You are on page 1of 23

Marine and Petroleum Geology 146 (2022) 105926

Contents lists available at ScienceDirect

Marine and Petroleum Geology


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

Geochemical investigation and basin modelling of the Al renk shale


formation in the Melut Basin, south Sudan: Implications for estimation of
thermogenic gas generation potential
Mohammed Hail Hakimi a, **, Atif N. Abass b, Aref Lashin c, i, Abbas F. Gharib d,
Ahmed E. Radwan e, *, Afikah Rahim f, Adeeb Ahmed g, Lanre Asiwaju h, Wafaa E. Afify i
a
Geology Department, Faculty of Applied Science, Taiz University, 6803 Taiz, Yemen
b
Rawat Petroleum Operation Company, Khartoum, Sudan
c
King Saud University, College of Engineering-Petroleum and Natural Gas Engineering Department, Riyadh, 11421, Saudi Arabia
d
Department of Oil and Gas Economics, College of Administrative and Financial Science, Imam Ja’afar Al-Sadiq University, Baghdad, Iraq
e
Institute of Geological Sciences, Jagiellonian University, Gronostajowa 3a, 30-387, Kraków, Poland
f
School of Civil Engineering, Faculty of Civil Engineering University of Teknologi Malaysia, Johor, Malaysia
g
Department of Earth and Environmental Sciences, Bahria University, Islamabad, 44000, Pakistan
h
Department of Geology, University of Malaya, 50603, Kuala Lumpur, Malaysia
i
Geology Department, Faculty of Science, Benha University, P.O. Box 13518, Benha, Egypt

A R T I C L E I N F O A B S T R A C T

Keywords: This study uses organic geochemical and basin modelling results to examine the potential for thermogenic gas
Al renk shale Formation generation from oil-to-gas conversion of the organic-rich Lower Cretaceous Al Renk shales in the Melut Basin,
1-D Basin modelling South Sudan. The overall geochemical results show that the Al Renk shales are considered good gas-source rocks
Geochemistry
with total organic carbon (TOC) content >1 wt % and a low hydrogen index (i.e., HI < 150 mg HC/g TOC;
Thermogenic gas potential
Melut Basin
mainly Type III kerogen). This finding contrasts with the source of organic matter and biomarkers related to the
South Sudan depositional environment. Based on biomarker indicators, the Al Renk shales contains a mixture of aquatic
organic matter with some terrigenous land plants and was deposited under suboxic to relatively oxic settings in a
lacustrine environment. The presence of mixed aquatic and terrigenous organic matter in lacustrine shale sed­
iments suggests a mix of Types II and III kerogen as the original organic matter input during deposition. The
geochemical and optical maturity indicators represent late-mature oil to gas generation windows, implying that
the Al Renk shale has substantial thermogenic gas generation potential. Therefore, high thermal maturation is a
crucial factor in drastically altering the original organic matter (OM) and decreasing HI values (conversion of
kerogen to hydrocarbons). The kerogen to oil conversion and the subsequent cracking of the retained oil into
thermogenic gas have been simulated by the burial/thermal/generation models in the Agordeed-1 well. These
models show that nearly 10–50% of kerogen was transformed into oil during the Late Cretaceous (78–68 Ma),
correlating with the early to peak mature of oil generation window (0.62–0.85 %Ro). Moreover, oil was expelled
from the Al Renk source rock system during the Late Cretaceous to Late Eocene. From the Late Eocene until the
present, significant amounts of residual oil were cracked into gas due to the maturation of the gas generation
window (>1.30 Easy % Ro), which is ideal for wet gas generation. Consequently, the findings of this study
advocate for conventional gas exploration in the Melut Basin’s deeper stratigraphic succession, where the Al
Renk Shale Formation has reached a high maturity level of the gas generation window.

