Vous êtes sur la page 1sur 26

W.A. SONIBARE, D. MIKE AND D.I.

COLE 303

FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA,


TANQUA-KAROO BASIN (SOUTH AFRICA): IMPLICATIONS FOR
FACIES ANALYSIS AND MODELLING

W.A. SONIBARE
Department of Earth Sciences, Stellenbosch University, Private Bag X1, Matieland 7602, Stellenbosch,
South Africa
e-mail: shoniwas@yahoo.com, 15022161@sun.ac.za

D. MIKE
Department of Earth Sciences, Stellenbosch University, Private Bag X1, Matieland 7602, Stellenbosch,
South Africa
e-mail: mikes@sun.ac.za

D.I. COLE
Council for Geoscience, 3 East Street, Bellville, 7535, South Africa
e-mail: dcole@geoscience.org.za
2011 December Geological Society of South Africa

ABSTRACT
In recent years, the construction of accurate geological models and subsequent upscaling for reservoir simulation are commonly
built from an integration of subsurface and surface datasets. Surface data can supply additional information on facies variability and
small scale heterogeneity needed for a robust reservoir model, yet impossible to attain from subsurface data. This study therefore
presents an outcrop characterisation study of the Permian Kookfontein deltaic succession using a combination of detailed
sedimentary log, gamma-ray log and photopanel analysis.
Based on texture and sedimentary structures, twelve depofacies are recognised which are broadly grouped into four lithofacies
associations i.e. sandstone facies, heterolithic facies, mudstone facies and soft-sediment deformation facies. Following a hierarchical
procedure, these depofacies and lithofacies form the basic building blocks for the flooding surface-bounded facies succession
(i.e. cycle). Lateral juxtaposition of observed vertical facies variations across each cycle in an inferably basinwards direction exhibits
upward change in features, i.e. decrease in gravity effects, increase in waves and decrease in slope gradient of subsequent cycles.
This systematic upward transition in features, grading vertically from distal to proximal, represents a deposition of mid-slope to
shelf-margin succession exhibiting an overall upward thickening and coarsening with progradational stacking pattern typical of a
normal regressive prograding delta. However, in detail, cycles show some anomalies from a purely thickening and coarsening
upward succession. Deposition of each cycle is believed to result from:
1. primary deposition by episodic and probably sporadic mouthbar events governed by stream flow dynamics; and
2. secondary remobilisation of sediments under gravity.
The architecture and geometry of the ensuing depo-system is interpreted to have been a river-dominated, gravitationally reworked
and wave-influenced shelf edge Gilbert-type delta. Widespread distribution of soft-sediment deformation structures, their
growth-style and morphology within the studied succession are empirically related to progradation of Gilbert-type mouthbars over
the shelf break as well as the slope gradients of the Kookfontein deltaic clinoformal geometry. Apparently, low-slope gradients may
have favoured mostly loading and dewatering structures rather than large-scale slumps.
Although shale content (shaliness) estimation from gamma-ray logs offers a quick and straightforward way for assessing
reservoir potentiality of encountered facies in the field, this is rather crude and therefore its reliability requires detailed petrographic
characterisation for confirmation. The described internal heterogeneity in this work is below the resolution (i.e. mm-scale) of most
conventional well-logs, and therefore could supplement well-log data especially where there is no borehole image and core data.
The combination of descriptive facies model and schematic geological model for our specific delta make the results of this study
applicable to any other similar ancient depo-system and particularly subsurface reservoir analogue.

Introduction and Background inherent complexity of deltaic regimes, which in turn is


Conceptual sedimentary models of deltaic depositional largely due to their position at the transition between
settings have advanced through field and laboratory continental and marine influences. Hence the need to
observations as well as digital computer simulation. develop standard facies models that attempt to simplify
However, lacking in many of these models is the this complexity, yet at the same time reflect the temporal
accurate prediction of the three-dimensional geometric and spatial variability of facies and their three-
behaviour of sedimentary bodies, partly due to the dimensional architecture.

SOUTH AFRICAN JOURNAL OF GEOLOGY, 2011, VOLUME 114.3-4 PAGE 303-328


doi:10.2113/gssajg.114.3-4.303
304 FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA

Figure 1. Chart showing subsurface data types that sample the reservoir volume plotted against their corresponding vertical resolution.
Both outcrop analogues and physical-based (experimental) modelling can help bridge the gaps between subsurface data types (modified after
Keogh et al., 2007).

Many previous studies, e.g. Elliot (1989), Postma lateral continuity (Figure 1). Therefore, subsurface
(1990), Muto and Steel (2002), Longhitano (2008), datasets lack the potential to incorporate all scales of
Catuneanu (2009) and Nichols (2009), have attempted to sedimentary heterogeneity in a geological-reservoir
explain in detail the various sedimentological factors model (Mike and Geel, 2006; Phillips and Wen, 2007).
such as climate, sediment supply rate, tectonic Outcrop-based facies analysis has better spatial
subsidence and uplift, global eustasy, delta-lobe resolution and can therefore provide adequate insights
switching, sediment transport and nature of discharge into sedimentary processes and can also reveal three-
systems that govern the evolution of deltaic regimes. dimensional facies geometry and architecture,
The interplay between these factors and base-level e.g. Keogh et al. (2007) and Catuneanu et al. (2009).
changes (i.e. creation of accommodation space) The characterisation and modelling of outcrop
determines the morphology, relative position and analogues allow:
sediment stacking pattern of deltas in both recent 1. the documentation of quantitative key reservoir
and ancient settings. For instance, Postma (1990) parameters, which can be applied to similar
grouped deltas into: subsurface depositional settings (Mayer and Chapin,
1. Gilbert-type or mouthbar-type, based on the nature 1991; Cabello et al., 2010);
of the active growth front of the delta at the mouth 2. the testing and validation of subsurface facies-
of the river; dependent geological modelling approaches
2. gravelly or alluvial fan delta, based on the nature of (Falivene et al., 2006; Keogh et al., 2007);
the river feeder system; 3. better constraining of subsurface deterministic facies-
3. shelf or shelf-edge delta, based on the position of the dependent modelling approaches over probabilistic
basin with respect to shoreline or shelf break roll- approaches;
over point; 4. reduction of uncertainties in predicting sand-body
4. shallow-water or deep-water delta, based on the geometries, thus improving net-to-gross estimation
water depth at the receiving basin. An understanding and fluid flow prediction.
of the interplay of criteria is crucial to three-
dimensional reconstruction of deltaic systems in In this paper, we attempt to characterise the Permian
surface and subsurface. Tanqua-Karoo deltaic sequence (Kookfontein
Formation) (Figures 2a, 2b) using an approach that
Seismic data acquisition and interpretation are important combines outcrop logs, gamma ray logs and
tools for proper understanding of the shelf-slope-basin photopanels. The Kookfontein Formation is restricted to
profile, e.g. Catuneanu et al. (2009). However, due to a small region in the southwestern part of the main
their limitations in terms of vertical and lateral Karoo basin, where it outcrops over a lateral extent
resolution, they cannot resolve sedimentary facies. of approximately 80 km in an area southwest of
Conventional well-logs on the other hand, lack the Sutherland (Figure 2b). The formation is mainly

SOUTH AFRICAN JOURNAL OF GEOLOGY


W.A. SONIBARE, D. MIKE AND D.I. COLE 305

argillaceous and its lower boundary is defined by the 1983; Cole and Vorster, 1999). Similar thorough
underlying arenaceous Skoorsteenberg Formation descriptions of the lithology and depositional
(Wickens, 1996). The Kookfontein Formation overlies environment are given in Wickens (1994, 1996) and
the Skoorsteenberg Formation with a sharp contact on Johnson et al. (2006). The studies by Wild (2005)
the uppermost prominent sandstone bed and grades up and Wild et al. (2009) appear to be the first attempt to
into the arenaceous Waterford Formation, where the describe in more detail the sedimentology, stratigraphic
upper boundary is defined as a horizon above which evolution and depositional setting of the Kookfontein
sandstone is relatively common (Johnson et al., 2006). deltaic succession.
It has a thickness of approximately 350 m at its type Wild (2005) and Wild et al. (2009) describe thirteen
locality on the farm Kook Fontein 76, 60 km southwest sedimentary cycles for the Kookfontein deltaic sequence
of Sutherland, and is a lateral equivalent of the upper observed at the Pienaarsfontein locality (Figure 2a), and
part of the Tierberg Formation towards the north and employ a sequence stratigraphic approach to interpret
south (Figure 2b; Wickens, 1996; Johnson et al., 2006). their sedimentological and stratigraphic evolutionary
The Kookfontein Formation is confined to a sub-basin or trend. Within these 13 cycles, they identify twelve
depocentre, namely the Tanqua-Karoo depocentre flooding surfaces separated by two erosional sequence
(Wickens, 1996). boundaries as the sequence stratigraphic surfaces that
The deltaic sequence of the Kookfontein Formation bound the sedimentary cycles of the Kookfontein
is composed of thirteen sedimentary cycles, which we Formation. They also use terms such as sediment
group into three units, i.e. Lower (1 to 3), Middle accretion (i.e. deposition and shelf construction) and
(4 to 6) and Upper (8 to 13). Quantitative data collection bypass (erosion and slope failure) to describe
is based on the lowermost unit (i.e. Cycles 1 to 3) the sedimentological and stratigraphic evolution of the
observed at the Pienaarsfontein locality. The aims of this Kookfontein deltaic succession and their relationships to
paper are as follows: the underlying basin-floor fan systems of the
1. Describe and analyse various sedimentary facies and Skoorsteenberg Formation. Although, these studies have
interpret their depositional environments. improved our understanding of the Kookfontein shelf
2. Develop a hypothetical descriptive facies model for edge sediment-body geometry, a proposition of standard
the delta. facies model for the delta is lacking. Only
3. Describe and interpret its facies architecture and reconnaissance petrophysical studies have been
geometry. previously conducted by Van Lente (2004) on the
4. Evaluate the shale content (shaliness) of the Tanqua-Karoo deltaic succession. According to Van
sedimentary facies. Lente (2004), the sandstones of the Tanqua
and Laingsburg depocentres have no visible porosity
Previous work: Tanqua-Karoo Sub-basin and permeability, primarily due to a narrow range in
The earliest studies on the Tanqua-Karoo depocentre grain size varying from very fine- to lower medium-sand,
(Figures 2a, 2b), e.g. Bouma and Wickens (1991), and the formation of authigenic quartz cement and
Wickens (1994), Johnson et al. (2001) and Van der Werff secondary chlorite and illite. However, preliminary
and Johnson (2003), concentrated on the fine-grained, results obtained from grain size-based petrophysical
sand-rich submarine fan systems of the Skoorsteenberg studies (Sonibare, 2011- unpublished MSc thesis) suggest
Formation, which underlie the Kookfontein deltaic that porosity reduction is predominantly due to
sequence. Most of these studies including some recent mechanical compaction.
ones, e.g. Bouma and Wickens (1991), Andersson et al. At present, there is neither a detailed
(2004), Hodgson et al. (2006) and Wild (2005), aimed at lithostratigraphic map nor a robust facies model that
application in similar deepwater reservoir analogues. describes Tanqua-Karoo deltaic succession in terms of
Although these studies have improved our under- geometry and internal architecture. Moreover, the
standing of ancient deepwater systems, their value to Kookfontein deltaic sequence has not been studied and
characterisation and modelling of subsurface hydrocarbon documented as an outcrop analogue to subsurface
reservoirs is still difficult to quantify. It can also be examples. This study is the first attempt to carry out a
argued that the degree of outcrop applicability will reservoir characterisation of the Kookfontein deltaic
depend on how similar they are to subsurface reservoirs. succession through a hierarchical-based facies analysis.
The deltaic sequence of the Kookfontein Formation,
which is up to 350 m thick (Wickens, 1994, 1996) with Geological Setting
a north to south lateral extent of about 72 km between The Late Carboniferous to Middle Jurassic cratonic Karoo
seven localities, i.e. Katjiesberg, Syfer, Skoorsteenberg, Basin (Figure 3) represents one of the best preserved
Bitterberg, Vaalberg, Pienaarsfontein and Roosterberg Gondwanan sequences, i.e. Paran, Karoo, Huab and
(Wild et al., 2009), has not been previously studied in Bowen Basins, e.g. Faure and Cole (1999), and forms
detail. The distribution of the Kookfontein Formation is one of the most complete stratigraphic successions in
shown on the 1:250 000 scale, 3220 Sutherland the world that span this time (Johnson et al., 2006).
geological and metallogenic maps, which have thorough The Karoo Supergroup covers an area of approximately
descriptions in the explanations to these maps (Theron, 700 000 km2 with the bulk of the sedimentation

