You are on page 1of 21

Marine and Petroleum Geology 111 (2020) 240–260

Contents lists available at ScienceDirect

Marine and Petroleum Geology


journal homepage: www.elsevier.com/locate/marpetgeo

Research paper

Sedimentology, geochemistry and reservoir potential of the organic-rich T


Qusaiba Shale, Tabuk Basin, NW Saudi Arabia
Mohamed Abouelresha,e,∗, Lamidi Babalolab, Abdaseed Bokharic, Mohammed Omerd,
Thomas Koithand, Daniel Boyded
a
Center for Environment & Water, RI, King Fahd University of Petroleum & Minerals, Saudi Arabia
b
CIPR, CPG, King Fahd University of Petroleum & Minerals, Saudi Arabia
c
Faculty of Petroleum Engineering, Peace University, Sudan
d
Weatherford, Dhahran Techno Valley, King Fahd University of Petroleum & Minerals, Saudi Arabia
e
Faculty of Petroleum & Mining Engineering, Suez University, Egypt

ARTICLE INFO ABSTRACT

Keywords: The Qusaiba Shale Formation of the Qalibah Group is the most prolific source rock for the Paleozoic petroleum
Sedimentology system in Saudi Arabia and is recently considered as a potential unconventional shale gas reservoir. Detailed
Organic rich shale sedimentological, petrographical, and geochemical analyses were undertaken on an ~31 m thick outcrop of this
Qusaiba organic-rich shale exposed in the northwest of Tayma City, NW of Saudi Arabia to provide a better under-
Unconventional resources
standing of the vertical and lateral lithofacies variations and the effects of such variations on reservoir prop-
Geochemistry
erties.
X-ray diffraction analysis reveals that clay minerals (average: 50%) and quartz (average: 28%) dominate the
bulk mineralogical composition while K-feldspar, plagioclase, and pyrite occur as common to minor components
in the analyzed samples. An integration of the sedimentological description, petrography and mineralogical
composition allows the subdivision of the investigated outcrop section into six lithofacies. These are: Laminated
organic-rich shale, Graptolitic concretion, Massive organic-rich shale, Organic-lean shale, Siltstone-shale interbedded,
and Cross-stratified siltstone/sandstone lithofacies. TOC content in the laminated and massive organic-rich shale
lithofacies ranges between 0.5 and 6.1 wt% (average 2.52 wt%) then it decreases upward (0.04–0.35 wt%,
average 0.18 wt%) in the overlying organic-lean shale and siltstone lithofacies. Trace elements (Cu, Mo, As, Ni,
Co, U and V) show positive excursions in the lowermost 20 m of the organic-rich, dark-grey shale interval and
display abrupt depletions at the contact with overlying organic-lean shale. The enrichment factors of these trace
elements are characterized by similar distribution patterns as their measured concentrations. Like the elements,
these factors also exhibit a significant depletion at the contact between the organic-rich shale and the overlying
organic-lean shale. Out of all the redox indices, V/(V + Cr), V/(V + Ni), Th/U and [Mo]s/TOC were found to
best distinguish between the organic-rich and the overlying organic-lean intervals.
The laminated organic-rich shale lithofacies indicate deposition of hemipelagic sediments by a relatively low
energy quiet anoxic bottom water, most likely in a sulfidic conditions that were associated with a strongly
restricted bottom water mass circulation at the time of its deposition. This restricted hydrographic condition was
likely enhanced in the study area due to the paleotopography that was created by the deglaciation of the Early
Silurian time. The TOC richness, common occurrence of pyrite and high enrichment of redox elements Mo, U and
V in the organic-rich shale intervals support the existence of anoxic bottom water at the time of deposition. The
abrupt decrease of the TOC and the redox sensitive elements and their associated indices from the underlying
organic-rich lithofacies to the overlying organic-lean shale lithofacies is likely due to a gradual increase in
oxygen level. This improved bottom water oxygenation condition might have led to a reduction in organic
productivity and limited organic matter preservation in the organic-lean shale lithofacies. The graptolitic con-
cretion lithofacies found in the sequence, refers to an in-situ, syngenetic to early diagenetic origin formed close
to the sediment-water layer when there was little or no deposition or at periods characterized by low sedi-
mentation rates. The siltstone - fine sandstone lithofacies in the topmost part of the studied section, is interpreted
to indicate the dominance of high energy and highly oxygenated shallow bottom water conditions during de-
position, probably in a middle to inner shelf depositional settings.


Corresponding author.
E-mail address: omaradh@yahoo.com (M. Abouelresh).

https://doi.org/10.1016/j.marpetgeo.2019.05.001
Received 30 November 2017; Received in revised form 14 February 2019; Accepted 2 May 2019
Available online 16 August 2019
0264-8172/ © 2019 Elsevier Ltd. All rights reserved.
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

The kerogen in QS is a mixture of type III (gas prone) and type IV (inert kerogen), Tmax values with an average
of 430 °C, and VRo values (average 0.58) calculated from the Tmax, assigned the studied section as immature.
Qualitative evaluation of porosity allows the identification of three different types of porosity including in-
organic porosity, organic porosity and natural fracture porosity. Although the ductile clay minerals are the most
abundant minerals in the lower part of the studied QS (average 56%), the brittle minerals (Quartz, Feldspar, and
Pyrite) form an average of 42% and exhibit a uniform distribution of brittleness in the organic-rich interval.

1. Introduction source rocks in the Middle East. It is equivalent to the other formations
such as the Mudawwara in Jordan, Sahmah in Oman, Akkas in Iraq,
Tabuk Basin located in northwest Saudi Arabia, extends into Jordan Dadas in southeast Turkey, Ghakum in Iran, Tanf in Syria, and Ta-
and Iraq to the north. The Precambrian rocks of the Arabian Shield from nezzuft in Libya (Konert et al., 2001; Lüning et al., 2005). The type and
the west and the south border Tabuk Basin while the northerly plunging maturity level of the organic content in the QS has been deeply in-
Ha'il-Rutbah Arch which marks the east side of the basin and separates vestigated and evaluated (e.g., Mahmoud et al., 1992; and Cole et al.,
it from the Widyan Basin (Al-Laboun, 1986, 2009) (Fig. 1A–B). These 1994). The most recent studies showed a significant variation of ma-
two basins (Tabuk and Widyan basins) are considered as parts of a turity from immature to over mature levels within the QS in the
broad Gondwana stable Paleozoic shelf, which was covered by the northwestern region of Saudi Arabia (Inan et al., 2016; and
Tethys Sea transgression during the Early Silurian time (Fox and MorCheshire et al., 2017).
Ahlbrandt, 2002). In 2008, Saudi Aramco initiated the exploration of unconventional
The Qusaiba Shale Formation (QS) of the Qalibah Group (Fig. 1C) is shale gas in the northwest area of Saudi Arabia by drilling the Shale-1
the most prolific source rock for the Paleozoic petroleum system in well to evaluate the reservoir quality of the Silurian shale (Alexeyenko
Saudi Arabia (Abu-Ali et al., 1991, 1999; McGillivray and Husseini, et al., 2012). This exploration activity addressed the occurrence of
1992; Jones and Stump, 1999; Konert et al., 2001; and Arouri et al., economic reserves in QS at shallow depths. The same formation also
2010). Additionally, QS is one of the widely distributed lower Silurian showed good evidence of gas reserves in the Empty Quarter (Rub’ Al-

Fig. 1. Geographic location of the studied section to SE of Tabuk City, NW Saudi Arabia. (B) Tectonic map shows the location of Tabuk Basin and its tectonic borders
(modified after Al-Laboun, 1986). (C) Lithostratigraphic classification of Qalibah Group in Saudi Arabia, the geologic time scale is from International stratigraphic
chart, 2004 (modified after Halawani et al., 2013). (D) Field photo of the studied section of Qusaiba Shale shows the organic richness in the basal unit. (E) Structure
map shows the Tayma Graben where the Qusaiba Shale escarpment represents its northeast side. Red circle is the location of the studied section (Structural map
modified after, Vaslet et al., 1994). (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

241
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

Khali Basin) in the southwest of Saudi Arabia (Stewart et al., 2016). To by fluvial channels, which are filled with moderately to well-sorted
date, production from these reserves has not started, likely due to the Sharawra sandstone. The Sharawra Fm. is unknown in the outcrop in
lack of infrastructure and limited shale gas production experience. the Wajid Plateau, but it is present in sections drilled at sites in
The reservoir properties of the Lower Silurian QS which is con- southwestern Saudi Arabia, where Qusaiba/Sharawra contacts also
sidered as a potential unconventional shale gas reservoir, have not been appear to be conformable. In the subsurface, the Qusaiba/Sharawra
fully understood. Therefore, an outcrop study was undertaken to pro- contact is usually gradational over a vertical distance of between 20 and
vide a better understanding of the vertical and lateral lithofacies var- 61 m.
iations and the effects of such variations on reservoir properties of Megafaunas including graptolites, molluscs, brachiopods, trilobites
potential oil and gas shale reservoirs. The main objective of an outcrop and euryptic fragments are diverse and abundant in the QS (Jones and
study is usually to answer traditional geological questions regarding Stump, 1999). This faunal content is also present in the upper shale
lithofacies, depositional environments, stratigraphy, and mineralogy. unit, but with less abundance. Based on the recovery of macrofauna,
The depositional environments of QS in the subsurface of the mainly graptolites, and the stratigraphic position of the Qusaiba and
eastern area and outcrops at central regions of Saudi Arabia, respec- Sharawra formations below the Devonian-age Tawil Formation, the
tively, have been well documented in the literature (e.g., Mahmoud Qalibah Group is assigned Early Silurian age. The recovery of graptolite
et al., 1992; Aoudeh and Al-Hajri, 1995; and Jones and Stump, 1999). fauna such as Glyptograptus, Lagrarograptus, Monograptus convolutes, and
However, prior to the present study, the QS in the northwest region Coronograptus gregarious zones in outcrops from the Tabuk Basin and
(Tabuk Basin) has not been thoroughly investigated in terms of de- wells from the eastern Saudi Arabia (e.g., Rickards and Koren, 1974;
positional setting. The current work focuses on full sedimentological McClure, 1988) allows the assignment of the QS to the Convolutus zone
and geochemical characterizations of a QS outcrop section in the Tabuk belonging to the Llandovery, Aeronian stage. Chitinozoans of the
Basin, northwestern of Saudi Arabia. Subsequent experimental analyses Middle Aeronian stage were reported in a water well (NWA10) from
provide information on depositional environments, rock mechanical Qasim. Similar occurrences and high diversity of Llandovery chit-
properties, organic richness, kerogen types, and reciprocal impacts in inozoas have been found in cores and ditch cutting samples obtained
order to evaluate the reservoir potential for shale gas. from wells that penetrated both the Qusaiba and Sharawra formations
in central Saudi Arabia (Paris et al., 1995).
2. Geologic setting The organic-rich shales at the base of the QS rarely exceed 50 m in
thickness and are usually about 9–31 m thick. Likewise, the richest
The Phanerozoic rocks of Saudi Arabia have been subjected to ex- organic beds are composed of micro-laminated shale and occur im-
tensive stratigraphic studies to explore their hydrocarbon content, mediately above, and have sharp, well-defined contacts with the basal
ground water and mineral resources as well. (Powers et al., 1966; sandstone of QS. In the most organic-rich shale beds, the terrigenous
Power, 1968; Simmons et al., 2007). Consequently, the classification, sediment is restricted to beds lying immediately above and below the
definition and nomination of the lithostratigraphic units have varied shale. Sandstones interbeds are rare within the most organic-rich source
considerably over the past decades. Recently, Saudi stratigraphic beds (Jones and Stump, 1999). The upper QS is mostly composed of
committee published a generalized column for the Paleozoic strati- shale but thin interbeds of coarse siltstone and very fine-grained
graphic rocks (Halawani et al., 2013) (Fig. 1). sandstone are also present throughout the unit. The interbedding
During the Late Ordovician period, polar glaciers expanded across sandstones are composed of 1.5–3 m thick beds, described as ripple
Gondwana and covered most of Saudi Arabia (Melchin et al., 2013). laminated, moderate to well sorted, fine to very fine-grained beds with
The geologic record shows glacial striations in present-day outcrops, erosional bases and gradational tops (Janjou et al., 1996a,b).
and glacially deposited formations (Vaslet et al., 1993). The advance Gamma-ray log response in the basal “hot shale” portion of the QS
and retreat of the glaciers resulted in the deposition of the Sarah and of the Silurian-age Qalibah Group exceeds 150° API units (Abu-Ali
Zarqa formations (Cole et al., 1994). These two formations represent et al., 1991; Mahmoud et al., 1992; Cole et al., 1994). The thickness of
glacial and periglacial sequences and consist of tillites and proglacia1 the “Qusaiba hot shale” varies from a few tens of meters (m) to 70 m
sands (Zarqa Formation) and finer sands in the Sarah Formation. By (Jones and Stump, 1999). The basal part of the QS “hot shale” having
Early Silurian time, the retreat and melting of the glaciers resulted in an up to 14 wt% of TOC, is the most effective hydrocarbon source-rock
abrupt sea-level rise, which led to an important regional marine facies for the Paleozoic reservoirs in Saudi Arabia (Mahmoud et al.,
transgression and consequently the deposition of the upward-coar- 1992; Cole et al., 1994; Bishop, 1995; Milner, 1998; Abu-Ali et al.,
sening sequence of the Qalibah Group in Saudi Arabia (Vaslet, 1987; 1999; Jones and Stump, 1999; Konert et al., 2001). Based on regional
Husseini, 1991). basin models, Qusaiba hot shale “QHS” cover a wide maturity range
The type locality and type sections of QS are located in the Tabuk (Inan et al., 2016). Across the eastern Arabian subcontinent, the QHS is
Basin, northwestern part of Saudi Arabia (Vaslet et al., 1993; Janjou mostly thermally mature for gas generation; however, the QS is over
et al., 1996a,b). At the type locality in the Qalibah area, the QS is ap- mature for gas generation in the deeper parts of some basins and mature
proximately 290 m thick and it reaches up to 482 m in thickness at its for oil generation along some basin margins (Fox and Ahlbrandt, 2002).
reference section located near Tabuk City “at the northwest of Saudi
Arabia”. Its outcrop near the old Qusaiba village in the Qasim region is 3. Methodology
about 83 m thick.
Stratigraphically, the QS is subdivided into an upper and a lower 3.1. Outcrop methodology
part. The lower part formed by regional marine transgression deposits
while the upper part comprises a coarsening upward prodeltaic shale The 31 m thick studied outcrop exposure is located at the northwest
(Jones and Stump, 1999; Lüning et al., 2000). QS represents the delta- of Tayma City (27° 52\ 10\ N 38° 34\ 35\ E) (Fig. 1A) where it extends
toe clays which dominates the Silurian-Carboniferous depositional laterally as a long ridge for ~3 km to form the NW side of Tayma
setting of the Arabian Plate (Sharland et al., 2001). Graben (Fig. 1D and E). The outcrop was selected based on its acces-
The QS is mainly composed of light to dark grey shales (mudstone) sibility, thickness, and documented lithological variation. A detailed
and thin beds of siltstone and sandstone within shales (Cole et al., field sedimentological description was conducted to document the
1994). The overlying Sharawra Formation is mostly composed of mi- outcrop's lithological variation, beds thicknesses, textural attributes and
caceous sandstone and subordinate siltstone and shale (Stump and Van visible sedimentary structures. A total of 68 samples were collected at a
Der Eem, 1995). On a regional basis, the Qusaiba/Sharawra contact is regular sampling interval of 0.5 m from the basal section up to the top
mostly conformable. However, in the Qasim area, QS is locally eroded of the outcrop (Fig. 1D). The sample set includes 43 samples obtained

