Vous êtes sur la page 1sur 46

Precambrian Research 100 (2000) 235280 www.elsevier.

com/locate/precamres

Reconnaissance sedimentology and hydrocarbon biomarkers of Ediacarian microbial mats and acritarchs, lower Ungoolya Group, Ocer Basin
K. Arouri *, P.J. Conaghan, M.R. Walter, G.C.O. Bischo , K. Grey
Department of Earth and Planetary Sciences, Macquarie University, Sydney, NSW 2109, Australia

Abstract Ediacarian sediments of the lower Ungoolya Group (~580565 Ma) in exploration-well Munta 1 of the Ocer Basin of South Australia accumulated towards the base of a ramp on the southeastern ank of an axial foredeep, mostly in relatively deep water hemipelagic and turbiditic environments, and at palaeoequatorial latitude. Rapid subsidence at ~580 Ma created a marine basin of 200300 m depth that was starved of sediment at the Munta 1 site until shortly after the Acraman Impact Event when the rst of four successive clastic slope-aprons prograded northwestward across the Munta 1 locality. The individual relief of these four slope-aprons averaged at least 175 m. An epiclastic silt-dominated prograding frontal slope progressively built up into warm surface waters above the thermo-/pycno-cline where accretion of shoalwater carbonate culminated in peritidal platform deposits and local evaporites. Renewed crustal subsidence and rapid marine transgression across the carbonate-platform caused the depositional system to founder, backstep southwestward on the structural ramp, and recommence progradation of a new slope-apron. These depositional systems formed during a prolonged period of arid climate that intensied in the period ~575570 Ma, and possibly again near the end of the preserved lower Ungoolya Group record in Munta 1. Chemostratigraphic excursions of d13C and d13C that occur at the stratigraphic base of each platform-carbonate org carb at the top of the slope-aprons coincide with the bathymetric intersection of the palaeo-pycnocline with the seaoor, suggesting that these excursions might at least partly result from an oceanic reservoir that was not isotopically homogeneous and was partitioned for long periods of time at this physical boundary. Re-deposited Acraman Impact ejecta occurs throughout at least 108 m of section in the sediments of the lowermost slope-apron. The reworked ejecta is of wind-borne origin in the hemipelagites of the lower slope-apron and of mass-ow origin in the turbidites and debrites of the upper slope-apron, and its presence above the 1810.5 m level in Munta 1 conrms an earlier prediction (based on comparative stratigraphy) as to the likely stratigraphic level at which the primary (but as yet undetected ) Acraman ejecta-layer should occur in Munta 1. The sediments contain well-preserved microbial mats. They occur variously as: (1) autochthonous mats within epiclastic silts in the peritidal platform carbonates, and in lower slopeapron deposits where they probably grew in water depths of ~180200 m, probably below the lowermost reaches of the photic zone; (2) allochthonous structurally dismembered mats within debris-ow deposits of the upper slopeapron that were sourced from upslope areas in the vicinity of the pycnocline; (3) tempestite deposits within the peritidal carbonates that contain large intraclasts of biolaminated siltstone; and (4) detrital fragments and small intraclasts of mat-bound substrate that occur most abundantly within low-concentration turbidites and pelagites of
* Corresponding author. Present address: Organic Geochemistry in Basin Analysis Group, University of Adelaide, SA 5005, Australia. Tel.: +61-8-8303-3870; fax: +61-8-8303-4347. E-mail address: karouri@geology.adelaide.edu.au ( K. Arouri) Deceased. 0301-9268/00/$ - see front matter Crown copyright 2000 Published by Elsevier Science B.V. All rights reserved. PII: S0 3 0 1- 9 2 68 ( 9 9 ) 0 00 7 6 -5

236

K. Arouri et al. / Precambrian Research 100 (2000) 235280

both the upper and lower slope-apron. There are also abundant acritarchs within ne-grained turbidites, hemipelagites and marls, and in relatively deep-water carbonates. Disruption, fragmentation, dislodgement, and removal of autochthonous mat material from the upper slope-apron in the vicinity of and above the pycnocline was accomplished by various mechanisms, both ambient and catastrophic. Episodic disturbance of the pycnocline by seiches was probably a major cause of disruption and downslope removal and re-deposition of mat material from the shallower parts of the photic zone on the upper slope-apron. Such disturbance caused slumping of in situ mats and generated slope-hugging debris-ows and low-concentration turbidity currents. These displaced abundant mat fragments to lower parts of the slope-apron, and trapped interows of suspension-load sediment containing abundant mat fragments at the pycnocline from where it spread laterally to generate long-continued rain of shallow-water mat kerogen to the distal slope-apron and deep-basin oor beyond. The kerogen of these mats occurs as either structurally discrete organic tissue or as gel-like homogeneous sheets, both with and without discernible internal structures. The organic tissue consists of laments and dense clusters of smooth spheres (~1 mm diameter) interpreted as fossil coccoid microbes. Very nely crystalline carbonate, intimately associated with the coccoids, probably formed by microbial mediation in peritidal environments characterised by warm to elevated water temperatures. Biomarker distributions of bitumen extracted from the Munta 1 microbial mats are characterised by abundant monomethyl and dimethyl alkanes of variable carbon-chain length and isomer distributions, indicating a predominantly cyanobacterial input, probably of diverse species populations at stratigraphically dierent levels. These cyanobacterial mats alternate with non-cyanobacterial microbial mats that probably included anoxygenic photosynthetic green non-sulfur bacteria, with minor terminal anaerobic sulfate-reducers and Archaea. Some turbidite-siltstone samples, containing only allochthonous mat detritus, have a cyanobacterial biomarker signature that is strongly overprinted by other bacterial signals, including those of terminal consumers. Abundant algae (acritarchs) are present in many palynological preparations throughout the section and show marked upward increase in taxonomic diversity. Sterane biomarker distributions are consistent with chlorophytes, rhodophytes, prymnesiophytes and/or dinoagellates, or their precursors. The association of such diverse algal sterane biomarkers with abundant acanthomorph acritarchs suggests a close genetic relationship and demonstrates the great diversity of planktonic algae during the terminal Proterozoic. Crown copyright 2000 Published by Elsevier Science B.V. All rights reserved.
Keywords: Acraman Impact ejecta; Acritarchs; Biomarkers; Cyanobacteria; Lithofacies analysis; Microbial mats; Neoproterozoic; Ocer Basin; Slope-aprons

1. Introduction 1.1. Preamble This report documents the occurrence and sedimentological and geochemical aspects of Ediacarian organic mats in the Munta 1 drillcore from the eastern Ocer Basin, Centralian Superbasin (Figs. 13). The presence of fossil mats is indicated by lms of kerogen and sheet-like fragments of organic matter. Microbial mats and their fragments have been recorded from drillcores in both the Ocer and Amadeus Basins of central Australia ( Fig. 1; e.g. Calver, 1995; Grey, 1998; Cotter, 1999; Logan et al., 1999). Mat material is present in marine, rhythmically laminated and graded-bedded, calcareous mudstone and dolomitic siltstone and sandstone that comprise successions up to several hundreds metres thick, as well

as in shallow-marine ( lagoonal/peritidal ) carbonates. In addition to microbial mat fragments, these rocks also contain well-preserved acritarchs (Zang and Walter, 1992; Grey, 1998). Two groups of acritarchs have been recorded: the lower part of the Ungoolya Group in the Ocer Basin ( Fig. 2; Table 1) is dominated by small, simple sphaeromorph acritarchs, mainly attributable to species of Leiosphaeridia, whereas the remainder of the lower Ungoolya Group is dominated by large, complex acanthomorph acritarchs [Simple Leiosphere Palynoora and Ediacarian Complex Acritarch Palynoora respectively of Grey (1998); Fig. 3 herein]. 1.2. Organic mats and their molecular fossils Like most modern mats, the majority of Proterozoic mats are interpreted to have formed

K. Arouri et al. / Precambrian Research 100 (2000) 235280

237

Fig. 1. Location of exploration-well (Comalco) Munta 1, eastern Ocer Basin, South Australia (SA), and geographic context of the Ocer Basin as part of the Centralian Superbasin. Diagram adapted from Walter et al. (1995, g. 1) with additions from Hill et al. (2000, g. 1). AR=Ammaroodinna Ridge; BS=Birksgate Sub-basin; MP=Murnaroo Platform; MT=Munyarai Trough.

in marine or hypersaline depositional environments (e.g. Walter et al., 1992). Both extant and fossil mats contain a diversity of taxa, generally dominated by bacteria (mainly cyanobacteria or other photosynthetic bacteria). Other mats are formed predominantly by algae. Modern examples of submerged cyanobacterial mats, probably similar to Proterozoic shallow-marine benthic microbial mats, include those at Solar Lake, Sinai (DAmelio et al., 1989) and at Guerrero Negro saltern, Baja California Sur, Mexico (Des Marais et al., 1992b). Mat surfaces are built essentially by the lamentous cyanobacterium Microcoleus chthonoplastes; overlain by unicellular, mucus-producing, cyanobacteria (commonly Synechococcus sp.) and diatoms, and underlain (i.e. in the chemocline where oxygen and sulde coexist) by Chloroexus sp. (green anoxygenic photosynthetic non-sulfur bacteria) and Beggiatoa sp. (colourless sulfur bacteria), and sparser non-photosynthetic bacteria and fragments of bacterial mucilage (DAmelio et al., 1989; Des Marais et al., 1992b). Such mats are widely distributed because they can grow at variable oxygen and sulde levels ( Ward

et al., 1992) and salinities, including that of normal seawater (Bauld, 1984). Marine (sabkha/upper-intertidal ) cyanobacterial mats are typied by those from Abu Dhabi ( United Arab Emirates) and Shark Bay ( Western Australia), built mainly by M. chthonoplastes and Lyngbya aestuarii, that overlie laminae of lamentous bacteria and purple sulfur bacteria (Golubic, 1976; Kenig et al., 1995). At higher salinities, unicellular cyanobacteria predominate and form a more gelatinous, transparent, and dispersed mat that normally retards oxygenic photosynthesis (e.g. Des Marais et al., 1992b). M. chthonoplastes mats, like other cyanobacterial mats, are oxygenic but, less commonly, can intermittently shift from oxygenic to anoxygenic photosynthesis. In contrast, mats built by purely anoxygenic photosynthetic bacteria are relatively rare (e.g. Ward et al., 1992). Bacterial mats are characterised by specic biomarkers such as hopanoids and methylalkanes, whereas eucaryotic mats are distinguished by abundant steranes and other compounds. Short-chain (C C ) monomethylalkanes (MMAs), detected 15 20

238

K. Arouri et al. / Precambrian Research 100 (2000) 235280

Fig. 2. Simplied Neoproterozoic to Early Cambrian stratigraphy of: (1) the eastern Ocer Basin in South Australia, after Lindsay and Leven (1996, g. 2); sea-level curve and sequence-stratigraphic framework are from Moussavi-Harami and Gravestock (1995, g. 2); and (2) the Amadeus Basin of the Northern Territory (NT ), from Grey and Cotter (1996); absolute time-scale is from Walter et al. (2000), except for the 650 Ma age of the base of the Tarlina Sandstone, which is from Morton (1997b, g. 6.2).

in most cyanobacterial cultures [see reviews by Shiea et al. (1990) and Kenig et al. (1995)], are probable cyanobacterial biomarkers. Of the MMA isomers, the 7- and 8-methylheptadecanes, and the less abundant isomers (4- and 6-methyl ), have been reported from cyanobacterial mats of recent shallow-marine (typically sabkha) sediments (e.g. Boon et al., 1985; de Leeuw et al., 1985; Kenig et al., 1995), and of hot-springs (Robinson and Eglinton, 1990; Shiea et al., 1990, and references cited therein). However, according to Shiea et al. (1991, p. 228), these compounds have not been

found in microbial mats which lack cyanobacteria [n]or [have they been found ] in open marine sediments. Hence, cyanobacteria appear to be the only group of microorganisms capable of producing these branched alkanes (Shiea et al., 1990). However, controversy exists regarding the origin of these MMAs in ancient sediments [see review by Shiea et al. (1990)]. Fowler and Douglas (1987), for example, suggest an unspecied procaryotic origin; others argue for a diagenetic origin via: (1) the transformation of functionalised lipid precursors (Summons et al., 1988b); (2) long-term

K. Arouri et al. / Precambrian Research 100 (2000) 235280

239

equilibration (Hoering, 1981; Klomp, 1986); or (3) acid-catalysis of alkenes formed by thermal cracking ( Kissin, 1987). Several isomers of C dimethylalkanes (DMAs) 19 have been identied in Phormidium luridum cyanobacterial cultures (Summons et al., 1998), in the Gavish Sabkha cyanobacterial mats (de Leeuw et al., 1985), and in hot-spring cyanobacterial mats (Shiea et al., 1990; Robinson and Eglinton, 1990). Longer-chain (C ) DMAs have been reported 21+ from the cyanobacterium Spirulina platensis (Paoletti et al., 1976) and from Abu Dhabi cyanobacterial mats along with other long-chain MMAs ( Kenig et al., 1995). The presence of certain branched alkanes in oils and sediments as old as the Proterozoic (Hoering, 1981; Jackson et al., 1986; Klomp, 1986; Fowler and Douglas, 1987; Summons, 1987; Summons et al., 1988a,b) implies a direct cyanobacterial contribution (Robinson and Eglinton, 1990; Shiea et al., 1990). However, cyanobacteria may not have been the only source of these compounds in ancient materials because they generally have carbon chains (C C ) longer 15 33 than those found in living and modern cyanobacteria (C C ), in some instances with dierent 15 21 odd- or even-carbon-number preferences ( Kenig et al., 1995, and references cited therein). Such dierences could reect compositional variability of cyanobacterial populations (e.g. Robinson and Eglinton, 1990). Accordingly, additional or alternative sources have been suggested, which include other organisms such as insects (e.g. Nelson, 1978; Kenig et al., 1995), by derivation from functionalised lipid precursors (Summons, 1987; Summons et al., 1988b), or as diagenetic products from alkanes (Hoering, 1981) and alkenes ( Kissin, 1987).

2. Stratigraphic context, methods, and materials studied 2.1. Regional setting and basin-margin location context of Munta 1 The Ocer Basin is a complex, polyphase basin (Lindsay and Leven, 1996, p. 421). It has a westeast extent of ~1400 km within Western

Australia and South Australia, occupies an area of at least 450 000 km2 (Morton, 1997a; Perincek, 1998), and is the largest and southernmost structural remnant of its mid-Neoproterozoic giant sagbasin precursor: the 840500 Ma Centralian Superbasin ( Fig. 1; Walter et al., 1995). Tectonic dismemberment of the precursor superbasin into smaller, interconnected, structural basins commenced at about the beginning of the Ediacarian (~595 Ma) with the emergence of the meridional Musgrave Block and consequent partitioning of the central part of the superbasin into the Amadeus Basin to the north and the Ocer Basin to the south ( Fig. 1; Lindsay and Leven, 1996; Walter and Veevers, 1997). Thereafter (through to the end of the Devonian), the Ocer Basin took the asymmetric form of a foreland basin, comprising an arcuate foredeep in the north closely paralleling the southern boundary of the Musgrave Block, and, in the south, beyond a gentle ramp, an adjoining shallow platform across the Archaean and Palaeoproterozoic Gawler and Yilgarn Cratons ( Fig. 1; Lindsay and Leven, 1996). Palaeomagnetic studies of late Neoproterozoic formations of the Adelaide Rift Complex of South Australia, coeval to those of the nearby Ocer Basin, indicate that the region occupied low to equatorial palaeolatitudes during this time (Schmidt and Williams, 1996). The Munta 1 well is located within the South Australia part of the eastern Ocer Basin where its foredeep or structural axis comprises the Birksgate Sub-basin in the west and the southwest northeast-trending Munyarai Trough in the east ( Fig. 1). These sub-basins are bordered on the south by the structural ramp that leads up to the Murnaroo Platform, the broad region of shallower basin cover that thins to the south and southeast over the Gawler Craton. Munta 1 occupies a structurally intermediate position towards the base of this ramp 35 km southeast of the axis of the Munyarai Trough as dened by depth to basement (Gravestock, 1997, g. 5.3; Calver and Lindsay, 1998, g. 2). To the west of Munta 1 (i.e. in the area of the eastern Birksgate Sub-basin) the structural ramp leading up to the Murnaroo Platform is wide and very gentle (0.3), but along strike to the northeast it steepens progressively through the

240

Table 1 Overview of the Neoproterozoic succession penetrated by Munta 1, Ocer Basin, South Australia. Compiled from descriptions in Calver (1995), Morton (1997b), Calver and Lindsay (1998), and Calver and Grey (1993). Informal lower/upper subdivision of Ungoolya Group follows Lindsay and Leven (1996) Stratigraphic thickness (m) 900 Lower and Middle Cambrian succession. Not relevant herein. Lithological characteristics Inferred depositional environments

Stratigraphic unit

Boundary depths (m)

MARLA GROUP

0900

UNGOOLYA GROUP 252 Divisible into three informal units in Munta 1 [see Calver (1995), p. 139, g. 3.23]. Upper unit (900 943 m): grey, ne-grained limestone. Middle unit (943 997 m): thin diuse limestone beds in dark-grey calcareous mudstone. Lower unit (9971152 m): thick succession of dark greygreen silty shale with minor (<10%) thin beds or laminae of paler calcareous siltstone or very ne-grained sandstone; many of these layers are internally plane-laminated, ripple-cross-laminated or slumped; thin planar siltstone laminae in many places exhibit a bundled distribution. Thick, intercalated succession of grey and red thinly laminated siltstone/mudstone and medium- to thick-bedded ne-grained limestones containing thin interbeds of dark-grey mudstone, typically with diuse/gradational bedding-plane contacts; mudrock intervals also commonly contain thin, and less commonly thick (0.5 m) graded and planar-parallel-laminated beds of siltstone and sandstone with rare low-angle, possibly hummocky, cross-lamination; limestones commonly feature stylobedding and the mudrock intervals above ca 1300 m show much disturbed bedding (see Fig. 3) and low- and high-angle syndepositional erosion surfaces against which overlying thinly bedded graded and laminated very-ne sandstones and siltstones terminate laterally with slight stratigraphic drape [see Calver (1995), pl. 3.3; Morton (1997b), photo 42411c, p. 63]; medium- to thick-bedded ne-grained limestone between 1438 and 1448 m contains black shale as sporadic, thinly-laminated layers and wispy intraclasts, interpreted as benthic microbial mats. Also present in this limestone unit is a 40 mm bed of at-pebble breccia containing a 1020 mm anhydrite nodule (1441.33 m) [which] has a mudstone coating with a rounded outer prole (Calver, 1995, pp. 132, 134) Deep, sub-wave-base, outer shelf canyon environment. Regionally the unit comprises two canyon-ll successions, the lower of which consists of debris-ow deposits grading to tidal-at deposits at the top. The upper (turbiditic) succession was deposited on a transgressive to highstand deeper-water submarine-fan, shallowing to marine shelf limestone and mudstone.

upper

Narana Fm.

9001152

K. Arouri et al. / Precambrian Research 100 (2000) 235280

lower

Tanana Fm.

11521614

462

A fractured, intraclastic limestone at the top of the formation contains intraclasts, and may have been formed in intertidal to upper subtidal shelf settings (Morton, 1997b, pp. 6364). The upper (above ca 300 m) graded and laminated siltsone/mudstone and very-ne graded sandstone interval with syndepositional erosion surfaces and disturbed bedding is suggestive of a deeper-water, perhaps pro-delta or basinslope turbiditic setting. Tepee-like structures are present in red mudstone associated with the carbonate units on structural highs, suggesting shallow water and subaerial conditions [see Calver and Lindsay (1998), pp. 522523].

Karlaya 16141675 Limestone

61

Divisible into upper and lower informal units separated by an intervening calcirudite. Upper unit (1614 1651.2 m): medium- to thick-bedded micritic limestone with common (ca 3040%) dark-grey shale interbeds with diuse/gradational bedding-plane contacts. Marker unit (1651.21651.7 m): intraclastic limestone conglomerate. Lower unit (1651.71675 m): thin- to medium-bedded micritic limestone with common (ca 30%) thin dark-grey shale/marl interbeds, commonly anastomosing to form nodular-bedded limestone [see Calver (1995), pl. 3.2 and p. 131] in which the interbeds have diuse/gradational contacts; ne- to medium-grained quartzose sandstone at base. Divisible into two informal units. Upper unit (1675 1850 m): predominantly dark greengrey, mostly nely (and, at some levels, rhythmically) laminated dolomitic or calcareous siltstone and minor ne sandstone, in some cases with starved ripples; a 0.8 m thick green mudstone bed at 1810.21811.0 m contains a 15 cm thick organic-rich black shale. Lower unit (1850 1973 m): redbrown, thinly laminated to massive mudstone; locally conglomeratic at base. Very coarse quartzo-feldspathic sandstone, ne-pebble feldspathic conglomerate, and red and green shale interbeds (upper 22 m); underlain by 83 m of ne to medium quartzose, cross-bedded sandstones (containing sporadic mudstone intraclasts), rare cosets of ripple-cross-laminated ne sandstone, and (below 2047 m) red siltstone. Upper unit: marine pro-delta, turbidite-dominated. Lower unit: possibly uvial in base, passing up to prodelta and delta-front environments subject to accumulation of rhythmites or tempestites.

Marine subtidalneritic, probably deeper than fairweather wave-base.

Dey 16751973 Dey Mudstone

298

K. Arouri et al. / Precambrian Research 100 (2000) 235280

Murnaroo Fm.