SOUTH AFRICAN JOURNAL OF GEOLOGY


306 FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA

2a

Figure 2. ( a) Location maps of the study area showing the southwest Karoo Basin, outline of its depocentres (i.e. Tanqua and Laingsburg)
and Pienaarsfontein se Berg locality. Red box in upper photo shows the outline of Kookfontein formation. See Figure 9 for basinwards
cross-sectional profile A-A.

occurring in the main basin with a maximum thickness delineated in the south-western part of the basin
of ~12 km for the south-eastern portion of the basin (Wickens, 1992, 1994; Figures 2a and 2b).
(Johnson et al., 2006). However, the study area lies The Karoo Basin is generally believed to represent a
within the Cape Fold Belt and the thickness has retro-arc foreland basin behind an inferred magmatic arc
probably been enhanced as a result of tectonic and fold-thrust belt with subsidence due to loading by
duplication (Stankiewicz et al., 2007; Lindeque et al. this the Cape Fold Belt, which lies along the southern and
volume). Two depocenters separated by a basement south-western margins of the basin, e.g. Lock (1980),
high, namely the Tanqua and Laingsburg, have been Johnson (1991), Veevers et al. (1994) and Catuneanu

SOUTH AFRICAN JOURNAL OF GEOLOGY


W.A. SONIBARE, D. MIKE AND D.I. COLE 307

2b

Figure 2. ( b) Geological map of the Tanqua depocentre showing the stratigraphic properties, distribution and boundaries of the Kookfontein
Formation (modified from Van Lente, 2004; based on 1:250 000 Geological series by Geological Survey of South Africa).

et al. (1998). Visser and Praekelt (1996) proposed a zone named the Atlantic Fracture Zone to the southwest
modification to the retro-arc foreland basin model. They of South Africa and along a major north-striking fracture
suggested that oblique subduction of the Panthalassan zone named the Falkland Fracture Zone, between the
(paleo-Pacific) plate underneath Gondwana led to the restored Falkland Islands to the southeast of South Africa
creation of a large interconnected basin stretching from and Antarctica. More recently, Tankard et al. (2009)
South America to Antarctica, including the main Karoo proposed an alternative tectonic model, i.e. early Karoo
Basin. Accommodation space in the main Karoo Basin dynamic subsidence due to mantle flow and late
was created by dextral transpressional strike-slip Karoo transtensional foreland system, based on the
movements along a major northwest-striking fracture following arguments:

SOUTH AFRICAN JOURNAL OF GEOLOGY


308 FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA

1. There is no geophysical evidence for the nearby Group begins with the basal Prince Albert Formation
magmatic arc; (predominantly shale; 289.6 +/- 3.8 Ma and 288 +/- 3 Ma
2. The Cape Fold Belt is a strike-slip orogen dated to in basal part; Bangert et al., 1999; Cole, 2005), the
late Karoo time; and Whitehill Formation (black, carbonaceous shale with
3. Absence of typical flexural foreland basin onlapping pelagic organisms; Visser, 1992) and the Collingham
stratigraphic features and the fact that the lithosphere Formation (alternating beds of siliceous mudrock and
is not laterally uniform. soft yellowish shale (volcanic tuffs) overlain by very
Therefore, it can be categorically stated that the tectonic fine-grained, turbiditic sandstone and minor intercalated
evolution of the Karoo basin is controversial and yet to tuffs; 270 +/- 1 Ma; Viljoen, 1994; Turner, 1999)
be fully understood (e.g. Milani and de Wit 2008; see (see Figure 4). The Collingham Formation is overlain by
also Lindeque et al., this volume and Booth, this several hundreds of metres of dark grey, basinal shale of
volume). the Tierberg Formation (600 m in Soekor borehole
KL1/65, 32 km southwest of Sutherland; Viljoen, 2005)
Stratigraphy followed by up to 445 m of sandstone-rich deepwater
The stratigraphy of the Karoo Supergroup is divided into sediments of the Skoorsteenberg Formation (Wickens,
four groups namely, the Dwyka Group (Westphalian to 1994). The overlying deltaic sequence (Kookfontein
early Permian glacial deposits), the Ecca Group Formation), shoreface/delta plain (Waterford Formation)
(Permian), the Beaufort Group (Permo-Triassic terrestrial and fluvial (Abrahamskraal Formation, Beaufort Group)
sediments) and the Stormberg Group (Late Triassic-Early successions mark the overall progradation of the
Jurassic terrestrial and desert sediments). The Permian sedimentary system to the north and east during the mid
Ecca Group comprises a total of 16 formations (Johnson to late Permian.
et al., 2006) interpreted mainly as shallow marine and A U/Pb date of 289.6 +/- 3.8 Ma from zircons in tuff
deltaic with subordinate deep water and fluvial facies. beds from the basal part of the Prince Albert Formation
This variety of depositional settings reflects the lateral (Bangert et al., 1999) and a 255 Ma date from early
facies changes that characterise the Ecca Group from the reptile fossils in the basal Beaufort Group (Rubidge
south-west to the north-east. In the Tanqua area of et al., 1999, 2000) bracket the whole Permian Ecca
the southwest Karoo Basin, the ~1400 m-thick Ecca Group deposits in the southwestern part of the Karoo

Figure 3. Geological map of the preserved Karoo Basin, showing the outcrop distribution of the main lithostratigraphic units of the Karoo
Supergroup. The Adelaide and Tarkastad subgroups together form the Beaufort Group (adapted from Catuneanu et al., 2002).

SOUTH AFRICAN JOURNAL OF GEOLOGY


W.A. SONIBARE, D. MIKE AND D.I. COLE 309

Figure 4. Simplified stratigraphy of the Tanqua depocenter indicating a 35 My period for the deposition of the Permian Ecca Group in the
southwest Karoo Basin

basin to a ca. 35 My period (Figure 4). This time bracket 19 590E and 20 0730E and Latitudes 32 440S and
was later reduced to ca. 24 million years, as a result of 32 4830S. The outcropping part of the Kookfontein
the earliest reptile fossils in the basal Beaufort Group deltaic succession selected for this study represents an
being assigned an age of ~266 Ma by Rubidge (2005). aerial extent of about 13 km2 (6.5 x 2.1 km) with a lateral
The studies by Fildani et al. (2007, 2009) subsequently coverage of 6.5 km (Figure 2a), and ranges in
expanded this period to ca. 47 million years based on stratigraphic thickness from 100 to 120 m. Four vertical
U/Pb dates of 254.4 +/- 1.8 Ma, 255.4 +/- 2 Ma and sections, VS1, VS2, VS3 and VS4 (Figure 2a), were
256 +/- 2 Ma from zircons in tuff beds from the logged at mm to m scale for complete coverage of
Skoorsteenberg Formation. The aforementioned sedimentary heterogeneities and facies distribution
contrasting absolute dates suggest that lithostratigraphic (from proximal to distal parts) within the studied
correlation within the Ecca Group requires a better stratigraphic column. Total gamma ray-count profiles
constraining of time-lines (i.e. temporal and spatial obtained with a hand-held scintillometer were logged
facies distribution and boundaries) for sedimentary along two (i.e. VS1 and VS2) of these outcrop logs at a
facies of the Ecca successions. 5 s count rate (North and Boering, 1999) and 50 cm
vertical sample spacing. The results of the gamma ray
Methodology logs were used for lithological correlation and
The Kookfontein Formation is well exposed along the evaluation of the shale content. This shaliness was
slopes of a prominent ridge named Pienaarsfontein se obtained by normalising the gamma ray values using the
Berge (Figure 2a) and this locality is the focus of our mathematical expression below:
study. It lies approximately between Longitudes

SOUTH AFRICAN JOURNAL OF GEOLOGY


310 FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA

Figure 5. Example of a log (VS2-C1) through cycle 1. These vertical facies successions are overlain by abrupt fining and thinning up
heterolithic sandstone and siltstone intercalations. Note the overall upward coarsening and low GR values for this succession.

Vsh = (Grlog Grmin/Grmax Grmin)* 100% Photo mosaics were derived from ground-based
(Equation 1) photographs and differential GPS. Lithofacies
boundaries were established through physical walking-
Where, Vsh = the shale content expressed as a out on outcrop exposures and using photo mosaics.
percentage, The integration of these datasets together with the
incorporation of existing conceptual facies models for
Grlog = gamma ray log, river-dominated shelf edge deltas enabled a
deterministic approach for spatial distribution of facies,
Grmax = the maximum gamma ray log reading, and establishing the relationships between temporal
facies successions and delta slope geometry, and
Grmin = the minimum gamma ray log reading. subsequent interpretation of depositional environments.