242
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

from the basal 20.5 m thick organic-rich shale interval and 25 samples for petrographic analysis and examination of diagenetic features. The
from the overlying 10.5 m thick overlying organic-lean shale and silt- thin sections were examined using a polarizing petrographic micro-
stone-sandstone interval. In order to obtain fresh unweathered samples, scope to identify their grain size, matrix, mineralogy and the dominant
the sampling site surfaces were excavated for about 20–30 cm inward. petrofacies types. Scanning electron microscopic analysis was also
These field data were integrated with sample lithological description, carried out using a JEOL JSM-6610LV to identify the matrix, miner-
petrographic and bulk mineralogical data to construct the stratigraphic alogy, grain to grain relationships and porosity. Energy Dispersive
column of the QS. Spectroscopy (EDS) was used to determine the elemental compositions
of the investigated samples.
3.2. Laboratory investigations All the studied samples were analyzed using X-ray diffraction fa-
cility at the Weatherford Research Center Laboratories to quantitatively
The samples collected during the field investigation were sedi- determine their bulk mineralogy. The X-ray fluorescence (XRF) spec-
mentologically described using a stereoscopic microscope and sub- trometry equipment at Weatherford Research Center Laboratories was
sampled for different laboratory analyses. Using the standard thin also used to determine the major elemental oxides and trace elemental
section preparation technique, twenty-two thin sections were prepared composition of all the 68 samples. Elemental concentrations of the

Fig. 2. Lithofacies description along with the bulk mineralogy and total organic carbon of the studied Qusaiba Shale section, Tabuk Basin, NW Saudi Arabia.

243
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

investigated samples were normalized with Al concentration, because Redox proxies are those elements, which have different chemical
this element is highly immobile and resistance to weathering and di- behavior in oxic and anoxic conditions according to the solubility of
agenetic processes (Brumsack, 1989). This helps to factor out biogenic their oxy anions (Morford et al., 2001), therefore, they may be in-
and authigenic minerals dilution. The normalized values were subse- corporated with the organic matter or adhered to the sulfide minerals
quently compared to their corresponding average shale values to de- (Lewan and Maynard, 1982; Morford and Emerson, 1999). The most
termine their enrichment factors (EF) using the following equation common trace elements that are used as redox proxies include V, Cr, Ni,
(Wedepohl, 1971; Taylor and McLennan, 1985; McLennan, 2001; U, Mn, Fe, and Mo (Dean et al., 1997; Algeo and Maynard, 2004;
Tribovillard et al., 2006). Tribovillard et al., 2012).
Sedimentary molybdenum [Mo]s to the total organic carbon cov-
EF = (Element/Al) sample/(Element/Al) Average Shale (1) ariation [Mo]s/TOC developed by Algeo and Lyons (2006) to in-
vestigate the variations of water mass restriction in anoxic silled marine
This ratio has been used in several studies (e.g. Arthur et al., 1990; systems was additionally used to assess the hydrographic variations in
Calvert and Pedersen, 1993; Morford et al., 2001; Algeo and Maynard, the water mass during the deposition of the investigated outcrop sec-
2004; Goldberg et al., 2007). In the present study, the enrichment factor tion.
(EF) for each element is presented in Table 3.

Fig. 3. Laminated organic-rich shale lithofacies


(A) Field photo shows the highly fissile and thinly
lamination style of the lithofacies (< 1 cm). (B) Thin
section photo (PPL) shows the lamination of silty-
size (bright) and clay-size shale (dark) (C) thin sec-
tion photo (CN) shows the disseminated organic
matter in the clay groundmass with few scattered
silty-size quartz grains. (D) SEM image of euhedral
pyrite crystals (white arrows) within clay flakes (red
arrows). Graptolitic concretion lithofacies (E)
Field photo shows the relatively hard denser lenti-
cular concretion, note the bending shaley laminae
below and above the concretion. (F) Photo shows
different fracture types inside the concretion. The
black color is due to the organic richness. (G) Photo
shows graptolites species preserved between the
bedding planes inside the concretion. Graptolite
branches (yellow arrows) (H) thin section photo
(PPL) shows the branched longitudinal calcareous
graptolite (white arrows) preserved in the organic-
rich clay ground mass (blue arrows). (For inter-
pretation of the references to color in this figure le-
gend, the reader is referred to the Web version of this
article.)

244
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

Total Organic Carbon (TOC) contents of the samples were measured few samples. Micro fractures are very common in this interval. XRD and
with the LECO carbon analyzer equipped with a HF-100 Induction SEM analyses indicate that this facies is dominated by clay minerals
Furnace on carbonate-free shale. Standard Rock-Eval (kerogen pyr- (range: 39–71%; average: 56%), the facies is also characterized by a
olysis) analysis was performed according to the method of Espitalie considerable content of quartz (23% average) and pyrite (7% average)
et al. (1977) using a Delsi Rock-Eval OSA on a 100 mg aliquot. The (Table 1; Fig. 3D). Feldspars and carbonate minerals (calcite and do-
Rock-Eval Pyrolysis analysis provided the main organic geochemical lomite), respectively, occur as minor and trace components within this
parameters (S1, S2, S3, Tmax) that are used to interpret the hydrocarbon facies.
potential of the QS (Table 5). Other geochemical parameters were
calculated, plotted and interpreted using the following formulae; [(S2/
TOC)*100] for hydrogen index (HI), [(S3/TOC)*100] for oxygen index 4.1.2. Graptolitic concretion lithofacies (10.5–11.5 m)
(OI), [S1/(S1+S2)] for production index (PI) and [(0.0180 x Tmax - This 1-m thick interval consists of graptolite-bearing concretions,
7.16) (Jarvie et al., 2001)] for the percentage of vitrinite reflectance which occur as discontinuous lenses within silty, clayey organic-rich
index (VRo) (Table 4). shale beds (Fig. 3E). It represents a good marker between the basal
laminated organic-rich shale interval and the overlying massive shale
4. Results lithofacies unit of QS in the outcrop. The concretions are mainly com-
posed of calcite (68%) with minor amounts (< 10%) of quartz, clay
4.1. Lithostratigraphy minerals and pyrite (Table 1). Graptolites are abundant and well pre-
served in the cores of the concretions. They are mainly monograptids,
Thirty-one m-thick sequence described at the outcrop is subdivided with minor occurrences of diplograptids (Fig. 3G). The shale beds above
into two broad units based on the color and sand content (Fig. 1.E). The and below the concretion interval exhibit soft sediment deformation
basal 20 m thick unit of the outcrop consists of a dark grey to black, structures such as micro-fold and budge-out structures (Fig. 3E). The
hard, fissile and thinly laminated shale. This lower unit is subdivided concretion interval displays multiple fracture sets ranging between
into two units separated by a graptolite-bearing concretion interval. bedding-parallel, vertical and inclined fractures (Fig. 3F–H) which are
The overlying top 10–12 m thick unit of the studied outcrop is com- filled mainly by calcite and pyrite. The concretions are likely of sep-
posed of laminated, organic-lean, variegated shale intercalated with tarian concretion type (Boles et al., 1985; Hudson et al., 2001). Thin
siltstone to fine-grained sandstone beds (Fig. 2). The integration of section petrography of the concretions shows clay matrix with dis-
petrographic analysis and field description allowed the subdivision of seminated silt-size quartz grains. Branched calcareous graptolitic net-
the two broad units of QS into six different lithofacies types, which are works, which cut-across the samples, are occasionally filled with or-
briefly characterized as follows: ganic matter (Fig. 3H).

4.1.1. Laminated organic-rich shale lithofacies (0–10.5 m)


The lowermost 10.5 m thick of the investigated outcrop consists of 4.1.3. Massive organic-rich shale lithofacies (11.5–20.5 m)
green, dark grey, black, laminated, organic-rich, silty and clayey shale This massive lithofacies, overlies the graptolitic concretion interval
(Fig. 3A). The shale, which is generally fissile at the base, becomes and it consists of massive to moderately lithified, dark grey to black
highly indurated towards the top of this lithofacies. Thickness of the shale with some nodular intraformational mud clasts. This interval is
laminae exhibits an upward thinning trend without evidence of bio- highly dissected by natural fracture network mainly filled with gypsum
turbation. This part of the outcrop is intensively fractured with gypsum (Fig. 4A). Cross lamination and soft sediment deformation of laminae
filling mineralization. The occurrence of gypsum decreases toward the are observed immediately above the concretion-bearing interval. Micro-
top of this interval. cross lamination is visible in some samples (Fig. 4B). In thin sections,
Petrographic examination of representative samples from this li- fine silt size quartz particles are disseminated in the organic matrix
thofacies indicates that it is composed of organic-rich silty claystone to where traces of graptolites are also observed (Fig. 4C). Clay content
siltstone (Fig. 3B). The organic rich sediments are characterized by (29–71% range and average of 56%) is highly enriched in this litho-
alternation between dark organic rich and light silt-rich laminae facies. In contrast, the quartz content is low in the facies (17–40% range
(Fig. 3B). Silt-size quartz grains are sparsely disseminated within the with average of 26%). Pyrite content (5% average) is less than in the
organic rich matrix in some of the samples (Fig. 3C). Graptolite net- basal interval. Both dolomite and siderite occur as minor components
works and organic matter debris (palyno-macerals) are common in a within this interval while calcite is only present in one sample (Fig. 4D).

Table 1
XRD Bulk mineralogy of the identified lithofacies in Qusaiba Shale and their equivalent brittlenes. N: number of samples in lithofacies; Car. Carbonate minerals; Cla.
Clay minerals.