19732078

105 (partial thickness; base not reached )

Locally aeolian in southwestern parts of basin, with distal alluvial-fan, sheet-outwash, and channeliseduvial and marine tidal environments progressively to the north.

241

242

K. Arouri et al. / Precambrian Research 100 (2000) 235280

Fig. 3. Neoproterozoic stratigraphy of (Comalco) Munta 1, eastern Ocer Basin, South Australia [from Morton (1997b), but with modications and additions based on data in Calver and Grey (1993) and the present study] showing also the number and stratigraphic location of the petrographic and biomarker analyses undertaken in the present study, and the results thereof. Drillcore colour

K. Arouri et al. / Precambrian Research 100 (2000) 235280

243

(simplied) is from Calver and Grey (1993) and Cucuzza (1987). Lithofacies codes are those of Pickering et al. (1989); lithofacies specications in parentheses are inferred (in intervals of no sample control ) from drillcore descriptions of Calver and Grey (1993). Palynoora zones are from Grey (1998, g. 6.6). Absolute time-scale is from Walter et al. (2000).

244

K. Arouri et al. / Precambrian Research 100 (2000) 235280

area of Munta 1 (Gravestock and Morton, 1997, g. 11.1) and where it borders the central and eastern parts of the Murnaroo Trough it is much narrower and culminates upslope in a northeasterly trending basement extension of the Murnaroo Platform called the Ammaroodinna Ridge (Lindsay and Leven, 1996; Gravestock, 1997). Lindsay and Leven (1996) divide the basin-ll into basin phase megasequences (M1, M2, etc.) bounded by erosion surfaces ( Fig. 2), stating (p. 410) that each megasequence represents subsidence and accumulation of sediments during a major tectonic event. Seismic sequence boundaries within the megasequences have distinctive stacking patterns that provide the basis of the sealevel curve shown in Fig. 2. Megasequence M1 (Callanna Group correlate of the Adelaide Rift Complex; ~840760 Ma) represents the sag-basin ( Willouran) phase of sedimentation. It is widespread and relatively uniform in thickness (200300 m) and was deposited in continental and shallow-marine (including evaporitic) environments. Rocks of Megasequence M2 (~700565 Ma) overlie a major erosion surface cut into Megasequence M1 and are bounded upwards by a regional erosion surface characterised by deep palaeo-canyons cut into and subdividing the Ungoolya Group into lower and upper sections ( Fig. 2; Lindsay and Leven, 1996). Throughout most of the Ocer Basin the erosion surface separating Megasequences M1 and M2 represents a time interval of ~100 m.y. and separates the early sag-basin phase of basin development from a subsequent compressional phase that witnessed the partitioning of the sag basin into the complex of smaller structural basins. In addition to sporadic Sturtian (~700690 Ma) glacigenic sediments at its base, Megasequence M2 comprises two major Ediacarian intervals of dierent stratigraphic and lithological character (Lindsay and Leven, 1996). The lower interval (Lake Maurice Group; Fig. 2) consists of a marine mudrock wedge (Meramangye Formation) enveloped by coarser marine and non-marine clastics ( Tarlina Formation below and Murnaroo Formation above) that collectively record renewed subsidence, and the establishment of the axial foredeep in the north where the thickness of the interval exceeds

1000 m. Isopachs of the Lake Maurice Group show a pattern of northwestward thickening across the Murnaroo Platform and marginal ramp toward and into the Munyarai Trough (Moussavi-Harami and Gravestock, 1995, g. 3). Seismic reection proles of this interval are said to show progradation and downlap in the same (i.e. basinward) direction (Lindsay and Leven, 1996), indicating sediment provenance from the craton to the south and southeast. The upper interval of Megasequence M2 comprises the marine lower Ungoolya Group (Dey Dey Mudstone to Munyarai Formation succession; Fig. 2), the focus of the present study (except for the Munyarai Formation, which is absent in Munta 1). The lower Ungoolya Group consists of intercalations of thick mudrock (mainly siltstone) and thin ne-grained carbonates ( Fig. 3). According to Calver and Lindsay (1998, p. 521, gs. 8 and 10a), this interval forms a relatively simple wedge that onlaps the Murnaroo Platform ramp to the southeast and thickens northward into the Munyarai Trough where it reaches a thickness of at least 4300 m. Lindsay and Leven (1996) and Calver and Lindsay (1998; g. 7) interpret the lower Ungoolya Group succession to comprise ve stacked basinward-prograding cyclothems in which the thick siltstones represent shallowingupward highstand systems tracts deposits and the thin carbonates represent the deposits of transgressive systems tracts. They interpret basinward prograding wedges that occupy the axial part of the Munyarai Trough and pinch out upslope on the ramp to the southeast as the deposits of lowstand systems tracts. As with the lower (Lake Maurice Group) interval of Megasequence M2, isopach patterns, facies architecture, and sequencestratigraphic evidence indicate that sediment inux during accumulation of the lower Ungoolya Group was from the south, with the Murnaroo Platform largely being bypassed (Lindsay and Leven, 1996, p. 420; Calver and Lindsay, 1998). Northsouth compression associated with the evolution of the foreland basin initiated high-angle reverse faults and southward-directed thrusts within the basinll on the Murnaroo Platform ramp, mobilised salt from the underlying sag-basin stage sediments (Alinya Formation), and produced localised

K. Arouri et al. / Precambrian Research 100 (2000) 235280

245

topographic/bathymetric (structural ) highs and lows during the period of accumulation of the lower Ungoolya Group, including a large area east of and bordering on Munta 1 (Lindsay and Leven, 1996, p. 415, g. 15b; see also Gravestock and Morton, 1997). Megasequence 3 comprises the late Ediacarian upper Ungoolya Group (Narana Formation) that overlies the canyoned, regional erosion surface and completes the Neoproterozoic succession in the Ocer Basin (Fig. 2). 2.2. Stratigraphy and cyclicity of the Neoproterozoic succession in Munta 1 The Neoproterozoic section in Munta 1 extends downwards from the late Ediacarian Narana Formation to the early Ediacarian Murnaroo Formation (Figs. 2 and 3; Table 1). Except for the Murnaroo Formation, all the Neoproterozoic formations intersected in Munta 1 are subdivisions of the Ungoolya Group (formerly known as the Rodda Beds). In upward stratigraphic order these are the Dey Dey Mudstone, Karlaya Limestone, Tanana Formation [all referred to the informal lower Ungoolya Group of Lindsay and Leven (1996)], and Narana Formation [referred to the Lindsay and Leven (1996) informal upper Ungoolya Group]. Regional stratigraphic studies (e.g. Sukanta et al., 1991; Lindsay and Leven, 1996; Morton, 1997b) show that the upper part of the lower Ungoolya Group, including the upper part of the Tanana Formation, has been erosionally removed from the Munta 1 location by the late Ediacarian regional canyon-cutting event. In Munta 1 this is marked by an unconformity overlain by the canyon-ll deposits of the Narana Formation. All underlying formations penetrated in Munta 1 are conformable. The Munta 1 succession is a key section through the lower Ungoolya Group and is described by Cucuzza (1987), Calver (1995) and Morton (1997b). Chemostratigraphy (Calver, 1995; Walter et al., 1995; Calver and Lindsay, 1998) and palynostratigraphy (Grey, 1998) have been documented for the Ungoolya Group succession and underlying Murnaroo Formation in Munta 1. The Narana Formation is predominantly barren of palynomorphs, most

probably as a result of high-energy conditions during deposition (Grey, 1998). A sequence-stratigraphic framework and inferred sea-level curve for the Ocer Basin Neoproterozoic succession in South Australia are illustrated in Fig. 2. A summary of the Munta 1 Neoproterozoic succession, logged by Cucuzza (1987) and Calver and Grey (1993), including predominant rock types of each formation and postulated depositional environments, is given in Table 1. A graphic log of the 821 m thick lower Ungoolya Group interval in Munta 1, the focus of the present study, is illustrated in Fig. 3. Sample coverage of the lower Ungoolya Group for the present petrographic and biomarker study is uneven among its individual formations (Fig. 3). The majority of the samples come from the Tanana Formation (22 thin-sections; 13 biomarker analyses). Neither the Narana nor Murnaroo Formations has been studied here petrographically or geochemically. The Dey Dey Mudstone conformably overlies the Murnaroo Formation in Munta 1, and comprises 298 m of section divisible into two units ( Table 1; Fig. 3). The basal 123 m consists of uniform redbrown slightly ssile mudstone. The thicker (175 m) upper unit consists predominantly of dark greengrey, mostly parallel-laminated dolomitic or calcareous siltstone, and rhythmically interlaminated very ne sandstone and siltstone with rare starved ripples and disturbed (slumped) bedding. A 15 cm thick organic-rich black shale interval occurs at 1810.5 m towards the base of the upper unit. The overlying 61 m thick Karlaya Limestone is also divisible into two units across a 0.5 m thick intraclastic calcirudite [see Calver and Grey (1993); Table 1; Fig. 3]. The 23.3 m thick lower unit comprises mainly thin- to medium-bedded micritic limestone with common, thin dark-grey shale or marl interbeds that typically have diuse boundaries with the limestone [Fig. 4(D)] and an irregular anastomosing (nodular-bedded) geometry (Calver, 1995, plate 3.2; Morton, 1997b, photo. 42411b). The 37.2 m thick upper unit consists of medium- to thick-bedded micritic limestone, also with common dark-grey shale/marl interbeds with diuse bedding-plane contacts with the limestone host and a 0.7 m thick interval of in situ [sedi-

246

K. Arouri et al. / Precambrian Research 100 (2000) 235280

K. Arouri et al. / Precambrian Research 100 (2000) 235280

247

mentary?] breccia at 1605.31606.0 m (Calver and Grey, 1993; Fig. 3). The overlying Tanana Formation is the thickest (462 m) of the lower Ungoolya Group units intersected by Munta 1 ( Fig. 3; Table 1). In the Munta 1 area it sustained relatively little denudation as a consequence of the late Ediacarian regional canyon-cutting event (Calver, 1995, pp. 128130), and hence is more intact than in wells located elsewhere in structurally higher positions on the Murnaroo Platform to the south (Calver, 1995, g. 3.19). The formation comprises alternations of two major lithological associations: (1) hectometre-scale intervals of marly mudrock, siltstone, and minor ne sandstone (i.e. a predominantly epiclastic association); and (2) decametrescale intervals of micritic limestone containing thin interbeds of siltstone and/or marl (i.e. a predominantly chemogenic association). The latter micritic carbonate intervals are said to be macroscopically similar to the underlying Karlaya Limestone

(Calver, 1995, pp. 131132; Calver and Lindsay, 1998, p. 521). Each hectometre-scale epiclastic interval coarsens upward subtly from predominantly drab-coloured (dark-grey/greygreen) mudrock or marly mudrock in its lower half or more (and from silty drab marl and drab limestone in the case of the lowermost interval; Fig. 3), into thin-bedded siltstone and minor ne sandstone in its topmost part. Intercalations of both drab- and red-coloured sediment are characteristic of the upper (coarser) part of the epiclastic intervals. In cyclic fashion, the epiclastic interval is then followed by a decametrescale chemogenic interval ( Fig. 3), variably paleor mid-grey, pinkish, or pale-red, and locally associated with dark-red mudstone and in situ breccia (e.g. at 1160 m; Calver and Grey, 1993; Fig. 3). The siltstone and very ne sandstone typically occur in very-thin/thin graded and parallellaminated beds with sporadic trains of starved ripples and rare ?oscillation ripples. Rather thick

Fig. 4. Thin-section photomicrographs of lower Ungoolya Group core samples from Munta 1, Ocer Basin, South Australia. Sedimentary facies codes are those of Pickering et al. (1989). Carbonate phases in images (E ), ( F ), and (H ) show darkish tones because they are of stained thin-sections. Images (G) and (H ) photographed in cross-polarised light; all others in plane-polarised light. Stratigraphic younging direction is upward in all images. (A) Very-thinly bedded, graded and rhythmically laminated, medium to ne siltstone containing abundant large, bedding-aligned, detrital sheets (arrowed ) and smaller akes of microbial mat; contact at mid-height of image separates the graded and rhythmically laminated top part of one bed and the graded basal part of another. Tanana Fm., 1324.8 m, Facies D2.3 (upper slope-apron, cycle 3), Microbial Facies 1A. (B) Poorly sorted, coarse to medium, clayey structureless silt with sporadic, crudely dened planar-parallel laminae (arrowed ), and containing very abundant, large (up to 1 2 mm long) detrital sheets and smaller akes of microbial mat (not conspicuous at this magnication); note crude upward sizegrading manifested by increasingly darker tone. Tanana Fm., 1234.0 m, Facies D1.1 (upper slope-apron, cycle 4), Microbial Facies 2. (C ) Very thinly laminated marl characterised by varve-like, planar-parallel, semi-discontinuous laminae of coarse-to-ne siliciclastic silt ( light-toned layers; probably resulting from episodic/?seasonal aeolian inux), and intervening ner silty dolomitic marl (darkertoned layers); abundant small detrital akes of microbial mat are present, but not resolvable at this magnication. Tanana Fm, 1561.9 m, Facies E2.2 ( lower slope-apron, cycle 2). (D) Kerogen stylocumulate concentrated from abundant, background dispersed organic matter (DOM ) in rock due to the pressure-dissolution of: (1) ne-grained limestone ( lighter-toned, lower four-fths of image), and (2) dolomitic marl (darker-toned, top one-fth of image); Karlaya Limestone, 1670.0 m, Facies E2.2 (upper, carbonate slope-apron, cycle 1). ( E ) Geometrically irregular, thin (<1 mm), varve-like couplets of coarsene/very-ne epiclastic siltstone containing autochthonous microbial mats that increase upward in degree of development within each varvic couplet from the coarse into the ne/very-ne silt, commonly capped in turn by a veneer of clay (thin, dark layers in basal two-thirds of image); 1438.3 m (from peritidal carbonate complex, extreme top of slope-apron cycle 2), Tanana Fm., Microbial Facies 2. ( F ) Mixed siliciclastic calcisilt with thin epiclastic silt laminae (arrowed ) containing autochthonous microbial mats and concentrations of authigenic pyrite; 1443.6 m (from peritidal carbonate complex, top of slope-apron cycle 2), Tanana Fm., Microbial Facies 2. (G) Very thinly and irregularly laminated ne epiclastic siltstone ( light-toned laminae/lenses), and very-ne, micaceous and clayey siltstone (grey-toned laminae/lenses) containing autochthonous microbial mats; black layers are zones of intense authigenic pyrite development. Tanana Fm., 1443.6 m (from peritidal carbonate complex, top of slope-apron cycle 2), Microbial Facies 2. (H ) Peloidal, intraclastic calcirudite (upper, wide part of sample) and laminated clayey-siltstone ( lower, narrow part of sample) mutual contact dened by arrows. The clayey-siltstone contains autochthonous microbial mats, and displacive anhydrite (white areas); 1441.33 m (from tempestite deposit, peritidal carbonate complex at top of slope-apron cycle 2), Tanana Fm., Microbial Facies 1A.

248

K. Arouri et al. / Precambrian Research 100 (2000) 235280

(0.5 m) graded ne sandstone beds within the upward-coarsening and -bed-thickening interval near the top of the formation (12501190 m) contain, or are intimately associated with, low-angle cross-lamination in hummocky bedforms (Calver and Grey, 1993; Fig. 3). The same upward increase in bed-thickness evidently accompanies the coarsening-upward pattern in the other epiclastic hectometre-scale intervals of the Tanana Formation, but this is subtle and not well documented. Above the 1300 m level in the uppermost of the epiclastic intervals, disturbed (slumped) bedding and prominent ( locally overhanging, and probably gullylike) penecontemporaneous erosion surfaces, some with at least decimetre-scale relief, are common (Calver, 1995, pl. 3.3; Morton, 1997b, photo. 42411c; Fig. 3 herein). Thin graded siltstone beds terminate laterally against these erosion surfaces. The decametre-scale chemogenic intervals are poorly documented but appear to comprise medium- to thick-bedded micritic limestone with sporadic decimetre-scale beds of calcarenite, calcirudite (containing sporadic intraclasts of detrital anhydrite Table 1; Section 3.1.3.3), and in situ [sedimentary?] breccia. Intercalated laminasets and thin interbeds of epiclastic siltstone and ne sandstone also occur. Sporadic cross-lamination is present in the epiclastic siltstone and ne sandstone and also in the calcarenites (Fig. 3). Some of these carbonate units are reported to contain tepee-like structures in red mudstone-rich intervals [Sukanta, 1993 cited in Calver and Lindsay (1998), p. 522]. The cyclic pattern of alternating upward-coarsening hectometre-scale ne epiclastics and decametre-scale micritic carbonates present in the Tanana Formation is also evident in the underlying Dey Dey Mudstone to Karlaya Limestone interval (Fig. 3). Thus, four such cycles are recognizable within the lower Ungoolya Group. These four separate hectometre-scale sedimentary cycles are numbered in upward order of stratigraphic succession in Fig. 3. They broadly correspond to the Calver and Lindsay (1998) sequence-stratigraphic sequences LU1LU4 with the qualications that: (1) Calver and Lindsay (1998) place their sequence boundaries at the base of each of the decametre-scale carbonate intervals, whereas the

cycle boundary in the present scheme is placed at the top these carbonates; (2) unlike sequence LU1 (which begins at the base of the Dey Dey Mudstone), the base of cycle 1 in the present scheme is placed at the upsequence red-to-drab colour change at 1850 m. Elaboration of the rationale for these four cycles and their boundaries is given in later sections. 2.3. Sampling and methodology Core samples were collected from the Ungoolya Group in Munta 1 by Kath Grey and Clive Calver in 1993 for palynological, isotopic and geochemical studies (Calver, 1995; Grey, 1998). Supplementary samples were collected by K. Arouri in 1997. The initial samples collected by Grey and Calver mainly comprised 2550 g of quartered core. Palynological and isotope chemostratigraphic preparations were carried out on splits of each sample, ensuring that the sample subsplits were stratigraphically equivalent. For palynology, the darker horizons were selected for processing using HCl/HF maceration techniques as described in Grey (1998). For the purpose of the present petrological and biomarker study, a subset of 27 samples was selected from the interval 1156.351810.50 m, of which 13 representative samples were analysed for their hydrocarbon biomarkers ( Fig. 3). Samples that contained light- and dark-coloured lithofacies (or lithofacies enclaves: e.g. dark-coloured intraclasts) were, in most cases, separately analysed for each of these two phases. However, sample bias was still inevitable because subsets of the core samples that were removed for palynological preparations prior to the biomarker analyses preferentially targeted the potentially most organic-rich material (i.e. the darker-coloured phases of the rock samples). As a consequence, the biomarker analyses cannot necessarily be expected to be representative of precisely the same organic matter present in the palynological yields and, therefore, some discrepancies can be expected between these two data-sets. Thin-sections were made of 27 core samples for petrographic analysis by transmitted-light and Hg-arc epiuorescence microscopy. Most were

K. Arouri et al. / Precambrian Research 100 (2000) 235280

249

stained with potassium ferricyanide and alizarin red-S to dierentiate carbonates. Clayey samples were cut and polished using kerosene and then washed in white spirit and heated on a hot-plate to ensure removal of the kerosene before coverslipping. A few polished-sections were prepared for study under reected white-light and epiuorescence excitation. Epiuorescence microscopy was undertaken using a high-pressure Hg-arc lamp, and excitation lters that generated excitation wavelengths of 365 nm ( long-wavelength UV ), 495 nm (blue) and 545 nm (green). Only semiquantitative visual estimates of the relative proportions of petrographic components were made in this study (see Fig. 3). Scanning electron microscopy (SEM ) was undertaken on bulk-rock subsamples and on >25 mm kerogen-concentrates using a Jeol JSM840 microscope. Preparation for SEM involved splitting dark phases of the sample along the bedding plane and etching the surface briey (23 min) with HF to remove silicate minerals and expose organic structures. After dehydration (using the CO critical-point technique), both 2 kerogen-concentrates and etched-surface subsamples were gold-coated. Energy-dispersive spectrometry ( EDS) analyses were made on an Oxford Link EXL-II unit attached to a Jeol 5400-LV SEM. Solvent-extractable organic matter (bitumen) was Soxhlet-extracted from 13 rock samples with an azeotropic mixture of dichloromethane and methanol (93:7) for 72 h. The bitumen was then fractionated by silica-gel column-chromatography to obtain the saturated hydrocarbon fraction which was analysed by a gas chromatographmass spectrometer (GCMS). A DB-1 (60 m0.32 mm i.d.; 0.1 mm lm thickness) capillary column was used. The MS was used in full-scan mode (m/z 50500). In addition to the total-ion-current ( TIC ) chromatogram, selected-ion-monitoring (SIM ) chromatograms were acquired for compound identication and integration; these included m/z 82 and 97 (alkyl- and methyl-alkylcyclohexanes); m/z 85 (n-alkanes), m/z 127 (3,7-DMAs), m/z 191 and 205 [tricyclic terpanes ( TTs), hopanes and methylhopanes], and m/z 217 and 231 (steranes and methylsteranes). Other operational details are described in Hill et al. (2000). The identication of methylalkanes is based on their GC retention

times and their diagnostic mass-fragmentograms, assisted by comparisons with the literature (e.g. Klomp, 1986; Fowler and Douglas, 1987; Summons, 1987; Kenig et al., 1995). Solvent-extracted rock powder was exhaustively macerated with HCl/HF acids to remove carbonate and silica and to concentrate kerogen for d13C analysis. Samples were sealed with CuO in evacuated quartz tubes and heated at 950C. The resulting CO was analysed in a Finnigan MAT 2 252 MS. The resulting d13C values are expressed as parts per mil relative to the PeeDee Belemnite standard.