SOUTH AFRICAN JOURNAL OF GEOLOGY


W.A. SONIBARE, D. MIKE AND D.I. COLE 311

Figure 6. Example of a log (VS1-C2) through cycle 2. These vertical facies successions give overall progradational stacking pattern overlain
by abrupt fining and thinning up retrogradational stacking trend. Also, note the overall upward decrease in GR values indicating an overall
coarsening upward succession.

Results The vertical succession of these elements, their


Stratigraphic elements and depositional sequence sedimentary structures, and their grain size trends in
hierarchy time and space depict all scales of cyclicity. The word
The stratigraphic elements for the Kookfontein deltaic cyclicity is used to indicate a rather regular repetition
sequence from the studied Pienaarsfontein sections of sedimentary features. The term depofacies denotes
reflect a depositional hierarchy that ranges from a body of rock with internally uniform sedimentary
cycle (stratigraphic-scale/facies succession 35 to features and forms the basic element for facies
40 m), bedsets (lithofacies-scale 0.2 to 30 m), modelling, i.e. reconstruction of an ancient depo-
beds (depofacies 0.01 to 1.2 m) and laminae (<0.01 m) system.
over a total stratigraphic thickness of 100 to 120 m.

SOUTH AFRICAN JOURNAL OF GEOLOGY


Table 1. Sedimentary facies described and interpreted depositional characteristics.
312

Depofacies Lithology Sedimentary structures Facies Lithofacies Bioturbation Sand: Silt Notes
thickness association ratio
1 Mudstone Structureless 0.2 m to 1 m Mudstone Prodeltaic hemipelagic deposition probably by
long period of suspension settling
2 Interbedded very fine- Parallel lamination with occasional 5.2 m at Western Heterolithic 55% : 45% Proximal prodeltaic sedimentation by alternating
grained sandstones current ripple lamination section suspension settling and probably gravity-driven
and siltstones Bedding (very thin to thin) hyperpycnal flow/low-density turbidity current
3 Interbedded very fine- Horizontal lamination to unidirectional 2.4 m to 3.9 m Heterolithic Weak Same as This deposition occurred below storm wave base
grained sandstones current ripples; Bedding (very thin above interpreted as the transitional between proximal
to thin) prodelta and distal undeformed delta front
4 Interbedded very fine Unidirectional current ripples with 4.5 m to 8.8 m Heterolithic Moderate 60% : 40% Same as above; but with more reworking by current
to fine-grained occasional parallel lamination; ripples and organic activities
sandstones and siltstones Bedding (very thin to thin)
5 Interbedded very fine Current to wave ripple lamination with 2.6 m to 14.25 m Heterolithic Moderate Same as Distal delta front sedimentation with variable
to fine-grained small-scale swaley cross-stratification; above reworking by current and wave ripples as well as
sandstones and siltstones Bedding (very thin to thin) organic activities; sand deposition is by gravity-driven
hyperpycnal flow/inertia and buoyancy
6 Fine-grained sandstones Current to wave ripple lamination with 2.7 m to 5.95 m Heterolithic Moderate 70% : 30% Mid-delta front sedimentation below fair weather
with siltstones low-angle planar lamination and wave base and above storm wave base. Variation in
small-scale swaley cross-stratification wave amplitude and sinuosity on different bed
(SCS); Bedding (very thin to thin) surfaces is indicative of fluctuation in wave energies
7 Fine-grained sandstones Planar cross-stratification with wave to 0.8 m to 30.7 m Heterolithic 65% : 35% Same as above
with siltstones current ripples and occasional parallel
lamination and SCS; Bedding (thin to thin)
8 Fine to medium-grained Low-angle planar lamination with wave 1.5 m to 5 m Heterolithic Moderate 80% : 20% Same as above
sandstones with siltstones ripples Bedding (thin to medium) Intense 80% : 20%
9A Soft-sediment deformed Homogenous fine sand/silt with 4 m to 6.5 m Soft-sediment 60% : 40% This represents product of intense soft-sediment
sandstones slump fold deformation deformation involving horizontal gliding of sediments

SOUTH AFRICAN JOURNAL OF GEOLOGY


siltstones (Slump) due to delta front instability and mass transport of
sediments under gravity over the shelf break
9B Soft-sediment deformed Homogenous sand and silt with load 3.35 m to 18.2 m Soft-sediment Same as Intense and localised soft-sediment deformation
sandstones and siltstones casts (i.e. fine to medium sand bulbous deformation above with no horizontal movement, resulting from rapid
(Loading/Dewatering structures) and flame structures loading of coarser sediments over finer sediments
FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA

structures) (i.e. very fine sand to silt injectites) under gravity effect
10 Fine to medium-grained Low-angle planar cross-stratification with 0.51 m to 1.55 m Sandstone 90-100% Delta front sandstones deposited by friction i.e.
bedded sandstones wave bedforms; Bedding (medium sand bedload features at or just below fair
to thick) weather wave base
11 Medium-grained Massive to planar cross-bedding with 3 m to 6.2 m Sandstone Same as Proximal river-dominated delta front sandstones
amalgamated bedded wave bedforms and very rare above probably deposited as bedload to gravity-driven
sandstones horizontal lamination hyperpycnal flow in subaqueous setting; also
sediments are being reworked by wave ripples
12 Medium-grained Massive to planar cross-bedding with 5m Sandstone Moderate 90% : 10% Same as above
amalgamated sandstones wave bedforms and very rare (northeastern
interbedded with sand/ horizontal lamination section)
silt interbeds
W.A. SONIBARE, D. MIKE AND D.I. COLE 313

Figure 7. Example of a log (VS1-C3) through cycle 3. These vertical successions are predominantly heterolithic facies indicating
progradational to aggradational sediment stacking pattern. Note the high GR values for this succession.

Facies distribution from measured outcrop logs 3. mudstone facies (depofacies 1); and
Based on texture and sedimentary structures twelve 4. soft-sediment deformation facies (depofacies 9A and
sedimentary depofacies have been identified and 9B) (Figure 8).
described (see Table 1; Figures 5, 6 and 7). These are Depofacies are thus the building blocks for each
grouped into four lithofacies associations based on lithofacies association while the later constitute the
lithological characteristics (using texture, and sand/clay building blocks for each facies succession (i.e. cycle).
ratio) and inferred depositional processes (using Sandstone facies, heterolithic facies, soft-sediment
sedimentary structures). The four associations are: deformation and mudstone facies represent 9%, 70%,
1. sandstone facies (depofacies 10, 11 and 12); 20% and 1%, respectively, of the described facies in the
2. heterolithic facies (i.e. alternations of siltstone and studied stratigraphic interval for this work. Figure 9
sandstone: depofacies 2, 3, 4, 5, 6, 7 and 8); shows the proportion of each of the four lithofacies

SOUTH AFRICAN JOURNAL OF GEOLOGY


314 FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA

Figure 8. Distribution of facies associations for cycles 1, 2 and 3 for the measured four vertical profiles (i.e. VS1, VS2, VS3 and VS4).

associations for each facies succession (i.e. cycle) along surface marked by mudstones and thinly laminated
the four vertical profiles. siltstones. The geometry of each clinothem is
reconstructed by detailed lateral correlation of facies
Sedimentology and depositional environments from west to east (Figure 9) along the dominant
Facies model northeast paleoflow direction measured from
The combination of detailed facies analysis (Figures 5, 6, unidirectional current ripples. If the vertical facies
7 and 9) and existing facies models for universal variations across subsequent cycles reflect lateral facies
classification of delta systems (Postma, 1990; equivalents within a cycle, then the upward change in
Bhattacharya, 2006) forms the basis for a hypothetical features decrease in gravitational effects, increase
descriptive facies model that depicts the temporal and in wave reworking, and decrease in slope gradient of
spatial distribution of facies, and their geometry and subsequent cycles reflects a natural delta progradation
architecture in three dimensions (Figure 10). This model (see Figures 10 and 11). The overall decrease in
is based on the generally accepted criteria for classifying Kookfontein cycle thicknesses at the Pienaarsfontein
deltas i.e. locality would then indicate that the vertical sequence
1. feeder system; covers only the top part of the Kookfontein clinoforms,
2. depth ratio; i.e. from mid-slope to shelf-margin (Figure 11).
3. river-mouth processes, particularly effluent or jet- This hypothesis is further supported by vertical facies
types; and successions through cycles 1 to 3 (Figures 5, 6 and 7),
4. basin dynamics (i.e. reworking processes by waves, which reflect a lateral succession from distal (i.e. deeper
tides and gravity). facies starting at mid-slope) to proximal (i.e. shallower
Each facies succession (i.e. cycle: C1, C2 and C3; facies ending at top-slope/shelf-margin). The shelf-
Figure 9) in the studied stratigraphic interval represents margin to top-slope is then dominated by: coarser
a clinothem bounded above and below by a flooding grains, frictional forces, in situ soft-sediment deformation

SOUTH AFRICAN JOURNAL OF GEOLOGY


W.A. SONIBARE, D. MIKE AND D.I. COLE 315

structures (with virtually no horizontal movement dewatering structures), basinwards and eastwards
(Oliviera et al., 2010), i.e. loadcasts, flames dipping foreset (i.e. fine-grained sandstone and siltstone
and dewatering structures); alternation of rounded- and interbeds with interlayered slumps, loadcasts, flames
sharp- crested wave reworked bedforms; current ripples; and dewatering structures) and parallel laminated
and moderate to intense bioturbation (Figure 11). bottomset (i.e. mudstones and thinly laminated siltstones
The mid-slope comprises: finer grains, inertia/buoyancy, and possibly low-density turbidites). According to the
slumping (i.e. resulting from some horizontal downslope estimated water-depth for the Permian Tanqua-Karoo
movement of sediments), in-situ soft-deformation and the observed sedimentary structures and textures in
structures; less wave rippled bed-tops; hummocky; the field, Postma (1990)s gravitationally modified
horizontal/current ripple laminations; and weak to deepwater Gilbert-type delta is the closest norm to the
moderate bioturbation (Figure 11). Kookfontein delta system. The delta front is therefore a
This organisation of facies suggests that deposition of prograding Gilbert-type mouthbar which builds out over
these clinothems results from: the shelf break, and consists of confined channelized
1. primary sedimentation of prograding mouthbars flow deposits (i.e. bedload features governed by
governed by stream flow dynamics, and frictional forces) at the river outlet and channelized to
2. secondary remobilisation of sediments immediately unconfined sheet flow deposits downstream
basinwards of river outlets due to rapid loading of (i.e. bedload and suspended-load features governed by
sediments under gravity (Figure 11). inertial and buoyant effluent dynamics) (Figure 11).
The prograding mouthbars for each clinothem coalesce A low-gradient (i.e. fine sand and silt- dominated) feeder
to form a seemingly uniform delta front (i.e. both system D of Postma (1990) is observed in this formation.
the topset and foreset strata), which constitutes of the The estimated basin-margin clinoform profile with
primary deposition by means of traction. The prodelta is relatively low slope angle of 0.5 to 1, water depth of
governed by secondary deposition through means of 150 to 200 m (Wild et al., 2009) and slope length
suspension beyond the influence of gravitational and of approximately 17 to 20 km corroborates this
wave reworking processes. The ensuing Kookfontein hypothesis. All general delta classifications for ancient
clinoforms would then consist of a topset (i.e. massive delta systems (i.e. the ternary river-, wave- and tide-
to planar cross-bedded coarse mouthbar sandstones dominated diagram of Galloway (1975) or its version
interlayered with localised loadcasts, flames and extended with grain size proportion by Orton and

Figure 9. Interpreted correlation panel of the lower sedimentary cycles of the Kookfontein deltaic succession (i.e. cycles 1 to 3) observed at
the Pienaarsfontein locality oriented parallel to the main paleocurrent direction (i.e. NE). This correlation is based on the four facies
associations observed within each facies succession (i.e. cycle).