Lithofacies (n) Total Calcite Dolomite Quartz K-spar Plag. Pyrite Gypsum Q+F Car. Others Cla. Brittleness
Clays
Cross-stratified siltstone/sandstone lithofacies Maxi. 53.00 1.00 14.00 56.00 11.00 10.00 2.00 17.00 77.00 14.00 18.00 53.00 86.68
(n = 4) Min. 8.00 0.00 1.00 18.00 5.00 3.00 1.00 0.00 26.00 1.00 1.00 8.00 24.07
Aver. 35.33 0.33 6.67 35.00 8.33 6.50 1.33 6.33 49.83 6.83 7.50 35.33 49.77
Siltstone-shale Interbedded lithofacies (n = 3) Maxi. 52.00 0.00 2.00 52.00 19.00 10.00 3.00 0.00 81.00 2.00 3.00 52.00 75.00
Min. 16.00 0.00 1.00 28.00 11.00 5.00 1.00 0.00 44.00 1.00 0.00 16.00 33.37
Aver. 31.80 0.00 1.60 42.20 15.20 7.60 2.00 0.00 65.00 1.60 1.40 31.80 56.72
Organic-lean shale lithofacies (n = 18) Maxi. 79.00 1.00 49.00 59.00 23.00 13.00 4.00 6.00 85.00 49.00 8.00 79.00 79.37
Min. 13.00 0.00 0.00 13.00 5.00 1.00 1.00 0.00 19.00 0.00 1.00 13.00 13.16
Aver. 45.00 0.25 6.05 32.50 9.95 6.70 2.15 0.60 48.65 6.25 2.65 45.00 42.69
Massive organic-rich shale lithofacies Maxi. 71.00 1.00 3.00 40.00 16.00 9.00 12.00 3.00 65.00 3.00 12.00 71.00 57.33
(n = 19) Min. 29.00 0.00 0.00 17.00 3.00 2.00 1.00 0.00 22.00 0.00 1.00 29.00 16.05
Aver. 56.00 0.14 1.38 25.71 7.52 4.19 4.81 0.29 37.43 1.48 4.95 56.00 29.25
Graptolitic concretion facies (n = 1) 9.00 68.00 2.00 9.00 2.00 2.00 8.00 0.00 13.00 70.00 8.00 9.00 79.32
Laminated organic-rich shale lithofacies Maxi. 71.00 3.00 3.00 32.00 11.00 8.00 15.00 0.00 49.00 5.00 15.00 71.00 40.82
(n = 23) Min. 39.00 0.00 0.00 17.00 4.00 2.00 0.00 0.00 25.00 0.00 0.00 39.00 18.41
Aver. 56.08 1.04 1.04 23.48 7.12 3.92 7.28 0.00 34.52 2.04 7.28 56.08 27.21

245
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

Fig. 4. Massive organic-rich lithofacies (A) Field


photo shows dark grey to black shale with nodular
structures (red arrow), gypsum-filled natural fracture
network (white arrows). (B) Thin section photo (PPL)
shows traces of longitudinal graptolites in clay
groundmass and the lamination pattern between
clay-dominated and silt dominated with high occur-
rence of organic traces in the clay dominated la-
minae. (C) Thin section photo (CN) shows a bran-
ched longitudinal graptolite (white arrows)
preserved in the clay (Cl). (D) SEM image in grap-
tolite-bearing thin section shows the calcite crystal
[C] filling the graptolite branches. Organic-lean
shale lithofacies (E) Field photo shows the colorful
purple and brownish shale with low-angle natural
fractures filled with gypsum (blue arrows). Note the
sharp boundary with the overlying silt dominated
lithofacies (dotted black line). (F) Thin section photo
(CN) shows the angular to sub-angular, well sorted
silt-size quartz grains. (G) Thin section photo-
micrograph (PPL) shows the clay-rich (dark brown)
and silt-rich (light yellow) lamination structure. Note
the horizontally aligned mica flakes (white arrows).
(H) SEM image showing calcite (c), dolomite (d),
quartz (q) and muscovite (m) as the main mineral
constituents of this lithofacies. (For interpretation of
the references to color in this figure legend, the
reader is referred to the Web version of this article.)

4.1.4. Organic-lean shale lithofacies (20.5–28 m) more frequently than in the underlying lithofacies. This ~1 m thick
This lithofacies is very distinctive by its colorful olive green, purple, interval shows three fining upward cycles (Fig. 5A). The siltstone beds
brownish shale laminae that are intercalated with thin siltstone to very occur at the base of each of the cycles and grade into greyish shale beds
fine sandstone beds (Fig. 4E). The shale laminae are generally fissile but at the top (Fig. 5B).
occasionally become highly indurated at some intervals. The silt-size The sediments are composed of angular to sub-angular, well sorted
quartz grains content of the samples within this facies, are sub-angular silt to fine sand size quartz grains with substantial amounts of musco-
to angular, and well sorted (Fig. 4F). Bands of clay-dominated and silt- vite grains that are dominantly aligned along the long axes of the mi-
dominated layers are also present (Fig. 4G). Mica grains are aligned neral grains (Fig. 5C–D). Interparticle porosity accounts for an average
along their long axes (Fig. 4G). Clay minerals (average 45%) and quartz of 6% while about 2% fracture porosity also exists. XRD analyses show
(average 33%) are the dominant minerals while K-feldspars and pla- almost equal proportion of quartz (38% average) and clay minerals
gioclase (< 10%) occur in minor amounts (Fig. 4H and Table 1). (34% average). The feldspar content in the facies also accounts for an
average 19% of the bulk mineralogical composition in this lithofacies.
4.1.5. Siltstone-shale interbedded lithofacies (28–29.05 m)
The sequence changes to laminated grey siltstone/sandstone and 4.1.6. Cross-stratified siltstone/sandstone lithofacies (29.05–31 m)
shale intercalation. The siltstone/sandstone beds are thicker and occur The topmost part of the studied section mainly consists of

246
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

Fig. 5. Silt shale interbedded lithofacies (A) Field


photo shows the hard resistant silt to fine-grain
sandstone layer and the soft eroded shale layers in
between (blue arrows). (B) Thin section photo (PPL)
shows the boundary between clay-rich (red arrows)
and silt-rich laminae (blue arrows). (C) Thin section
photo (CN) shows the silt to fine sand fraction of this
lithofacies. Note mica flakes characterized by high
interference color (red arrows). (D) Thin section
photo (CN) shows the silt fraction of this lithofacieas.
Cross-stratified siltstone/sandstone lithofacies
(E) Field photo shows the whole unit, note the up-
ward thickening of sandstone layers. (F) Field photo
shows the sigmoidal S and hummocky H cross stra-
tification in this unit. (G) Field photo shows trough
cross stratification (white arrows). Scale is the pen in
the middle of the photo above thin diagenetic
gypsum layer (Red arrows). (H) Thin section photo
(CN) shows the sub-angular to angular well sorted
quartz grains. Note the porosity with light blue color.
(For interpretation of the references to color in this
figure legend, the reader is referred to the Web ver-
sion of this article.)

intercalation of siltstone and sandstone beds (Fig. 5E). The facies 4.2. Geochemistry
characterized by thick sigmoidal to trough-cross stratification, fining
upwards sediments at the base and hangs to hummocky-cross stratifi- 4.2.1. Inorganic geochemistry
cation at the top (Fig. 5F–G). Additionally, a very thin gypsum layer 4.2.1.1. Major elements and oxides. The major element oxides in the QS
occurs in the facies (Fig. 5G). show uniform distribution in the lower 25 m of the outcrop (Table 2).
Petrographic analysis indicates that the detrital grains in this li- Some of the elements (Si, Al, Ti, Na and P) exhibit peak excursions
thofacies are predominantly sub-angular to angular, well-sorted, silt towards the top of the organic-lean shale through the cross-bedded
and very fine sands (Fig. 5H). Hydrocarbons occur as infilling in the siltstone-sandstone lithofacies. XRF analysis reflects the dominance of
interparticle pore spaces and as patches on grain surfaces. Interparticle SiO2 (average: 43%), which is consistent with the abundant quartz
porosity is the dominant porosity type in the facies and it ranges from content identified by the XRD and SEM–EDX analyses (Fig. 3D). Al2O3
~6% in the siltstone sample to 10% in the fine sandstone sample. By (average 20.86%) is the second most abundant major element oxide in
average, quartz and total clay equally exist in this lithofacies (35% for the studied section. The Fe2O3 concentrations vary from 2.6 to 12.2%
each) while dolomite, k-feldspar and plagioclase are less dominant (average: 8.43%) and those of K2O range from 1.13 to 3.44% (average:
(~ < 8%) (Table 1). 2.56%). The potassium enrichment likely corresponds to the
occurrences of K-feldspar, clay mineral and mica. Sulfur is highly

247
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

concentrated in the lower organic-rich shale interval, likely due to the


S (%)

3.63
0.06
1.50
0.31
0.07
0.19
3.95
0.24
1.24
4.02
1.31
2.58
1.36
5.51
abundance of pyrite. Other major elements (Mg, Na, Ti, Ca, and Mn)
exist in low concentrations close to or below 1.0% (Table 2).
Ti (%)

0.92
0.39
0.73
0.97
0.60
0.78
0.98
0.43
0.78
0.82
0.64
0.75
0.35
0.83
4.2.1.2. Trace elements. The distribution patterns of Cu, Mo, As, Ni, Co,
U and V are similar to that of the TOC, particularly, in the organic-rich
K (%)

2.81
2.07
2.38
3.26
2.64
2.99
3.34
1.64
2.59
2.86
1.97
2.22
0.94
2.33
intervals of the investigated outcrop section. Similar to TOC, the trace
Maximum, Minimum and average values of the major element oxides of the identified lithofacies in Qusaiba Shale and their equivalent brittlenes. N is the number of samples in each lithofacies.

element distribution patterns exhibit positive excursions in the


Ca (%)

10.42

23.45
lowermost 20 m thick of organic-rich, dark-grey shale interval and
4.43
0.34
2.42
0.25
0.10
0.19

0.06
1.45
1.77
0.22
0.51

0.51
display abrupt decrease at the contact with overlying organic-lean shale
Mg (%)

(Fig. 6A). Other elements including the lithogenic related Ba, Rb, Th,
2.56
1.82
2.14
1.17
0.89
1.05
5.71
0.75
1.73
2.02
0.75
1.16
0.81
2.12
Ga, Nb, and Y are characterized by more or less uniform distribution
patterns in the organic-rich shale interval, and exhibit an upward
Fe (%)

6.60 increasing trend at the contact with the overlying organic-lean shale
7.07
2.69
5.46
4.18
2.94
3.66
8.40
2.68
5.52
8.54
3.46
6.51
4.24

facies (Table 3). Except for occasional peak excursions, Sr, Zr and Pb are
uniformly distributed throughout the outcrop section.
Al (%)

12.63

10.08
13.45
12.37
12.94
16.94

11.68
12.14

10.99

11.74
6.01

5.94

9.49

4.55

The enrichment factor (EF) values for V, Cu, As, Mo, Ni, and U are
characterized by similar distribution patterns as their measured con-
Si (%)

24.94
16.47
20.70
28.11
22.62
25.18
29.49
12.04
21.40
22.64
15.54
19.20

23.81

centrations by exhibiting significant depletion at the contact between


8.70

the organic-rich shale and the overlying organic-lean shale (Fig. 6B). In
SO3 (%)

the basal 20.5 m of the outcrop, represented by the organic-rich lami-


10.04

13.77
3.40
9.06
0.16
3.75
0.77
0.17
0.46
9.86
0.61
3.09

3.28
6.45

nated and massive shale lithofacies, the average values are in the excess
of 1 suggesting that they are enriched above their average shale values
K2O (%)

(Table 4). MoEF and UEF in particular, are in the excess of 3 in most of
3.38
2.50
2.87
3.93
3.18
3.60
4.03
1.97
3.11
3.44
2.37
2.67
1.13
2.80

the samples within this basal section and MoEF is occasionally in excess
of 10 at few intervals (Fig. 6B). Except for AsEF and MoEF which are
Na2O (%)

consistently above 1 throughout the outcrop section, the EF's values for
most of the aforementioned elements are highly depleted in the over-
1.25
0.36
0.70
1.47
0.64
0.99
1.88
0.37
0.76
0.68
0.29
0.40
0.06
0.53

lying organic-lean shale through the siltstone/sandstone lithofacies.


Only As, Zn, and Mo are characterized by average values that are higher
CaO (%)

14.58

32.80

than their shale values in these lithofacies. The lithogenic related ele-
6.20
0.47
3.38
0.35
0.13
0.26

0.09
2.03
2.48
0.31
0.71

0.72

ments Rb, Sr, Y, Zr, and Nb are poorly enriched in the outcrop section.
Except for the associated subtle excursions at the concretion interval
MgO (%)

and in the uppermost siltstone-sandstone lithofacies, their EF's are


4.25
3.02
3.54
1.94
1.48
1.74
9.47
1.25
2.87
3.34
1.24
1.92
1.34
3.51

generally, uniformly distributed in the outcrop section (Table 4). GaEF is


enriched with values higher than 1 in almost all the lithofacies (except
Fe2O3 (%)

in the concretion facies). ThEF characterized by low enrichment in the


10.11

12.01

12.21
3.85
7.81
5.97
4.20
5.23

3.83
7.90

4.95
9.31
6.06
9.43

lower 20.5 m of the outcrop, exhibits increased enrichment trend at the


contact between the organic-rich massive shale and the organic-lean
Al2O3 (%)

shale and continues through the upper section of the outcrop (Table 4).
23.86
11.35
19.04
25.42
23.38
24.46
32.01
11.23
22.06
22.94
17.94
20.77

22.18
8.59

4.2.1.3. Redox indices. In addition to the use of the redox-sensitive


elements, redox indices such as [Mo]s/TOC, Ni/Co, Th/U, V/Cr, V/Mo,
TiO2 (%)

V/(V + Cr) and V/(V + Ni) were also determined and used in this
1.53
0.65
1.22
1.62
1.00
1.30
1.64
0.72
1.30
1.36
1.07
1.25
0.58
1.39

study. Out of all the indices, V/(V + Cr), V/(V + Ni), Th/U and [Mo]s/
TOC were found to best distinguish between the organic-rich and the
SiO2 (%)

overlying organic-lean intervals (Table 4, Fig. 6B). The values of the V/


53.36
35.23
44.28
60.13
48.38
53.87
63.09
25.75
45.77
48.43
33.25
41.08
18.61
50.94

(V + Ni) ratio range from 0.54 to 0.79 (average: 0.69) in the organic-
rich shale lithofacies, 0.49–0.81 (average: 0.70) in the organic-lean
Maxi.
Maxi.

Maxi.

Maxi.

Maxi.
Aver.

Aver.

Aver.

Aver.
Min.

Min.

Min.