3. Analyses and results 3.1. Petrographic, lithofacies and environmental analyses 3.1.1. Procedural comments Results of the petrographic and facies analyses of the Munta 1 core samples are consolidated in Fig. 3, which records: (1) the petrogenetic anities of each sample; (2) the various sedimentary facies present according to the Pickering et al. (1989) deep-water facies classication scheme (where applicable); (3) the respective sediment-emplacement processes; (4) the corresponding regime of either background- or event-dominated sedimentation (explained below). Also recorded in Fig. 3 are inferred palaeodepths [relative to storm wavebase ca 200 m in present-day oceans, see Pickering et al. (1989), p. 1] reconstructed primarily on the basis of the thin-section petrographic and lithofacies evidence, and extrapolated throughout the intervals of no sample control by reference to the macroscopic lithofacies descriptions of Calver and Grey (1993) of both sampled and unsampled intervals. Deep-water as used here refers to lithofacies that can be inferred to have accumulated in water depths dominated by suspension-settling and massow emplacement processes (as opposed to sediments that accumulated under conditions dominated by traction currents). The depth boundary that separates deep-water from shallow-water environments in practice is the mean depth of the

250

K. Arouri et al. / Precambrian Research 100 (2000) 235280

pycnocline. Sedimentary structures, sediment composition and colour (reecting oxidative state and relative rate of substrate accretion), among other criteria, allow reconstruction of the approximate level of the fossil pycnocline within each of the hectometre-scale cycles (Fig. 3). The level of the pycnocline as reconstructed here corresponds closely to the Calver and Lindsay (1998) sequence boundaries LU25, as underpinned by chemostratigraphic excursions of stable isotopes of carbon. As used here, the concepts background-dominated and event-dominated sedimentation regimes follow Einsele et al. (1991; and references cited therein); and are briey claried as follows in the context of the Neoproterozoic sedimentary realm. Background sediment is deemed to constitute autochthonous ne-grained, mainly suspension-load material which is produced in the area of deposition itself ( Einsele et al., 1991, p. 10). In contrast, event-deposits are deemed to constitute laterally transported sediment that is allochthonous to its environment of deposition, and that has been rapidly shed into and re-deposited in that environment by briey active geological phenomena. Event-deposits include turbidites, debrites, and tempestites, among others. Turbidites typically contain materials of distant provenance (commonly mixed with entrained basinal sediment), but the material of tempestites is either entirely autochthonous or derived from nearby sources ( Einsele et al., 1991, p. 10). A single sedimentary rock among those studied here [sample 1441.33 m an intraclastic calcirudite; Table 1; Fig. 4(H )] is attributed to tempestite origin and, for the reasons explained above, this rock is attributed (together with the predominantly carbonate samples of shallow-water anity that stratigraphically bracket it) to background-dominated rather than to eventdominated sedimentation ( Fig. 3). Among the gravel-sized carbonate intraclasts of this sample is a non-carbonate intraclast comprising displacive anhydrite in biolaminated siliciclastic mud that, on the basis of bulk-sample organic-geochemical analysis, is the subject of the biomarker study herein. Hence, interpretation of the resulting biomarker signal must be made cognisant of the samples petrographic complexity and possibly mixed-environmental provenance.

3.1.2. Preliminary interpretative comments The sedimentological characteristics of the hectometre-scale epiclastic intervals of the lower Ungoolya Group in Munta 1 indicate that these sediments are of deep-water pelagic and turbiditic origin. Among the genetically important evidence that allows this conclusion is: (1) the prevalence of very thin parallel-laminae in the mudstone and marl, indicative of very low energy conditions [Fig. 4(C )]; (2) the prevalence of graded bedding in much of the siltstone and sandstone [Fig. 4(A) and (B)]; (3) the complete absence of large-scale cross-bedding and the relative rarity of small-scale cross-lamination throughout the epiclastic intervals, except in their upper, coarser part where cross-lamination occurs either in starved ripples or hummocky bedforms ( Fig. 3); (4) the relative rarity of scour-and-ll structures, except in the upper, coarser part of the intervals where there is evidence of bottom scour in the form of truncation surfaces, especially in the top of cycle 4 ( Fig. 3). That this deep-water environment was marine is indicated by the presence throughout the succession of palynomorphs considered to be marine ( Walter et al., 1995, p. 183; Grey, 1998). Other evidence suggests that above 1850 m (the base of cycle 1) the environments in which these individual epiclastic intervals accumulated were slope-dominated, either episodically or on a sustained basis, and especially throughout the periods of accumulation of their upper, coarser parts. This is indicated by the presence of: (1) disturbed (slumped ) graded beds in the top part of cycles 1 and 4 (Calver and Grey, 1993; Fig. 3); (2) the gully-like, penecontemporaneous erosion surfaces incised into gradedbedded silts in the upper part of cycle 4 ( Table 1; Fig. 3); (3) the prevalence of micro-scale (including growth-) faults, both normal and reverse (micro-thrusts), throughout much of the succession, especially in the upper, coarser parts of cycles 1 and 4 and in the chemogenic base of cycle 2 ( Fig. 3); and circumstantially, (4) by the predominance of drab colours throughout all but the very top part of each epiclastic interval, signifying rates of substrate accretion sucient to preclude oxidation of the surcial sediment, and thus implying gravitationally enhanced (slope-controlled massow) sedimentation.

K. Arouri et al. / Precambrian Research 100 (2000) 235280

251

A relatively slow rate of epiclastic sedimentation in deep water can be inferred for the Dey Dey Mudstone below 1850 m (i.e. below the base of cycle 1). This basal part of the lower Ungoolya Group comprises slightly ssile mudrock of uniformly redbrown colour (Table 1), implying very slow substrate accretion and complete oxidation of the disseminated iron. The gradational colour change in the vicinity of 1850 m from redbrown below to predominantly dark greygreen above (Fig. 3) signies the onset of faster rates of substrate accretion and is the basis for placing the lower boundary of cycle 1 at this level ( Fig. 3). The sedimentological characteristics of the decametre-scale carbonate intervals indicate that, in general, they formed in low-energy shallow-water. Among the criteria that suggest this are: (1) their predominantly micritic texture coupled with the sporadic presence of pinkish and red colours; (2) the presence of sporadic beds of intraclastic rudite [Fig. 4(H )] and in situ ?sedimentary breccia (Fig. 3); (3) the presence of ripple cross-lamination (and possibly cross-bedding) within sporadic beds of calcarenite and interbeds of epiclastic siltstone (Fig. 3); (4) the relative frequency of in situ microbial (predominantly cyanobacterial ) mats within epiclastic silt interbeds and laminasets in these carbonates [Figs. 3 and 4( E )(G)]; and (5) the sporadic presence of tepee-like structures within red mudstone-rich intervals within these carbonates (Sukanta, 1993). Carbonates of intermediate depth aspect are also present. These were probably deposited in only moderate water depths, for the most part probably above the pycnocline, but under tranquil conditions. They occur in both the base and the top of the Karlaya Limestone [Figs. 3 and 4(D)] and in the immediately overlying basal part of the Tanana Formation. Their colour is typically neutral ( light-coloured; Calver, 1995, plate 3.2; Calver and Grey, 1993); or various shades of grey (Cucuzza, 1987; Calver and Grey, 1993). These carbonates exhibit compelling evidence that they accumulated on a primary slope. Such evidence takes the form of complex founder-structuring (uidal, asymmetric nodular bedding; Calver, 1995, plate 3.2; Fig. 3) and en-echelon complexes

of normal (micro-) faults with ubiquitous extensional gapes (e.g. in sample 1581.3 m; Fig. 3). 3.1.3. Petrogenesis of the non-organic sediment components 3.1.3.1. Epiclastics. Most of the lower Ungoolya Group is of epiclastic origin ( Fig. 3), predominantly siliciclastic (quartz, feldspar, mica, physils, etc.). Clay-sized physils are volumetrically minor. Variable amounts of possibly extraclastic dolosilt and accessory calcisilt are also present, which, on the basis of textural and compositional evidence, are epiclasts and/or reworked penecontemporaneous diagenetic sediment or carbonate substrate. Epiclastic feldspar and mica are ubiquitous throughout the lower Ungoolya Group succession, predominantly as silt-sized grains. On the basis of twinning patterns and ex-solution textures, the feldspar comprises numerous varieties (plagioclase, ?orthoclase, microcline, perthite, etc.) and is invariably fresh, as is most of the mica. This, together with (1) the common presence of the ?epiclastic dolosilt; (2) the volumetrically minor amount of clay throughout the succession; and (3) the ubiquity of varvic sandy-silt laminae of windborne origin within the background-dominated lithofacies of all three constituent formations [Figs. 3 and 4(C ) and (D)], suggests that the climate was arid. 3.1.3.2. Impact ejecta. Compositionally and texturally distinctive impact ejecta is present in the Dey Dey Mudstone in cycle 1 and is tentatively identied here as Acraman Impact ejecta [see Gostin et al. (1986)]. Acraman Impact ejecta has been identied within the Dey Dey Mudstone in other Ocer Basin drillcores ( Wallace et al., 1989; Gravestock et al., 1997), but not previously in Munta 1 (Grey, 1998, p. 290). In these other drillcores (Observatory Hill 1, and Lake Maurice West 1, located on the Murnaroo Platform respectively to the south and southwest of Munta 1), the ejecta layer occurs within monotonous shale near the base of the Dey Dey Mudstone, is very thin (<1 mm and 7 mm respectively in the former two wells; Wallace et al., 1989; Morton, 1997b, photo 42408a), and is not reworked. In Munta 1

252

K. Arouri et al. / Precambrian Research 100 (2000) 235280

the impact ejecta occurs in all four samples of the Dey Dey Mudstone examined, covering 107.6 m of section in the upper half of the formation (Fig. 3). It increases upward in relative proportion and grain-size from trace amounts of wind-borne ne sand grains of reddish glass and exotic minerals at 1810.5 m, to common concentrations of wind-borne ne sand grains of similar character in sample 1736.8 m, to more common nemedium mass-ow-emplaced sand grains in sample 1728.4 m, and to conspicuous concentrations of mass-ow-emplaced sand- and gravel-sized microtectites and clasts of suevite at 1702.9 m ( Fig. 3). The variably dispersed to locally concentrated occurrence of these exotic clasts in Munta 1 and their presence over such a thick interval in the upper half of the Dey Dey Mudstone indicate that they are re-deposited. This nding is consistent with the Grey (1998, p. 291 and g. 6.6) prediction based on comparative stratigraphy, that the primary ejecta layer in Munta 1 should occur between 1987.6 m and 1810.5 m (Fig. 3). 3.1.3.3. Chemogenics. Among the chemogenic components of the lower Ungoolya Group are abundant non-allochemical carbonate grains whose petrogenetic anities are not clear, because: (1) their mineralogy and textures overlap with probable epiclastic carbonate grains and potentially also with re-worked penecontemporaneous diagenetic sediment and carbonate substrate; (2) they are predominantly of ne crystal size, and hence are prone to have been aected by diagenetic overprints; and (3) orthochemical carbonate replacements of labile detrital grains (e.g. feldspar, other carbonates, etc.) can be dicult to distinguish from unaltered carbonate grains. Throughout the epiclastic (including marly) intervals of all four cycles, orthochemical carbonate is present in variable amounts as cement and authigenic replacements (mainly of feldspar), but it is omitted in the plot of petrogenetic composition in Fig. 3 unless it shows evidence of displacive growth (as it does in the marls). It comprises predominantly ferroan dolomite and, less commonly, ferroan calcite. The ferroan dolomite is present as euhedral and subhedral pore-lling rhombs and as overgrowths on non-ferroan dolo-

silt cores. It is particularly abundant in the marls as small rhombs that probably began growth in micropores within the initial clayey silt and continued to grow displacively once the micropores had become occluded. These dolomite crystals, and especially their rims, uoresce strongly under Hg-arc epiuorescence excitation, showing yellowish colour under both long-wavelength UV and blue light, and reddish-orange colour in green light. This suggests the probability that the crystals are concentrically zoned with respect to outwardly increasing Mn2+ and outwardly decreasing Fe2+ content as a function of continued crystal growth under changing geochemical conditions with progressive burial [see Machel and Burton (1991)]. The only signicant non-carbonate orthochemical mineral present in all host-sediment types (i.e. epiclastics, marls, and limestones) is pyrite. It occurs in variable, but generally small, proportions, mainly as framboids/polyframboids, but also as euhedra, in association with organic matter [Fig. 4(F )(H )], and locally as a pore-lling phase in clean silts and in subvertical water-escape? ssures in the marls. Minor syndepositional anhydrite occurs in sample 1441.33 m, removed from a 40 mm thick bed of intraclastic calcirudite [an inferred tempestite deposit in the top of cycle 2; Figs. 3 and 4( H )]. The anhydrite in this sample occurs within a 1020 mm2 coarsely crystalline anhydrite intraclast [see the Calver (1995) description of the calcirudite quoted herein in Table 1]. However: (1) small enclaves of the calcirudite host sediment are present within the argillaceous mud of the anhydritic intraclast due to soft-sediment involution of their mutual boundary; (2) the anhydrite crystals overprint displacively, not only the biolaminated argillaceous mud, but also the enclaves of calcirudite and their mutual boundaries with the argillaceous mud; and (3) the anhydrite crystals overprint replacively some parts of the calcirudite enclaves. This suggests that the ?gypsum-precursor of the anhydrite continued to grow both displacively and replacively after deposition of the calcirudite but prior to sediment lithication. It further suggests that the pore-water of the sediment was probably sulfate-rich. Nine of the 27 core samples studied here come

K. Arouri et al. / Precambrian Research 100 (2000) 235280

253

from among the various generalised limestone and silty limestone/limy siltstone intervals depicted in the log of Fig. 3; some other samples studied here that also contain abundant carbonate phases (variously of both orthochemical mainly post-depositional authigenic and detrital, possibly epiclastic origin) come from stratigraphic horizons adjacent to these decametre-scale carbonate intervals. Of these nine samples, two are marls [Fig. 4(C )], ve [one of which contains a marl interbed; Fig. 4(D)] consist predominantly of limestone that is variously (1) mainly of allochemical origin [Fig. 4( F ) and (H )], (2) mainly of orthochemical origin (e.g. samples 1156.3 m and 1581.3 m not illustrated here), and (3) of mixed origin [Fig. 4(D) and (F )], and two samples (1324.8 m and 1347.4 m) are carbonate-cemented epiclastic siltstone turbidites [Fig. 4(A)], with marly ne-grained interbeds. In limestone samples that texturally are now microspars, or that contain microspar enclaves [e.g. Fig. 4( F )], the precursor primary carbonate phases appear to have originated as very ne orthochemical precipitates (carbonate mud), variously within both shallow[Fig. 4( F )] and intermediate-depth settings [Fig. 4(D)]. The predominant carbonate allochems in the limestone samples are: (1) silt-sized, formerly micritic, peloids; and (2) sand- and gravel-sized, formerly micritic, tabular intraclasts [Fig. 4(H )]. The silt peloids are subspherical/ellipsoidal and have the appearance of faecal pellets, but they could be small rounded intraclasts. Those that occur in an intermediate-depth setting (i.e. sample 1670.0 m, basal Karlaya Limestone) are cryptic in thin-section and are evident only in very bright transmitted light. The notion that faecal pellets occur in these limestones is problematical because their age (~575 Ma, Fig. 3) pre-dates the earliest known palaeontological evidence of metazoans with muscular, unidirectional guts capable of producing such pellets [see Walter (1995)]. Post-depositional orthochemical carbonate phases occur as interstitial cement in the limestones of mainly allochemical origin [e.g. Fig. 4(H )], and they occlude bedding-parallel and bedding-normal planar micro-fractures that may be desiccation features in limestone sample 1443.6 m. Depending on the sizes of the cavities occluded, these pore-

lling carbonates have a cement stratigraphy that begins with non-ferroan calcite, changes to ferroan calcite (terminal phase in small pores), and (in large pores/cavities) ends with ferroan dolomite (e.g. sample 1443.6 m). This mineralogical progression from non-ferroan to ferroan phases suggests a cementation history that began under surface (oxidising) conditions, and continued during progressive shallow burial (i.e. under reducing conditions). The ferroan-dolomite cement and ne displacive ferroan-dolomite euhedra and subhedra that characterise the epiclastic and marly intervals are much less common in the predominantly limestone intervals, presumably reecting a paucity of Fe2+ in the absence of an adequate diagenetic source from signicant proportions of admixed siliciclastic minerals. 3.1.4. Lithofacies 3.1.4.1. Facies specication, sedimentation-process and palaeodepth signicance. As elaborated above, the lower Ungoolya Group in Munta 1 comprises four sedimentary cycles of upward-shallowing nature underlain by an interval of deep-water red brown mudrock ( Fig. 3). The bulk of each cycle comprises a hectometre-scale interval of epiclastics of deep-water aspect followed by a decametrescale carbonate interval of shallow-water aspect ( Fig. 3). The deep-water sediments can be categorised within the Pickering et al. (1989) lithofacies scheme. All but four of the 27 samples studied here can be accommodated within this classication as either: (1) Facies Group D1 [disorganized silts and silty muds; i.e. the product of highconcentration, silt-dominated turbidity currents, or highly uid, silty debris-ows; Fig. 4(B)]; or (2) Facies Group D2 [organized silts and muddy silts; i.e. the product of low-concentration turbidity currents with or without accessory weak bottom currents; Fig. 4(A)]; or (3) Facies E2.2 [i.e. laminated muds and clays; the product of slow continuous settling from the water column of background suspension-load nes, punctuated by episodic inux of suspension-load mud in lowconcentration turbidity currents and wind-borne silt and accessory ne sand to form very-thin, sandy-silt laminae that are commonly semi-discon-

254

K. Arouri et al. / Precambrian Research 100 (2000) 235280

tinuous; see Schieber (1990), gs. 3 and 4; Fig. 4(C ) and (D) herein]. These various lithofacies accumulated at depths that exceeded the pycnocline, and probably that of storm wave-base, and possibly that of the photic zone. The four samples that defy accommodation within the Pickering et al. (1989) scheme are from two separate, predominantly carbonate, intervals of shallow-water aspect in the Tanana Formation, specically (Fig. 3): the interval 14551430 m (samples 1438.3, 1441.33, and 1443.6 m; cycle 2); and the interval 11701152 m (sample 1156.35 m; cycle 4). These samples, with one exception (i.e. sample 1438.3 m), are predominantly carbonates, and all have sedimentological characteristics suggestive of accumulation as background carbonate and (minor) epiclastic silt in shallow-water photiczone environments subject to mild traction-currents, occasional tempestite events, probable episodic inux of wind-borne epiclastic silt, possibly rare/local subaerial exposure, and possibly episodic evaporitic conditions [Figs. 3 and 4( E) (H )]. 3.1.4.2. Deep-water facies and inferred environments. The lowermost two samples of the Dey Dey Mudstone (1810.5 m and 1736.8 m; cycle 1) belong to Facies E2.2 (+D2.1 in the very base of sample 1810.5 m). In these (siliciclastic) samples, Facies E2.2 manifests hemipelagic rain of siliciclastic silt and physils, punctuated by episodic (?seasonal ) inux of wind-borne silt and ne sand (which includes sporadic Acraman Impact debris in both samples). Facies D2.1 represents an inux of suspended-load silt and clay in low-concentration turbidity currents. This facies association indicates a regime of deep-water background sedimentation on a low-gradient distal turbidite fan or apron ( Fig. 3). Organic matter is abundant in both samples: rstly, as autochthonous mats at 1810.5 m ( Fig. 3; TOC=6.56 mg/g, Calver, 1995); secondly, as abundant detrital akes at 1736.8 m (Fig. 3), similar to the deep-water carbonaceous streak shale facies of the Mesoproterozoic Belt Basin, Montana (Schieber, 1990, g. 3). No biomarker data are available from these autochthonous organic mats at 1810.5 m. These deep-water microbial mats show hardly any red epiuores-

cence under green-light UV-excitation ( Fig. 3) in contrast to the moderately bright to bright red epiuorescence of most other autochthonous mats of inferred shallow-water anity (as well as much of the detrital mat material in the turbidites, including within a thin Facies D2.1 bed in sample 1810.5 m; Fig. 3). In reference to the black shale of sample 1810.5 m, Calver (1995) states that it is similar to the black shale beds in the [coeval ] Bunyeroo Formation (Adelaide Rift Complex), and that, like them, it also lacks the anomalous 13C-depletion of the black shales of the benthic microbial mat facies. This evidence, together with the sedimentological lithofacies evidence, suggests that the in situ benthic mats of sample 1810.5 m probably represent non-phototrophic organic accumulation. The upper two samples of the Dey Dey Mudstone (1728.4 m and 1709.2 m; cycle 1) are referable to Facies Group D2 and Facies D1.1 respectively (Fig. 3). They manifest emplacement by low-concentration turbidity currents, and by (slightly sandy) silt-dominated high-concentration turbidity currents or silty debris-ows. Coarse (including gravel-grade) and relatively abundant Acraman Impact debris occurs in this interval. This upper part of the Dey Dey Mudstone (above ca 1730 m) reects higher energy (and higher gradient) conditions than the underlying section and can be interpreted as having accumulated on a turbidite slope dominated by silt-sized sediment inux, but coarsening upward with time. The unsampled redbrown mudrock interval of the Dey Dey Mudstone below 1850 m is interpreted [from Calver and Grey (1993)] to comprise Facies E2.2 (and possibly minor D2.1), but its overall slightly ner-grained character and red brown colour compared with the immediately overlying drab-coloured interval indicates that it has tranquil basin-oor anity (Fig. 3) and accumulated very slowly (as testied by the near absence of preserved palynmorphs in this entire interval up to 1820 m; elaborated in Section 3.2). One sample from the basal Karlaya Limestone (cycle 1) and three samples from the basal part of the Tanana Formation (cycle 2) are pelagites ( Facies E2.2; extrapolated in the case of basal cycle 2 up to ~1500 m; Fig. 3). These pelagites