SOUTH AFRICAN JOURNAL OF GEOLOGY


316 FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA

Figure 10. Hypothetical descriptive facies model for Kookfontein river-dominated, gravitationally modified and wave-influenced shelf edge
Gilbert-type delta. This model forms the basis for describing and interpreting internal heterogeneity and facies architecture of the studied
lower Kookfontein succession (i.e. cycles 1, 2 and 3). Detailed description of longitudinal section A-A is given in Figure 12.

Reading (1993), and Postmas (1990) prototype deltas- Table 1), into heterolithic and sandstone facies that are
based on feeder system, water depth and gravitational dominated by soft-sediment deformation structures and
reworking) offer a useful guide for field observations symmetrical rippled bed-tops. The sandstone beds of
and interpretations. Our descriptive model (Figures 10 heterolithic facies become thicker and coarser upward
and 11) presents a refinement of these existing models with prevalence of unidirectional climbing ripples,
for our specific delta. The studied stratigraphic swaley-cross stratification, slumps and in-situ soft-
succession is bottomed at the centre (i.e. thickest) part sediment deformation structures (i.e. load casts, flames
of the Kookfontein clinoforms (Figure 11). In the and dewatering structures) (Figures 12 and 13). Slump
following section we will describe the three cycles that structures are commonly homogenised layers and their
we studied in detail from the lowest (cycle 1) to the proportion is generally less than load-casts and
uppermost (cycle 3). dewatering structures. Large scale slump folds within
massively deformed unit are encountered towards the
Cycle 1 middle of this cycle at VS2 (Figure 12). The cycle has
This sedimentary cycle consists mainly of heterolithic the highest proportion of soft-sediment deformation
facies and soft-sediment deformation facies with a mean structures of the three cycles studied (Figure 8).
vertical thickness of about 41 m (Figures 5 and 8). Heterolithics with no signs of wave reworking
The cycle is underlain by a claystone on top of Unit 5 of processes (i.e. depofacies 2, 3 and 4) grade upward
the Skoorsteenberg Formation (Figure 5). This claystone into heterolithics with wave reworked bed-tops
has been previously described by Wild (2005) as a 12 m (i.e. depofacies 5, 6, 7 and 8; Figures 14 and 15).
regional claystone marker within the Tanqua The proportion of the former are relatively higher than
depocentre. Along a typical vertical profile, the the latter. Loadcasts of in-situ soft-sediment deformation
lithofacies association passes from mudstones, thinly facies exhibit preservation of swaley cross stratification.
laminated siltstones and undeformed heterolithics Amalgamated massive to planar cross-laminated
(i.e. horizontally to current ripple-laminated sandstones sandstones with wave reworked bed-tops and weak
interbedded with siltstones), with no signs of wave to moderate bioturbation (i.e. 30 to 85 cm thick) close to
reworking processes, (i.e. depofacies 1, 2 and 3; the top of this cycle contains more deformed layers at

SOUTH AFRICAN JOURNAL OF GEOLOGY


W.A. SONIBARE, D. MIKE AND D.I. COLE 317

the most distal profile (VS1) than at the most proximal reactivated by gravitational processes and transported as
profile (VS2) within the studied interval. The stacking slumps (Figure 10). This cycle probably corresponds to
pattern of sediments gives an overall character of the steepest part of the delta slope (Figure 11).
progradation, thickening and coarsening upward The 1 m- thick mudstones at the base of this succession
(Figure 5). However, the vertical thickening- and and the previously described underlain Unit 5 claystone
thinning-upward trends of bedsets (i.e. lithofacies marker by Wild (2005) would then be the prodeltaic
associations) are somewhat irregular (Figure 16); the top facies which are probably deposited by secondary
of this cycle is characterised by a 3.5 m thick unit of deposition through hemipelagic suspension.
thinning- and fining-upward, unidirectional, current
ripple-laminated sandstones with wave reworked bed Cycle 2
forms interbedded with siltstones (Figure 5). Typical The average thickness of this sedimentary succession is
sand/silt ratios from bottom to top within the succession 39 m, and it consists of prograded packages of
vary from 55:45%, 60:40%, 65:35% to 70:30%. heterolithic, i.e. unidirectional ripple, low-angle cross-
laminated and moderately bioturbated very fine to fine
Interpretation: This cycle is interpreted to have been sandstone with siltstone interbeds (depofacies 2, 3, 4, 5,
deposited on the mid-slope margin of the basin with 6, 7 and 8; Table 1; Figures 6 and 8) in a sheet-like
sub-depositional facies ranging from (i.e. proximal to geometry. Towards the upper part of the cycle, these
distal) distal deformed mouthbar sands, intermediate depofacies are overlain by packages of bedded and
to distal and gravitationally reworked delta front and amalgamated massive coarser sandstones (depofacies
distal undeformed delta front (Elliot, 1989; Postma, 10, 11 and 12; Table 1; Figures 6 and 8). The lateral
1990; Figures 10 and 11). These sediments reflect tracing of a 21 cm-thick silt to very fine carbonaceous
sands deposited as bedload by mouthbars that sandy-siltstone bed (see Figure 6) on outcrop windows
prograded beyond the shelf break (Postma 1990; Olariu for approximately 50 to 100 m, and its presence at about
and Bhattacharya, 2006) and were subsequently 13 m (for both VS1 and VS2) from the base of Cycle 2

Figure 11. Detailed description of a longitudinal section A-A through hypothetical facies model for Kookfontein river-dominated,
gravitationally modified and wave-influenced shelf edge Gilbert-type delta showing Kookfontein clinoformal geometry, its facies stacking and
its hydrodynamic zonation. Hydrodynamic Zone A belongs to primary mouthbar deposition governed by stream flow dynamics while
hydrodynamic Zone B and Zone C are governed by secondary remobilisation of sediments under gravity.

SOUTH AFRICAN JOURNAL OF GEOLOGY


318 FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA

Figure 12. Schematic outcrop view of soft-sediment deformation structures at Pienaarsfontein locality. (a) and (b) show large-scale slump
folds observed within highly deformed unit of cycle 1 at Pienaarsfontein vertical section VS2 (c) Dewatering of sandstone layer within
relatively undeformed thinly bedded sandstones interbedded with siltstones.

proved a good tool for resolving and describing the northeast, northeast to east-northeast. Bedded and
facies successions of this cycle. The proportion of soft- amalgamated massive sandstone lithofacies are thicker,
sediment deformation layers (i.e. slumps, loadcasts, cleaner and better sorted upward in this cycle than in
flames and dewatering structures; Figures 12 and 13) is the cycles below and above it (Figure 9). Amalgamated
approximately the same in all the measured vertical massive sandstones are thicker and relatively
profiles (Figure 8). Slump folds oriented in southeast to undeformed at the most proximal profile (VS2) whereas
northwesterly direction are encountered towards the they are deformed with mottled appearance and
base of deformed heterolithic units. Lithofacies interbedded with sand/silt interbeds (Figure 14) at the
associations vary vertically from heterolithic facies most distal profile (VS1). This cycle gives an overall
(i.e. alternation of fine to very fine grained current ripple thickening- and coarsening- upward sequence that is
laminated sandstones interbedded with siltstones overlain by a 2.5 to 3 m thick, thinning- and fining-
influenced by wave processes) to bedded and upward unit of unidirectional current ripple laminated
amalgamated massive to planar cross-laminated and moderately bioturbated sandstones with wave
sandstone facies with wave reworked bed forms reworked bed-tops interbedded with siltstones at the top
(Figure 6). The bed contacts are usually sharp between (Figure 6). Variation in sand/silt ratios for the cycle from
lithofacies associations, whereas they are mainly bottom to top of vertical profiles ranges from 55:45%,
gradational within depofacies. The proportion of 60:40%, 75:25% to 80:20%.
undeformed heterolithics without evidence for wave
rework processes (i.e. depofacies 2, 3 and 4) is less than Interpretation: This cycle is interpreted to represent
in the overlain heterolithics with wave reworked bed- deposition of upper mid-slope to top-slope, with sub-
tops. The wave ripples exhibit symmetrical profiles with depositional facies ranging from proximal undeformed
alternation of broad (i.e. rounded) and sharp crested to distal deformed mouthbar sands, intermediate
wavelengths and are trending mainly northwest to deformed delta front and distal undeformed delta front
southeast, east-west and northeast to southwest (Elliot 1989; Postma, 1990; Olariu and Bhattacharya,
(Figure 15). Typical wave crest to crest distance ranges 2006; Figures 10 and 11). The proximal undeformed
from 8 tp 10 cm. Paleocurrent data measured from mouthbar sands are bedload features (i.e. rolling and
unidirectional current ripples range from north- saltating sediment transport processes) governed

SOUTH AFRICAN JOURNAL OF GEOLOGY


W.A. SONIBARE, D. MIKE AND D.I. COLE 319

Figure 13. Detailed view of loading structures within cycle 2 observed at Pienaarsfontein vertical section VS4. (a) Large-scale loadcasts,
flames and pseudonodules; yellow arrows show fine grained sand/silt flames in between elongated coarser sandstones (i.e. loadcasts)
(b) Loadcasts and flame overlain by deformed layer with extensive pseudonodules (c) White dotted lines show large-scale elongation of
coarser sandstones (i.e. loadcasts) separated by flame structures.