Min.

shale lithofacies and 0.65–0.82 (mean 0.72) in siltstone-shale and


siltstone-sandstone lithofacies. V/(V + Cr) ranges from 0.6 to 0.7
Cross-stratified Siltstone/Sandstone lithofacies (n = 4)

(averages 0.6) in the organic-rich shale interval while its values are
Laminated organic-rich shale lithofacies (n = 23)

between 0.3 and 0.6 (average 0.5) in the overlying organic-lean shale,
Massive organic-rich shale lithofacies (n = 19)
Siltstone-Shale Interbedded lithofacies (n = 3)

siltstone-shale and siltstone-sandstone lithofacies. Th/U values range


from 0.38 to 1.6 (average 0.97) in the laminated interval and from 1.21
Graptolitic concretion lithofacies (n = 1)
Organic-lean shale lithofacies (n = 18)

to 2.23 (average: 1.57) in the massive organic-rich shale lithofacies.


The only Th/U values above 2 recorded in the massive organic-rich
shale lithofacies were in the topmost samples which are likely a
transition from the organic-rich shale to the organic-lean shale
intervals. Additionally, Th/U values in the organic-lean shale (range:
1.77–13, average: 4.06) and the siltstone/sandstone lithofacies (range:
2.21–16, average: 6.14) are characterized by higher values than the
Lithofacies (n)

organic-rich shale facies.


The basal organic-rich shale section of the outcrop is characterized
Table 2

by low and uniformly distributed [Mo]s/TOC values (Table 4). With the
exception of the samples near the contact between the organic-rich and

248
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

Fig. 6. A) The stratigraphic distribution of the Al-normalized trace elements correlated with the TOC. B) The stratigraphic distribution of the enrichment factor EFs
values of trace elements and redox indices ratios.

organic-lean shales, the values of this ratio range between 3.5 and 9.54 (Table 4). The basal laminated organic-rich shale and massive organic-
in the organic-rich shale interval. However, it has an average of 17.4 in rich shale lithofacies are characterized by Hydrogen Index (HI) in the
the organic-lean shale and 25.5 in the shale/siltstone and siltstone/ ranging between 4 and 219 mg HC/g TOC with an average of 70 mg
sandstone facies, respectively. The stratigraphic plot of the [Mo]s/TOC HC/g TOC. The calculated VRo values for the investigated samples
ratio shows uniform distribution pattern from the base down to 19.5 m. range from 2.77 to 0.28 (average 0.58) (Table 4).
This trend was punctuated by minor excursions at 19.5–21 m and
22.5–23.5 m. An upward increase characterized the distribution of this
ratio from 23.5 through the top of the outcrop section (Fig. 6B). 5. Discussion

QS shows significant vertical variability in texture, composition,


4.2.2. Organic geochemistry
primary sedimentary structure and organic content due to the interplay
The stratigraphic distribution trend of the TOC is characterized by a
effects of sedimentation processes (e.g., supply, transport, and deposi-
sharp up-section decrease from the basal black to dark grey laminated
tion), water chemistry (e.g., oxic, anoxic, and euxinic), water energy
organic-rich shale to the topmost cross-stratified siltstone-sandstone
(e.g., high, or low) in a wide range of depositional environments (Lazar
lithofacies (Fig. 2). Generally, TOC in the dark-grey laminated and
et al., 2015). In order to develop an appropriate depositional model, the
massive organic-rich shale lithofacies ranges between 0.5 and 6.1 wt%
field observations, lithofacies, petrographic, bulk mineralogical and
(average 2.52 wt%) then it decreases (range: 0.04–0.35 wt%, average:
geochemical data were integrated (Fig. 6). Geochemical data is of cri-
0.18 wt%) in the overlying organic-lean shale and siltstone interval
tical importance in understanding the key factors of organic richness
(Table 5). Among the identified lithofacies, the lower three lithofacies
and mineralogical heterogeneity (Tribovillard et al., 2001, 2006; 2008,
only provide reliable Rock-Eval Pyrolysis results. The values of the re-
2012; Wignall and Newton, 2001; Piper and Perkins, 2004; Rimmer,
maining hydrocarbon potential (S1) vary between 0.00 and 0.91 mg
2004; Meyers et al., 2006; Algeo and Tribovillard, 2009; Scott and
HC/g rock (average: 0.11 mg HC/g rock). Similarly, the values of S2
Lyons, 2012).
range from 0.06 to 9.78 mg HC/g rock (average of 2.13 mg HC/g rock)

249
Table 3
Trace Elements concentrations (ppm) and their enrichment values in the identified lithofacies of Qusaiba Shale and redox ratios used for paleooceanography interpretation. N: number of samples in each lithofacies.

Lithofacies (n) Total V ppm VEF Crppm CrEF Coppm CoEF Nippm NiEF Cuppm CuEF Znppm ZnEF Asppm
Clays (%)
M. Abouelresh, et al.

Cross-stratified Maxi. 53.00 163.40 0.95 110.20 0.91 24.00 0.91 69.30 0.83 32.80 0.53 265.70 2.02 7.10
siltstone/ Min. 8.00 18.10 0.21 34.10 0.75 9.60 0.75 9.60 0.21 6.70 0.22 25.60 0.40 0.60
sandstone Aver. 35.33 110.58 0.67 81.95 0.82 17.78 0.82 47.00 0.57 22.13 0.40 159.82 1.30 4.37
lithofacies
(n = 4)
Siltstone-shale Maxi. 52.00 152.70 0.78 117.50 0.96 25.50 0.96 36.50 0.39 28.30 0.42 73.60 0.56 6.50
Interbedded Min. 16.00 49.00 0.26 55.60 0.47 13.60 0.47 24.80 0.25 13.50 0.20 49.80 0.36 2.70
lithofacies Aver. 31.80 99.04 0.51 85.18 0.75 20.60 0.75 31.20 0.32 20.54 0.31 63.72 0.47 4.36
(n = 3)
Organic-lean shale Maxi. 79.00 169.60 0.90 133.40 0.96 24.70 1.59 72.60 0.91 46.80 0.75 274.50 4.32 8.40
lithofacies Min. 13.00 23.10 0.14 37.40 0.32 7.80 0.22 23.80 0.26 15.10 0.20 67.00 0.37 1.70
(n = 18) Aver. 45.00 118.51 0.69 96.35 0.79 16.10 0.70 49.07 0.56 31.69 0.54 126.68 1.21 5.53
Massive organic- Maxi. 71.00 191.40 1.15 102.50 0.93 31.70 1.27 120.30 1.35 108.50 1.87 166.80 1.63 31.50
rich shale Min. 29.00 120.40 0.68 81.60 0.76 11.50 0.50 43.30 0.47 24.40 0.40 82.90 0.72 4.60
lithofacies Aver. 56.00 168.39 1.06 95.06 0.86 19.00 0.79 75.79 0.89 82.26 1.51 117.68 1.00 22.37
(n = 19)
Graptolitic 9.00 70.50 1.06 23.30 0.51 1.50 0.15 20.50 0.59 20.50 0.89 28.20 0.58 9.60
concretion
facies (n = 1)
Laminated Maxi. 71.00 213.60 1.40 108.40 0.92 47.60 2.14 123.30 2.08 180.60 3.43 206.00 2.36 36.40
organic-rich Min. 39.00 143.90 0.93 62.80 0.70 17.20 0.70 53.60 0.63 57.10 0.98 77.00 0.65 11.60
shale Aver. 56.17 171.07 1.10 88.41 0.81 26.05 1.17 79.70 1.01 92.90 1.79 133.44 1.20 22.10
lithofacies
(n = 23)

250
Lithofacies (n) AsEF Brppm BrEF Rbppm RbEF Srppm SrEF Yppm YEF Zrppm ZrEF Nbppm NbEF Moppm MoEF

Cross-stratified 3.02 0.30 0.09 158.30 0.80 157.20 0.66 43.10 1.39 170.80 1.58 16.70 0.75 3.90 2.22
siltstone/ 0.83 0.10 0.02 57.90 0.61 90.80 0.21 24.40 0.46 111.20 0.57 9.10 0.54 1.95 1.73
sandstone 2.21 0.22 0.05 116.95 0.72 123.37 0.41 33.28 0.82 142.92 0.93 13.27 0.65 2.98 1.99
lithofacies
(n = 4)
Siltstone-shale 2.59 0.70 0.12 150.10 0.71 128.10 0.29 31.40 0.54 183.10 0.82 21.50 0.79 3.10 1.71
Interbedded 1.24 0.25 0.04 111.20 0.54 108.20 0.24 26.50 0.43 143.00 0.59 14.60 0.55 2.20 1.12
lithofacies 1.92 0.43 0.07 130.78 0.63 119.10 0.27 29.32 0.49 167.00 0.72 17.46 0.66 2.72 1.45
(n = 3)
Organic-lean shale 2.71 3.80 0.68 154.40 0.94 217.10 0.54 75.50 2.01 227.10 1.10 23.70 0.92 6.10 6.56
lithofacies 1.06 0.05 0.01 87.70 0.52 65.50 0.16 22.90 0.29 70.20 0.42 7.80 0.54 0.70 0.38
(n = 18) 2.20 0.72 0.13 133.05 0.74 108.57 0.29 38.18 0.79 135.01 0.66 15.29 0.66 3.05 2.11
Massive organic- 2.83 1.60 0.33 160.20 1.07 150.60 0.41 41.00 0.79 1056.00 5.04 20.30 0.83 21.20 15.27
rich shale 2.10 0.30 0.05 126.90 0.72 63.70 0.17 27.40 0.54 88.30 0.44 0.00 0.00 7.00 0.15
lithofacies 2.56 0.67 0.13 140.36 0.81 84.67 0.22 32.22 0.63 192.26 0.76 11.65 0.53 11.70 7.18
(n = 19)
Graptolitic 1.45 0.30 0.15 47.20 0.66 408.10 2.66 33.40 1.59 96.30 1.18 7.00 0.76 10.40 15.64
concretion
facies (n = 1)
Laminated 2.75 3.30 0.70 139.00 0.84 109.40 0.29 58.10 1.54 150.00 0.82 21.60 1.02 31.80 22.31
organic-rich 2.27 0.30 0.06 93.40 0.60 64.30 0.18 25.00 0.49 92.60 0.44 11.60 0.52 6.50 2.97
shale 2.50 1.10 0.24 124.67 0.74 78.27 0.22 35.95 0.75 114.54 0.60 14.17 0.66 16.28 10.56
lithofacies
(n = 23)
Marine and Petroleum Geology 111 (2020) 240–260

(continued on next page)


M. Abouelresh, et al.

Table 3 (continued)

Lithofacies (n) Bappm BaEF Hfppm HFEF Pbppm PbEF Thppm ThEF Uppm UEF Th/U Mo/U MoEf/UEF [Mo]s/ V/ (U+V)/
TOC (V + Ni) (U + V-
+ Mo)

Cross-stratified 606.20 1.38 4.60 1.16 22.30 0.85 13.60 0.90 4.10 0.98 16.00 0.83 1.92 44.02 0.73 0.98
siltstone/ 384.90 0.48 1.90 0.70 10.80 0.59 3.20 0.39 0.20 0.10 3.24 4.07 9.39 14.68 0.65 0.90
sandstone 504.42 0.81 3.10 0.92 17.23 0.74 10.02 0.70 2.43 0.62 7.97 0.78 1.80 30.09 0.69 0.95
lithofacies
(n = 4)
Siltstone-shale 1054.00 1.10 7.60 1.95 24.80 0.82 14.00 0.77 4.40 0.97 3.18 0.57 1.32 18.34 0.82 0.99
Interbedded 540.70 0.55 4.40 1.04 18.30 0.62 6.40 0.36 2.90 0.66 2.21 1.03 2.39 14.71 0.66 0.95
lithofacies 820.22 0.86 5.78 1.43 21.26 0.72 10.24 0.58 3.60 0.81 2.77 0.50 1.15 16.43 0.74 0.97
(n = 3)
Organic-lean shale 1138.00 1.35 10.90 2.78 32.60 1.27 16.20 0.95 6.40 1.73 13.00 14.25 32.88 32.62 0.81 1.00
lithofacies 333.60 0.41 1.00 0.28 8.70 0.47 3.90 0.25 0.40 0.20 1.77 0.17 0.38 3.45 0.49 0.88
(n = 18) 604.63 0.73 3.63 0.97 20.79 0.82 11.46 0.72 3.81 0.96 4.06 1.81 4.17 17.38 0.70 0.97
Massive organic- 680.70 0.77 8.10 2.30 39.40 1.84 14.30 0.98 11.20 2.85 2.23 3.79 8.74 16.56 0.74 1.00
rich shale 327.80 0.40 1.00 0.28 18.30 0.67 9.10 0.56 4.50 1.10 1.21 0.02 0.05 3.88 0.61 0.88

251
lithofacies 422.48 0.52 4.37 1.25 23.61 0.94 11.49 0.77 7.75 2.09 1.54 1.51 3.48 7.45 0.69 0.94
(n = 19)
Graptolitic 315.60 0.95 1.00 0.70 7.40 0.72 4.20 0.68 0.20 0.13 21.00 52.00 120.00 9.54 0.77 0.87
concretion
facies (n = 1)
Laminated 826.10 1.05 7.90 2.37 21.70 0.98 12.60 0.99 33.20 9.45 1.58 2.72 6.28 9.47 0.79 0.97
organic-rich 345.50 0.41 1.00 0.27 12.80 0.55 9.80 0.72 7.40 1.87 0.38 0.47 1.09 3.49 0.54 0.88
shale 412.09 0.53 4.99 1.47 19.14 0.80 11.38 0.80 13.41 3.90 0.97 1.28 2.95 5.35 0.68 0.92
lithofacies
(n = 23)
Marine and Petroleum Geology 111 (2020) 240–260
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

Table 4
Maximum, Minimum and average values of the TOC and Rock Eval Pyrolysis results of the organic-rich lithofacies in Qusaiba Shale; S1: amount of free hydrocarbons
in sample (mg/g), S2 = amount of hydrocarbons generated through thermal cracking (mg/g), S3: amount of CO2 (mg of CO2/g of rock), Tmax: the temperature of
maximum pyrolytic degradation of kerogen., HI: calculated hydrogen index, OI: calculated oxygen index, PI: the production index.