K. Arouri et al. / Precambrian Research 100 (2000) 235280

255

are variously either chemogenic (sample 1581.3 m), marly [samples 1596.2 and 1561.9 m; Fig. 4(C )], or comprise thin-section-scale interbeds of each of these phases [1670.0 m; Fig. 4(D)]. These sediments represent low-energy and/or deepish-water background-sedimentation, starved of bedload inux of epiclastics ( Fig. 3). There is evidence of episodic wind-borne sandy silt throughout this interval [Figs. 3 and 4(C ) and (D)], probably signifying extreme aridity in the sourcelands. The intervals 15001455 m (cycle 2), 1380 1320 m (cycle 3), and 13051255 m and 1216 1170 m (both in cycle 4) are dominated by lowconcentration turbidity current deposits [Facies D2.1 and D2.3; Figs. 3 and 4(A)], and the intervening interval within the middle part of cycle 4 (12551216 m) by high-concentration turbiditycurrent deposits or silty debris-ow deposits (Facies D1.1) containing abundant detrital and allochthonous (slumped) organic mat material [Figs. 3 and 4(B)]. The upper half of the turbiditeslope environment in cycle 4 (above ~1250 m) was evidently of relatively high-gradient character and experienced: (1) episodic gravitational failure (as testied by the frequency of disturbed bedding; Fig. 3); (2) penecontemporaneous gully erosion (as indicated by the frequency of high-relief, turbidite-lled scours; Calver, 1995; Morton, 1997b; Fig. 3 herein); and (3) an upward increase in bedthickness and grain-size associated with the development of hummocky bedforms (?antidunes in pycnocline seiche-generated turbidites). All this evidence signies relatively high-energy, high-gradient conditions and active physical disturbance of the substrate at these levels and above, presumably as a consequence of the intersection of the pycnocline with the seaoor above this level and because of the physical consequences of this for bottom-water movement. Some samples of the deep-water lithofacies contain micro-faults, mostly sub-vertical normal growth faults (including one demonstrating that part of the fracture functioned as a water-escape conduit), and less common micro-thrusts ( Fig. 3). In our thin-sections of the lower Ungoolya Group succession, micro-faults are present discontinuously in the deep-water lithofacies of cycles 1, 2, and 4. In cycle 4 they are present throughout the

entire epiclastic interval, together with macroscopically discernible disturbed bedding (Fig. 3). Sample 1581.3 m, a ne-grained limestone from the basal part of the Tanana Formation (cycle 2), shows a complex of en-echelon normal microfaults with tensional gapes healed by carbonate cement. The presence of these micro-fractures in both turbiditic and hemipelagic lithofacies presumably indicates ongoing slope-driven soft-sediment deformation of the surcial sediment involving down-slope creep. Above ca 1730 m in the Dey Dey Mudstone, and above ca 1300 m in the Tanana Formation, the presence of macroscopic-scale softsediment disruption ( Fig. 3) presumably indicates larger-scale slumping and sliding, which, in turn, implies increasing slope gradient as a consequence of the progradation of coarser-grained turbiditic aprons or lobes. 3.1.4.3. Shallow-water facies and environmental constraints. The basal part of the Dey Dey Mudstone has not been studied here in thinsection, but its uvial to ?deltaic lithofacies anity has been argued by others (Table 1; Fig. 3). Red colour dominates this basal part of the formation, as well as the underlying Murnaroo Formation (Calver and Grey, 1993), and together with the arkosic nature of the sandstones (Morton, 1997b) indicates arid conditions (Fig. 3). Only cycles 2 and 4 are represented by samples of shallow-water anity in the present study (intervals 14551435 m and 11701152 m). These samples are predominantly carbonates that contain thin siliciclastic silt laminae and/or laminasets that host in situ microbial mats [Figs. 3 and 4( E ) (G)]. These samples are petrologically complex and contain evidence of probable peritidal deposition. This includes: (1) indications of gentle to strong traction currents [e.g. small-scale scourand-ll structures, irregular and lenticular bedding, ripple cross-lamination, and imbricated concentrations of well-rounded intraclasts; Fig. 4( E ) (H )] and presence of a probable tempestite [Figs. 3 and 4(H )]; (2) manifestations of abrupt pressure changes and possibly of subaerial exposure [e.g. water-escape structures, and various pre-lithication deformation structures, including possible desiccation cracks; Fig. 4(F )]; (3) the ubiquity of

256

K. Arouri et al. / Precambrian Research 100 (2000) 235280

benthic microbial mats, both autochthonous [Figs. 3 and 4( E )(G)] and para-autochthonous [Fig. 4(H )], most of which show bright to moderately bright red epiuorescence in green-light UV-excitation (Fig. 3); (4) the presence (in sample 1441.33 m) of mechanically reworked (intraclastic, and syndepositionally regenerated) anhydrite, suggestive of an evaporitic strandline source and possible syndepositional hypersalinity [Fig. 4(H ); Table 1; see Section 3.1.3.3]; (5) the predominantly cyanobacterial biomarker anities of the benthic microbial mats (possibly including thermophiles) present in these samples ( Fig. 3; Section 3.4), suggestive of shallow, photic-zone palaeodepths; (6) the presence in the microbial mats of sample 1443.6 m of probable microbially mediated carbonate crystals [Fig. 5(E ) and (F ); see also Sections 3.3 and 3.4.3.3] indicative of water temperatures typical of tropical/subtropical salterns [see Gerdes et al. (1991)], sabkhas, and hot-springs and associated lakes [see Folk et al. (1985)], in turn suggesting very-shallow photic-zone depths; (7) in the benthic microbial mats of sample 1438.30 m, the presence of a sub-millimetre-scale varvic stratigraphy characterised by alternating bedding-oblique and bedding-parallel domains of interstitial kerogen, interpreted [see Gerdes et al. (1991)] to indicate periodic (?seasonal or ?diurnal ) alternation of vertical and horizontal growth increments respectively in a sediment-starved environment [Fig. 4( E )]; and (8) the relative abundance in these limestones of subspherical/ ellipsoidal peloids grains that, while not diagnostic of any particular environment, are the main components of very shallow-water subtidal sediments, especially lagoonal sands and muds ( Wright and Burchette, 1996, p. 327). These two intervals of inferred shallow-water, predominantly carbonate deposition, represent periods of epiclastic-starved background sedimentation (Fig. 3): the lower interval concluding cycle 2, and the upper interval concluding cycle 4. The carbonate intervals that conclude cycles 1 and 3, though unrepresented here by samples, are inferred to have analogous facies and environmental anities to those of cycles 2 and 4 on the basis of their description in the Calver and Grey (1993) log.

3.1.5. Organic petrography and sedimentology There is considerable and abrupt stratigraphic variation in the nature and relative abundance of organic matter in the Munta 1 samples (Figs. 3 and 4). The following types of organic matter are optically detectable. (1) Abundant-to-rare akes and sheets of detrital organic tissue; present in samples from all four cycles (and especially noticeable in the epiclastic intervals; Fig. 3), and transported and emplaced both as part of the suspended-load that resulted in background (pelagic) sediments [Fig. 4(C ) and (D)] and within turbiditic mass-ows that generated turbidites and debrites [Fig. 4(A) and (B)]. Most palynological samples contain abundant nely disseminated organic matter consisting of particles <1 mm in diameter. (2) Allochthonous organic mats that have been transported either by granular (incoherent), high-concentration mass-ows (category 2A), or within structurally coherent, large intraclasts (category 2B). The distinction between allochthonous and detrital as used here is qualitative: compared with category 1, category 2A is characterised by high concentrations of large, semi-continuous, bedding-aligned mat sheets, as well as by intraclastic organically bound enclaves. Among our samples, category 2A is represented only from the upper part of the Tanana Formation at 1216.0 m (i.e. from the upper, high-energy part of the epiclastic interval in cycle 4; Fig. 3), and category 2B is represented only at 1441.33 m in the shoalwater carbonate interval that terminates cycle 2 [Figs. 3 and 4(H )]. (3) Pervasive kerogen, present as wavy, anastomosing lms that occur preferentially within siliciclastic mudrock, especially siltstone, and which demonstrably constitute in situ benthic organic mats. The thickness of individual layers of kerogen range from 3 to 100 mm. This category, referred to as autochthonous (or in situ) mat, is represented among our samples only in the Tanana Formation and the Dey Dey Mudstone [Figs. 3 and 4( E )(G)]. In the Dey Dey Mudstone the in situ mats occur in sample 1810.5 m within the lower part of cycle 1 (i.e. lowenergy turbidite apron; Fig. 3). Those from the Tanana Formation occur in the shoalwater carbonates that terminate cycles 2 and 4 (Fig. 3). Various other types of more dispersed kerogen

K. Arouri et al. / Precambrian Research 100 (2000) 235280

257

are also optically discernible in many samples. These include: (4) Brown kerogen dispersed as background organic matter in ne-grained carbonates where its presence is demonstrated by its concentration as stylocumulate. This category is represented among our samples by the Tanana Formation and the Karlaya Limestone [i.e. in cycles 1, 2, and 4; Figs. 3 and 4(D)]. (5) Typically patchy, glossy gel-like (?mobile) organic matter that is variously colourless, pale yellowish, or pale orange, and evidently occurs as a graincoating/interstitial or perhaps, in some cases, substrate-pervasive phase. This last type of kerogen occurs in samples of cycles 2, 3 and 4, especially in the epiclastic intervals. The general petrographic nature and microscale bedding characteristics of the structurally discrete types of kerogen in the autochthonous benthic mats, and comprising the redeposited detrital fragments presumably derived therefrom, appear to be similar to other Proterozoic and Phanerozoic counterparts, such as those described by Shieber (1986, 1989, 1990, 1999) and OBrien (1990). The autochthonous mats appear to be substrate-specic, preferring to colonise only epiclastic silt, and are most prevalent in silt laminae and laminasets within the shoalwater carbonates (Fig. 3). Only in the basal part of cycle 1 (sample 1810.5 m) do we nd autochthonous mats in a deep-water setting. An estimate of the minimum water depth at which the latter mats grew can be made on the basis of the constructional sedimentary relief embodied in the down-to-basin (shallow-to-deep) palaeoenvironmental prole of cycle 1 at the Munta 1 site ( Fig. 3; i.e. ~220 m, making no allowance for sediment compaction, and reduced by 40 m to allow for the stratigraphic dierence between the base of cycle 1 and the level of the mats). This gives a minimum water depth for the mats of ~180 m. The sedimentology of the allochthonous and detrital categories of microbial mats is best understood in relation to their respective environmental settings within the overall upward-shallowing hectometre-scale cycles, as summarised in Fig. 3. Of foremost signicance among these considerations are the palaeophysiographic locations of autochthonous mat growth in relation to the depth of

the photic zone and, in particular, the depth of the pycnocline. Given the equatorial palaeolatitudinal position of the basin during the late Neoproterozoic (Section 2.1), and the evident starved-basin sedimentation regime that prevailed at the time of the Dey Dey Mudstone and beyond ( Fig. 3), it is probable that the photic zone extended down more than 100 m and so probably did not pose a critical barrier to microbial mat growth at depth. Nonetheless, the petrographic evidence shows that the in situ mats grew most abundantly in two places: (1) on thin (possibly mainly wind-blown) epiclastic silt substrates within the shallow-water carbonate platform deposits that terminate each of the hectometre-scale cycles; (2) on the upper shallower (and higher energy) parts of the epiclastic turbiditic slope deposits that lay just downslope from the edges of these carbonate shelves. Mechanical reworking and redeposition of the in situ mats took place in three major ways. (1) On the carbonate platforms, occasional storms fragmented the mats together with their immediate epiclastic substrate and the muddy carbonate sediment that underlay it. The storms generated tempestite deposits such as are represented in sample 1441.33 m [Figs. 3 and 4(H )], categorised here as allochthonous mat type 2B. (2) On the upper highenergy epiclastic turbidite slope, mats evidently grew abundantly, as testied by their slumped and high-concentration mass-owed deposits present as Facies D2.1 (categorised here as allochthonous mat type 2A) in the middle part of cycle 4, where they occur in intimate association with spectacular slope-gully erosion surfaces, much disturbed bedding, and graded beds with hummocky bedforms [Figs. 3 and 4(B); see also Calver (1995), plate 3.3, and Morton (1997b), photo. 42411c]. That the mats in these high-concentration debris-ow deposits grew on the upper parts of the epiclastic turbidite slope and not on the adjacent carbonate platform (nor on spill-over upper-slope carbonates, where they existed, such as in the top of cycle 1; Fig. 3) is proven by: (i) the complete absence of carbonate sediment within these debrites; (ii) the near-structurally intact nature of some of these debrites (e.g. sample 1216.0 m), implying a limited distance of movement from their source; and (iii) by the absence of autochthonous mats in sample

258

K. Arouri et al. / Precambrian Research 100 (2000) 235280

1670.0 m in the basal Karlaya Limestone, which can be reliably interpreted to have accumulated in this palaeobathymetric location during a period of intense epiclastic starvation such that the upper slope deposits failed to capture appreciable epiclastics. Mechanisms that likely triggered the slumping of these autochthonous mats and their downslope mobilisation in high-concentration debris-ows include storm-generated bottom currents above the pycnocline and (probably most signicantly) episodic seiches at the pycnocline. The collision of these internal waves would have caused seismic ground disturbance of the seaoor sediments of the upper slope as well as uid-stressing of the substrate, and gravity-driven surges of both dense and less-dense water up onto the higher parts of the slope above the pycnocline [see Woodrow (1985)], with subsequent catastrophic return. (3) Such downslope return of this water (and especially the denser water) would have caused bottomsediment and in situ mat dislodgement and entrainment, resulting in downslope-driven sandy and silty turbidity currents (and probable slope-gully incision), and entrapment of the ner suspensionload particles, including myriad fragments of the formerly autochthonous mats, as interows along the pycnocline, from which they would have long continued to settle as pelagites ( Facies E2.2) in the deeper basin areas, including on the distal slope. This latter phenomenon presumably accounts for the observation (evident in the data summarised in Fig. 3) that detrital mat fragments (category 3 kerogen) are, on average, equally abundant in the pelagites deposited under a regime of background sedimentation as they are in the lowconcentration turbidites ( Facies D2.1 and D2.3) deposited during regimes of event-dominated sedimentation. A further corollary of this observation would probably require that, during regimes of background-dominated sedimentation, much fragmentation and mechanical reworking of the autochthonous mats on the slope areas above the pycnocline was accomplished by processes related to more ambient ( lower-energy) phenomena [surface waves, tidal currents, and, in intertidal environments, mat desiccation, cracking, curl-up, and otation etc., see Schieber (1999), g. 11].

3.2. Palynology Palynology of the Munta 1 core is consistent with stratigraphic distribution patterns recorded from other drillholes in the Ocer and Amadeus Basins (Grey, 1998). In Munta 1, samples from the Murnaroo Formation and Dey Dey Mudstone below 1810.5 m mostly lack identiable acritarchs, but contain a few simple sphaeromorph acritarchs including rare Leiosphaeridia jacutica, Leiosphaeridia crassa, Leiosphaeridia tenuissima and Leiosphaeridia minutissima. From the available evidence these samples are assigned to the Simple Leiosphere Palynoora of Grey (1998). The rest of the Dey Dey Mudstone is characterised by poor preservation and low, or somewhat sporadic, yields. However, from 1810.5 m upwards, large acanthomorph acritarchs typical of the Ediacarian Complex Acanthomorph Palynoora of Grey (1998) make their rst appearance, and taxa become more abundant as preservation improves upwards through the Dey Dey Mudstone. Palynomorphs are abundant in most samples from the Karlaya Limestone and Tanana Formation. The transition from the leiosphere-dominated assemblage to the acanthomorph-dominated assemblage occurs in a barren interval between 1987 and 1810.5 m and in Fig. 3 is placed high in this interval above the predicted level of the Acraman Impact Event and just above the base of cycle 1. Typical acritarchs include Alicesphaeridium medusoidum, Multifronsphaeridium pelorium, species of Appendisphaera, Cavaspina, Papillomembrana, Schizofusa, and Tanarium (some of which are new), together with several other new genera and species (Grey, 1998). Tanarium conoideum and Tanarium irregulare, previously recorded from Siberia (Moczydlowska et al., 1993), occur respectively from 1326.0 m to 1198.6 m and 1234.6 m to 1198.6 m at about the middle of the Tanana Formation. From evidence elsewhere in the Centralian Superbasin, the large complex acanthomorph palynoora may have been short-lived, and had become extinct before the end of the Neoproterozoic, at about the time of the main appearance of the Ediacaran fauna (Grey, 1998).

K. Arouri et al. / Precambrian Research 100 (2000) 235280

259

3.3. SEM and EDS analyses Selected palynological residues and freshly broken, slightly HF-etched rock surfaces were

studied under the SEM, resolving structures of biological or probable biological origin not routinely detected by conventional light microscopy. They include mineralised laments and coccoid

Fig. 5. Scanning electron micrographs of microorganisms and possibly microbially mediated structures from the Tanana Formation, Munta 1 drillcore, eastern Ocer Basin. All images are from the autochthonous microbial mat micro-lithofacies of sample 1443.6 m [i.e. the dark phase of the sample; and Microbial Facies 2 in Table 2; see also Fig. 4(G)]. Bar-scale in all images is 1 mm, except in (C ) (10 mm). (A), (C ) Coccoidal structures on slightly etched bedding-plane surface. (B) Densely packed mass of possibly bacterially mediated granules on etched bedding-plane surface. (D) Filaments incorporated into kerogen from palynological residue. (E ), ( F ) Slightly etched carbonate crystals (on bedding-plane surface) enclosing clusters of granules or capsules that have possibly formed around bacteria.

260

Table 2 Overview of the biomarker signatures in core-samples of the Tanana Formation in Munta 1, Ocer Basin, South Australia, categorised on the basis of the three separate organic-facies dened at left. See text for detailed commentary
n-Alkanes n-Alkylcyclohexanes Range (maximum) Odd-overeven-C-number predominance Abundance 5-Me18/ nC 18 Monomethylalkanes (MMAs) Dimethylalkanes (DMAs)

Organic facies & sample number (depth, m) Range (maximum)

Macroscopic appearance in hand-specimen (TOC %; measured )

Only 3,7-DMAs (m/z 127 trace)

3,7di21/ nC 20

Microbial Facies 1A 1324.8 dark-coloured; laminated (0.06) C C (C ) 14 34 18 EOP (C ); 1620 OEP (C ) 2533 no OEP C C (EOP: C ) 15 26+ 18 C C ( EOP; C ) 16 28 18 abundant (mainly 3- & 5-); (C C ; low 16 33 MW>high MW ); even MMAs> odd MMAs abundant (mainly 3- & 5-); (C C ; low 16 33 MW>high MW ); even MMAs> odd MMAs

0.80

C C major 19 35

0.9

1441.33

intraclastic calcirudite with dk-coloured biolaminated mdst intraclast (bulk analysis, 0.23)

C C (C ) 14 33 16

0.60

C C abundant 19 35

0.31

Microbial Facies 1B 1264.8 dark-coloured; laminated (0.10) slight OEP (>C ) 25 slight OEP (>C ) 25 C C (C ) 16 26+ 16 C C ; 14 33 C minimal; (C ) 17 18 C C (OEP: C , C ) 15 25 17 19 C C 15 29+ (OEP: C , C ) 17 19

0.31 0.80

0.70 1.9

1347.4

slightly laminated (0.10)

abundant (C C ) 16 24 very abundant (C C ) 16 24

C C moderate 17 24 C C major 17 24

Microbial Facies 2 1234.0 dark-coloured; laminated (0.15) C C ; 16 26 C minimal; (C ) 17 18 C C (C , C ) 16 21 16 18 C C (C ) 17 25+ 18 EOP (C ) 1620 no OEP EOP (<C ) 20 EOP (C ) 1620 C C (C ) 16 19 18 C C (OEP: C , C ) 16 30 19 21 C C (C ) 16 26+ 19 C C (C ) 16 26+ 17 EOP (C ) 1726 C C 16 21+ (C ; C /nC =0.06) 18 18 18

C , C , C (mainly 2-, 3-, & 5-isomers) 16 18 20 C & C MMAs only (mainly 2-, 3-, 16 18 & 5-isomers) C C : EOP (mainly 5-isomers) 18 24 C , C , C (traces) (mainly 5-isomers) 16 18 20 mainly C & C ; but C , C and C 16 18 17 19 20 also traceable

1.0

not detectable

0.0

K. Arouri et al. / Precambrian Research 100 (2000) 235280

1252.2

(i) dark-coloured; laminated (0.15)

1.0 1.0 0.46 0.22 0.09 0.36

not detectable C C minor 19 31 C C minor 19 35 not detectable

0.0 0.1 0.1 0.0 0.06 0.07

(ii) light-coloured; non-laminated (0.07)

1438.3

(i) dark-coloured; laminated (n.d.)