predominantly by frictional forces, and deposited at the thinning upward successions with an overall coarsening
shelf-margin above the shelf break (Wright, 1977; Orton upward trend (Figures 7 and 16). The dominant
and Reading, 1993; Nemec, 1995; Figure 11). The sheet- lithofacies associations are heterolithic (depofacies
like and rhythmic delta front heterolithics are dominated 4, 5, 6, 7 and 8) and soft-sediment deformation facies
by bedload and suspensionload features, and are (depofacies 9B). This cycle has the highest proportion of
interpreted as being deposited by periodic heterolithic facies of the three cycles studied in this work
(and probably some sporadic) mouthbar events (Figure 8). The highly rhythmical and sheet-like layering
followed by little or no river input and subsequent of interbedded unidirectional current ripple laminated
reworking by wave activities. Secondary remobilisation sandstones interbedded with siltstones gradually change
of sediments due to gradient break at the shelf edge upward from thicker siltstone and thinner sandstone
resulted in gravity-driven inertia/buoyancy-dominated packages (i.e. very thin to thin beds) to thinner siltstone
deposits and in-situ soft-sediment deformation and and thicker sandstone packages (i.e. thin to medium
slumping structures (Figure 11). The widespread beds) (Figure 7). Wave ripples are encountered less
and sheet-like geometry of these rhythmites are frequently in the lower part of this cycle than in the
indicative of a stable progradation of the delta front. upper part. Observed wave ripples generally exhibit
The prevalence of wave activities and reduction in symmetrical profiles with alternating sharp-crested
slumping events are indicative of this cycle being ripples and rounded-crested ripples, and trend
shallower than the underlying cycle 1. The alternation dominantly northeast to southwesterly and northwest to
of sharp-crested and rounded-crested wave ripples soiutheasterly (Figure 15). Typical wave crest to crest
on bed-tops (Figure 15) of delta front heterolithics distance ranges from 10 to 12 cm while their amplitude
suggests fluctuations in wave energies (Nichols, ranges from 0.5 to 0.7 cm. Also, the heterolithic facies
2009). are weakly to moderately bioturbated. The dominant
paleocurrent data measured from unidirectional current
Cycle 3 ripples within the cycle range from easterly, east-
The overall thickness of this succession is 35 m, and it northeasterly to northeasterly. All observed soft-
exhibits an asymmetrical profile of thickening and sediment deformation facies are loadcasts, flames and

SOUTH AFRICAN JOURNAL OF GEOLOGY


320 FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA

Figure 14.(a) Thinly bedded, unidirectional current ripple laminated, fine grained sandstones with wave reworked bed-tops interbedded
with siltstones interpreted as intermediate delta front facies; ( b) Amalgamated massive to planar cross laminated sandstones interlayered by
sand and silt interbeds (i.e. white dotted lines) interpreted as proximal mouthbar sands.

Figure 15. View of wave ripples on bed-tops of thinly bedded sandstones. (a) Different orientations and crest-types (i.e. rounded and sharp)
of wave ripples within same unit of heterolithic (i.e. sand and silt interbeds) facies. Red and Yellow arrows show orientation i.e. northwest-
southeast and northeast-southwest respectively (b) Sharp crested wave ripples.

dewatering structures (Figures 12 and 13), and they are Interpretation: These successions represent sediment
variably interbedded heterolithic facies in all the deposition of top-slope to shelf-margin successions with
measured vertical profiles. Near the top, the overlying sub-depositional facies ranging from proximal
deformed heterolithics are 2 to 3 m thick amalgamated undeformed mouthbar sands (i.e. amalgamated massive
massive to planar cross laminated, cleaner well-sorted to planar cross laminated sandstones with wave
fine to medium sandstones with wave reworked and reworked and weakly bioturbated bed-tops) and
moderately bioturbated bed-tops. The top of this cycle is intermediate deformed delta front facies (i.e. wave
capped by a 1 to 1.5 m thick heterolithic unit with wave reworked and moderately bioturbated heterolithics and
reworked and weak bioturbated bed-tops (Figure 7). in situ loading and dewatering structures) (Elliot, 1989;
Variation in sand/silt ratios for the cycle ranges from Postma, 1990; Olariu and Bhattacharya, 2006; Figures 10
bottom to top of vertical profiles from 60:40%, 70:30% and 11). The widespread and well-developed rhythmic
to 80:20%. heterolithics with intense wave reworking processes

SOUTH AFRICAN JOURNAL OF GEOLOGY


W.A. SONIBARE, D. MIKE AND D.I. COLE 321

(Figure 8) is indicative of stable progradation and deformed mouthbar sands (i.e. with loading and
progressive shallowing of the Kookfontein Gilbert-type dewatering structures), to proximal undeformed
delta front. Also, the presence of only in-situ loading mouthbar sands (i.e. without soft-sediment deformation
and dewatering structures, and undeformed mouthbar structures) (Figures 10 and 11). This transition
sands, are suggestive of a decrease in slope gradient of corresponds to decrease and increase in gravitational
this cycle compared to the underlying cycles 2 and 3. and wave reworking processes, respectively, an increase
in bedload/total-load ratios and with decrease in slope
Summary of Cycles 1, 2 and 3 gradient of subsequent cycles. This upward change of
The sedimentary facies scheme obtained for this study is depositional features indicates a progressively
summarised in Table 1. The lowermost unit of the shallowing and natural progradation of the delta system,
Kookfontein deltaic sequence (i.e. Cycles 1, 2 and 3) and represents the deposition of mid-slope to shelf-
represents an upward transition of depositional facies, margin successions. Each cycle is interpreted to have
from prodelta (i.e. top of Unit 5 claystone marker (Wild, been deposited by both episodic and sporadic mouthbar
2005) and depofacies 1) to distal undeformed delta front events governed dominantly by friction-dominated
facies (i.e. weakly bioturbated heterolithics with no bedload features (Figure 11) followed by wave
wave reworked bed-tops), to distal and intermediate reworking of bed surfaces. Secondary deposition is by
deformed delta front facies (i.e. slumps weakly wave remobilisation of sediments under gravity, governed
reworked and moderately bioturbated heterolithics, by slope instability processes and gravity-driven
in situ loading and dewatering structures), to distal inertia/buoyancy-dominated hyperpycnal flows

Figure 16. Variation in bedset (i.e. lithofacies association) thickness across cycles 1, 2 and 3 for measured sections VS1 and VS2 showing
the irregular thickening and thinning upward profile.

SOUTH AFRICAN JOURNAL OF GEOLOGY


322 FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA

Figure 17. Vertical shaliness distribution for the studied stratigraphic interval (i.e. Cycle 1 to Cycle 3) depicting the overall coarsening up
trend of each cycle. Red box = sandstone facies % (0 to 30%), Green box = heterolithic % (30 to 70%) and mudstone facies % (60 to100%);
note the ~10% overlap between heterolithic and mudstone facies.

(Figure 11). The style of soft-sediment deformations soft-sediment deformation structures within the studied
generally exhibits a higher proportion of loading stratigraphic succession. This evidence corroborates the
structures (i.e. in the form loadcasts, flames and hypothesis that cycle 1 was probably located
pseudonodules) and dewatering structures (i.e. water- approximately at the centre (i.e. thickest part with
escape structures) rather than slumping features maximum slope gradient) of the Kookfontein clinoforms
(i.e. lateral/horizontal gravitational gliding of sediments) (Figure 11). Also, loadcasts commonly show
(Figures 12 and 13). Loadcasts comprise coarse preservation of ripples and swaley cross stratifications.
sediments (i.e. usually fine to medium cleaner sands) The alternation of different wave-forms (i.e. rounded to
that sink into the underlying fine sediments (i.e. mud sharp crested waves; Figure 15) on different bed tops
and silts). Slumps are mostly encountered within cycle 1, suggests a hydrodynamic condition with fluctuating
thereby suggesting these to be the most distal (deeper) wave energies (Nichols, 2009).

SOUTH AFRICAN JOURNAL OF GEOLOGY


W.A. SONIBARE, D. MIKE AND D.I. COLE 323

The vertical succession of lithofacies associations for exceptions where the gamma-ray measurements do not
each sedimentary cycle gives a somewhat irregular relate to changes in grain size. For example, a 0.21 m-
thickening- and thinning- upward, progradational thick bed of silt to very fine organic-rich silty-sand with
stacking trend, followed by a retrogradational profile at a sharp base and top contacts (Figure 6), gives low
the top (Figures 5, 6, 7 and 16). However, the vertical gamma radiation similar to some sandstone facies.
progression of the three studied cycles gives an overall This inverse occurrence could be due to low
coarsening- and thickening-upward stacking pattern. concentration of radioactive minerals in the silty-sand
The thinning and fining upward succession on top of bed and possibly better sorting of fine-grained sands
each cycle signifies a change of sediment dispersal (Evans et al., 2007). The approximately higher GR values
mechanisms from friction-dominated bedload for cycle 3 than cycles 1 and 2 (Figures 5, 6 and 7) is
deposition into inertia/buoyancy-dominated bedload probably be due to high concentration of heavy minerals
and suspendedload deposition. The lateral progradation e.g. radioactive, thorium- bearing monazite mineral
of facies and breaks in depositional events within each (Myers and Bristow, 1989).
cycle are believed to be due to a lack of sediments input Potentiality of different facies associations for
from river flow processes, and a possibly local switching reservoir properties was roughly assessed through
of distributary mouthbars. The interpreted depo-system shaliness evaluation (i.e. Equation 1). The results
for the studied successions (i.e. Cycles 1, 2 and 3), based obtained from this evaluation give the following
on vertical facies variations across each cycle and their approximate range of clay content (Vsh) for the main
lateral facies equivalents, is one of a river-dominated, lithologies: clay/silty-clay (60- to 100%), silty/shaly sand
gravitationally reworked and wave-influenced delta front (30 to 70%) and sand (0 to 30%). Variations in shale
of a Gilbert-type mouthbar prograding over the shelf content distribution with depth for different lithofacies
break (Elliot, 1989; Postma, 1990; Olariu and are shown in Figure 17. The overlap between
Bhattacharya, 2006; Figures 10 and 11). heterolithic facies and mudstone (Figure 17) suggests
an effect of background sedimentation, or initial
Gamma ray log characteristics depositional environment, on reservoir quality.
The outcrop gamma-ray logs for this study (Figures 5, 6 Background sedimentation tends to favour fine-grained
and 7) were generated by measurement at a 5 s count deposition as the system naturally transits laterally from
rate and 50 cm sample spacing along each vertical that dominated by bedload/suspendedload features into
sedimentary log. 200 cps is used as the midpoint that dominated by hemipelagic suspension settling
between the maximum (shaded red is 200 cps to downslope. Generally, cleaner and coarser sandstones
maximum) and the minimum (shaded yellow is 200 cps of the proximal mouthbars are characterised by low GR
to minimum) gamma radiations. The first point to note values, while delta front heterolithics and prodelta
is that there is an upward decrease in GR readings mudstones are characterised by high GR values. This
within each cycle, and an overall upward decrease in GR serves as a rather crude estimation for sand to clay ratio,
from cycle 1 to cycle 2. Although the cycle 3 GR values and therefore further analyses requires detailed
are higher than those of cycle 1 and cycle 2, both trends petrographic characterisation for confirmation. It is
show an upward decrease in GR values punctuated by worth mentioning that the GR log offers a useful tool for
peaks which correlate usually with changes in lithology, correlating stratigraphic and lithofacies boundaries, as
grain-size and possibly mineralogy (Myers and Bristow, well as for identifying sandbody architecture in the
1989; Evans et al., 2007). The GR profiles were then studied stratigraphic succession.
plotted against the respective measured sections as
adjacent vertical profiles, and depicted a broadly similar Discussion
trend between lithofacies associations, grain size and Depofacies architecture and geometric elements
GR logs (Figures 5, 6 and 7). The GR peaks mostly The hypothetical descriptive facies model for our
coincide with boundaries between facies associations specific delta, together with the detailed vertical
(i.e. sandstone, heterolithic, mudstone and soft-sediment sedimentary logs and basinwards cross-sectional
deformation facies). However, it should be noted that it profiles (Figure 9), form the basis for characterising the
is very difficult to determine the boundaries between internal architecture and geometry of the studied
heterolithic facies and soft-sediment deformation facies succession. The observed architectural elements seem to
from the GR logs alone, without referring to detailed be typical of a river-dominated, gravitationally reworked
outcrop logs as these two facies give more or less similar Gilbert-type delta deposited in a deepwater setting
GR readings. Also, comparing the GR profiles with (i.e. a prograding Gilbert-type mouthbar over the shelf
detailed outcrop logs depicts the boundaries between break) (Postma, 1990; Olariu and Bhattacharya, 2006;
mouthbar sandstones and delta front heterolithics facies Figures 10 and 11).
(Figure 6). As an expected rule, the gamma radiation The trend of facies successions for each cycle, which
increases with decreasing grain size. This GR trend grade vertically from distal facies (i.e. prodelta and distal
normally conforms to the depositional sequence undeformed delta front) to proximal facies (i.e. distal to
(i.e. overall coarsening-upward succession) that are intermediate deformed delta front, distal deformed
observed on the outcrop logs. However, there are few mouthbar sands and proximal undeformed mouthbar