Eleva. (m) Sample no. Lithofacies (n) TOC S1 S2 S3 Tmax HI OI Calc. %VRo PI

11.5–20.5 m 19 Maxi. 3.09 0.18 2.85 3.22 552 104 274 2.77 0.09
Min. 1.14 0.00 0.06 1.66 420 4 84 0.40 0.00
Aver. 1.87 0.02 0.87 2.60 434 44 146 0.73 0.01
10.5–11.5 m 1 Graptolitic concretion facies 1.09 0.07 2.39 0.88 428 219 81 0.54 0.03
0–10.5 m 23 Maxi. 6.10 0.91 9.78 4.48 430 171 150 0.73 0.09
Min. 1.77 0.00 0.36 1.79 414 18 57 0.28 0.00
Aver. 3.17 0.15 2.87 2.83 422 80 95 0.44 0.04

5.1. Paleoenvironmental marine conditions above 2 in the lower part (0.97 for the laminated and 1.57 for massive
organic-rich shale lithofacies, respectively). In particular, Th/U vs V/
The distribution patterns of the redox sensitive elements (e.g. Mo, U, (V + Ni) best discriminates these two major shales at Th/U cutoff value
V, Cu, As, and Ni) distinctively differentiate the organic-rich dark grey 2 (Fig. 7 B). These average values indicate deposition in a low oxygen
shale lithofacies in the lower part of the studied QS outcrop section environment, likely in anoxic conditions (Wignall and Twitchett, 1996;
from the overlying organic-lean shale, siltstone and sandstone litho- Eltom et al., 2016; Baioumy and Lehmann, 2017). The values are much
facies in the upper part (Fig. 6A). Similar to the stratigraphic dis- higher in the overlying organic-lean shale (average: 4.06) and in the
tribution of TOC, the redox sensitive elements exhibit a dramatic de- siltstone-shale and siltstone/sandstone lithofacies (6.14), indicating
pletion at the contact between the organic-rich and the organic-lean that the bottom water conditions changed from anoxic to oxic through
shales (Fig. 6B). time.
In the organic-rich shale lithofacies, the enrichment factors (EFs) of The relationships between TOC and metal/Al ratios and EFs of the
V, CU, Mo, Ni, As and U, are above 1 (Table 4 and Fig. 6B), suggesting redox elements were examined to identify the redox conditions during
that they are more enriched than in an average shale (Tribovillard the deposition of the QS (Fig. 7C–F). Their cross-plots reveal that there
et al., 2006). Their co-enrichments in this part of the studied outcrop is no covariance between TOC and the redox sensitive elements Mo, Cu,
section indicate that these sediments were deposited under anoxic to U and V in the organic-lean facies but weak to moderate covariance in
euxinic environment. Except for Mo and As, the EFs of these elements the underlying intervals of the organic-rich lithofacies (Fig. 7C–F).
are depleted in the organic-lean shale and in the siltstone-shale and According to Tribovillard et al. (2006) and Algeo and Maynard (2004),
siltstone-sandstone facies indicating better oxygenated bottom water. TOC values less than 2 in association with non-covariance of the redox
Because of their high affinity for euxinic redox conditions, concurrent elements indicate oxic-dysoxic conditions while TOC values ranging
enrichment of U, V and Mo as is the case in the organic-rich facies of the from 2 to 7 wt% with moderate to strong covariance with U, and V and
QS outcrop, is often used to infer euxinic conditions in water column or moderate EFs values are an indication for anoxic conditions. Euxinic
at the sediment-water interface (e.g., Algeo and Maynard, 2004; conditions are normally indicated by TOC values above 7 wt%, and
Tribovillard et al., 2004, 2006). high EFs for U, V and Mo exhibiting non-covariance with the TOC
In addition to using elemental concentrations, their enrichment (Tribovillard et al., 2006; Algeo and Maynard, 2004). The value of these
factors and TOC to evaluate the QS redox conditions, some of the elements’ EFs and TOC in the organic-rich shale samples plot within the
commonly used redox ratios and indices [e.g. V/Cr, Ni/Co, Th/U, V/ bottom water with low oxygen (euxinic) region while those of the
(V + Ni) and (U + V)/(U + V + Mo)] were also utilized in this study. overlying organic-lean shale and siltstone/sandstone facies plot in fairly
The V/(V + Ni), V/Cr and Ni/Co ratios did not provide any strong oxic region (Fig. 7C–F).
evidence to differentiate the two major units, the organic-rich and or- Anoxic basins with weak water mass restriction, are characterized
ganic-lean shales from each other. However, the values and cross-plots by high [Mo]s/TOC ratios whereas highly stratified anoxic silled basins
of Mo versus (U + V)/(U + V + Mo) and Th/U versus V/(V + Ni) with strong water mass restriction and limited deep-water upwelling,
differentiated the organic-rich from the overlying organic-lean litho- are typified by low [Mo]s/TOC ratios (e.g., Algeo and Lyons, 2006;
facies (Fig. 7A and B). The organic-rich facies are characterized by Algeo et al., 2007; Algeo and Rowe, 2012). In the lower portion of the
lower (U + V)/(U + V + Mo) ratios (0.87–0.97) and higher Mo QS outcrop, the [Mo]s/TOC recorded an average of 5.3 in the organic-
(5–23.1 ppm) concentrations than the overlying organic-lean shale and rich laminated shale and 7.5 in the massive organic-rich shale litho-
the siltstone-sandstone facies indicating anoxic-euxinic conditions facies (Table 5). These low ratios in the lithofacies suggest that they
(Table 4). The high (U + V)/U + V + Mo) ratios (0.93–1) and less Mo were deposited under anoxic conditions, associated with strongly re-
(0.7–~6 ppm) concentrations in the organic-lean and siltstone-sand- stricted bottom water mass circulation. This restricted hydrographic
stone facies show that U and V were enriched in the sediments in- condition may have been enhanced in the study area, particularly, on
dependent of Mo and likely an indication of oxic to suboxic conditions paleotopography which was created due to the deglaciation of the Early
(Pi et al., 2013). Silurian time (Sharland et al., 2001; Hayton et al., 2017).
The ratio of V/(V + Cr) in most of the samples within the organic- Similar to the QS, the average [Mo]s/TOC value of 2 observed in the
rich facies exceed the 0.6 threshold that is associated with strong re- organic and pyrite-rich Barnette Shale in the US was attributed to
ducing conditions (Pi et al., 2013). The average value of 0.5 in both the strong water mass restriction (Loucks and Ruppel, 2007; Rowe et al.,
organic-lean shale, siltstone-shale and siltstone-sandstone facies suggest 2008). As in the Tabuk Basin, the Fort Worth Basin which was thought
less reducing conditions at the time of their deposition. to have been isolated from the Rhei and Panthalassic Ocean during the
The Th/U ratio developed by Adams and Weaver (1958), to examine late Mississippian hindered an effective water mass circulation
redox conditions, clearly shows a transition from a more intense re- (Gambacorta et al., 2016). Higher values of [Mo]s/TOC in the upper
ducing condition in the lower part of the studied outcrop section (or- 10 m of the outcrop compared to its basal section with an average of
ganic-rich lithofacies) to an improved and better oxygenated bottom 17.4 in the organic-lean shale and 25.5 in the siltstone-shale through
water condition in the upper part (organic-lean shale and overlying siltstone-sandstone facies indicate that the water mass became less re-
siltstone and sandstone intervals). The values of this ratio are rarely stricted and showed improved deep-water renewal.

252
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

Fig. 7. A) Bivariate plot between Mo and (U + V)/(U + V + Mo) ratio shows the separation between organic-rich and organic-lean samples, the same is also shown
in Figure (B) using Th/U to V/(V + Ni) ratios. (C) Bivariate plot shows the difference between organic-rich and organic-lean samples using V normalized to Al and
VEF. (D) Bivariate plot shows the distinguish between organic-rich from organic-lean samples using Cu normalized to Al and CuEF. (E) Bivariate plot shows the
difference between organic-rich and organic-lean samples using Mo normalized to Al and MoEF. (F) Bivariate plot shows the difference between organic-rich and
organic-lean samples using U normalized to Al and UEF. (G) Plot of Al2O3 vs TiO2 to show the provenance rock type of QS (modified after Schieber, 1992). (H) Plot
of SiO2 vs K2O/Na2O (after Roser and Korsch, 1986) show that QS was deposited entirely in passive continental margin setting.

253
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

5.2. Depositional model sediments were deposited in a deep, confined to semi-confined, low
energy environment (Cole, 1994).
The QS is mainly composed of siliciclastic sediments with traces of The deposition of the organic-rich laminated shale lithofacies in-
carbonates. As indicated by the bivariate relationship between the filled the accommodation space, which was created by the underlying
detrital influx indicators, Ti/Al and Th/Al, the sediments were sourced basin floor relief (Sharland et al., 2001; Hayton et al., 2017) (Fig. 8).
from the proximal granitic to basaltic rocks in the Precambrian Arabian Accordingly, the water depth gradually decreased as indicated by the
shield (Schieber, 1992; Riquier et al., 2006) (Fig. 7 G). The bivariate retraction of organic content coupled with the dilution effect of the
plot of SiO2 vs K2O/Na2O (quartz and k-feldspar to plagioclase ratio increased sediment influx. The water energy slightly increased to the
relationship) (Fig. 7 H) suggests that the shale sediments were de- turbidity current level and this associated with some changes in the
posited on a passive continental margin setting (Roser and Korsch, chemical conditions at sediment-water interface as evidenced by the
1986). existence of free calcium cations enough to form the underlying car-
The lower laminated organic-rich shale lithofacies characterized by bonate concretions (Aplin and Macquaker, 2011).
well-preserved laminations, fine-grain size, high clay content, and ab- The changes that accompanied the gradual decrease of the TOC and
sence of bioturbation. These evidence indicate a deposition in relatively pyrite in the massive organic-rich shale lithofacies may suggest a de-
low energy, quiet water by settling down from hemipelagic sediments position in the lower shoreface to inner shelf and close to the turbidity
(Loucks and Ruppel, 2007; Abouelresh and Slatt, 2011, 2012). How- current zone as inferred from the existence of the intraformational mud
ever, the terrestrial inputs into this lithofacies are predominantly sub- clasts. This lithofacies can be related to the less-organic rich shale
angular, sub-rounded silt-size quartz grains (Mulder and Alexander, samples described by Jones and Stump (1999) from the central and
2001; Liang et al., 2012). This lithofacies contains the highest average eastern subsurface of Saudi Arabia.
of pyrite (Table 1) in association with relatively high MoEF (10 com- The sharp decrease of the TOC from the underlying massive organic-
pared to the standard value for shale) indicating that the bottom-water rich lithofacies to the overlying organic-lean shale lithofacies (Fig. 2)
conditions were most likely sulfidic (Tribovillard et al., 2004, 2006). was likely due to improved bottom water oxygenation. The depletion of
Although concretions are very common in organic-rich shale as an redox elements and indices in this interval likely limited the preserva-
early diagenetic feature, they have not been discussed before in the QS, tion of organic matter in the upper part which include organic-lean
particularly in Tabuk Basin. In the study area, the graptolitic concre- shale and silty/sand stone lithofacies. Additionally, the increase of silt
tions occur as discontinuous lenses of calcareous concretions exhibiting and sand in these lithofacies might have also contributed to the re-
a similar appearance throughout the entire outcrop (Fig. 1D). The soft duction of preserving the organic matter particles.
sediment deformation structures in the shale laminae above and below Planar, trough and hummocky-cross stratifications are present in
the concretion lenses, internal fracturing and the preserved graptolite the siltstone-fine sandstone lithofacies in the topmost part of the studied
traces in the cores of these concretions support an in-situ, syngenetic to section. These cross stratifications are interpreted to indicate the
early diagenetic origin formed close to the sediment-water interface dominance of high energy shallow water and highly oxygenated bottom
during times of no or low sedimentation rates (Raiswell, 1988). The water conditions during deposition, probably in a middle to inner shelf
observed soft sediment deformation of the laminae in the immediate depositional settings (Fig. 8). The occurrences of trough and hummocky
vicinity of the concretions might have resulted from the differences in cross stratification in the facies also indicate the existence of a very
the bulk density in the concretions’ calcite composition, and the organic dynamic water setting during the deposition. The abundance of mi-
matter and silica richness of the surrounding shale laminae. The oc- caceous minerals and low TOC content (< 0.5 wt%) in the facies sug-
currence of graptolites as nuclei in these concretions indicates that the gests deposition close to the provenance areas.