(ii) light-coloured (0.12)

C C ; 15 34 C minimal; (C ) 17 18 C C (C ) 15 33 16

1443.6

(i) dark-coloured; laminated (0.30)

(ii) light-coloured (0.12)

C C (C ) 14 33 16 C C ; 15 33 C minimal; (C ) 17 16

slight OEP (C ) 24+ EOP (<C );slight 20 OEP (C ) 24+

C C (C ) 14 26+ 17 C C 15 26+ (OEP: C , C , C ) 17 19 21

C C (minor) 15 20 C & C MMAs only (mainly 2-, 3-, 16 18 & 5-isomers)

C C traces 19 25 C C traces 17 25

Algal Facies 1209.5 C C (C ) 15 22 20 C C (C ) 13 22 14

light-coloured; non-laminated (0.08)

0.03 0.01

0.02 0.02

1216.0 (complex facies)

light-coloured; non-laminated (0.09)

EOP (C ) 1520 EOP (C ) 1317

C C (C ) 15 21 17 C C (C ) 13 21 16

nearly absent (<C ) 21 nearly absent (<C ) 21

nearly absent (C +C ) 19 21 nearly absent (?C C ) 17 23

Isoprenoids Pr/nC 17 18 Ph/nC Range and abundance C hopane/ 30 nC or nC 17 18 0.30 1.0 Regular; irregular; methylsteranes C sterane/ 29 C hopane 30 0.05 0.04

Triterpanes

Steranes

Organic facies & sample number (depth, m)

Abundance

Pr/Ph

Unidentied branched compounds

d13C kerogen ()

Microbial Facies 1A 1324.8 <C 0.67 0.63 0.71 0.79 not detectable C C (major; C >C ); TTs minor; 27 34 30 29 2a-methylhopanes traces not detectable C C (major; C >C ); TTs traces; 27 35 30 29 2a-methylhopanes traces C >C >C >C ; dia<sterane; 2a>4a&3b; 29 27 28 30 abundant 4a C &C >C >C ; dia%sterane; 2a>4a>3b; 29 27 28 30 abundant 4a

22

0.6

27.47 28.53

1441.33

<C

22

1.0

Microbial Facies 1B 1264.8 <C 0.52 0.68 0.79 0.09 0.55 not detectable C C (C >C ); TTs; 2a-methylho27 32 29 30 panes abundant 0.02 not detectable C C (C >C ); 2a-methylhopanes 27 32 29 30 abundant

22

1.5

C >C >C >C ; 2a>3b>4a-methylsteranes 29 27 28 30 C >C >C >C ; 2a>3b>4a-methylsteranes; 29 27 28 30 4a ?absent

0.03 0.02

26.23 27.40

1347.4

<C

22

1.0

Microbial Facies 2 1234.0 not detectable no Pr no Pr abundant 0.07 no Ph major C C ; TTs trace; 2a-methylhopanes 27 31 abundant C C ; TTs; 2a-methylhopanes abun27 33 dant 0.11 0.07 0.13 no Ph major C C ; TTs; 2a-methylhopanes traces 27 33 0.30

no Pr; no Ph

C C >C >C ; dia traces; 29 27 28 30 methylsteranes

2a>4a>3b-

0.03 0.06 0.06

29.88 29.84 n.d.

1252.2

(i) not detectable

no Pr; no Ph 0.63; 0.12 Pr traces no Pr 0.24 0.16 moderate 0.12 major

C >C >C >C ; 2a>3b>4a-methylsteranes 29 27 28 30 C >C >C >C ; 2a>3b>4a-methylsteranes 29 27 28 30

(ii) <C

22

traces only

0.5

1438.3

no Pr

0.03 0.03

27.45 n.d.

(i) <C traces only 22 (ii) only Pr & Ph (minor) C C ; TTs; 2a-methylhopanes present 27 34 C C ; TTs; 2a-methylhopanes present 27 33 (major); TTs traces; no methylho0.36 0.40 0.11 abundant 0.26 traces C C 27 35 panes

1.3

C >C >C >C ; diasteranes<steranes 29 27 28 30 C >C >C >C ; dia%sterane; 2a>4a&3b; 27 29 28 30 abundant 4a 0.74 1.05 C &C >C >C ; no dia; 4a>2a>3b; abun29 28 27 30 dant 4a C &C C >C ; no dia; 4a>2az3b; abun29 27 28 30 dant 4a

K. Arouri et al. / Precambrian Research 100 (2000) 235280

1443.6

(i) <C

22

(minor)

1.4

0.03 0.02

28.18 27.84

(ii) only traces of Pr & Ph

0.9

C C ; TTs; no methylhopanes 27 34

Algal Facies 1209.5 0.30 0.32 0.18 minor 0.11 minor C C 27 33 traces

minor Pr, Ph

1.0

; TTs traces; 2a-methylhopanes

0.03 0.05

C >C >C 29 27 28 methylsteranes

>C ; dia traces; 30 C >C >C >C ; dia traces; 29 27 28 30 methylsteranes

2a>4a>3b2a>4a>3b-

0.30 0.02

27.79 28.54

1216.0 (complex facies)

minor Pr, Ph

2.2

C C ; TTs traces; 2a-methylhopanes 27 33 traces

261

262

K. Arouri et al. / Precambrian Research 100 (2000) 235280

microfossils, non-mineralised problematica reminiscent of fungal sclerotia, and minute carbonate crystals containing substructures that might constitute bacterially mediated granules (Fig. 5). Filaments seen on the slightly etched bedding planes are either organically preserved [Fig. 5(D)] or, more commonly, mineralised. Some of the mineralised laments show signs of partial dissolution of the mineral crust due to the etching with HF. Coccoid structures, seen on acid-etched rock surfaces, occur as irregularly distributed patches or densely packed masses [Fig. 5(A)(C )]. They have not been detected in thin-sections. Individual cells are globular to ellipsoidal, ranging in diameter from 1.0 to 2.2 mm. They occur variously as individuals, in pairs, in short curved chains of three or four individuals, or in clusters that in some cases have a spiral pattern. Budding is very rare; most cells divide by what appears to be binary ssion. Mineralised threads, with stretch (extension) patterns at the periphery of, or between, adjacent cells are not uncommon. EDS analyses indicate that the mineralised coccoids typically have moderate to high amounts of Al (1017%), Si (2972%) and Ca (926%), low amounts of K (6% maximum) and Fe (5% maximum), and traces of Na, Mg, Ti and Zn. However, in one area the specimens contain relatively high amounts of Fe (1462%) and Ca (up to 46%), and in one case K (24%), and only moderate amounts of Si (924%). These structures are tentatively interpreted as mineralised coccoid bacteria. The non-mineralised problematica reminiscent of fungal sclerotia occur on fragments of apparently structureless kerogen. They have a globular form ranging from 7 to 10 mm across, a structure comprising branching vermicular elements, and appear to be attached to laments. Carbonate crystals that enclose clusters of globular granules occur either singly (320 mm in diameter) or in aggregates on the etched rock surfaces [Fig. 5( E ) and (F )]. These crystals are similar to those described by Folk and Chafetz (1983) and illustrated by Folk et al. (1985, g. 10C and D) from bacterially mediated hot-spring travertines (Chafetz and Folk, 1984) and from carbonates in Recent peritidal cyanobacterial mats (Buczynski and Chafetz, 1993).

3.4. Biomarkers 3.4.1. Introductory statement Biomarker data of 13 Tanana Formation samples examined here are summarised in Table 2, and representative GCMS TIC chromatograms are shown in Fig. 6. Biomarker distributions, including abundant methylalkanes and hopanes, reect the major cyanobacterial contribution to the Tanana Formation mats. Methylalkanes comprise MMAs and DMAs, most importantly the 3,7-DMAs or 3,v7-DMAs. At least three major biomarker associations can be recognised in the Tanana Formation in Munta 1 ( Table 2). Variations in biomarker content are primarily governed by microbial diversity, but can also be caused by dierent biosynthetic pathways resulting from dierent environmental conditions (e.g. Robinson and Eglinton, 1990; Des Marais et al., 1992a). Three organic facies can be recognised (Fig. 6): (1) Microbial Mat Facies 1; (2) Microbial Mat Facies 2; and (3) Planktonic Algal Facies. 3.4.2. Microbial Mat Facies 1 This facies is divisible into two subfacies, 1A and 1B, both characterised by biomarker assemblages typical of cyanobacterial mats resembling modern examples in Abu Dhabi and the Gavish Sabkha. Microbial reworking is also evident. 3.4.2.1. Subfacies 1A. A representative GCMS trace ( TIC ) of this microbial subfacies is shown in Fig. 6(a). This mat assemblage is represented ( Fig. 3) by samples from 1324.8 m [a Facies D2.3 turbidite from cycle 3; Fig. 4(A)] and 1441.33 m [a shallow-water tempestite deposit from the top of cycle 2: Fig. 4(H )]. The latter sample includes part of a small pebble-sized intraclast comprising dark biolaminated clayey siltstone with displacive nodular anhydrite (described in Sections 3.1.1 and 3.1.3.3), possibly indicating evaporitic conditions in its ( likely nearby, peritidal ) source environment. The TIC chromatogram of the saturated hydrocarbon fraction [Fig. 6(a)] shows a bimodal distribution covering the range n-C to at least n-C . 14 33 The rst cluster covers the low molecular-weight range (n-C ) and maximises at n-C or n-C . 1620 16 17

K. Arouri et al. / Precambrian Research 100 (2000) 235280

263

Fig. 6. Partial TIC chromatograms of saturated hydrocarbons of extracts from representative organic facies of the Tanana Formation: (a) Microbial Facies 1A, sample 1441.33 m; (b) Microbial Facies 2, sample 1443.6 m; and (c) Planktonic Algal Facies, sample 1209.5 m. The numbers above major peaks refer to the carbon numbers of n-alkanes or isoprenoids (i16; i18). Pr and Ph indicate pristane and phytane respectively. M indicates a suite of MMAs (eluting before their corresponding n-alkanes) showing examples of branching positions (3-, 5-, etc.). Open circles indicate 3,7-dimethyl alkanes (only with odd total-carbon-number), eluting before their corresponding n-alkanes and MMAs. See also Table 2.

264

K. Arouri et al. / Precambrian Research 100 (2000) 235280

The higher molecular-weight compounds form a second cluster with a strong odd-over-even carbonnumber predominance most clearly at n-C and 29 n-C ( Table 2). As shown in Table 2, and as 31 explained in detail below, MMAs and DMAs occur over the entire carbon-number range, in contrast to the acyclic isoprenoids, which are restricted to the low molecular-weight range, and to the regular hopanes (C - and/or 29 C -dominant), which are a major component of 30 the higher alkane range. Similar bimodal distributions of saturated hydrocarbons have been reported from the Abu Dhabi cyanobacterial mats ( Kenig et al., 1995). However, whether or not these longer chain n-alkanes have a cyanobacterial source is questionable because cyanobacteria lack fatty acids higher than C (e.g. Paoletti et al., 18 1976). The possibility that these higher molecularweight n-alkanes constitute contaminants from either modern insects or land plants is remote because of stringent laboratory discipline. It is, therefore, possible that these n-C C alkanes 29 33 have subsidiary, or alternative, sources from anoxygenic photosynthetic bacteria, such as species of Chloroexus. These green bacteria, commonly reported as an undermat-community within hypersaline cyanobacterial mats (e.g. DAmelio et al., 1989; Des Marais et al., 1992b) and hotspring cyanobacterial mats (e.g. Ward et al., 1987), produce abundant C C di- and triunsaturated 29 33 alkenes (Des Marais et al., 1992a, g. 6.8.6B). MMAs are present over a large range of carbon number (C up to at least C ), with even-carbon16 33 numbered compounds dominating their oddcarbon-numbered counterparts [Fig. 6(a); Table 2]. MMAs are more abundant in the C 1620 range. All possible structural isomers are present but are dominated by 3- and 5-methyl isomers (5-methylheptadecane predominating; 5Me18/ n-C =0.60.8; Table 2). These latter isomers 18 form the bulk of the MMAs in the higher molecular-weight range. Short-chain MMAs also include the terminally branched (2-methylpentadecane, 2-methylhexadecane, 2-methylheptadecane) besides the commonly reported 7- and 8-methylheptadecanes [Fig. 6(a)]. The terminally branched MMAs are mainly of bacterial origin and the midchain branched homologues, specically the 7- and

8-methylheptadecanes, are particularly from a cyanobacterial source (e.g. Gelpi et al., 1970; see also Section 1.2 herein). Distributions of shortchain MMAs, similar to those found in some of these Tanana Formation mat facies, are reported from the cyanobacterial mats of the Gavish Sabkha (de Leeuw et al., 1985) and Abu Dhabi ( Kenig et al., 1995). In the Abu Dhabi cyanobacterial mats, for example, all the possible isomers (6- and 7-dominant, and to a lesser extent 5- and 4-) have been reported in the short-chain MMA range, whereas the longer-chain isomers carry the methyl substituents at odd-numbered carbon atoms (mainly at C-3 and C-5), with the total even-numbered MMAs slightly dominant. This is very similar to the MMA series identied in these Tanana Formation mats, suggesting a major cyanobacterial input. A part of the biomarker signature of Microbial Facies 1 (namely: evencarbon-numbered MMAs predominating over odd-carbon-numbered MMAs, and branching at the odd-carbon sites such as 3, 5, 7, etc.) has been reported previously in other Proterozoic sediments (Mycke et al., 1988; Logan et al., 1999). Mycke et al. (1988) did not suggest a specic biological source for these MMAs with this particular distribution, but Logan et al. (1999) argue on the basis of a combination of evidence (including depleted carbon and sulfur isotopic values: d13C=34 to 32; 34S =4050), that their progenitor pyrite is likely to have included sulde-oxidising bacteria. Therefore, it is possible that sulde-oxidising bacteria formed part of the consortia of Microbial Facies 1. However, the absence of even-over-odd carbon-number preference in the C n-alkanes 20+ of this facies [in contrast to the pattern observed by Logan et al. (1999)], renders this uncertain. Ten Haven et al. (1988) suggested a heterotrophic bacterial source for the 2-methylpentadecanes and 2-methylhexadecanes compounds also found in the Tanana Formation mats. However, no 2-methyloctadecane is evident in the Microbial Facies 1 extracts. This compound is from a halophytic (particularly methanogenic) source (Brassell et al., 1981). Its absence here probably excludes the Archaea (including halophiles, methanogens, and thermoacidophiles) from being a major contributor to the organic matter in this

K. Arouri et al. / Precambrian Research 100 (2000) 235280

265

mat facies; this inference is supported by signicant concentrations of lower molecular-weight acyclic isoprenoids dominated by pristane and phytane [Fig. 6(a)]. In such cyanobacterial mats, chlorophyll-a is the dominant pigment, giving rise to abundant phytol (e.g. Shiea et al., 1991). No isoprenoids >C (including regular and irregular 22 C and C ) were detected ( Table 2), suggesting 25 30 a photoautotrophic (chlorophyll/tocopherol ) rather than Archaeal (methanogens/halophiles) source for these isoprenoids. DMAs occur exclusively as 3,7-DMAs (or 3,v7-DMAs) with odd-numbered carbon atoms in the range C C [Fig. 6(a)]. This homologous 17 35 series occurs in varying abundances and maximises at C , C and C [Fig. 6(a)]. A similar series 29 31 33 was previously reported as unidentied branched compounds in Proterozoic carbonaceous shales and suldes (Mycke et al., 1988). These unidentied compounds were tentatively assigned to 3,7-DMAs in: (1) Permian carbonaceous shales (Dill et al., 1988); (2) bacterial macromolecules (bacterans) from extant bacteria and marine kerogen ( Flaviano et al., 1994); and (3) Holocene cyanobacterial mats ( Kenig et al., 1995). As for the methylalkanes, the vast majority of DMAs identied in the Abu Dhabi mats are methylated at C-6 (or v6) and C-7 (or v7). A common biosynthetic cyanobacterial source for such MMAs and DMAs has, therefore, been suggested ( Kenig et al., 1995). The association of these odd-carbonnumbered 3,7-DMAs with slightly predominant even-carbon-numbered MMAs (particularly 3and 5-methyl isomers) in the Tanana Formation mats is also peculiar and suggests that these homologous series are biosynthetically related. The oddover-even carbon-number predominance with abundant C and C in the C n-alkanes, and 29 31 20+ C and C in the 3,7-DMAs in these mats is also 31 33 noteworthy ( Table 2). Cuticular waxes of insects are also known to contain 3,7-DMAs (amongst other DMAs) and trimethylalkanes (e.g. Nelson and Blomquist, 1995). Similar (i.e. animal ) sources for these compounds in the Ediacarian Ungoolya Group mat facies is, however, unlikely because insects evolved after the Devonian (Robinson and Kaesler, 1987). Lower, but considerable, concentrations of n-

alkylcyclohexanes were detected in the range C 15 to at least C , with the even-carbon-numbered 28 compounds predominating and maximising at C (Table 2). The n-alkylcyclohexanes are well 18 known from algal-rich rocks (e.g. Hoering, 1976). A bacterial (thermoacidophilic bacterium Bacillus acidocaldarius) origin has also been suggested as a possible source for the C - and 17 C -n-alkylcyclohexanes (de Rosa et al., 1972; 19 Oshima and Ariga, 1975). Similar odd-carbonnumber preference was subsequently reported from Ordovician oils and sediments (Fowler and Douglas, 1984; Fowler et al., 1986) and from Proterozoic sediments (Mycke et al., 1988), implying possible contributions from comparable bacteria. Therefore, the unusual predominance of evencarbon-numbered n-alkylcyclohexanes observed here probably precludes such bacteria from being signicant contributors. This could be a marker for, as yet, unreported cyanobacteria or extinct bacteria. Microbial Mat Facies 1A is highly enriched in C C hopanes, which are dominated by the 27 35 normal C homologues ( Table 2). This is sugges30 tive of a signicant cyanobacterial component amongst other bacterial inputs [see reviews by Summons and Walter (1990) and McKirdy and Imbus (1992)]. Hence, Archaea, purple sulfur bacteria, and most Gram-positive and Gram-negative bacteria (all of which lack hopanoids) were either not signicant contributors to the biomass of this facies or else their signature has been strongly diluted by input from hopanoid-producers. The presence of 25,28,30-trisnorhopane and enhanced relative abundances of homohopanes are indicative of intermittent episodes of anoxic and/or evaporitic conditions in the microbial mat facies. The predominance of C hopane over their C homologues 30 29 is also consistent with the clayey-siltstone character of the host sediment of these biomarkers. Minimal eucaryotic (algal ) input to Microbial Facies 1A is evidenced by low sterane concentrations ( Table 2). The steranes in this mat facies are almost exclusively ethylsteranes, although other steranes (C >C >C ) and diasteranes do 27 28 30 occur. The methylsteranes (including 4-methyl isomers) are relatively abundant, suggesting that marine algal input comprised a mixture of primi-

266

K. Arouri et al. / Precambrian Research 100 (2000) 235280

tive chlorophytes/prymnesiophytes and putative dinoagellates (see also: Summons and Walter, 1990; Summons, 1992). The dinoagellate biomarker dinosterane is reported from the ~260 m.y. older Alinya Formation ( Fig. 2) of the Ocer Basin (Zang and McKirdy, 1994), and several newly recognised acritarchs from approximately coeval rocks of Arctic Canada are also ascribed to dinoagellate cysts based on morphological similarities (Buttereld and Rainbird, 1998). Low sterane/hopane ratios are typical of the Tanana Formation samples ( Table 2). Ratios <0.05 are characteristic of high microbial (including cyanobacterial ) input in carbonateanhydrite facies of anoxic sabkha environments (e.g. Connan et al., 1986). Higher ratios are not very common in pre-Phanerozoic rocks, but have been reported from Neoproterozoic sediments and are attributed to an early evolutionary radiation of the algae (e.g. McKirdy and Imbus, 1992). 3.4.2.2. Subfacies 1B. Subfacies 1B occurs at 1264.8 m (cycle 4) and 1347.4 m (cycle 3; Fig. 3; Table 2). These samples are Facies Group D2 turbidites and contain only detrital fragments of mats (Fig. 3). In comparison with Subfacies 1A, Subfacies 1B contains a similar, but weak cyanobacterial biomarker signal, overwhelmed by other bacterial and Archaeal signatures. The TIC traces of the saturated hydrocarbons are largely unimodal, maximising at n-C or n-C , and decreasing 16 18 dramatically with increasing carbon number. MMAs and particularly DMAs are generally of low molecular weight (C C ), maximising, 16 24 respectively, at 3- and/or 5-methylnonadecanes, and 3,7-dimethylnonadecane ( Table 2). Traces of 2-methyloctadecane are also present along with enhanced abundances of acyclic isoprenoids and C - and C -dominant n-alkylcyclohexanes 17 19 ( Table 2). Other biomarkers, present in higher abundances than in Subfacies 1A, are: C >C hopanes, 29 30 ba-hopanes (moretanes), and 2a-methylhopanes ( Table 2). The 2a-methylhopanes are believed to be derived from methylotrophs (Bisseret et al., 1985; Summons and Jahnke, 1992), but can also be cyanobacterially derived (Summons et al.,