SOUTH AFRICAN JOURNAL OF GEOLOGY


324 FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA

Figure 18. Schematic geological model for the lower Kookfontein shelf edge Gilbert-type delta (i.e. cycles 1, 2 and 3) showing facies
distribution, stacking and boundaries. At reservoir-scale, we assume flooding surface-bounded parasequence (i.e. facies succession/cycle) to
be a flow unit consisting of flow cells (i.e. facies associations/depofacies). Two important boundaries for flow behaviours are: parasequence
boundary (i.e. flow barrier) and facies/depofacies boundary (i.e. flow conductor).

sands), depicts a naturally prograding delta. Assuming The prograding mouthbars then combine to form a
vertical facies variations across subsequent cycles to be seemingly uniform Gilbert-type delta front that builds
equivalent to lateral facies variations within a cycle, then out over a shelf break (Postma, 1990) (Figure 10).
upward changes such as a decrease in gravity effects, The prevalence of mouthbar sands, heterolithics and
an increase in waves, and a decrease in slope gradient, soft-sediment deformation facies in the studied interval
would indicate that each facies succession (i.e. cycle) is suggests that the Kookfontein Formation represents the
a deltaic clinothem (Figure 11). This study has not been active depositional area (i.e. delta front) of the Tanqua-
able to follow a complete clinothem laterally, as this is Karoo delta system.
difficult to do from the studied outcrops alone; however, The greatest proportion of highly rhythmical and
they can be confidently reconstructed through the sheet-like heterolithic facies (i.e. sands and silts)
upward changes in depositional features, and the overall (Figure 8) suggests a stable progradation of the delta
decrease in cycle thicknesses. front facies. On the whole, the ratio of bedload/total-
The studied successions (cycles 1, 2 and 3) are load decreases from proximal to distal mouthbar sands,
believed to be bottomed within the thickest part of the and from intermediate delta front facies to distal delta
foreset of the Kookfontein clinoforms, i.e. from mid- front facies, thereby reflecting the overall shallowing and
slope to shelf-margin (Figure 11). Juxtaposing observed progradation of the Gilbert-type delta front. Supporting
sedimentary features suggest that each clinothem is evidence for overall shallowing of the delta system are
deposited by: the occurrence of more wave reworked processes at
1. primary mouthbar deposition governed by stream top-slope to shelf-margin successions (i.e. cycles 2 and
flow dynamics; and 3) than at mid-slope succession (i.e. cycle 1) (Figure 11).
2. secondary remobilisation of sediments governed by Wave ripples usually exhibit alternations of sharp-
gravity processes. crested and rounded-crested waves, indicative of

SOUTH AFRICAN JOURNAL OF GEOLOGY


W.A. SONIBARE, D. MIKE AND D.I. COLE 325

fluctuating wave energies. Detailed studies on these upward succession with progradational sediment
wave ripples may yield more information on stacking trend. The overall progradational pattern is
hydrodynamic conditions. Lateral facies variations within indicative of normal regressive shelf edge trajectory path
each cycle are probably related to breaks in depositional (Muto and Steel, 2002; Catuneanu et al; 2009) through
events, due to either no input of sediments through river time (Figure 11). Thus, each cycle is a flooding surface-
processes, or due to switching of distributary mouthbars. bounded parasequence which architecture and
Distribution of soft-sediment deformation structures geometry are governed by the interplay between
throughout this succession (i.e. Cycles 1, 2 and 3) internal (i.e. local depositional processes) and external
suggests extensive gravitational reworking processes (i.e. regional control on sediment supply,
through progradation of distributary mouthbars over the accommodation and base-level changes, e.g. Muto and
shelf break. Large scale soft-sediment deformations Steel (2002); Labourdette et al. (2008)) forces.
structures have also been recorded on the clinoform The observed cyclicity (i.e. regular repetition of
gradients of shelf edge deltas in the Gulf of Mexico sedimentary features) within each cycle suggests that
(average 4 to 8; Suter and Berryhill, 1985) and Rhone this repetitiveness is governed primarily by internal
delta (average ~1; Bhattacharya, 2006). Without further forcing, such as the switching of prograding distributary
study, it is very difficult to determine the influence mouthbars. Such cycles would be akin to Van Wagoner
between the observed soft-sediment deformation et al.s (1990) definition of classic shelf parasequences
structures (i.e. slumps, loading and dewatering that are internally prograding without external forcing.
structures) on one another. However, their spatial At sequence scale, regular repetition of flooding
distribution suggests that they are related to variations in surfaces is more influenced by external depositional
slope gradients. Slump structures such as slump folds controls, i.e. climate, tectonics and eustatic fluctuation.
(Figure 12) are encountered within the lower unit Thus, if the overall progradation of the Kookfontein
(i.e. cycles 1 and 2) of the studied succession, Gilbert-type delta system was filled to base-level, then
representing the deepest soft-sediment deformation the overall reduction in cycle thickness within the
structures at maximum slope gradient. Upward formation was accommodation-driven (i.e. base-level
reduction in slope gradient of subsequent cycles changes governed by tectonic subsidence, with eustasy
coincides with a change in style of soft-sediment responsible for long term decrease in accommodation
deformation from slumps into loading and dewatering space and the overall reduction in cycle thickness).
structures (Figures 10 and 11). According to By contrast, if the system was not filled to base-level,
mechanical models of soft-sediment deformation then the overall reduction in cycle thickness was
development, slumps are governed by gravitational sediment supply-driven (i.e. climate and tectonics may
gliding of sediments involving horizontal movement be responsible for decrease in supply of clastic
(e.g. Martinsen, 1989; Plink-Bjrklund et al., 2001), while sediments to the available accommodation). Testing of
loading and dewatering structures are due to rapid these hypotheses from the rock record alone will be
deposition of sand on mud that may induce density further limited if base-level was never fully filled-up.
instabilities and subsequent rapid escape of pore-
water, i.e. collapse of coarser sediments and vertical Implications for reservoir modeling
sediment motion of finer sediments (Lowe, 1975). The construction and upscaling of accurate geological
As such, loading and dewatering structures tend to be models for reservoir simulation is an important tool
more localised and without horizontal movement within the oil industry for reservoir management and
(e.g. Martinsen, 1989; Bhattacharya, 2006; Oliviera et al., making decisions on drainage strategy (e.g. Lasseter
2010). Rapid loading of sediments and increase in grain et al., 1986; Mayer and Chapin, 1991; Flint and Bryant,
size, due to stable progradation of delta front and 1993; Qi and Hesketh, 2005; Mike and Geel, 2006;
upward reduction in slope gradient of subsequent Keogh et al., 2007; Labourdette et al., 2008; Cabello
cycles, may have been responsible for higher proportion et al., 2010). These models, with the objective to
of in-situ loading and dewatering structures than incorporate all scales of sedimentary heterogeneities, are
slumping structures in the studied successions. built from an integration of subsurface (i.e. seismic, well
The proposed low-slope gradient (i.e. 0.5 to 1) for the log, core and borehole image) and analogue datasets
Kookfontein shelf edge delta system by Wild et al. (i.e. outcrops of palaeosystems and modern systems).
(2009) and Oliviera et al. (2010) is consistent with these The success of any reservoir modelling procedure
observations (Figures 13 and 15). Bhattacharya (2006) depends on how well facies variability is represented in
also noted that even lower slope gradient in the Alberta the geological model. Therefore, detailed facies analysis
shelf edge restrict soft-sediment deformation features and sedimentological models from outcrops can provide
mostly to loading structures rather than large-scale additional information on facies variability and small
slumps, slides or growth faults. scale heterogeneity needed for a robust reservoir model
Each facies succession (i.e. cycle) exhibits a (Flint and Bryant, 1993; Falivene et al., 2006).
somewhat irregular (i.e. asymmetrical) thickening The schematic geological model, consisting of facies
pattern of bedsets (i.e. lithofacies) (Figure 16) but on the successions (i.e. cycles or parasequences) as flow units,
whole depicts overall thickening- and coarsening- and facies associations/depofacies as flow cells, for the