Fig. 8. Sketch diagram (not to scale) shows the suggested depositional model for the Qusaiba Shale in Tabuk Basin. The model shows the upward shallowing in water
depth creates vertical facies change.

254
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

5.3. Reservoir potential only be produced by the application of appropriate technologies (e.g.
horizontal drilling and hydraulic fracture). The QS act as the main
The hydrocarbons generated in source rocks are not completely source rock for the gas condensate produced from the Permo-Carboni-
expelled and migrated into the associated conventional reservoirs, a ferous Unayzah Sandstone and the overlying Permo-Triassic Khuff
proportion of 20–40% of the hydrocarbon generated, remains in the carbonate reservoirs in Saudi Arabia (Pollastro et al., 2002.), therefore,
source rocks, probably adsorbed on pore walls or on kerogen particles it has the potentials to act as an unconventional reservoir. Subse-
(McKenzie et al., 1988). The retained hydrocarbons in these source quently, the QS outcrop section under investigation was suggested as an
rocks accumulate to form potential unconventional resources that could analogue for its subsurface equivalent.

Fig. 9. Rock-Eval pyrolysis data from Qusaiba Shale (n = 68): (a) Total Organic Carbon TOC (wt%) vs Residual hydrocarbon potential S2 (mgHC/g rock). (B) Plot of
oxygen index OI (mg CO2/g TOC) vs Hydrogen index HI (mg HC/g TOC) to confirm the Type III and IV for the kerogen content of QS. (C) Plot of Maturity (based on
Tmax oC) vs Hydrogen index HI (mg HC/g TOC) shows that the studied samples mainly immature with few samples in mature window.

255
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

In the geologic contexts, the unconventional reservoir quality refers content in these lithofacies (Sageman et al., 2003; Bohacs et al., 2005).
to a combination of properties that provide high storage capacity and The higher surface areas in clay-rich sediments facilitates organic
good production potential (Gamero-Diaz et al., 2013). Bed thicknesses, matter adsorption hence, they are naturally more enriched in organic
porosity, brittleness, kerogen type and amount, as well as its thermal matter than in corresponding coarser grained, silty or sandy sediments
maturity are the most important properties in the characterization of (Yu et al., 2009; Barber et al., 2017). Consequently, the study results
shale gas reservoirs (Curtis, 2002; Engelder et al., 2009; Hammes et al., indicate that the depositional settings, sediment influx and vertical li-
2011; Slatt, 2011). thological variation of the QS in Tabuk Basin as well as the paleo-to-
The TOC content in the basal laminated shale (10.5 m thick) ranges pography, control the vertical distribution of the organic content.
between 1.77 wt% and 6.1 wt%, with an average of 3.23 wt% while it The bivariate plot of TOC versus S2 shows that the kerogen in the
ranges between 0.5 wt% to 3.09 wt% with an average of 1.8 wt% in the studied QS is a mixture of type III (gas prone) and type IV (inert
overlain massive shale (9 m thick). Therefore, the lower part of the kerogen), although it was identified as type II in the subsurface core
studied QS section has the highest source rock potential and conse- samples (Mahmoud et al., 1992; Wender et al., 1998; Abu-Ali et al.,
quently it matches with the regionally distributed “Qusaiba Hot Shale” 1999; Abu-Ali and Littke, 2005). The contradiction here is most likely
in the central and eastern Saudi Arabia (Jones and Stump, 1999). The due to the low hydrogen content in the graptolites, which makes it
vertical distribution of TOC data shows an upward decreasing trend, chemically similar to type III kerogen (Inan et al., 2016). In addition,
which is consistent with the coarsening-upward sedimentation cycle with the increasing maturity, the relative inert kerogen content will
suggested for the QS in the study area (Pollastro et al., 2002). The low increases as the reactive organic matter is used in hydrocarbon gen-
values of TOC in certain intervals of the massive organic-rich shale li- eration (Dembicki, 2008, 2016). Furthermore, the reworking and oxi-
thofacies suggests that the bottom waters were exposed to periodic dizing of the organic material may also contribute to the inert kerogen
oxygenation during deposition. On the other hand, oxygenated bottom proportion (Tissot and Welte, 1984). HI in the organic-rich lithofacies
water might be the reason for the extremely low TOC values in the ranges between 4 and 219 mg HC/g TOC with average 70 mg HC/g
organic-lean, siltstone-shale and siltstone-sandstone intervals. High in- TOC) and high O content (OI up to 274 with 57 in average) (Fig. 9A–C).
flux of clastic detritus might have also negative influence on the organic The Tmax values ranges between 414 °C and 552 °C with an average of

Fig. 10. (A) SEM image show the inorganic porosity created as a result of random orientation of platelet-like grains. (B) SEM image shows a diagenetic pyrite
frameboide host some microporosity (C) SEM image show an immature organic particle free of porosity. (D) SEM image of mature organic particles host many
microporosity. (E) Thin section image show microfracture filled with hydrocarbon (Black arrows) in the laminated organic-rich shale lithofacies. (F) SEM image
shows a long narrow space between clay (mica ?) particles which might work as permeability channels.

256
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

430 °C. Excluding the anomalously high value in one sample (2.77), the contribute to the storage capacity of an organic-rich shale (Curtis,
calculated VRo from the measured Tmax, indicate that the QS in the 2002). The existence, development and propagation of mm to cm nat-
study area is immature to marginally mature. This combined with the ural fractures in an organic-rich shale are essentially controlled by type,
low Tmax and PI values suggest that the QS in the study area is immature percentage and distribution of mineral composition (Gale et al., 2014)
(Table 4). Based on graptolite vitrinite reflectance from subsurface, and (Fig. 10E–F). Therefore, identifying the spatial orientation of natural
at outcrops near to the current studied locality, Inan et al. (2016), fractures is the key factor of hydraulic fracture processes that will be
concluded that the QS is immature to overmature. needed for the reservoir completion (Walton and McLennan, 2013).
The storage capacity of an unconventional shale gas system is From the mineralogical point of view, although the ductile clay
mainly controlled by porosity as a storage index, and natural fractures mineral is the dominant mineral (average 56%) in the lower part of the
as permeability pathways (Curtis, 2002). Qualitative evaluation of studied QS, the brittle minerals (Quartz, Feldspar, and Pyrite) form an
porosity has been carried out on the QS using thin section petrography average of 42% and exhibit a uniform distribution of brittleness in the
and SEM imaging. Accordingly, different types of porosity including organic-rich interval (Fig. 11). In correlation with the worldwide shale
inorganic porosity, organic porosity and natural fracture porosity have gas, QS show typical matching with the Barnett Shale based on the bulk
been identified (e.g., Wang and Reed, 2009; Slatt and O'Brien, 2011; mineralogy (Fig. 11). On the other hand, pyrite has a distinctive var-
Loucks et al., 2012, Abouelresh, 2017). Inorganic porosity is created iation between 5 wt% and 15 wt% and shows a strong positive re-
between the main rock components. For instance, the random or- lationship with the vertical distribution of TOC. Interestingly, this
ientation of platelet-like mineral grains (clay and mica) is responsible particular interval has a natural fracture density that reaches up to 7.7
for a large number of pore size ranges between 2 and 20 μm (Fig. 10 A). fracture/m2, which has likely resulted from the high organic richness
Additionally, pyrite framboids also contribute to the storage capacity of and the high pyrite content (Abouelresh et al., 2016). Based on these
QS by creating diagenetic-related micro-porosity with an average size characteristics, the lower part may have the highest reservoir potenti-
less than 1 μm (Fig. 10 B). Organic-related porosity occurs within and/ ality of QS in the study area.
or on the surface of the mature organic particles as a result of genera-
tion and expulsion of hydrocarbons (e.g. Jarvie et al., 2007; Loucks
6. Conclusion
et al., 2009; Milliken et al., 2013). Organic particles of the QS in the
Rub’ Al-Khali Basin, Saudi Arabia host about 25% of the total porosity,
The current work focused on full sedimentological and geochemical
although the organic richness is less than 2 wt% (Abouelresh, 2017). In
characterization of QS sections that crop out in the Tabuk Basin in the
the studied samples, organic-related porosity does not have that high
northwest of Saudi Arabia.
fraction of the total porosity due to the low maturity levels of these
The integration of petrographic analysis and field description al-
samples (Fig. 10 C). However, in a few cases, the organic particles with
lowed the subdivision of the QS into six different lithofacies; (1)
relatively (high?) maturity show a significant micro-porosity with pore
Laminated organic-rich shale (2) Graptolitic concretion lithofacies (3)
size ranging between 0.2 and 0.6 μm, and show a high level of pore
Massive organic-rich shale lithofacies (4) organic-lean shale lithofacies
connectivity, which is expected to substantially contribute to the po-
(5) Siltstone-shale interbedded lithofacies and (6) Cross-stratified silt-
tential permeability (Fig. 10 D).
stone/sandstone lithofacies. The major element oxides in the QS show
The formation's natural fracture network can significantly
uniform distribution in the lower 25 m of the outcrop. Whilst in the

Fig. 11. Ternary plot of the mineral composition of the QS in the study area correlated with the worldwide shale gas (Barnett Shale, Eagle Ford Shale, Haynesville
Shale and Marcellus Shale). In order to normalize the three variables (Quartz, carbonate and clay minerals), K-feldspar and plagioclase were added to quartz.