1998). Recent analyses of cultured cyanobacteria, as well as of Proterozoic oils and rocks, reveal a strong correlation between high abundances of these compounds and low molecular-weight normal and branched alkanes, suggesting a common cyanobacterial source for both groups of compounds (Summons et al., 1998, and written communication, 1999). Both assumptions (cyanobacteria and methylotrophs), in association with the absence of C C homohopanes ( Table 2), 33 35 are consistent with biomass accumulation in a more oxic depositional environment. A large unresolved complex mixture-envelope evident in the TIC traces (not shown) of this subfacies may be related to extensive microbial reworking of the pre-existing (presumably cyanobacterial ) substrate, thereby diluting their original biomarker signature. Framboidal pyrite is present sporadically within the detrital kerogen of this subfacies, but it is unclear whether this formed prior to, or subsequent to, emplacement of the detrital mat fragments in the sediment. In kerogen strew-slides of the two samples with this biomarker anity, lamentous structures and fewer acritarchs are present but are subordinate to large amorphous organic sheets (Grey, 1998). Sterane algal biomarkers are minor in this subfacies. Their dominance by ethylsteranes (C ) and 29 the absence of 4-methylsteranes is accompanied by low acritarch diversity (and a dominance of the species M. pelorium, especially in sample 1347.4 m). This conrms earlier suggestions that this particular acritarch is of chlorophycean anity and is not a dinoagellate (Arouri et al., 1999). 3.4.3. Microbial Mat Facies 2 3.4.3.1. Introductory remarks. The lighter-coloured phases of the core samples that yield this biomarker facies: (1) exhibit cyanobacterial signatures similar to those that characterise some modern hot-spring and hypersaline lagoonal cyanobacterial mats; and (2) lack authigenic pyrite. In contrast, the darker-coloured phases contain bacterial and Archaeal mats, which, based on the relative abundance of authigenic pyrite, suggest the involvement of sulfate-reducing bacteria. In one sample (1438.3 m; a shallow-water

K. Arouri et al. / Precambrian Research 100 (2000) 235280

267

siltstone from the very top of cycle 2), autochthonous cyanobacterial mats are associated with a biomarker characteristic of modern green nonsulfur anoxygenic photosynthetic bacterial mats. The biomarker association of Microbial Mat Facies 2 is best represented by sample 1443.6 m, a shoalwater carbonate from towards the top of cycle 2, but is also in evidence at various other levels and in dierent lithofacies: 1234.0 m (Facies D1.1; cycle 4); 1252.2 m ( Facies D1.1; cycle 4); and 1438.3 m (a shallow-water siltstone from the top of cycle 2; Fig. 3). In thin-section, samples 1443.6 m and 1438.3 m are seen to contain autochthonous mats [Figs. 3 and 4( E)(G)], as well as dispersed gel-like kerogen in the case of sample 1443.6 m. SEM images of slightly HF-etched bedding-planes within the dark-coloured phase of the latter sample reveal abundant laments, as well as nely disseminated organic matter, large organic amorphous fragments, acritarchs, and ne carbonate crystals of probable microbial-mediation origin (Fig. 5). As summarised in Fig. 3, samples yielding this biomarker assemblage represent both shallowwater (probably peritidal ) settings on the one hand (i.e. samples 1443.6 m and 1438.3 m), and deeper-water, upper-turbidite-slope environments on the other hand (samples 1252.2 m and 1234.0 m). The latter samples are debrites sourced from autochthonous mats of the high-energy upper turbidite slope and contain only detrital mat fragments and mat-bound silt-substrate intraclasts [Fig. 4(B)]. 3.4.3.2. Biomarker characteristics of the light-coloured phases. The lighter-coloured phases of these samples, as seen in hand-specimens ( Table 2), are characterised by restricted MMA distribution, where only C and C (and probably C ) MMA 16 18 20 isomers are detected, which are dominated by 2-, 3-, and 5-methyl isomers. Unlike Facies 1, no C , C , or C MMAs are evident [ Table 2; 15 17 19+ Fig. 6(b)]. Similar MMAs with restricted distributions were reported from the Orakei Korako (New Zealand ) hot-spring cyanobacterial mats, which are built predominantly by the cyanobacterium Chlorogloeopsis sp. (Shiea et al., 1990, 1991). Similar hydrocarbons were also found in other hot-spring (Dobson et al., 1988; Robinson and

Eglinton, 1990) and lagoonal (e.g. Boon et al., 1985) cyanobacterial mats. The short-chain oddcarbon-numbered 3,7-DMAs are minor or not detectable ( Table 2). Other unidentied branched compounds also occur [marked with asterisks in Fig. 6(b)], with evidence for multiple branching, as inferred from their retention times and mass spectra. Isoprenoids (<C ) are present in only 22 trace amounts except for minor amounts of pristane and phytane. A strong predominance of evencarbon-numbers is evident in the n-alkanes <C , 20 maximising at n-C and/or n-C , and an 16 18 odd-over-even carbon-number preference in the higher range [Fig. 6(b)]. Hopanes (C - or 29 C -dominant) are abundant, along with traces of 30 TTs ( Table 2) and varying concentrations of 2a-methylhopanes (Table 2). Steranes are minor and are dominated by ethylsteranes (C 29 sterane/C hopane<0.07; Table 2) and commonly 30 include abundant 2- and 4-methylsteranes (e.g. 4aS methylsterane/C aaS sterane=0.7). 29 The implied major cyanobacterial and minor algal biomarker signatures are accompanied by other strong bacterial signals (elaborated below). This is in contrast to hydrocarbons from purely cyanobacterial mats and cyanobacterial cultures (e.g. Shiea et al., 1990; Robinson and Eglinton, 1990), but is similar to hydrocarbons from anoxygenic photosynthetic bacterial mats (e.g. those of the mesophilic green non-sulfur Chloroexus aurantiacus) which lack mid-chain branched alkanes and contain n-octadecanol superseding phytol (Shiea et al., 1991; see Section 1.2 herein). This probably parallels the dominance of n-C over phytane in this mat 18 assemblage [the low relative abundance of n-C 17 accompanying n-C zphytane; cf. the opposite 18 relationships in Microbial Facies 1A; see Fig. 6(a) and (b)], and may indicate the involvement of green non-sulfur bacteria in building these mats. An example of the occurrence of Chloroexus sp. in association with cyanobacterial population is the Icelandic hot-spring (4366C ) microbial mats, commonly dominated by Mastigocladus laminosus or Oscillatoria/Phormidium sp. (Robinson and Eglinton, 1990). Although purely anoxygenic photosynthetic bacterial mats are known from modern hot-springs only, their accumulation in

268

K. Arouri et al. / Precambrian Research 100 (2000) 235280

other anoxic, suldic- or H -rich environments, 2 including the Proterozoic examples, is possible ( Ward et al., 1992). Terminal anaerobic food-chain consumers in samples of this facies seem to have been in low relative abundance. This notion is evidenced by: (1) the relative rarity/absence of framboidal pyrite (suggesting little bacterial sulfate-reduction), at least in respect of samples 1234.0 m, 1252.2 m, and 1438.3 m (which are either entirely or predominantly siliciclastic and hence would not have constituted iron-poor diagenetic systems, even though the timing of the pyrite formation is indeterminable, especially where the mat material is detrital see also Fig. 3); and (2) the minor presence of C - and C -dominant n-alkylcyclohexanes 17 19 ( Table 2) currently known only from thermoacidophilic bacteria (de Rosa et al., 1972; Oshima and Ariga, 1975; see also Mycke et al., 1988). An algal origin was also inferred in Phanerozoic sediments and oils for n-alkylcyclohexanes and methyl-nalkylcyclohexanes, with the former series possessing an odd-carbon-number preference, but these are commonly accompanied by a similar oddcarbon-number predominance in the n-alkanes (e.g. Fowler and Douglas, 1984; Fowler et al., 1986) a feature not observed in the Tanana Formation mat samples examined ( Table 2). 3.4.3.3. Biomarker characteristics of the dark-coloured phases. The darker-coloured phases in samples 1443.6 m and 1438.3 m comprise silt laminae with autochthonous mats, whereas the lightercoloured phases comprise mainly ne-grained carbonate without optically observable mats [Fig. 4( E )(G)]. Thin-section observations reveal that the dark coloration of these horizons is due to the presence of pervasive kerogen, but that it is enhanced in the case of sample 1443.6 m by the presence of abundant authigenic pyrite [Fig. 4(F ) and (G)]. The biomarker signatures from the darker phases of the Microbial Facies 2 samples are suggestive primarily of non-cyanobacterial mats. The molecular signatures are dierent than those observed in the lighter phases, and include: (1) MMAs (C C ) and 3,7-DMAs (mainly C ), 16 20 25 present as traces only; (2) no predominance of

odd- or even-carbon number in the n-alkylcyclohexanes; and (3) no methylhopanes among the other abundant hopanes and homohopanes. Collectively, these biomarkers indicate a primarily non-cyanobacterial community (morphological details of which are illustrated in Fig. 5). The abundance of framboidal pyrite in these dark phases [Fig. 4( F ) and (G)] indicates a signicant contribution from sulfate-reducing bacteria. As a general observation, the cores of many framboids can be seen to consist of organic material once the outer layer of the framboids has been dissolved during palynological preparations, suggesting that these structures were nucleated around saprophytic bacteria. Additionally, the minute carbonate crystals that are suggestive of a microbial origin [Fig. 5( E) and ( F )] are also present within the dark phase of sample 1443.6 m. The presence of such carbonate suggests a warm palaeoclimate [see Gerdes et al. (1991), p. 602; see also Buczynski and Chafetz (1993)]. Detailed studies of modern biogenic varvites led Gerdes et al. (1991) to conclude that: There is no in situ production of carbonates in biolaminations of the temperate humid climate zone, even not in biogenic varvites of salterns which grow under similar hypersaline conditions as in lower latitudes. 3.4.4. Planktonic Algal Facies The stratigraphically highest part of the Tanana Formation in Munta 1 examined for biomarkers (e.g. sample 1209.5 m; cycle 4) represents a predominantly non-bacterial organic facies. Thin-section study shows only dispersed organic matter (DOM ), predominantly detrital akes, whereas kerogen slides show that the DOM consists primarily of acanthomorph acritarchs. About 40 species have been identied, including a large number of new species (Grey, 1998). The TIC trace of the extracted saturated hydrocarbons [Fig. 6(c)] reveals a restricted range of compounds ranging between n-C and n-C with even-over-odd 15 22 carbon-number predominance in n-alkanes <C . 20 The predominance of n-C is consistent with a 20 non-bacterial input, and is a feature commonly seen in some hypersaline or carbonate environments (ten Haven et al., 1988). It is in this upper

K. Arouri et al. / Precambrian Research 100 (2000) 235280

269

(cycle 4) part of the Tanana Formation where the steranes are in their highest relative abundances (515 times higher than those in the lower part of the formation; Table 2). Steranes are dominated by C homologues, but C compounds are also 29 27 signicant. This possibly reects abundant chlorophytes with rare input from rhodophytes. A single lamentous colony of a probable rhodophyte alga was found by Grey (1998) in a palynological preparation from this depth range in Munta 1 (specically at 1211.10 m). The uppermost clastic part of the Tanana Formation comprises only the deposits of low-concentration, silt-dominated turbidity currents (Facies D2.1 and D2.3), emplaced at only moderate water-depths below the pycnocline on the high-energy upper slope ( Fig. 3). Samples from this interval (i.e. samples 1198.6 m and 1209.5 m) show bright red uorescence under green-UV excitation in the coarser basal silt phases of the turbidites, and only very dull or no red uorescence in the clay-caps (where present) of these same beds ( Fig. 3). This observation suggests that the source of this algal biomarker signal constitutes redeposited pelagic material sourced from shallower upslope environments above the pycnocline and did not accumulate through pelagic (inter-turbidite) rain, because these samples either lack, or have only extremely thin, clay-caps to the very-thin turbidite beds. Munta 1 sample 1216.0 m from the immediately underlying interval that comprises the deposits of silt-dominated debris-ows and high-concentration turbidity currents also has some biomarker anities with the Algal Facies ( Table 2; Fig. 3), and, as already explained (in Section 3.1.5), can be reliably interpreted to have been sourced from the upper epiclastic slope above the pycnocline. 3.5. d13C kerogen Kerogen of the Tanana Formation is somewhat enriched in 13C (d13C=26.2 to 29.9; Table 2) relative to other Neoproterozoic noncyanobacterial mats (cf. Calver, 1995; Logan et al., 1999). As reported by these authors, a larger depletion in 13C of the benthic mat carbon (d13C=32.1) relative to that of the background rock-matrix (d13C=26.5) is interpre-

ted by them as a sulde-oxidising bacterial signal. The lack of anomalous 13C depletion, accompanied by low TOC, led Calver (1995) to suggest that either (1) the mats comprise only a minor percentage of the total biomass, or (2) that they are isotopically the same as associated planktonic organic matter. Our current palynological and petrographical observations favour the rst assumption, i.e. the benthic microbial mats are found at only very few horizons (Fig. 3). There is very abundant, optically unresolvable (at least under ordinary white-light illumination) DOM present as background kerogen within at least some of the samples studied here. This is evidenced by (1) its abundance in the <25 mm palynological residues, and (2) its appearance as stylocumulate in the ne-grained limestones where these have been tectonically/diagenetically overprinted by the development of stylolites [Figs. 3 and 4(D)].

4. Discussion and conclusions 4.1. Slope-apron anity of the basin-margin palaeophysiography in the vicinity of Munta 1 Reconstruction of the likely palaeophysiography of the southern structural-ramp-margin of the Munyarai Trough is signicant in understanding the likely sedimentary history of the lower Ungoolya Group in Munta 1 and can be made on the basis of the following considerations. (1) The regional basin-architectural setting of Munta 1 is toward the base of the structural ramp that separates the Murnaroo Platform shelf on the south and southeast, and the Munyarai Trough to the north and northwest (Section 2.1). (2) Most of the lower Ungoolya Group succession in Munta 1 comprises relatively deep-water silt-dominated hemipelagites and ne-grained turbidites and debrites, as well as intercalated chemopelagites, many of which show evidence of micro-scale and macroscopic soft-sediment deformation suggestive of slope-induced penecontemporaneous movement at all scales (elaborated in Section 3.1.4.2). Moreover, at the regional scale, the lower Ungoolya Group (Rodda Beds) succession has been described as comprising predominantly

270

K. Arouri et al. / Precambrian Research 100 (2000) 235280

slumped carbonaceous siltstone in which the numerous slump-folded units (individually as much as 20 m thick) indicate syntaphral movement on a northerly palaeoslope (Preiss et al., 1993, pp. 200202). (3) Evidence of various kinds indicates that sediment provenance on the southern ank of the eastern Ocer Basin was from the south (i.e. from the Gawler Craton region) and that progradation of the developing sediment prism in the Munyarai Trough was regionally from south to north (Section 2.1). Indeed, there is little evidence of sediment inux into the eastern Ocer Basin from the north during its late Ediacarian history and Lindsay and Leven (1996, p. 420) state that during accumulation of the M2 Megasequence: The basin lled by progradation from south to north [and the] Murnaroo Platform appears to have been largely bypassed by sediments. (4) Compressional tectonics during accumulation of the lower Ungoolya Group produced faults within the crystalline basement and overlying basin-ll throughout the structural-ramp zone anking the Munyarai Trough, induced salt mobilization from the M1 Megasequence rocks and produced large areas of complex bathymetry superposed on the regional northwesterly palaeoslope that separated the Murnaroo Platform shelf on the southeast and the Munyarai Trough on the northwest (Section 2.1). These structural movements were probably accompanied by seismic events that would have caused sediment slumping. (5) No signicant bodies of coarser-grained clastics are known to occur within the lower Ungoolya Group such as might constitute more proximal deposits of slope-canyon feeder-conduits for the ne-grained turbidites; nor are such structures known to occur within the lower Ungoolya Group on the basis of reection-seismic studies (e.g. Lindsay and Leven, 1996). (6) The absence of coarse (sandy) epiclastics and the predominance of ne-epiclastics and chemopelagic carbonates throughout the lower Ungoolya Group succession (Fig. 3), spanning a time interval of some 15 m.y., indicate overall conditions of sediment starvation. These various lines of evidence suggest that the lower Ungoolya Group turbidites and pelagites were emplaced as sheets or aprons on an un-canyoned incline of gentle/moderate gradient

(but that must have steepened along sedimentary strike to the northeast on the anks of the Ammaroodinna Ridge), with sediment supply either from a linear source or from multiple sources located upslope on the Murnaroo Platform shelf to the south and southeast. The transfer of sediment from this/these source(s) perhaps involved numerous relatively small-scale slope-gullies that functioned as the main feeder conduits (such as the palaeogullies that appear to be common in cycle 4 in the upper part of the Tanana Formation; Fig. 3; Calver, 1995; Morton, 1997b). Moreover, gravitational failure and down-slope movement of these muddy slope sediments was evidently common, as indicated especially by the prevalence of disturbed beds in the upper part of cycle 4 in Munta 1 ( Fig. 3), as well as regionally within the lower Ungoolya Group succession (Preiss et al., 1993, pp. 200202, and g. 6.13). In summary, the collective evidence suggests that the lower Ungoolya Group in Munta 1 accumulated on a silt-dominated, sediment-starved, non-canyoned northwesterly palaeoslope of low to moderate gradient that constituted either a turbiditic submarine ramp or slope apron in the sense of Reading and Richards (1994) and Stow et al. (1996). Reading and Richards (1994) proposed a conceptual model and terminology for basin-margin turbidite sedimentation based on sediment calibre and feeder-system geography. This scheme recognises three dierent physiographic arrangements: (1) a pattern of multiple sources feeding a deepbasin margin, known as a submarine ramp; (2) a linear source feeding a deep-basin margin, known as a slope apron; and (3) a point-source feeding the base-of-slope on a deep-basin margin known as a submarine fan. Many point-sourced systems, particularly sand-dominated and sandmud-dominated systems, have a feeder-system that involves a moderate- to large-scale shelf-canyon, but the linear- and multiple-sourced systems do not (Reading and Richards, 1994, Table 2). As mentioned by these authors, distinguishing between a submarine ramp and a slope apron in the ancient record is not always easy, particularly within systems at the coarser end of the spectrum. Previous authors have used the terms clinoform (e.g. Woodrow and Isley, 1983) and turbidite slope

K. Arouri et al. / Precambrian Research 100 (2000) 235280

271

(e.g. Lundegard et al., 1985) for such linear/multiple-sourced, low- to moderately high-gradient, non-canyoned types of basin-margin. Turbidite submarine-ramps and slope aprons occur on some present-day continental margins, but they are also known to characterise the margins of intracontinental basins, including foreland basins (Stow et al., 1996). In foreland basins (such as was the Ocer Basin during most of its postsag-basin-phase history) the non-orogenic margin commonly evolves as a structural ramp, which, in the case of deep marine inundation of the foredeep, can be perpetuated as a physiographic and sedimentological ramp/slope-apron. Such a history would appear to apply in the case of the nonorogenic southern margin of the intracratonic Ocer Basin discussed here. Additionally, a physiographic ramp can also develop along the orogenic side of a deep-marine foreland basin through progradation of the piedmont wedge, whose submarine slope into the foredeep can, depending on individual circumstances, constitute either a turbiditic submarine-ramp or a slope-apron, as exemplied by the Devonian Catskill Delta margin of the Appalachian Basin of North America ( Woodrow and Isley, 1983; Lungegard et al., 1985; Woodrow, 1985). The Catskill Delta margin of the Appalachian Basin constituted a submarine-ramp in the terminology of Reading and Richards (1994), where multiple-fed turbidity currents owed downslope with a gradient estimated to be 2.5 m km1 (Stow et al., 1996, p. 435). The geographic width of slope aprons is said to range from 2 to 200 km with relatively high gradients (10150 m km1) (Stow et al., 1996, p. 438). In the case of the sedimentological ramp/slope-apron that was pregured by the Murnaroo Platform structural ramp, the likely width of the physiographic incline that separated the Murnarooo Platform shelf and the axial foredeep of the Munyarai Trough probably ranged from a few tens of kilometres in the northeast (Ammaroodinna Ridge area) to 60100 km in the southwest (i.e. southern anks of southwest Munyarai Trough and Birksgate Sub-basin). The slope-gradient of the constructional physiographic features of such basin-margin turbidite systems is

mainly a function of the sediment calibre produced by the feeder system. In mud/silt-dominated systems (such as characterised the lower Ungoolya Group) the gradient is said to range from moderate (2.525 m/km) in submarine-ramps to moderate high (40150 m/km) in slope aprons (Reading and Richards, 1994, table 2). Various sedimentological, facies-architectural and scalar attributes are used by Reading and Richards (1994, tables 2 and 3) to distinguish among the three major turbidite systems and their dierent sediment-calibre variants. Reference to these criteria in respect of the lower Ungoolya Group sediments suggests that they accumulated in a mud/silt-dominated slope-apron system (rather than in a submarineramp system), and that the feeder system was a wide, starved shelf (i.e. the Murnaroo Platform shelf ), not a shelf fed by deltas that reached the basin-margin ramp (as suggested by previous workers, e.g. Calver and Lindsay, 1998). According to Reading and Richards (1994, p. 813), although the hinterland source-areas of the shelves that border slope aprons may be large, the slope aprons lack coarse epiclastic sediment either because: (1) such sediment is trapped in canyon heads and transported down into the associated base-of-slope submarine fans (as in the typical California borderlands basins); or (2) because in the absence of shelf-canyons, such sediment starvation of broard shelves is eected primarily by arid climate in the neighbouring continental hinterland. The likelihood of an arid climatic regime during sedimentation of the lower Ungoolya Group was reached independently (in Section 3.1.3.1) on the basis of mineralogical criteria and elsewhere throughout this analysis on the basis of other evidence. 4.2. Sedimentary history of the lower Ungoolya Group succession in Munta 1 4.2.1. General overview Within the limitations of the stratigraphic sample control available in this study and the lack of a detailed sedimentological log of the lower Ungoolya Group succession, the sequence of inferred depositional environments of the lower Ungoolya Group in Munta 1 ( Fig. 3) allows provi-