SOUTH AFRICAN JOURNAL OF GEOLOGY


326 FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA

studied succession of the Kookfontein Formation, our deposystem. This led to the identification
is given in Figure 18. This model is populated with facies of twelve depofacies grouped into four facies
at different scales (i.e. facies succession, facies associations (i.e. sandstone, heterolithic, soft-
association, depofacies and laminae) as the basic sediment deformation facies and mudstone facies).
building blocks. Thus, it is ready to be transformed into Lateral juxtaposition of observed vertical facies
a reservoir model. The most important implications of variations across each cycle in a basinwards direction
this work for reservoir modelling of similar subsurface exhibits the following upward changes: decrease
analogues are as follows: in gravity effects, increase in waves and decrease in
1. A hierarchical description of heterogeneity, i.e. slope gradient of the successive cycles.
ranging in decreasing order from facies succession This upward transition represents a change from
(cycle), facies association (bedsets), depofacies deposition of mid-slope (i.e. distal consisting of finer
(beds) and laminae, offers a simple and grains, inertia/buoyancy deposits, less waves,
straightforward approach to describe internal horizontal/current ripples, swaley/hummocky,
architecture of facies, and in particular sandbodies, loading/dewatering structures and slumps) to shelf-
in the ensuing geological model (Figure 18). margin (i.e. proximal consisting of coarser grains,
This approach follows from the previous work by friction/bedloads, more wave reworking processes,
Mike and Geel (2006). current ripples and loading/dewatering structures).
2. Each flooding surface-bounded parasequence is a The studied sections exhibit an overall upward
flow unit, and each flow unit (consisting of flow thickening and coarsening succession with
cells, i.e. facies/depofacies) can be divided into progradational stacking pattern typical of a normal
rectangular grid cells (Mike et al., 2006; Labourdette regressive prograding delta. Though asymmetrical
et al., 2008; Cabello et al., 2010). thickening and thinning profiles are noted within
3. Two important boundaries that could determine fluid each cycle.
flow are parasequence (i.e. cycle) and depofacies Deposition of each facies succession (i.e. flooding
boundaries. Parasequence boundaries as flooding surface-bounded cycle) is a result of primary
surfaces, may be flow barriers, while depofacies deposition by episodic and probably sporadic
boundaries may be flow conductors (particular if mouthbar events governed by stream flow dynamics,
they are substantially gradational). Determination of and secondary remobilisation of sediments under
the degree of facies interfingering might also be gravity.
crucial to predicting facies boundary behaviour to Lateral facies variations within each cycle are due to
fluid flow. breaks in depositional events resulting from: either
4. The reworking of sediments by gravity processes, no sediment input from river flow processes, or
waves and bioturbation could also have an important switching of distributary mouthbars.
effect on porosity and permeability reduction. The basinwards transition of depositional facies
5. Recognition of facies variations within facies ranges from undeformed mouthbar sands into
succession (i.e. cycle) is very difficult from GR log deformed mouthbar sands, intermediate and distal
alone without performing Outcrop log-GR log tie. deformed delta fronts, and distal undeformed delta
6. The described internal heterogeneity in this work is fronts and prodeltas. Based on this transition, the
below the resolution (i.e. mm-scale) of most architecture and geometry of our depo-system is
conventional well-logs, and therefore can interpreted to be river-dominated, gravitationally
supplement well-log data, especially where there is reworked and wave-influenced shelf edge Gilbert-
no borehole image and core data. type delta.
Widespread distribution of soft-sediments
Conclusion and recommendation deformation structures, their growth-style and
An outcrop study, which combines detailed outcrop morphology within the studied succession are
logs, gamma-ray logs, photopanel interpretation and empirically related to progradation of Gilbert-type
incorporation of existing delta facies models, yields the mouthbars over the shelf break as well as the slope
necessary information required for description and gradients of the Kookfontein clinoforms. Low-slope
interpretation of facies and their sub-environments gradients may have favoured mostly loading and
of deposition. This facilitates the construction of a dewatering structures rather than large-scale slumps.
hypothetical descriptive facies model, describing and Outcrop log-GR log tie offers a useful tool for
interpreting facies architecture, estimating shale (clay) identifying and correlating trends in facies
content and the construction of schematic geological succession, facies boundaries and particularly
model for the Kookfontein shelf edge delta system. sandbody architectures.
The aforementioned allows the following conclusions: Shale content estimation from GR logs is a rather
A hierarchical approach to facies description depicts crude but fast way for assessing reservoir potentiality
all levels of heterogeneity (i.e. facies succession/ of encountered facies in the field. Its reliability
cycle, facies association/bedsets, depofacies/beds requires detailed petrographic characterisation for
and laminae) that form the basic building blocks for comfirmation.

SOUTH AFRICAN JOURNAL OF GEOLOGY


W.A. SONIBARE, D. MIKE AND D.I. COLE 327

The described internal heterogeneity in this work is Deltas: Sites and Traps for Fossil Fuels. The Geological Society, London,
Special Publication, 41, 2146.
below the resolution (i.e. mm-scale) of most
Evans, R., Mory, A.J. and Tait, A.M., 2007. An outcrop gamma ray study of
conventional well-logs, and therefore could the Tumblagooda Sandstone, Western Australia. Journal of Petroleum
supplement well-log data, especially where there is Science and Engineering, 57, 3759.
no borehole image and core data. Faure, K. and Cole, D., 1999. Geochemical evidence for lacustrine microbial
From this work, a recommendation to study the blooms in the vast Permian Main Karoo, Paran, Falkland Islands
and Huab basins of southwestern Gondwana. Palaeogeography
remaining upper 180 to 230 m of Kookfontein Formation
Palaeoclimatology Palaeoecology, 152, 189213.
would be useful to further test our hypothetical facies Falivene, O., Arbues, P. J., Howell, J., Fernandez, O., Cabello, P., Munoza,
model and advance our understanding of the J.A. and Cabrera, L., 2006. A FORTRAN program to introduce field-
Kookfontein clinoformal geometry. measured sedimentary logs into reservoir modelling packages. Computers
and Geosciences, 32, 15191522.
Fildani, A., Drinkwater, N.J., Weislogel, A., McHargue, T., Hodgson, D.M. and
Acknowledgements
Flint, S.S., 2007. Age controls on the Tanqua and Laingsburg deep-water
We are grateful to the Department of Earth Sciences, systems: new insights on the evolution and sedimentary fill of the Karoo
Stellenbosch University, for providing the full logistical Basin, South Africa. Journal of Sedimentary Research, 77, 901908.
and financial support for the fieldwork in 2009, and Fildani, A., Weislogel, A., Drinkwater, N.J., McHargue, T., Tankard, A.,
Inkaba yeAfrica for providing the 2010 research funds Wooden, J., Hodgson, D.M. and Flint, S.S., 2009. U-Pb zircon ages from
the southwestern Karoo Basin, South Africa implications for the
for the project. We thank the anonymous reviewers for
Permian-Triassic boundary. Geology, 37, 719722.
their critical comments. This is Inkaba yeAfrica Flint, S.S. and Bryant, I.D., 1993. The geological modelling of hydrocarbon
contribution 43. reservoirs and outcrop analogues. Special Publication of the International
Association of Sedimentologists, 15, unpaginated.
References Galloway, W.E., 1975. Process framework for describing the morphologic
Andersson, P.O.D., Worden R.H., Hodgson, D.M. and Flint, S., 2004. and stratigraphic evolution of deltaic depositional systems. In: M.L.
Provenance evolution and chemostratigraphy of a Palaeozoic submarine Broussard (Editor), Deltas, Models for Exploration. Houston Geological
fan-complex: Tanqua Karoo Basin, South Africa. Marine and Petroleum Society, Texas, U.S.A., 8798.
Geology, 21, 555577. Hodgson, D.M., Flint, S.S., Hodgetts, D., Drinkwater, N.J., Johannessen, E.P.
Bangert, B., Stollhofen, H., Lorenz, V. and Armstrong, R., 1999. The and Luthi, S.M., 2006. Stratigraphic evolution of fine-grained submarine
geochronology and significance of ash-fall tuffs in the glaciogenic fan systems, Tanqua depocentre, Karoo Basin, South Africa. Journal of
Carboniferous-Permian Dwyka Groupof Namibia and South Africa. Sedimentary Research, 76, 2040.
Journal of African Earth Sciences, 29, 3349. Johnson, S.D., Flint, S., Hinds, D. and Wickens, H. de V., 2001. Anatomy,
Bhattacharya, J.P., 2006. Deltas. In: H.W. Posamentier and R.G. Walker geometry and sequence stratigraphy of basin floor to slope turbidite
(Editors), Facies models revisited. Society for Sedimentary Geology, systems, Tanqua Karoo, South Africa. Sedimentology, 48, 9871023.
Special Publication, 84, 237292. Johnson, M.R., 1991. Sandstone petrography, provenance and plate tectonic
Bhattacharya, J.P. and Giosan, L., 2003. Wave-influenced deltas: setting in Gondwana context of the southeastern Cape-Karoo Basin. South
geomorphological implications for facies reconstruction. Sedimentology, African Journal of Geology, 94, 137154.
50, 187210. Johnson, M.R., Van Vuuren, C.J., Visser, J.N.J., Cole, D.I., Wickens, H. de V.,
Booth, P., 2011. Stratigraphic, structural and tectonic enigmas associated Christie, A.D.M., Roberts, D.L. and Brandl, G., 2006. Sedimentary rocks of
with the Cape Fold Belt: Challenges for future research. South African the Karoo Supergroup. In: M.R. Johnson, C.R. Anhaeusser and R.J.
Journal of Geology, 114, to complete Thomas (Editors), the Geology of South Africa. Geological Society of
Bouma, A.H. and Wickens, H. de V., 1991. Permian passive margin South Africa/Council for Geoscience, South Africa, 461499.
submarine fan complex, Karoo Basin, South Africa: possible model to Keogh, K.J., Martinius, A.W. and Osland, R., 2007. The development of
Gulf of Mexico. Transactions of the Gulf Coast Association of Geological fluvial stochastic modelling in the Norwegian oil industry: A historical
Societies, 41, 3042. review, subsurface implementation and future directions. Sedimentary
Cabello, P. O., Falivene, O., Lopez-Blanco, M., Howell, J., Arbues, P. and Geology, 202, 249268.
Ramos, E., 2010. Modelling facies belt distribution in fan deltas coupling Labourdette, R., Casas, J. and Imbert, P., 2008. 3D sedimentary modelling of
sequence stratigraphy and geostatistics: The Eocene Sant Lloren del Munt a Miocene Deltaic reservoir unit, Sincor Field, Venezuela: A new
example (Ebro foreland basin, northeast Spain). Marine and Petroleum approach. Journal of Petroleum Geology, 31, 2, 135152.
Geology, 27, 254272. Lasseter, T. J., Waggoner, J. R. and Lake, L. W., 1986. Reservoir
Catuneanu, O., Hancox, P.J. and Rubidge, B.S., 1998. Reciprocal flexural heterogeneities and their influence on ultimate recovery. In: Qi, D. and
behaviour and contrasting stratigraphy: a new basin development model Hesketh, T. (Editors), an analysis of upscaling Techniques for Reservoir
for the Karoo retroarc foreland system, South Africa. Basin Research, Simulation. Petroleum Science and Technology, 23, 827842.
10, 417439. Lindeque, A., de Wit, M.J., T Ryberg, T., M Weber,M. and L Chevallier, L.,
Catuneanu, O., Abreu, V., Bhattacharya, J.P., Blum, M.D., Dalrymple, R.W., 2011. Deep Crustal Geophysical Profile Across the Southern Karoo Basin
Eriksson, P.G., Fielding, C.R., Fisher, W.L., Galloway, W.E., Gibling, M.R., and Beattie Magnetic Anomaly, South Africa: an Integrated interpretation.
Giles, K.A., Holbrook, J.M.,Jordan, R., St.C., Kendall, C.G., Macurda, B., South Afrrican Journal of Geology, 114, to complete
Martinsen, O.J., Miall, A.D.,Neal, J.E., Nummedal, D., Pomar, L., Lock, B.E., 1980. Flat-plate subduction and the Cape Fold Belt of South
Posamentier, H.W., Pratt, B.R., Sarg, J.F.,Shanley, K.W., Steel, R.J., Strasser, Africa. Geology, 8, 3539.
A., Tucker, M.E. and Winker, C., 2009. Towards the standardisation of Longhitano, S.G., 2008. Sedimentary facies and sequence stratigraphy of
sequence stratigraphy. Earth-Science Reviews, 92, 133. coarse-grained Gilbert-type deltas within the Pliocene thrust-top Potenza
Cole, D.I., 2005. Prince Albert Formation, In: M. R. Johnson (Editor). Basin (Southern Apennines, Italy). Sedimentary Geology, 210, 87110.
Catalogue of South African Lithostratigraphic Units. South African Lowe, D.R., 1975. Water escape structures in coarse-grained sediments.
Committee for Stratigraphy, 8, 3336. Sedimentology, 22, 157204
Cole, D.I. and Vorster, C.J., 1999. The metallogeny of the Sutherland area. Martinsen, O.J., 1989. Styles of soft-sediment deformation on a Namurian
Explanation of Metallogenic Sheet 3220 (scale 1:250 000), Council for (Carboniferous) delta slope, Western Irish Namurian Basin, Ireland.
Geoscience, Pretoria, 41pp. In: M.K.G. Whateley and K.T. Pickering (Editors), Deltas: Sites and Traps
Elliott, T., 1989. Deltaic systems and their contributions to an understanding for Fossil Fuels. The Geological Society, London, Special Publication,
of basin-fill successions. In: M.K.G. Whateley and K.T. Pickering (Editors), 41, 2146.