257
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

upper 5 m of the outcrop, some of the elements (Si, Al, Ti, P) exhibit Al-Laboun, A.A., 1986. Stratigraphy and hydrocarbon potential of the paleozoic succes-
peak excursions towards the top of the organic-lean shale through the sion in both Tabuk and Widyan basins, Arabia. In: In: Halbouty, M.T. (Ed.), Future
Petroleum Provinces of the World: American Association of Petroleum Geologists
cross-bedded siltstone-sandstone units. Memoir, vol. 40. pp. 399–425.
Significant enrichments of V, Mo, and U together with low values of Al-Laboun, A.A., 2009. Tectono stratigraphy of the exposed Silurian deposits in Arabia.
Th/U (less than 2) in the organic-rich lithofacies of the investigated Arab. J. Geosci. 2 (2), 119–131.
Alexeyenko, V.A., Adebiyi, A.I., Bartko, M.K., Faraj, O., 2012. First Shale Gas Experience
outcrop section are indicative for low oxygen conditions, creating an in Saudi Arabia: Lessons Learned. SPE 163320.
anoxic environment during the deposition of organic-rich shale in- Algeo, T.J., Lyons, T.W., 2006. Mo-total organic carbon covariation in modern anoxic
terval. The low values of the [Mo]s/TOC ratio in this basal section of the marine environments: implications for analysis of paleoredox and paleohydrographic
conditions. Paleoceanography 21 PA1016 23 pp.
outcrop also reflects strongly stratified restricted anoxic water column Algeo, T.J., Maynard, J.B., 2004. Trace-element behavior and redox facies in core shales
and euxinic conditions during the deposition. of Upper Pennsylvanian Kansas-type cyclothems. Chem. Geol. 206, 289–318.
The paleo-topography, which was created during the Late Algeo, T.J., Rowe, H., 2012. Paleoceanographic applications of trace-metal concentration
data. Chem. Geol. 324–325 (2012), 6–18.
Ordovician deglaciation event is largely controls the deposition of QS,
Algeo, T.J., Tribovillard, N., 2009. Environmental analysis of paleoceanographic systems
in particular, the ponding of the lower organic-rich part. The lower based on molybdenum–uranium covariation. Chem. Geol. 268 (3), 211–225.
laminated organic-rich shale lithofacies deposited as suspended hemi- Algeo, T.J., Hannigan, R., Rowe, H., Brookfield, M., Baud, A., Krystyn, L., Ellwood, B.B.,
pelagic sediments. The water depth decreased gradually as indicated by 2007. Sequencing events across the permian–Triassic boundary, Guryul Ravine
(Kashmir, India). Palaeogeogr. Palaeoclimatol. Palaeoecol. 252, 328–346.
the retraction of organic preservation coupled with the dilution effect of Aoudeh, S.M., Al-Hajri, S.A., 1995. Regional distribution and chronostratigraphy of the
the sediment influx. The siltstone - fine sandstone facies in the topmost Qusaiba member of the Qalibah formation in the Nafud basin, northwestern Saudi
part of the studied section pointed to the dominance of high energy Arabia. In: In: Al-Husseini, M.I. (Ed.), The Middle East Petroleum Geoscience,
Selected Middle East Papers from the Middle East Geoscience Conference, April
shallow oxygenated (probably middle to inner shelf) depositional set- 25–27, 1994, vol. 1. Gulf Petrolink, Bahrain, pp. 143–154.
tings. The obtained results indicate that the depositional settings, se- Aplin, A.C., Macquaker, H.S., 2011. Mudstone diversity: origin and implications for
diment influx and the vertical lithological variation of the QS in the source, seal, and reservoir properties in petroleum systems. AAPG Bull. 95 (12),
2031–2059 (December 2011).
Tabuk Basin control the quality and distribution of the organic content. Arouri, R.K., Van Laer, J.P., Prudden, H.M., Jenden, D.J., Carrigan, J.W., Al-Hajji, A.A.,
TOC is dramatically decreased from the basal black and dark grey 2010. Controls on hydrocarbon properties in a Paleozoic petroleum system in Saudi
laminated organic-rich shale to the topmost cross-stratified silty/sandy Arabia: exploration and development implications. AAPG Bull. 94 (2), 163–188.
Arthur, M.A., Jenkyns, H.C., Brumsack, H.J., Schlanger, S.O., 1990. Stratigraphy, geo-
lithofacies. The Tmax and calculated VRo reflect the immaturity to chemistry, and palaeoceanography of organic carbon-rich Cretaceous sequences. In:
marginal maturity of the QS in the study area. Qualitative evaluation of Ginsburg, R.N., Beaudoin, B., Ser, C. (Eds.), Cretaceous Resources, Events and
porosity in the QS showed that there are three porosity types including, Rhythms. NATO (N. Atlantic Treaty Org.) ASI (Adv. Stud. Inst.) Kluwer, Dordrecht,
pp. 75–119.
inorganic porosity, organic porosity and natural fracture porosity. The
Baioumy, H., Lehmann, B., 2017. Anomalous enrichment of redox-sensitive trace ele-
brittle minerals (Quartz, Feldspar, and Pyrite) form an average of 42% ments in the marine black shales from the Duwi Formation, Egypt: evidence for the
and exhibit a uniform distribution of brittleness in the organic-bearing late Cretaceous Tethys anoxia. J. Afr. Earth Sci. 133, 7–14.
interval. Natural fracture density reaches up to 7.7 fracture/m2 in the Barber, A., Brandes, J., Leri, A., Lalonde, K., Balind, K., Wirick, S., Wang, J., Gélinas, Y.,
2017. Preservation of organic matter in marine sediments by inner-sphere interac-
basal organic-rich lithofacies, which indicate their high potential for tions with reactive iron. Sci. Rep. 7, 366. https://doi.org/10.1038/s41598-017-
effective hydraulic fracture. 00494-0.
Bishop, R.S., 1995. Maturation history of the lower paleozoic of the eastern arabian
platform. In: In: Al-Husseini, M.I. (Ed.), Geo-94, Middle East Petroleum Geosciences
Acknowledgement Conference: Gulf Petrolink, vol. 1. pp. 180–189 Manama, Bahrain.
Bohacs, K.M., Grawbowski, G.J., Carroll, A.R., Mankeiwitz, P.J., Miskell-Gerhardt, K.J.,
We thank Center for Integrative Petroleum Research (CIPR), CPG, Schwalbach, J.R., Wegner, M.B., Simo, J.A., 2005. Production, Destruction, and
Dilution – the Many Paths to Source-Rock Development, vol. 82. SEPM Special
KFUPM and Weatherford Center at Dhahran Techno Valley (DTV) for Publication, pp. 61–101.
the continuous support during this work. We appreciate the contribu- Boles, J.R., Landis, C.A., Dale, P., 1985. The Moeraki Boulders; anatomy of some septarian
tions from two anonymous reviewers whose comments and suggestions concretions. J. Sediment. Petrol. 55 (3), 398–406.
Brumsack, H.-J., 1989. Geochemistry of recent TOC-rich sediments from the Gulf of
greatly improved the manuscript. We acknowledge the great help and
California and the black sea. Geol. Rundsch. 78, 851–882.
enthusiasm we received from the editor. Calvert, S.E., Pedersen, T.F., 1993. Geochemistry of recent oxic and anoxic marine se-
diments: implications for the geological record. Mar. Geol. 113, 67–88.
Cole, A.G., 1994. Graptolite-chitinozoan reflectance and its relationship to other geo-
Appendix A. Supplementary data
chemical maturity indicators in the Silurian Qusaiba shale, Saudi Arabia. Energy
Fuels 8, 1443–1459.
Supplementary data to this article can be found online at https:// Cole, A.G., Abu-Ali, M.A., Aoudeh, S.M., Carrigan, W.J., Chen, H.H., Colling, E.L.,
doi.org/10.1016/j.marpetgeo.2019.05.001. Gwathney, W.J., Al-Hajji, A.A., Halpern, H.I., Jones, P.J., Al-Sharidi, S.H., Tobey,
M.H., 1994. Organic geochemistry of the Paleozoic petroleum system of Central
Saudi Arabia. Energy Fuels 8, 1425–1442.
References Curtis, J.B., 2002. Fractured shale-gas systems. AAPG Bull. 86, 1921–1938.
Dean, W.E., Gardner, J.V., Piper, D.Z., 1997. Inorganic geochemical indicators of gla-
cial–interglacial changes in productivity and anoxia on the California continental
Abouelresh, M.O., 2017. An integrated characterization of the porosity in Qusaiba Shale, margin. Geochim. Cosmochim. Acta 61, 4507–4518.
Saudi Arabia. J. Pet. Sci. Eng. 149, 75–87. Dembicki Jr., H., 2008. Three common source rock evaluation errors made by geologists
Abouelresh, M.O., Slatt, R., 2011. Shale depositional processes: example from the pa- during prospect or play appraisals. AAPG Bull. 93 (3), 341–356.
leozoic Barnett shale, Fort Worth Basin, Texas, USA. Cent. Eur. J. Geosci. 3 (4), Dembicki, H., 2016. Practical Petroleum Geochemistry for Exploration and Production.
398–409. Elsevier Science, Document, Internet resource, Saint Louis 9780128033517
Abouelresh, M.O., Slatt, R., 2012. Lithofacies and sequence stratigraphy of the Barnett 0128033517.
shale in the east-central Fort Worth Basin, Texas, USA. AAPG Bull. 96 (1), 1–22. Eltom, H.A., Abdullatif, O.M., Babalola, L.O., Bashari, M.A., Yassin, M., Osman, M.S.,
Abouelresh, M.O., Boyde, D., Babalola, L.O., Koithan, K., Omer, M., 2016. Factors con- Abdulraziq, A.M., 2016. Geochemical characterization of the Permian–Triassic
trolling natural fracture development in the Qusaiba hot shale, northwest Saudi transition at outcrop, central Saudi Arabia. J. Pet. Geol. 39 (1), 95–113.
Arabia; outcrop approach. In: AAPG Middle East Geosciences Technology Workshop Engelder, T., Lash, G.G., Uzcategui, R.S., 2009. Joint sets that enhance production from
Source Rocks of the Middle East 25-26 January 2016 Abu Dhabi. middle and upper devonian gas shales of the appalachian basin. AAPG Bull. 93 (7),
Abu Ali, M.A., Littke, R., 2005. Paleozoic petroleum systems of Saudi Arabia: A basin 857–889. https://doi.org/10.1306/03230908032.
modeling approach. Geo- Arabia 10 (3), 131–168. Espitalie, J., Laporte, L.J., Madec, M., Marquis, F., Leplat, P., Paulet, J., Boutefeu, A.,
Abu-Ali, M.A., Franz, U.A., Shen, J., Monnier, F., Mahmoud, M.D., Chambers, T.M., 1991. 1977. Methode rapide de caracterisation des roches mères, de leur potential petrolier
Hydrocarbon Generation and Migration in the Paleozoic Sequence of Saudi Arabia. et de leur degré d'evolution. Rev., Inst. Franc. Pétrole 32, 32–42.
Society of Petroleum Engineers SPE 21376, pp.34–356. Fox, J.E., Ahlbrandt, T.S., 2002. Petroleum Geology and Total Petroleum Systems of the
Abu-Ali, M.A., Rudkiewicz, J.L., McGillivray, J.G., Behar, F., 1999. Paleozoic petroleum Widyan Basin and Interior Platform of Saudi Arabia and Iraq. U.S. Geological Survey
system of central Saudi Arabia. GeoArabia 4 (3), 321–335. Bulletin 2202-E, pp. 25.
Adams, J.A., Weaver, C.E., 1958. Thorium to uranium ratios as indicators of sedimentary Gale, J.F.W., Laubach, S.E., Olson, J.E., Eichhubl, P., Fall, A., 2014. Natural fractures in
processes; example of concept of geochemical facies. AAPG Bull. 42 (2). shale: a review and new observations. AAPG Bull. 98 (11), 2165–2216. https://doi.