272

K. Arouri et al. / Precambrian Research 100 (2000) 235280

sional reconstruction of the history of the basinll and possibly also that of associated climate control. At the broadest scale, this history reveals a succession of four vertically stacked hectometrescale slope-apron cycles (cycles 14), each terminating in a cap of shoalwater/peritidal carbonates of variable thickness (Fig. 3). Subsidence and rapid marine transgression across the shallowwater carbonate platform that terminates each cycle caused retrogradation of the slope-apron system and the commencement of the next cycle (Fig. 3). These individual slope-apron cycles correspond broadly with the lower Ungoolya Group stratigraphic sequences (i.e. LU14) recognised by Calver and Lindsay (1998, gs. 5, 7) with the following qualications. (1) In the Calver and Lindsay (1998) scheme their sequence LU1 begins at the base of the Dey Dey Mudstone. In the scheme of slope-apron cycles identied here, the base of cycle 1 begins 123 m above the base of the Dey Dey Mudstone where the rst distal deposits of this slope-apron can be identied to prograde across the starved basin-oor deposits of the Munyarai Trough. (2) Calver and Lindsay (1998) place their sequence boundaries at the base of the carbonate units. In the present slope-apron cycle scheme, with the exception of cycles 1 and 2 (whose mutual boundary lies within a stratigraphically continuous carbonate interval, but one involving a change from probable peritidal platform carbonate below to deeper-water chemopelagics above), the boundaries that separate each slope-apron cycle are placed at the top of the carbonate units (and in the case of the boundary separating cycles 1 and 2, at the top of the peritidal carbonate interval ). This means that the boundary used here to separate slope-aprons 1 and 2 is placed at about mid-level within the carbonate interval that begins the Calver and Lindsay (1998) sequence LU2 (i.e. at about mid-level within the carbonate interval that constitutes the Karlaya Limestone and the basal carbonates of the Tanana Formation). (3) In the Calver and Lindsay (1998) sequence-stratigraphic scheme the sequence boundaries that they place at the base of each of the limestones of the lower Ungoolya Group succession are identied in part on the basis of carbon-isotope (d13C and org d13C ) chemostratigraphic excursions and/or carb

upsection breaks in the chemostratigraphic curves, and each limestone is interpreted to manifest a transgressive systems tract deposit. In the slopeapron cycle scheme proposed here, the transgressive systems tract deposits are constrained to occur at the base of each new slope-apron cycle, i.e. to the strata immediately above the shoalwater carbonate intervals. In this scheme the carbon-isotope excursions that Calver and Lindsay (1998) detect at the base of each shallow-water limestone are chemical and biochemical eects of the presence of the pycnocline. The attributes of the pycnocline can be expected to have varied with the mean atmospheric temperature and degree of intensity of aridity. The sedimentary basin-ll history of the lower Ungoolya Group in Munta 1 is summarised in Fig. 3 in terms of its ve major phases and is only very briey further elaborated here. The ve major stages are, in chronological order: (1) the initial starved marine-basin phase following rapid subsidence and marine transgression (interval 1973 1850 m; lower redbrown ssile mudrock part of the Dey Dey Mudstone); (2) slope-apron cycle 1 (interval 18501630 m; upper drab-coloured part of the Dey Dey Mudstone and overlying shallowwater part of the Karlaya Limestone); (3) slopeapron cycle 2 (interval 16301430 m; upper drabcoloured deeper-water silty part of the Karlaya Limestone through the basal one-third of the overlying Tanana Formation); (4) slope-apron cycle 3 (interval 14301305 m; mid-one-third of the Tanana Formation); (5) slope-apron 4 (interval 1305>1152 m; upper one-third of the Tanana Formation incomplete because of erosional truncation by the canyon-cutting event). The thickness of the individual slope-aprons ranges from 125 m (cycle 3) to 220 m (cycle 1) and their cumulative thickness in Munta 1 is 690 m ( Fig. 3). Their average thickness is ~175 m, a dimension that can be used as a minimum estimate of the average water depth associated with the slope-apron system as a whole (as distinct from that associated with each one separately; this would have varied from 125 m to 220 m, but more if one takes into consideration the slightly up-ank location of Munta 1 relative to the axis of the Munyarai Trough). The vertical stacking of these

K. Arouri et al. / Precambrian Research 100 (2000) 235280

273

separate slope-apron deposits evidently manifests an interplay of long-term basin-subsidence rates and eustatic movements in sea-level in a structuralramp setting, thus facilitating a pattern of alternate slope-apron progradation and aggradation followed by renewed transgression and retrogradation [as previously described by Calver and Lindsay (1998)]. 4.2.2. Polycyclic slope-apron history The upsequence progression of environments and palaeophysiographies recorded in Munta 1 begins with an arid coastal-plain (top of Murnaroo Formation and base of Dey Dey Mudstone) that was transformed by rapid subsidence and marine transgression into a moderately deep, sedimentstarved basin (water depth in the range of several hundred metres). During the early history of this basin the Munta 1 site occupied the bathymetrically distal slopes of the Munaroo Platform structural and physiographic ramp, northwest of a developing clastic slope-apron that prograded northwest towards the axial foredeep of the Munyarai Trough in which accumulated thoroughly oxidised (redbrown) ssile mudrock (sediment-starved marine basin phase in Fig. 3). This slope-apron (cycle 1) makes its rst appearance at the Munta 1 site very shortly after the Acraman Impact Event and deposits of both its distal ( lower slope) and proximal (upper slope) facies contain reworked Acraman Impact debris emplaced as wind-blown dust in the former, and in both lowand high-concentration mass-ows sourced from the Murnaroo Platform shelf in the latter (Fig. 3; Section 3.1.3.2). [ Upper and lower slope-apron facies are dierentiated on the basis of the criteria outlined in Pickering et al. (1989, p. 93).] By ~575 Ma this rst clastic slope-apron at the Munta 1 site had accreted upward to probable shoalwater depths above a deep pycnocline that had developed as an artifact of the coeval intensication of aridity and heightened epiclastic sediment starvation of the basin. These circumstances initiated carbonate deposition, at the Munta 1 site initially as the slope deposits of what had now become a carbonate slope-apron depositional system (see Pickering et al., 1989, table 5.1). This carbonate slope-apron shallowed up to peritidal

depths at the Munta 1 site, but just prior to ~570 Ma (top of the Karlaya Limestone) it had begun to subside to accumulate silty slope-carbonates belonging to the transgressive basal deposits of the next cycle (slope-apron 2; basal Tanana Formation). The extreme aridity and related tight epiclastic sediment budget that had initiated deposition of the Karlaya Limestone at the close of cycle 1 accompanied deposition of about one-third of cycle 2. Slope-apron 2 had much the same amount of bathymetric/physiographic relief as slope-apron 1 (~200 m), but the pycnocline that accompanied its progradational and aggradational build-up to peritidal levels had only half the depth (i.e. ~25 m; Fig. 3) that accompanied build up of the shoalwater deposits of slope-apron 1. Moreover, the evident absence of extensive sedimentological evidence of normal upper-slope higher-energy conditions in cycle 2 indicates that the same level of possibly noisy perturbation of the pycnocline that had characterised accumulation of the upper slope-apron deposits of cycle 1 did not recur during accumulation of the upper slope-apron deposits of cycle 2. Slope-apron 3 followed abruptly the demise of slope-apron 2, but its bathymetric/physiographic relief was only slightly half (~125 m) that of cycles 1 and 2 ( Fig. 3). Like cycle 2, its upper slope deposits lack compelling evidence of a noisy pycnocline (which was, also like that of cycle 2, also relatively shallow: ~25 m). Slope-apron 4 (the nal one preserved in the lower Ungoolya Group at the Munta 1 site) followed abruptly slope-apron 3, and had physiographic/bathymetric relief in the order of 160 m ( Fig. 3). Its pycnocline was at a palaeodepth of at least 30 m (minimum estimate due to erosional loss of the topmost part of its platform carbonate interval; Fig. 3). This evidently represents a depression of the pycnocline relative to those that accompanied the close of cycles 2 and 3, and that it was impressively noisy is witnessed by an abundance of sedimentological evidence in the upper slope-apron deposits of cycle 4, such as have been mentioned earlier in Sections 3.1.4 and 3.1.5 and are summarised in Fig. 3. These include the extensive occurrence of disturbed (slumped) bedding, the conspicuous slope-gully truncation surfaces, the upward increase in bed-thickness and

274

K. Arouri et al. / Precambrian Research 100 (2000) 235280

grain-size, the prevalence of Facies D1.1 massowed debrites with contained allochthonous (type 2A) microbial mats, and the relatively high frequency of occurrence of hummocky and other bedforms indicative of active sediment-laden bottom currents. Following completion of deposition of the peritidal carbonates and associated red mudstones that complete slope-apron 4, the last formation of the upper Ungoolya Group (i.e. the Munyarai Formation; Fig. 2) was possibly deposited as an additional slope-apron cycle [see Calver and Lindsay (1998), g. 7] prior to its erosional removal (together with the uppermost part of slope-apron 4) by the late Ediacarian canyoncutting event. 4.3. Organic sedimentology and biomarker geochemistry Organic matter is present in all 27 samples examined. It occurs predominantly as: (1) rare to very abundant detrital sheet-like fragments of kerogen tissue and mat-bound silty-sediment-substrate intraclasts; (2) allochthonous structurally intact mat-bound sediment in large intraclasts; (3) structurally disrupted, allochthonous mats and intraclastic mat-bound silt-substrate-enclaves within silt-dominated debrites and/or high-concentration turbidites; and, less commonly, (4) completely autochthonous mats that preferentially colonised siliciclastic silt substrates in both deepand shallow-water environments, and within thin siliciclastic silt laminae in ne-grained shallowwater (peritidal ) carbonates. DOM is ubiquitous in <25 mm palynological residues, and is also present as optically resolvable interstitial/graincoating kerogen and as stylocumulate in limestone. Detrital mat material appears, on average, to be just as abundant in sediment that accumulated from suspension-load rain on the lower slopeapron ( Facies E2.2) under regimes of background sedimentation as it does in turbiditic sediments (Facies Groups D1 and D2) deposited in the same parts of the slope-apron system under event-dominated regimes ( Fig. 3). The source of much of the allochthonous and detrital mat material that is re-deposited on the upper and lower slope-apron was in situ mats that grew in the shallow photic

zone at the top of the upper slope-apron at depths above and straddling the pycnocline. The depth of the pycnocline varied with time, but ranged between ~25 and 60 m. Deep pycnoclines were especially perturbed by seiches that collided with the seaoor in the zone of growth of microbial mats, causing extensive sediment disturbance, slumping, fragmentation, transfer and re-deposition of the dislodged mat fragments to lower parts of the slope-apron by downslope massow processes, and through widespread pelagic settling of suspension-load nes via interow dispersion along the pycnocline. Biological diversity in the Tanana Formation mat facies is reected by diverse biomarker associations. Based on biomarker and kerogen d13C data, the mats are inferred to be predominantly of cyanobacterial anities in Microbial Facies 1 and 2. Subfacies 1A is characterised by wide distributions of saturated hydrocarbons, including abundant n-C , n-C and n-C alkanes, monomethyl 17 29 31 alkanes (mainly 3- and 5-methyl isomers, but also include 7- and 8-methylheptadecanes), and oddcarbon-numbered 3,7-dimethyl alkanes. Contributions from other bacteria are indicated by abundant hopanes and even-carbon-numbereddominated n-alkylcyclohexanes, whereas minimal algal inputs are reected by low sterane abundance. Facies 1 mat-assemblages, in general, accumulated variously within both peritidal and deep-water ( lower and upper slope-apron) environments ( Fig. 3), in the latter of which the mat material is exclusively detrital, presumably derived from reworked, formerly autochthonous, shallow-water, predominantly cyanobacterial mats. The cyanobacterial biomarker signal of Subfacies 1A can also be detected weakly in the samples referred to Subfacies 1B, indicating that these latter samples are biomolecular palimpsests, as the signal is largely diluted and masked by other bacterial, and possibly by Archaeal, contributions. The kerogen in Subfacies 1B samples predominantly comprises large amorphous organic sheets with minimal lamentous structures and few acritarchs, associated with only sporadic authigenic pyrite. Modern analogous cyanobacterial mat examples with similar biomarker assemblages to those of Subfacies 1A described here are known from the Gavish Sabkha

K. Arouri et al. / Precambrian Research 100 (2000) 235280

275

and Abu Dhabi; in the latter examples the mats are typically built by the lamentous cyanobacterium M. chthonoplastes, commonly overlain by another lamina of gelatinous slime-producing cyanobacteria, and underlain by lamentous (usually green non-sulfur) bacteria and purple sulfur bacteria. The same species consortia are also known from mats at Solar Lake, Sinai, and Guerrero Negro saltern, Baja California Sur, Mexico. Other benthic microbial mat horizons ( Facies 2) within the Tanana Formation, as seen in thinsection and SEM images, reveal abundant coccoidal structures, laments, lamentous sheaths, ne DOM, large amorphous fragments and acritarchs, as well as possible microbially mediated ne carbonate crystals. Similar assemblages of organic remains are obtained in palynological preparations. Most samples that are categorised within this facies were separately analysed for their light- and dark-coloured phases, where present. The lighter-coloured phases in these samples variously comprise ne-grained carbonate, and ne siltstone. The dark-coloured phases comprise autochthonous mat-bound siltstone and intraclasts. Facies 2 exhibits a cyanobacterial biomarker signature dierent than that of Facies 1, suggesting dierent cyanobacterial species input. The light-coloured phases contain a dierent suite of biomarkers to Facies 1, and can be attributed to microbial mats that formed in shallow-marine (including peritidal ) environments as autochthonous mats, but which, in some samples, also have detrital fragments and mat-bound silt-substrate intraclasts re-deposited into deeper-water environments (upper slope-apron; Fig. 3) by high-sediment-concentration mass-ow processes. The most striking biomarker dierences of the light-coloured phases of this facies compared with Facies 1 are: (1) the restricted MMA distribution (only C , 16 C and, occasionally, C MMAs) similar to 18 20 some modern lagoonal and hot-spring cyanobacterial mats built predominantly by the cyanobacterium Chlorogloeopsis spp.; (2) the strong predominance of even-carbon-numbered nalkanes <C (e.g. predominant n-C with n-C 20 18 17 nearly absent); (3) the presence of a prominent series of unidentied branched compounds but with only minimal or no 3,7-dimethyl alkanes and

isoprenoids (including pristane and phytane). Collectively, these biomarker signatures suggest the involvement of green non-sulfur bacteria in building these mats. Biomarkers from terminal anaerobic consumers (e.g. sulfate-reducers and Archaea) are very faint, as witnessed circumstantially by the absence or rarity of authigenic pyrite and only minuscule presence of C - and 17 C -dominant n-alkylcyclohexanes. 19 The biomarker signature of the dark-coloured phases of Facies 2 is essentially non-cyanobacterial. The presence of abundant pyrite framboids in all samples of deep-water anity in this facies is circumstantially consistent with signicant contributions from sulfate-reducers. The organic core of these pyrite framboids may suggest their nucleation around saprophytic bacteria. The algal biomarker signature in the Tanana Formation organic mats is minimal, but ubiquitous throughout. This is evident (1) petrographically, as abundant acanthomorph acritarchs (presumably of phytoplanktonic algal anity), and (2) chemically, as minor steranes. Numbers of acritarch specimens vary from sample to sample, and this appears to be a taphonomic response to environment. Fewer specimens are preserved in the higher-energy environments, and the species present in such samples tend to be the more robust forms. However, acritarch species diversity increases steadily upsection throughout the lower Ungoolya Group, and the number of new species recognized shows no apparent relationship to facies variation. This suggests that the recorded radiation represents a true diversication of the palynoora, rather than one that is environmentally controlled. It is within the upper part of the Tanana Formation (much of which is still preserved beneath the canyon-ll deposits of the Narana Formation in Munta 1) where the algal (chlorophyte/rhodophyte/dinoagellate) steranes are relatively most abundant. These Algal Facies were deposited by low-concentration, silt-dominated turbidity currents in relatively deep water (upper slope-apron; Fig. 3) prior to progressive shallowing of the seaoor. Fluorescence microscopy of thin-sections of the Algal Facies samples suggests that the algal material is present mainly within the coarser silt bases of these turbidites and

276

K. Arouri et al. / Precambrian Research 100 (2000) 235280

not in their ner-grained tops, thereby discounting its accumulation in these sediments by inter-turbidite pelagic rain. The microbial-mat facies investigated here are evidently dierent than those reported from other Ediacarian microbial mats of the Centralian Superbasin, including mat samples from the Tanana Formation in Munta 1. For example, Logan et al. (1999) [see also Calver (1995)] ascribed the distinctive biomarker and depleted 13C and 34S isotopic signatures of mat samples from the Tanana Formation in Munta 1 (as well as samples from other Australian Neoproterozoic formations) to sulde-oxidising bacteria. The distinctive biomarker distributions that Logan et al. (1999) document include: (1) even-carbon-numbered MMAs predominating over odd-carbonnumbered MMAs, and branching at the oddcarbon sites; (2) even-over-odd carbon-number preference in the C n-alkanes; (3) depleted 20+ carbon and sulfur isotopic values (d13C = kerogen 32 to 34; 34S =4050). The rst of pyrite these three molecular signatures is present in microbial-mat Facies 1A reported in the present study ( Table 2), and a contribution from suldeoxidising bacteria to this facies is, therefore, possible. However, the absence of the second of these three molecular signatures in, and the more enriched 13C values of, our Facies 1A samples, render this uncertain. The relative abundances of cyanobacteria and acritarchs, as revealed in palynological preparations on the one hand and in the biomarker signal on the other hand, appear paradoxical. In the palynological strew-slides, acritarchs are conspicuously more abundant than are morphologically recognisable cyanobacterial remains. However, in the biomarker distributions, evidence of cyanobacteria overwhelms that of other palynomorphs, including acritarchs. This apparent contradiction may be due to a marked contrast in preservation potential between cyanobacteria and acritarchs, the latter being far more resistant to (bio)degradation. Given that acritarchs have dinosporin [see Sarjeant (1986), Kokinos et al. (1998), and Arouri et al. (2000)], algaenan [see Arouri et al. (1999), and references cited therein] or sporopollenin [see Sarjeant (1986)] wall composition, their preserva-

tion potential is considerably higher than the cyanobacteran of the cyanobacterial sheaths [see de Leeuw and Largeau (1993)].

Acknowledgements We thank Macquarie University technical sta for invaluable contributions to this project: Tom Bradley (petrographic services); Judy Davis (computer graphics); David Mathieson (dark-room services); Michael Engelbretson (miscellaneous technical support). K.A. thanks Graham Logan (Australian Geological Survey Organisation) for helpful discussion regarding biomarker topics to do with this work; P.J.C. thanks David Durney, Macquarie University, for helpful discussion regarding structural-diagenetic features of the Munta 1 core samples; and P.J.C. and M.R.W. thank Andrew Glikson, Canberra, for discussion regarding the impact ejecta in Munta 1. We thank David Gravestock, Graham Logan and Simon George for constructive reviews of the manuscript. K.G. publishes with the permission of the Director of the Geological Survey of Western Australia. This work was funded mainly by the Australian Research Council through grants to M.R. Walter and J.J. Veevers, including a post-doctoral fellowship to K.A. This is publication No. 201 of the ARC National Key Centre for the Geochemical Evolution and Metallogeny of Continents (GEMOC ).

References
Arouri, K., Greenwood, P.F., Walter, M.R., 1999. A possible chlorophycean anity of some Neoproterozoic acritarchs. Org. Geochem. 30, 13231337. Arouri, K., Greenwood, P.F., Walter, M.R., 2000. Biological anities of Neoproterozoic acritarchs from Australia. Org. Geochem., in press. Bauld, J., 1984. Microbial mats in marginal marine environments: Shark Bay Western Australia, and Spencer Gulf, South Australia. In: Cohen, Y., Castenholz, R.W., Halvorson, H.O. ( Eds.), Microbial Mats: Stromatolites. Liss, New York, pp. 3958. Bisseret, P., Zundel, M., Rohmer, M., 1985. Prokaryotic triterpenoids. 2. 2b-Methylhopanoids from Methylobacterium

K. Arouri et al. / Precambrian Research 100 (2000) 235280 organophilum and Nostoc muscorum, a new series of prokaryotic triterpenoids. Eur. J. Biochem. 150, 2934. Boon, J.J., de Leeuw, J.W., Krumbein, W.E., 1985. Biogeochemistry of Gavish Sabkha sediments. II. Pyrolysis mass spectrometry of the laminated microbial mat in the permanently water-covered zone before and after the desert sheetood of 1979. In: Friedman, G.M., Krumbein, W.E. (Eds.), Hypersaline Ecosystems: The Gavish Sabkha. Springer, Heidelberg, pp. 368380. Brassell, S.C., Wardroper, A.M.K., Thomson, I.D., Maxwell, J.R., Eglinton, G., 1981. Specic acyclic isoprenoids as biological markers of methanogenic bacteria in marine sediments. Nature 290, 693696. Buczynski, C., Chafetz, H.S., 1993. Habits of bacterially induced precipitates of calcium carbonate: examples from laboratory experiments and Recent sediments. In: Rezak, R., Lavoie, D.L. ( Eds.), Carbonate Microfabrics. Springer, Berlin, pp. 105116. Buttereld, N.J., Rainbird, R.H., 1998. Diverse organic-walled fossils, including possible dinoagellates, from the early Neoproterozoic of arctic Canada. Geology 26, 963966. Calver, C.R., 1995. Ediacarian isotope stratigraphy of Australia. Ph.D. Thesis, Macquarie University, Sydney (unpublished ). Calver, C.R., Grey, K., 1993. Lithologic log (1:400 scale) of the Neoproterozoic succession in Comalco Munta 1 drillcore, Ocer Basin, South Australia. Macquarie University, Sydney (unpublished ). Calver, C.R., Lindsay, J.F., 1998. Ediacarian sequence and isotope stratigraphy of the Ocer Basin, South Australia. Aust. J. Earth Sci. 45, 513532. Chafetz, H.S., Folk, R.L., 1984. Travertines: depositional morphology and the bacterially constructed constituents. J. Sed. Petrol. 54, 289316. Connan, J., Bouroullec, J., Dessort, D., Albrecht, P., 1986. The microbial input in carbonate-anhydrite facies of a sabkha palaeoenvironment from Guatemala: a molecular approach. Org. Geochem. 10, 2950. Cotter, K.L., 1999. Microfossils from Neoproterozoic Supersequence 1 of the Ocer Basin, Western Australia. Alcheringa 23, 6386. Cucuzza, J., 1987. Munta 1 Well completion report. PEL 23, northeastern Ocer Basin. Report for Comalco Aluminium Ltd, South Australia. Department of Mines and Energy. Open le envelope, 7090 (unpublished). DAmelio, E., Cohen, Y., Des Marais, D.J., 1989. Comparative functional ultrastructure of two hypersaline submerged cyanobacterial mats: Guerrero Negro, Baja California Sur, Mexico, and Solar Lake, Sinai, Egypt. In: Rosenberg, E. ( Ed.), Microbial Mats: Physiological Ecology of Benthic Microbial Communities. American Society for Microbiology, Washington, DC, pp. 97113. Chapter 9 de Leeuw, J.W., Largeau, C., 1993. A review of macromolecular organic compounds that comprise living organisms and their role in kerogen, coal, and petroleum formation. In: Engel, M.H., Macko, S.A. ( Eds.), Organic Geochemistry, Principles and Applications. Plenum, New York, pp. 2372.