SOUTH AFRICAN JOURNAL OF GEOLOGY


328 FACIES ARCHITECTURE OF KOOKFONTEIN SHELF EDGE DELTA

Mayer, D.F. and Chapin, M.A., 1991. A comparison of outcrop and subsurface Suter, J.H., and Berryhill, H.L. JR., 1985. Late Quaternary shelf-margin deltas,
geologic characteristics and fluid flow properties in the lower Cretaceous Northwest Gulf of Mexico: American Association of Petroleum Geologists,
muddy J. Sandstone. In: Bill Linville (Editor), Reservoir Characterisation Bulletin, 69, 7791.
III. PennWell Publishing Company, Oklahoma, U.S.A., 327352. Tankard, A., Welsink, H., Aukes, P., Newton, R. and Stettler, E., 2009.
Mike, D. and Geel, C.R., 2006. Standard facies model to incorporate all Tectonic evolution of the Cape and Karoo basins of South Africa. Marine
heterogeneity levels in a reservoir model. Marine and Petroleum Geology, and Petroleum Geology, 26, 13791412.
23, 943959. Theron, J.N., 1983. Die geologie van die gebied Sutherland (skaal 1:250 000).
Mike, D., Barzandji, O.H.M., Bruining, J. and Geel, C.R., 2006. Upscaling of Explanation of Geological Sheet 3220, Geological Survey of South Africa,
small-scale heterogeneities to flow units for reservoir modelling. Marine 29pp.
and Petroleum Geology, 23, 931942. Turner, B.R., 1999. Tectonostratigraphical development of the Upper Karoo
Milani, E.J. and de Wit., M.J., 2008. Correlations between classic Parana and foreland basin: orogenic unloading versus thermally-induced Gondwana
Cape-Karoo basins of South America and southern Africa and their basin rifting. Journal of African Earth Sciences, 28, 215238.
infills flanking the Gondwanides: Du Toit revisited. In: R.J. Pankhurst, Van der Werff, W. and Johnson, S., 2003. High resolution stratigraphic
R.A.J. Trouw, B.B. de Brito Neves and M.J. de Wit (Editors). West analysis of a turbidite system, Tanqua Karoo Basin, South Africa. Marine
Gondwana: Pre-Cenozoic Correlations Across the South Atlantic Region. and Petroleum Geology, 20, 4569.
The Geological Society, London, Special Publications, 294, 319342. Van Lente, B., 2004. Chemostratigraphic trends and provenance of the
Nichols, G., 2009. Sedimentology and Stratigraphy, Wiley-Blackwell Permian Tanqua and Laingsburg depocentres, Southwestern Karoo Basin,
Publication UK, 2nd edition, 179196. 419pp. South Africa. Unpublished PhD Thesis, University of Stellenbosch, South
North, C.P. and Boering, M., 1999. Spectral gamma-ray logging for facies Africa, 339pp.
discrimination inmixed fluvialeolian successions: a cautionary tale. Van Wagoner, J.C., Mitchum, R.M., Campion, K.M. and Rahmani, V.D., 1990.
American Association of Petroleum Siliciclastic sequence stratigraphy in well logs, cores, and outcrops:
Geologists Bulletin, 83, 155169. concepts for high resolution correlation for time and facies. American
Olariu, C. and Bhattacharya, J.P., 2006. Terminal distributary channels and Association of Petroleum Geologists, Methods in Exploration, 7, 55pp.
delta frontarchitecture of river-dominated delta systems. Journal of Veevers, J.J., Cole, D.I. and Cowan, E.J., 1994. Southern Africa: Karoo basin
Sedimentary Research, 76, 212233. and Cape Fold Belt. In: J.J. Veevers and C.McA. Powell (Editors), Permian-
Oliveira, C.M.M., Hodgson, D.M. and Flint, S.S., 2010. Distribution of soft- Triassic Pangean Basinsand Foldbelts along the Panthalassan Margin of
sedimentdeformation structures in clinoform successions of the Permian Gondwanaland. Geological Society of America, Memoir, 184, 223279.
Ecca Group, Karoo Basin, South Africa. Sedimentary Geology, Viljoen, J.H.A., 1994. Sedimentology of the Collingham Formation, Karoo
235, 314330. Supergroup. South African Journal of Geology, 97, 167183.
Orton, G. and Reading, H.G., 1993. Variability of deltaic processes in terms Viljoen, J.H.A., 2005. Tierberg Formation, In: M.R. Johnson, (Editor).
of sediment supply, with particular emphasis on grain size. Catalogue of South African Lithostratigraphic Units. South African
Sedimentology, 40, 475512. Committee for Stratigraphy, Volume 8, 3740.
Phillips, P. and Wen, R., 2007. Improving net-to-gross reservoir estimation Visser, J.N.J., 1992. Deposition of the Early to Late Permian Whitehill
with small-scale reservoir modelling. American association of Petroleum Formation during a sea-level highstand in a juvenile foreland basin. South
Geologists Search and Discovery Article #40252 delivered at AAPG annual African Journal of Geology, 95, 181193.
convention, California, April 2007, unpaginated. Visser, J.N.J. and Praekelt, H.E., 1996. Subduction, mega-shear systems and
Plink-Bjrklund, P., Mellere, D. and Steel, R.J., 2001. Turbidite variability and Late Paleozoic basin development of Gondwana. Geologische Rundschau,
architecture of sand-prone, deep-water slopes: Eocene clinoforms in the 86, 632646.
Central Basin, Spitsbergen. Journal of Sedimentary Research, 71, 895912. Wickens, H. de V., 1992. Submarine fans of the Permian Ecca Group in the
Postma, G., 1990. An analysis of the variation in delta architecture. Terra SW Karoo Basin: Their origin and reflection on the tectonic evolution of
Nova, 2, 124130. the basin and its source areas. In: M.J. de Wit and I. Ransome (Editors),
Qi, D. and Hesketh, T., 2005. An analysis of upscaling techniques for Inversion Tectonics of the Cape Fold Belt, Karoo and Cretaceous Basins
reservoir simulation. Petroleum Science and Technology, 23, 827842. of Southern Africa. Balkema, Rotterdam, 117125.
Rubidge, B.S., 2005. Re-uniting lost continents Fossil reptiles from the Wickens, H. de V., 1994. Basin floor fan building turbidites of the
ancient Karoo and their wanderlust. South African Journal of Geology, southwestern Karoo Basin, Permian Ecca Group, South Africa.
108, 135172. Unpublished PhD Thesis, University of Port Elizabeth, South Africa,
Rubidge, B.S., Hancox, P.J. and Catuneanu, O., 2000. Sequence analysis of 223pp.
the Ecca-Beaufort contact in the southern Karoo of South Africa. South Wickens, H. de V., 1996. Die stratigrafie en sedimentologie van die Groep
African Journal of Science, 103, 8196. Ecca wes van Sutherland. Bulletin of the Geological Survey of South
Rubidge, B.S., Modesto, S.P., Sidor, C. and Welman, J., 1999. Eunotosaurus Africa, 107, 49pp.
africanus from the Ecca-Beaufort contact in Northern Cape Province, Wild, R.J., 2005. Sedimentological and sequence stratigraphic evolution of a
South Africa implications for Karoo Basin development. South African Permian Lower Slope to Shelf Succession, Tanqua Depocentre, SW Karoo
Journal of Science, 95, 553555. Basin, South Africa. Unpublished PhD Thesis, University of Liverpool,
Sonibare, W.A., 2011. Sedimentary modelling and petrographic U.K., 350pp.
characterisation of a Permian Tanqua-Karoo shelf edge delta, SW Karoo Wild, R.J., Flint, S.S. and Hodgson, D.M., 2009. Stratigraphic evolution of the
Basin (South Africa). Unpublished MSc Thesis, University of Stellenbosch, upper slope and shelf edge in the Karoo Basin, South Africa. Basin
South Africa, 201pp. Research, 21, 502527.
Stankiewicz, J., Ryberg, T., Schulze, A., Lindeque, A., Weber, M.H. and de
Wit, M.J., 2007. Initial results from wide-angle seismic refraction lines in
the southern Cape. South African Journal of Geology, 110, 407418 Editorial handling: M.J. de Wit

SOUTH AFRICAN JOURNAL OF GEOLOGY

View publication stats

Vous aimerez peut-être aussi