belt, Karoo and Cretaceous Basins of Southern Africa. Balkema, Rotterdam, pp. 87-96. Gurnis, M. (1992). Rapid continental subsidence following the initiation and evolution of subduction. Science, 255, 15561558. Holt, W.E., Stern, T.A. (1994). Subduction, platform subsidence and foreland thrust loading: the late tertiary development of Taranaki basin, New Zealand. Tectonics, 13, 10681092. Johnson, M.R., van Vuuren, C.J., Visser, J.N.J., Cole, D.I., Wickens, H. de V., Christie, A.D.M., Roberts, D.L., Bandl, G. (2006). Sedimentary Rocks of the Karoo Supergroup. In: Johnson, M.R., Anhaeusser,C.R. and Thomas, R.J. (Eds). The Geology of South Africa. Geological Society of South Africa, Johannesburg/Council of Geoscience, Pretoria, 461-499. Johnson, M.R. (1991). Sandstone petrography, provenance and plate tectonic setting in Gondwana context of the south-eastern Cape Karoo Basin. South African Journal of Geology, 94 (2/3), 137154. Lock, B.E.(1980). Flat-plate subduction and the Cape Fold Belt of South Africa. Geology, 8, 3539. Mitrovica, J.X., Beaumont, C., Jarvis, G.T. (1989). Tilting of Continental Interiors by the Dynamical Effects of Subduction. Tectonics 8, 5, 10791094.

Mpodozis, C., Kay, S.M. (1992). Late Paleozoic to Triassic evolution of the Gondwana margin: evidence from Chilean frontal cordilleran batholiths (28S to 31S). Geological Society of America Buletinl, 104, 9991014. Pysklywec, R.N., Mitrovica, J.X. (1999). The role of subduction-induced subsidence in the evolution of the Karoo Basin. Journal of Geology, 107, 155164. Sinclair, H.D., Allen, P.A. (1992). Vertical versus horizontal motions in the Alpine orogenic wedge: stratigraphic response in the foreland basin. Basin Research, 4, 215232. Smellie, J.L. (1981). A complete arc-trench system recognised in Gondwana sequences of the Antarctic Peninsula region. Geology Magazine, 118, 139159. Visser, J.N.J. (1992). Basin tectonics in southwestern Gondwana during the Carboniferous and Permian. In: de Wit, M.J., Ransome, I.G.D. (Eds.), Inversion Tectonics of the Cape Fold Belt, Karoo and Cretaceous Basins of Southern Africa. Balkema, Rotterdam, pp. 109115. Visser, J.N.J. (1992). Deposition of the early to Late Permian Whitehill Formation during a sea-level highstand in a juvenile foreland basin. South African Journal of Geology, 95, 181193.

Vous aimerez peut-être aussi