277

de Leeuw, J.W., Sinninghe Damste, J.S., Klok, J., Schenck, P.A., Boon, J.J., 1985. Biogeochemistry of Gavish Sabkha sediments. I. Studies on neutral reducing sugars and lipid moieties by gas chromatographymass spectrometry. In: Friedman, G.M., Krumbein, W.E. ( Eds.), Hypersaline Ecosystems: The Gavish Sabkha. Springer, Berlin, pp. 350367. de Rosa, M., Gambacorta, A., Minale, L., Bulock, J.D., 1972. The formation of v-cyclohexyl-fatty acids from shikimate in an acidophilic thermophilic bacillus. Biochem. J. 128, 751754. Des Marais, D.J., Bauld, J., Palmisano, A.C., Summons, R.E., Ward, D.M., 1992a. The biogeochemistry of carbon in modern microbial mats. In: Schopf, J.W., Klein, C. ( Eds.), The Proterozoic Biosphere: A Multidisciplinary Study. Cambridge University Press, Cambridge, pp. 299308. Des Marais, D.J., DAmelio, E., Farmer, J.D., Jorgensen, B., Palmisano, A.C., Pierson, B.K., 1992b. Case study of a modern microbial mat-building community: the submerged cyanobacterial mats of Guerrero Negro, Baja California Sur, Mexico. In: Schopf, J.W., Klein, C. ( Eds.), The Proterozoic Biosphere: A Multidisciplinary Study. Cambridge University Press, Cambridge, pp. 325333. Dill, H., Teschner, M., Wehner, H., 1988. Petrography, inorganic geochemistry of lower Permian carbonaceous fan sequences (Brandschiefer series) Federal Republic of Germany: constraints to their paleogeography and assessment of their source rock potential. Chem. Geol. 67, 307325. Dobson, G., Ward, D.M., Robinson, N., Eglinton, G., 1988. Biogeochemistry of hot springs environments: extractable lipids of a cyanobacterial mat. Chem. Geol. 68, 155179. Einsele, G., Ricken, W., Seilacher, A., 1991. Cycles and events in stratigraphy basic concepts and terms. In: Einsele, G., Seilacher, A. ( Eds.), Cycles and Events in Stratigraphy. Springer, Berlin, pp. 119. Flaviano, C., Le Berre, F., Derenne, S., Largeau, C., Connan, J., 1994. First indications of the formation of kerogen amorphous fractions by selective preservation. Role of nonhydrolysable macromolecular constituents of Eubacterial cell walls. Org. Geochem. 22, 759771. Folk, R.L., Chafetz, H.S., 1983. Diagenetic calcite varieties from travertines of central Italy. Am. Assoc. Petrol. Geol. Bull. 67, 461 Folk, R.L., Chafetz, H.S., Tiezzi, P.A., 1985. Bizzare forms of depositional and diagenetic calcite in hot-spring travertines, central Italy. In: Scheiderman, N., Harris, P.M. ( Eds.), Carbonate Cements, SEPM Spec. Publ. 36, 349369. Fowler, M.G., Douglas, A.G., 1984. Distribution and structure of hydrocarbons in four organic-rich Ordovician rocksAdvances in Organic Geochemistry 1983, Schenck, P.A., de Leeuw, J.W., Lijmbach, G.W.M. ( Eds.), Org. Geochem. 6, 105114. Fowler, M.G., Douglas, A.G., 1987. Saturated hydrocarbon biomarkers in oils of Late Precambrian age from Eastern Siberia. Org. Geochem. 11, 201203. Fowler, M.G., Abolins, P., Douglas, A.G., 1986. Monocyclic

278

K. Arouri et al. / Precambrian Research 100 (2000) 235280 Kissin, Y.V., 1987. Catagenesis and composition of petroleum: origin of n-alkanes and isoalkanes in petroleum crudes. Geochim. Cosmochim. Acta 51, 24452457. Klomp, U.C., 1986. The chemical structure of a pronounced series of iso-alkanes in South Oman crudesAdvances in Organic Geochemistry 1985, Leythaeuser, D., Rullkotter, J. ( Eds.), Org. Geochem. 10, 807814. Kokinos, J.P., Eglinton, T.I., Goni, M.A., Boon, J.J., Martoglio, P.A., Anderson, D.M., 1998. Characterization of a highly resistant biomacromolecular material in the cell wall of a marine dinoagellate resting cyst. Org. Geochem. 28, 265288. Lindsay, J.F., Leven, J.H., 1996. Evolution of a Neoproterozoic to Palaeozoic intracratonic setting, Ocer Basin, South Australia. Basin Res. 8, 403424. Logan, G.A., Calver, C.R., Gorjan, P., Summons, R.E., Hayes, J.M., Walter, M.R., 1999. Terminal Proterozoic mid-shelf benthic microbial mats in the Centralian Superbasin and their environmental signicance. Geochim. Cosmochim. Acta 63, 13451358. Lundegard, P.D., Samuels, N.D., Pryor, W.A., 1985. Upper Devonian turbidite sequence central and southern Appalachian basin: contrasts with submarine fan deposits. In: Woodrow, D.L., Sevon, W.D. ( Eds.), The Catskill DeltaGeological Society of America Special Paper 201, 107121. Machel, H.G., Burton, E.A., 1991. Factors governing cathodoluminescence in calcite and dolomite and their implications for studies of carbonate diagenesis. In: Barker, C.E., Kopp, O.C. ( Eds.), Luminescence Microscopy: Quantitative and Qualitative Aspects. SEPM Short Course 25, 3757. McKirdy, D.M., Imbus, S.W., 1992. Precambrian petroleum: a decade of changing perceptions. In: Schidlowski, M., Golubic, S., Kimberley, M.M., McKirdy, D.M., Trudinger, P.A. ( Eds.), Early Organic Evolution: Implications for Mineral and Energy Resources. Springer, pp. 176192. Moczydlowska, M., Vidal, G., Rudavskaya, V.A., 1993. Neoproterozoic ( Vendian) phytoplankton from the Siberian Platform, Yakutia. Palaeontology 36, 495521. Morton, J.G.G., 1997a. Introduction and summary. In: Morton, J.G.G., Drexel, J.F. ( Eds.), Ocer Basin. The Petroleum Geology of South Australia vol. 3, 15, South Australia. Department of Mines and Energy Resources Report Book 97/19. Morton, J.G.G., 1997b. Lithostratigraphy and environments of deposition. In: Morton, J.G.G., Drexel, J.F. (Eds.), Ocer Basin. The Petroleum Geology of South Australia vol. 3, 4786, South Australia. Department of Mines and Energy Resources Report Book 97/19. Moussavi-Harami, R., Gravestock, D.I., 1995. Burial history of the eastern Ocer Basin, South Australia. APEA J. 35, 307319. Mycke, B., Michaelis, W., Degens, E.T., 1988. Biomarkers in sedimentary sulde of Precambrian age. In: Mattavelli, L., Novelli, L. ( Eds.), Advances in Organic Geochemistry 1987, Org. Geochem. 13, 619625. Nelson, D.R., 1978. Long-chain methyl branched hydro-

alkanes in Ordovician organic matter. Org. Geochem. 10, 815823. Gelpi, E., Schneider, H., Mann, J., Oro, J., 1970. Hydrocarbons of geochemical signicance in microscopic algae. Phytochemistry 9, 603612. Gerdes, G., Krumbein, W.E., Reineck, H.-E., 1991. Biolaminations ecological versus depositional dynamics. In: Einsele, G., Ricken, W., Seilacher, A. ( Eds.), Cycles and Events in Stratigraphy. Springer, Berlin, pp. 592607. Gostin, V.A., Haines, P.W., Jenkins, R.J.F., Compston, W., Williams, I.S., 1986. Impact ejecta horizon within late Precambrian shales, Adelaide Geosyncline, South Australia. Science 233, 198200. Golubic, S., 1976. Organisms that build stromatolites. In: Walter, M.R. ( Ed.), Stromatolites. Developments in Sedimentology 20. Elsevier, Amsterdam, pp. 113126. Gravestock, D.I., 1997. Geological setting and structural history. In: Morton, J.G.G., Drexel, J.F. ( Eds.), Ocer Basin.Petroleum Geology of South Australia vol. 3., 3545., South Australia. Department of Mines and Energy Resources Report Book 97/19 Gravestock, D.I., Morton, J.G.G., 1997. Potential traps and prospects. In: Morton, J.G.G., Drexel, J.F. (Eds.), Ocer Basin. Petroleum Geology of South Australia vol. 3., 129140., South Australia. Department of Mines and Energy Resources Report Book 97/19 Gravestock, D.I., Morton, J.G.G., Zang, W.-L., 1997. Biostratigraphy and correlation. In: Morton, J.G.G., Drexel, J.F. ( Eds.), Ocer Basin, Petroleum Geology of South Australia vol. 3, 8797, South Australia. Department of Mines and Energy Resources Report Book 97/19. Grey, K., 1998. Ediacarian palynology of Australia. Ph.D. Thesis, Macquarie University, Sydney (unpublished). Grey, K., Cotter, K.L., 1996. Palynology in the search for Proterozoic hydrocarbons. In: Western Australia Geological Survey Annual Review 199596, 7080. Hill, A.C., Al-Arouri, K., Gorjan, P., Walter, M.R., 2000. The geochemistry of marine and non-marine environments of a Neoproterozoic cratonic carbonate/evaporite: the Bitter Springs Formation central Australia. In: Grotzinger, J.P., James, N.P. ( Eds.), Precambrian Carbonates. SEPM Spec. Publ., in press. Hoering, T.C., 1976. Molecular fossils from the Precambrian Nonesuch Shale. Carnegie Inst. Wash. Yearbook 75, 806813. Hoering, T.C., 1981. Monomethyl, acyclic hydrocarbons in petroleum and rock extracts. Carnegie Inst. Wash. Yearbook 80, 389394. Jackson, M.J., Powell, T.G., Summons, R.E., Sweet, I.P., 1986. Hydrocarbon show and petroleum source rocks in sediments as old as 1.7109 years. Nature 322, 727729. Kenig, F., Sinninghe Damste, J.P., Kock-Van Dalen, A.C., Rijpstra, W.I.C., Huc, A.Y., de Leeuw, J.W., 1995. Occurrence and origin of mono-, di-, and trimethylalkanes in modern and Holocene cyanobacterial mats from Abu Dhabi, United Arab Emirates. Geochim. Cosmochim. Acta 59, 29993015.

K. Arouri et al. / Precambrian Research 100 (2000) 235280 carbons: occurrence, biosynthesis and function. Adv. Insect. Pyhsiol. 13, 133. Nelson, D.R., Blomquist, G.J., 1995. Insect waxes. In: Hamilton, R.J. ( Ed.), Waxes: Chemistry, Molecular Biology and Functions. The Oily Press, Dundee, pp. 1129. OBrien, N.R., 1990. Signicance of lamination in Toarcian (Lower Jurassic) shales from Yorkshire, Great Britain. Sediment. Geol. 67, 2534. Oshima, M., Ariga, T., 1975. v-Cyclohexyl fatty acids in acidophilic bacteria. J. Biol. Chem. 256, 69636968. Paoletti, C., Pushparaj, B., Florenzano, G., Capella, P., Lercker, G., 1976. Unsaponiable matter of green and blue green algal lipids as a factor of biochemical dierentiation of their biomasses: I. Total unsaponiable hydrocarbon fraction. Lipids 11, 258265. Perincek, D., 1998. A compilation and review of data pertaining to the hydrocarbon prospectivity of the Ocer Basin. Geological Survey of Western Australia Record 1997/6, 209 pp. Pickering, K.T., Hiscott, R.N., Hein, F.J., 1989. Deep Marine Environments: Clastic Sedimentation and Tectonics. Unwin Hyman, London. Preiss, W.V., Belperio, A.P., Cowley, W.M., Rankin, L.R., 1993. Neoproterozoic. In: Drexel, J.F., Preiss, W.V., Parker, A.J. ( Eds.), The Geology of South Australia. vol. 1: The Precambrian. Geological Survey of South Australia Bulletin 54, 170203. Reading, H.G., Richards, M., 1994. Turbidite systems in deepwater margins classied by grain size and feeder system. Am. Assoc. Petrol. Geol. Bull. 78, 792822. Robinson, N., Eglinton, G., 1990. Lipid chemistry of Iclandic hot spring microbial mats. Org. Geochem. 15, 291298. Robinson, R.A., Kaesler, R.L., 1987. Phylum Arthropoda part I. In: Boardman, R.S., Cheetham, A.H., Rowell, A.J. ( Eds.), Fossil Invertebrates. Blackwell, pp. 205269. Sarjeant, W.A.S., 1986. Review of Evitt, W.R., 1985, sporopollenin dinoagellate cysts: their morphology and interpretation. Microplaeontology 3, 282285. Schieber, J., 1986. The possible role of benthic microbial mats during the formation of carbonaceous shales in shallow midProterozoic basins. Sedimentology 33, 521536. Schieber, J., 1989. Facies and origin of shales from the midProterozoic Newfoundland Formation, Belt Basin, Montana, USA. Sedimentology 36, 203219. Schieber, J., 1990. Signicance of styles of epicontinental shale sedimentation in the Belt Basin, mid-Proterozoic of Montana, USA. Sedimentology 39, 297312. Schieber, J., 1999. Microbial mats in terrigenous clastics: the challenge of identication in the rock record. Palios 14, 312. Schmidt, P.W., Williams, G.E., 1996. Palaeomagnetism of the ejecta-bearing Buneroo Formation, late Neoproterozoic, Adelaide fold belt, and the age of the Acraman impact. Earth Planet. Sci. Lett. 144, 347358. Shiea, J., Brassell, S.C., Ward, D.M., 1990. Mid-chain branched mono- and dimethyl alkanes in hot spring cyanobacterial mats: a direct biogenic source for branched alkanes in ancient sediments? Org. Geochem. 15, 223231.

279

Shiea, J., Brassell, S.C., Ward, D.M., 1991. Comparative analysis of extractable lipids in hot spring microbial mats and their component photosynthetic bacteria. Org. Geochem. 17, 309319. Stow, D.A.V., Reading, H.G., Collinson, J.D., 1996. Deep seas. In: Reading, H.G. ( Ed.), Sedimentary Environments: Processes, Facies and Stratigraphy, third ed., Blackwell, Oxford, pp. 395453. Sukanta, U., 1993. Sedimentology, sequence stratigraphy and palaeogeography of Ediacarian sediments in the eastern Ocer Basin, South Australia. Ph.D. Thesis, Flinders University of South Australia, Adelaide (unpublished, not seen). Sukanta, U., Thomas, B., Von Der Borch, C.C., Gatehouse, C.G., 1991. Sequence stratigraphic studies and canyon formation. PESA J. 19, 6873. Summons, R.E., 1987. Branched alkanes from ancient and modern sediments: isomer discrimination by GC/MS with multiple reaction monitoring. Org. Geochem. 11, 281289. Summons, R.E., 1992. Abundance and composition of extractable organic matter. In: Schopf, J.W., Klein, C. ( Eds.), The Proterozoic Biosphere: A Multidisciplinary Study. Cambridge University Press, Cambridge, pp. 101115. Summons, R.E., Jahnke, L.L., 1992. Identication of the methylhopanes in sediments and petroleum. Geochim. Cosmochim. Acta 54, 247251. Summons, R.E., Walter, M.R., 1990. Molecular fossils and microfossils of prokaryotes and protists from Proterozoic sediments. Am. J. Sci. 290-A, 212244. Summons, R.E., Brassell, S.C., Eglinton, G., Evans, E., Horodysky, R.J., Robinson, N., Ward, D.M., 1988a. Distinctive hydrocarbon biomarkers from fossiliferous sediment of the Late Proterozoic Walcott Member, Chuar Group, Grand Canyon, Arizona. Geochim. Cosmochim. Acta 52, 26252637. Summons, R.E., Powell, T.G., Boreham, C.J., 1988b. Petroleum geology and geochemistry of the Middle Proterozoic McArthur Basin, Northern Australia: III. Composition of extractable hydrocarbons. Geochim. Cosmochim. Acta 52, 17471763. Summons, R.E., Jahnke, L.L., Hope, J.M., Logan, G.A., 1998. Geolipids from cyanobacteria: new studies of old rocks and hot spring microbial mats from Yellowstone National Park. In: Proc. Aust. Org. Geochem. Conf. 1998, 1011. Ten Haven, H.L., de Leeuw, J.W., Sinninghe Damste, J.S., Schenck, P.A., Palmer, S.E., Zumberge, J.E., 1988. Application of biological markers in the recognition of palaeo hypersaline environments. In: Kelts, K., Fleet, A., Talbot, M. ( Eds.), Lacustrine Petroleum Source Rocks, Geological Society of London Special Publication 40, 123140. Wallace, M.W., Gostin, V.A., Keays, R.R., 1989. Discovery of the Acraman impact ejecta blanket in the Ocer Basin and its stratigraphic signicance. Aust. J. Earth Sci. 36, 585587. Walter, M.R., 1995. Faecal pellets in world events. Nature 376, 1617. Walter, M.R., Veevers, J.J., 1997. Australian Neoproterozoic palaeogeography, tectonics, and supercontinental connections. AGSO J. Geol. Geophys. 17, 7392.

280

K. Arouri et al. / Precambrian Research 100 (2000) 235280 Biosphere: A Multidisciplinary Study. Cambridge University Press, Cambridge, pp. 309324. Woodrow, D.L., 1985. Paleogeography, paleoclimate and sedimentary processes of the Late Devonian Catskill Delta. In: Woodrow, D.L., Sevon, W.D. ( Eds.), The Catskill DeltaGeological Society of America Special Paper 201, 5163. Woodrow, D.L., Isley, A.M., 1983. Facies, topography, and sedimentary processes in the Catskill Sea (Devonian), New York and Pennsylvania. Geol. Soc. Am. Bull. 94, 459470. Wright, V.P., Burchette, T.P., 1996. Shallow-water carbonate environments. In: Reading, H.G. ( Ed.), Sedimentary Environments: Processes, Facies and Stratigraphy, third ed., Blackwell, Oxford, pp. 325394. Zang, W., McKirdy, D.M., 1994. Microfossils and molecular fossils from the Neoproterozoic Alinya Formation a possible new source rock in the eastern Ocer Basin. PESA J. 22, 8990. Zang, W., Walter, M.R., 1992. Late Proterozoic and Early Cambrian microfossils and biostratigraphy, Amadeus Basin, central Australia. Association of Australasian Palaeontologists Memoir 12. 132 pp

Walter, M.R., Bauld, J., Des Marais, D.J., Schopf, J.W., 1992. A general comparison of microbial mats and microbial stromatolites: bridging the gap between the modern and the fossil. In: Schopf, J.W., Klein, C. ( Eds.), The Proterozoic Biosphere: A Multidisciplinary Study. Cambridge University Press, Cambridge, pp. 335338. Walter, M.R., Veevers, J.J., Calver, C.R., Grey, K., 1995. Neoproterozoic stratigraphy of the Centralian Superbasin. Precam. Res. 73, 173195. Walter, M.R., Veevers, J.J., Calver, C.R., Gorjan, P., Hill, A.C., 2000. Dating the 840544 Ma Neoproterozoic interval by isotopes of strontium, carbon, and sulfur in seawater, and some interpretative models. Precam. Res. 100, 371433. Ward, D.M., Yayne, T.A., Anderson, K.L., Bateson, M.M., 1987. Community structure and interactions among community members in hot spring cyanobacterial mats. Symp. Soc. Gen. Microbiol. 41, 179210. Ward, D.M., Bauld, J., Castenholz, R.W., Pierson, B.K., 1992. Modern phototrophic microbial mats: anoxygenic intermittently oxygenic/anoxygenic thermal eukaryotic and terrestrial. In: Schopf, J.W., Klein, C. ( Eds.), The Proterozoic

Vous aimerez peut-être aussi