258
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

org/10.1306/08121413151. the Late Ordovician–early Silurian: review and new insights from black shales and
Gambacorta, G., Trincianti, E., Torricelli, S., 2016. Anoxia controlled by relative sea-level nitrogen isotopes. GSA Bull. 125 (11–12), 1635–1670. https://doi.org/10.1130/
changes: an example from the Mississippian Barnett Shale Formation. Palaeogeogr. B30812.1.
Palaeoclimatol. Palaeoecol. 495, 152–162. Meyers, P.A., Bernasconi, S.M., Forster, A., 2006. Origins and accumulation of organic
Gamero-Diaz, H., Miller, C., Lewis, R., 2013. sCore: A Mineralogy Based Classification matter in expanded Albian to Santonian black shale sequences on the Demerara Rise,
Scheme for Organic Mudstones. SPE 166284. South American margin. Org. Geochem. 37, 1816–1830.
Goldberg, T., Strauss, H., Guo, Q., Liu, C., 2007. Reconstructing marine redox conditions Milliken, K.L., Rudnicki, M., Awwiller, D.N., Zhang, T., 2013. Organic matter-hosted pore
for the Early Cambrian Yangtze Platform: evidence from biogenic sulphur and or- system, marcellus formation (Devonian), Pennsylvania, USA. AAPG Bull. 97,
ganic carbon isotopes. Palaeogeogr. Palaeoclimatol. Palaeoecol. 254 (1–2), 175–193. 177e200.
Halawani, M.A., Al-Duaiji, A.A., Bahabri, B.H., Basyoni, M.H., Al Dabbagh, M.E., Al Milner, P.A., 1998. Source rock distribution and thermal maturity in the Southern
Ramadan, K., Al Ajmi, H.F., Al Mahri, A.K., Saudi Stratigraphic Committee Members, Arabian Peninsula. GeoArabia 3, 339–356.
2013. Phanerozoic Stratigraphy of Saudi Arabia: Part 1 – Paleozoic Succession of the MorCheshirea, S., Craddockb, P.R., Xu, G., Sauerer, B., Andrew, E., Pomerantz, A.E.,
Arabian Shelf (Cover Rocks), first ed. Saudi Geological Survey Special Publication McCormick, D., Abdallah, W., 2017. Assessing thermal maturity beyond the reaches
SGS-SP-2012-1, pp. 69. of vitrinite reflectance and Rock-Eval pyrolysis: a case study from the Silurian
Hammes, U., Hamlin, H.S., Ewing, T.E., 2011. Geologic analysis of the upper Jurassic Qusaiba formation. Int. J. Coal Geol. 180, 29–45.
Haynesville shale in east Texas and west Louisiana. AAPG Bull. 95 (10), 1643–1666. Morford, J.L., Emerson, S., 1999. The geochemistry of redox sensitive trace metals in
Hayton, S., Rees, A.J., Vecoli, M., 2017. A punctuated Late Ordovician and early Silurian sediments. Geochim. Cosmochim. Acta 63, 1735–1750.
deglaciation and transgression: evidence from the subsurface of northern Saudi Morford, J.L., Russell, A.D., Emerson, S., 2001. Trace metal evidence for changes in the
Arabia. AAPG Bull. 101 (6), 863–886. redox environment associated with the transition from terrigenous clay to diato-
Hudson, J.D., Coleman, M.L., Barreiroà, B.A., Hollingworth, N.T.J., 2001. Septarian maceous sediment, Saanlich Inlet, BC. Mar. Geol. 174, 355–369.
concretions from the Oxford Clay (Jurassic, England, UK): involvement of original Mulder, T., Alexander, J., 2001. The physical character of subaqueous sedimentary
marine and multiple external pore fluids. Sedimentology 48, 507–531. density currents and their deposits. Sedimentology 48, 269–299.
Husseini, M.I., 1991. Tectonic and depositional model of the Arabian and adjoining plates Paris, F., Verniers, J., Al-Hajri, S., Al-Tayyar, H., 1995. Biostratigraphy and palaeogo-
during the Silurian-Devonian. Am. Assoc. Pet. Geol. Bull. 75, 108–120. graphic affinities of Early Silurian chitinozoans from Central Saudi Arabia. Rev.
Inan, S., Goodarzi, F., Schmidt Mumm, A., Arouri, K., Qathami, S., Ardakani, O.H., İnan, Palaeobot. Palynol. 89, 75–90. https://doi.org/10.1016/0034-6667(95)00042-V.
T., Tuwailib, A.A., 2016. The Silurian Qusaiba hot shales of Saudi Arabia: an in- Pi, D.H., Liu, C.Q., Graham, A., et al., 2013. Trace and rare earth element geochemistry of
tegrated assessment of thermal maturity. Int. J. Coal Geol. 159, 107–119. https://doi. black shale and kerogen in the early cambrian Niutitang formation in Guizhou pro-
org/10.1016/j.coal.2016.04.004. vince, south China: constraints for redox environments and origin of metal enrich-
Janjou, D., Halawani, M.A., Al-Muallem, M.S., Robelin, C., Brosse, J.-M., Courbouleix, S., ments. Precambrian Res. 225, 218–229.
Dagain, J., et al., 1996a. Explanatory Notes to the Geologic Map of the Al Qalibah Piper, D.Z., Perkins, R.B., 2004. A modern vs. Permian black shale- the hydrography,
Quadrangle, Kingdom of Saudi Arabia, Geological Map GM-135, Scale 1:250,000, primary productivity, and water-column chemistry of deposition. Chem. Geol. 206,
Sheet 28C: Jeddah, Kingdom of Saudi Arabia. Deputy Ministry for Mineral Resources, 177–197.
Ministry of Petroleum and Mineral Resources, pp. 44. Pollastro, R.M., Schenk, C.J., Ahlbrandt, T.S., 2002. Undiscovered Oil and Gas Resources
Janjou, D., Halawani, M.A., Brosse, J.-M., Al-Muallem, M.S., Becq-Giraudon, J.F., Dagain, of Lower Silurian Qusaiba-Paleozoic Total Petroleum Systems, Arabian Peninsula.
J., Genna, A., et al., 1996b. Explanatory Notes to the Geologic Map of the Tabuk U.S. Geological Survey Fact Sheet FS-008-02, 2 p.
Quadrangle, Kingdom of Saudi Arabia, Geological Map GM-137, Scale 1:250,000, Powers, R.W., 1968. Lexique stratigraphique international: Saudi Arabia, v. III, Asie, fasc
Sheet 28B: Jeddah Kingdom of Saudi Arabia. Deputy Ministry for Mineral Resources, 10b1. Centre National de la Recherché Scientifique, Paris, pp. 177.
Ministry of Petroleum and Mineral Resources, pp. 49. Powers, R.W., Ramirez, L.F., Redmond, D.D., Elberg Jr., E.L., 1966. Geology of the
Jarvie, D.M., Claxton, B.L., Henk, F., Breyer, J.T., 2001. Oil and shale gas from the Barnett Arabian Peninsula, sedimentary geology of Saudi Arabia. USGS Professional Paper,
Shale, Ft. Worth Basin, Texas (abs.). AAPG Ann. Meet. Progr. 10, A100. 560-D, pp. 150.
Jarvie, D.M., Hill, R.J., Ruble, T.E., Pollastro, R.M., 2007. Unconventional shale-gas Raiswell, R., 1988. Chemical model for the origin of minor limestone-shale cycles by
systems: the Mississippian Barnett Shale of north-central Texas as one model for anaerobic methane oxidation. Geology 16, 641–644.
thermogenic shale-gas assessment. AAPG Bull. 91, 475e499. Rickards, R.B., Koren, T.N., 1974. Virgellar meshworks and sicular spinosity in
Jones, P.J., Stump, T.E., 1999. Depositional and tectonic setting of the lower Silurian Llandovery graptoloids. Geol. Mag. 3 (3), 193–272.
hydrocarbon source rock facies, Central Saudi Arabia. AAPG Bull. 83 (2), 314–332. Rimmer, S.M., 2004. Geochemical paleoredox indicators in devonian–Mississippian black
Konert, G., Afifi, A.M., Al-Hajri, S.A., Droste, H.J., 2001. Paleozoic stratigraphy and hy- shales, central appalachian basin (USA). Chem. Geol. 206, 373–391.
drocarbon habitat of the Arabian Plate. GeoArabia 6, 407–442. Riquier, L., Tribovillard, N., Averbuch, O., Devleeschouwer, X., Riboulleau, A., 2006. The
Lazar, O.R., Bohacs, K.M., Macquaker, J.H.S., Schieber, J., Demko, T.M., 2015. Capturing late Frasnian Kellwasser horizons of the Harz Mountains (Germany): two oxygen-
key attributes of fine-grained sedimentary rocks in outcrops, cores, and thin sections: deficient periods resulting from different mechanisms. Chem. Geol. 233 (1–2),
nomenclature and description guidelines. J. Sediment. Res. 85, 230–246. 137–155.
Lewan, M.D., Maynard, J.B., 1982. Factors controlling the enrichment of vanadium and Roser, B.P., Korsch, R.J., 1986. Determination of tectonic setting of sandstone-mudstone
nickel in the bitumen of organic sedimentary rocks. Geochim. Cosmochim. Acta 46, suites using Si02 content and K20/Na20 ratio. J. Geol. 94, 635–650.
2547–2560. Rowe, H.D., Loucks, R.G., Ruppel, S.C., Rimmer, S.M., 2008. Mississippian Barnett for-
Liang, C., Jiang, Z.X., Yang, Y.T., Wei, X.J., 2012. Shale lithofacies and reservoir space of mation, Fort Worth Basin, Texas: bulk geochemical inferences and Mo-TOC con-
the Wufeng–longmaxi formation, Sichuan basin, China. Pet. Explor. Dev. 39, straints on the severity of hydrographic restriction. Elsevier Chem. Geol. 257, 16–25.
736–743. https://doi.org/10.1016/S1876-3804(12)60098-6. Sageman, B.B., Murphy, A.E., Werne, J.P., Ver Straeten, C.A., Hollander, D.J., Lyons,
Loucks, R.G., Ruppel, S.C., 2007. Mississippian Barnett shale: lithofacies and depositional T.W., 2003. A tale of shales: the relative roles of production, decomposition, and
setting of a deep water Shale-gas succession in the Fort Worth Basin, Texas. AAPG dilution in the accumulation of organic-rich strata, middle-upper Devonian,
Bull. 91, 579–601. https://doi.org/10.1306/11020606059. Appalachian Basin. Chem. Geol. 195, 229–273.
Loucks, R.G., Reed, R.M., Ruppel, S.C., Jarvie, D.M., 2009. Morphology, genesis, and Schieber, J., 1992. A combined petrographical-geochemical provenance study of the
distribution of nanometer scale pores in mudstones of the Mississippian Barnett Newland Formation, Mid-Proterozoic of Montana. Geol. Mag. 129, 223–237.
Shale. J. Sediment. Res. 79, 848–861. https://doi.org/10.2110/jsr.2009.092. Scott, C., Lyons, T.W., 2012. Contrasting molybdenum cycling and isotopic properties in
Loucks, R.G., Reed, R.M., Ruppel, S.C., Hammes, U., 2012. Spectrum of pore types and euxinic versus non-euxinic sediments and sedimentary rocks: Refining the paleo-
networks in mud rocks and a descriptive classification for matrix-related mud rock proxies. Chem. Geol. 324–325, 19–27.
pores. AAPG Bull. 96 (6), 1071–1098. Sharland, P.R., Archer, R., Caey, D.M., Davies, R.B., Hall, S.H., Heward, A.P., Horbury,
Lüning, S., Craig, J., Loydell, D.K., Storch, P., Fitches, B., 2000. Lower Silurian ‘hot shales’ A.D., Simmons, M.D., 2001. Arabian Plate Sequence Stratigraphy, vol. 2. GeoArabia
in north africa and Arabia: regional distribution and depositional model. Earth Sci. Special Publication, pp. 371.
Rev. 49, 121–200. Simmons, M.D., Sharland, P.R., Casey, D.M., Davies, R.B., Sutcliffe, O.E., 2007. Arabian
Lüning, S., Shahin, Y.M., Loydell, D., Al-Rabi, H.T., Masri, A., Tarawneh, B., Kolonic, S., Plate sequence stratigraphy: potential implications for global chronostratigraphy.
2005. Anatomy of a world-class source rock: distribution and depositional model of GeoArabia 12, 4.
Silurian organic-rich shales in Jordan and implications for hydrocarbon potential. Slatt, R.M., 2011. Important geological properties of unconventional resource shales.
Am. Assoc. Pet. Geol. Bull. 89 (10), 1397–1427. Cent. Eur. J. Geosci. 3 (4), 435–448. https://doi.org/10.2478/s13533-011-0042-2. •
Mahmoud, M.D., Vaslet, D., Al-Husseini, M.I., 1992. The lower Silurian Qalibah forma- 2011 •.
tion of Saudi Arabia: an important hydrocarbon source rock. Am. Assoc. Pet. Geol. Slatt, R.M., O'Brien, N.R., 2011. Pore types in the Barnett and Woodford gas shales:
Bull. 76, 1491–1506. contribution to understanding gas storage and migration pathways in fine-grained
McClure, H.A., 1988. Chitinozoan and acritarch assemblages, stratigraphy and biogeo- rocks. AAPG Bull. 95, 2017–2030. https://doi.org/10.1306/03301110145.
graphy of the early palaeozoic of northwest Arabia. Rev. Palaeobot. Palynol. 56, Stewart, A.S., Reid, C.T., Hooker, N.P., Kharouf, O.W., 2016. Mesozoic siliciclastic re-
41–60. servoirs and petroleum system in the Rub’ Al-Khali basin, Saudi Arabia. AAPG Bull.
McGillivray, J.G., Husseini, M.I., 1992. The Paleozoic petroleum geology of Saudi Arabia. 100 (5), 819–841.
Am. Assoc. Pet. Geol. Bull. 76, 1473–1490. Stump, T.E., Van Der Eem, J.G., 1995. The stratigraphy, depositional environments and
McKenzie, A.S., Leythaeuser, D., Muller, P., Quigley, T.M., Radke, M., 1988. Movement of periods of deformation of the Wajid outcrop belt, southwestern Saudi Arabia. J. Afr.
hydrocarbons in shales. Nature 331, 63–65. Earth Sci. 21 (3), 421–441.
McLennan, S.M., 2001. Relationships between the trace element composition of sedi- Taylor, S.R., McLennan, S.M., 1985. The Continental Crust: its Composition and
mentary rocks and upper continental crust. Geochem. Geophys. Geosyst. 2 Evolution. Blackwell Scientific Publication, Carlton, pp. 312.
(2000GC000109). Tissot, B.P., Welte, D.H., 1984. Petroleum Formation and Occurrence. Springer-Verlag,
Melchin, J.M., Mitchell, E.C., Holmden, C., Štorch, P., 2013. Environmental changes in New York, pp. 699.

259
M. Abouelresh, et al. Marine and Petroleum Geology 111 (2020) 240–260

Tribovillard, N., Bialkowski, A., Tyson, R.V., Lallier-Vergès, E., Deconinck, J.F., 2001. Ministry for Mineral Resources, Ministry of Petroleum and Mineral Resources, pp. 51.
Organic facies variation in the late Kimmeridgian of the Boulonnais area (northern- Walton, I., McLennan, J., 2013. The role of natural fractures in shale gas production,
most France). Mar. Pet. Geol. 18 (3), 371–389. Chapter 16. In: Bunger, A.P., McLennan, J., Jeffrey, R. (Eds.), Effective and
Tribovillard, N., Riboulleau, A., Lyons, T., Baudin, F., 2004. Enhanced trapping of mo- Sustainable Hydraulic Fracturing, https://doi.org/10.5772/45724.
lybdenum by sulfurized organic matter of marine origin as recorded by various Wang, F.P., Reed, R.M., 2009. Pore networks and fluid flow in gas shales. In: SPE Annual
Mesozoic formations. Chem. Geol. 213, 385–401. Technical Conference and Exhibition, vol. 8 Society of Petroleum Engineers, New
Tribovillard, N., Algeo, T.J., Lyons, T., Riboulleau, A., 2006. Trace metals as paleoredox Orleans, Louisiana (SPE 124253).
and paleoproductivity proxies: an update. Chem. Geol. 232, 12–32. Wedepohl, K.H., 1971. Environmental influences on the chemical composition of shales
Tribovillard, N., Bout-Roumazeilles, V., Algeo, T., Lyons, T.W., Sionneau, T., Montero- and clays. In: In: Ahrens, L.H., Press, F., Runcorn, S.K., Urey, H.C. (Eds.), Physics and
Serrano, J.C., Riboulleau, A., François Baudin, F., 2008. Paleodepositional conditions Chemistry of the Earth, vol. 8. Pergamon, Oxford, pp. 305–333.
in the Orca Basin as inferred from organic matter and trace metal contents. Mar. Geol. Wender, L.E., Bryant, J.W., Dickens, M.F., Neville, A.S., Al-Moqbel, A.M., 1998. Paleozoic
254 (1), 62–72. (Pre-Khuff) hydrocarbon geology of the Ghawar area, eastern Saudi Arabia.
Tribovillard, N., Algeo, T.J., Baudin, F., Riboulleau, A., 2012. Analysis of marine en- GeoArabia 3, 273–302.
vironmental conditions based on molybdenum–uranium covariation—applications to Wignall, P.B., Newton, R., 2001. Black shales on the basin margin: a model based on
Mesozoic paleoceanography. Chem. Geol. 324–325, 46–58. examples from the upper Jurassic of the Boulonnais, Northern France. Sediment.
Vaslet, D., 1987. Early Paleozoic glacial deposits in Saudi Arabia, a lithostratigraphic Geol. 144 (3), 335–356.
revision. In: Saudi Arabian Deputy Ministry for Mineral Resources Technical Record Wignall, P.B., Twitchett, R.J., 1996. Oceanic anoxia and the end Permian mass extinction.
BRGM-TR-07-1, pp. 24. Science 272, 1155–1158.
Vaslet, D., Janjou, D., Robelin, C., Al-Muallem, M.S., Halawani, M.A., Brosse, J.-M., Yu, B., Dong, H., Jiang, H., Lv, G., Eber, D., Li, H., Kim, J., 2009. The role of clay minerals
Breton, J.-P., Courbouleix, S., Roobol, M.J., Dagain, J., 1993. Explanatory Notes to in the preservation of organic matter in sediments of Qinghai lake, NW China. Clay
the Geologic Map of the Tayma Quadrangle, Kingdom of Saudi Arabia, Geological Clay Miner. 57 (No. 2), 213–226.
Map GM-134, Scale 1:250,000, Sheet 27C: Jeddah, Kingdom of Saudi Arabia. Deputy

260

You might also like