1. Introduction generate enough gas to form conventional gas reservoirs and/or shale
gas reservoirs through cracking of retained oils, in addition to gas-prone
Organic-rich shale source rocks containing type I and II kerogens can source rocks containing type III kerogen or coals (Curtis, 2002; Gai et al.,

* Corresponding author.
** Corresponding author.
E-mail addresses: ibnalhakimi@yahoo.com (M.H. Hakimi), radwanae@yahoo.com, ahmed.radwan@uj.edu.pl (A.E. Radwan).

https://doi.org/10.1016/j.marpetgeo.2022.105926
Received 27 December 2021; Received in revised form 14 September 2022; Accepted 16 September 2022
Available online 20 September 2022
0264-8172/© 2022 Published by Elsevier Ltd.
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

2018). The main distinct gas types in petroleum systems are biogenic gas onshore rift basins that are mostly exploited by conventional petroleum
(e.g., Antrim Shale of the Michigan Basin; Martini et al., 1998) and exploration activities in Sudan (Tong et al., 2004). It has gained atten­
thermogenic gas (e.g., Barnett Shale of the Fort Worth Basin and Bossier tion from the academic community and upstream operators owing to its
Shale of east Texas; Jarvie et al., 2007), and it can be a mixed system of excellent source rock characteristics and commercial hydrocarbon ac­
both two types as a third type (e.g., some New Albany Shale; Cluff and cumulations (Dou, 2004; Dou et al., 2007; Hakimi et al., 2021). Most of
Dickerson, 1982). The derivation of thermogenic gas in source rocks can the previous studies have focused on the organic geochemical charac­
be through thermogenic cracking of kerogen and in other cases, it can be teristics of the shale facies within the Mesozoic sedimentary rocks and
derived from the cracking of any retained oil (secondary cracking of concluded favourable source rock intervals in the Cretaceous Al-Gayger
petroleum) (e.g., Barnett Shale; Jarvie et al., 2007). In addition, and Galhak formations with oil generation potential (Dou et al., 2007;
recombination of these processes can occur, as indicated by Erdmann Hakimi et al., 2021). These organic-rich shale rocks of the Cretaceous
and Horsfield (2006) in mixtures of terrigenous and marine organic Al-Gayger and Galhak formations were deposited in a lacustrine depo­
matter from the Norwegian North Sea. In the oil and gas industry, it is sitional setting under suboxic to relatively oxic conditions (Dou et al.,
important to understand the various geochemical processes, basin 2007; Hakimi et al., 2021). Although previous studies by Dou et al.
modelling history, and source rock characteristics in order to retrieve (2007) and Hakimi et al. (2021) have discussed the organic matter
the source of gas and oil in the potential resources and to understand the characteristics of Cretaceous organic-rich clays in the Melut Basin, there
gas-oil generation and storage. In turn, accurate analysis of gas or oil has been a general lack of work on the organic-rich shales in the Lower
sources can be achieved through volumetric calculations, such as Cretaceous successions, particularly those of the Al Renk Shale Forma­
defining free-gas and adsorbed-gas (Radwan et al., 2022). tion. The global shale gas revolution is knocking on the Sudanese gas
The impact of temperature, paleo heat flow, thermal maturity, and market’s doors, capturing the attention of scientists, technologists, and
depositional environment on organic matter (OM) is one of the impor­ policymakers. Sudan’s strategic plans for the gas and oil sectors rely on
tant controlling factors in gas-oil generation. The tectonic system plays developing its shale gas resources. For Sudanese shale gas exploration,
an important role in the distribution of temperatures across the sedi­ investigations into the origin and formation mechanism of highly
mentary basins by forming structural lows and highs with different mature shale gas and the upper limit of maturation for gas generation
formation temperatures. Therefore, the distribution of temperatures due from kerogens are more critical and can lead to the exploration of new
to tectonic style or geological history together with geochemical and resources for energy. As a result, more evaluation criteria are required to
petrological characteristics are all collaborated into basin modelling and assess the shale gas generation potential in different Sudanese sedi­
used to give information for both conventional and unconventional mentary basins. The potential generation of thermogenic gas in the
resource exploitation. Melut Basin shales has attracted our attention to perform this research to
This study focuses on the oil-producing Melut Basin, which is located better understand the generation and primary source of mature ther­
in Sudan’s southern region (Fig. 1A). The Melut basin is one of the mogenic gas and residual gas generation potential of the Lower

Fig. 1. (A) Regional tectonic map of western and central African rifted basins (modified from Makeen et al., 2016), and (B) The Melut Basin, Sudan (modified after
Dou et al., 2007).

2
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

Cretaceous shales, which have not been tracked in a proper way prior to contain horst, tilted, and downward blocks (Fig. 2) that were initially
this research. Hence, organically rich shale horizons within the Lower created in the Early Cretaceous and further developed in the Cenozoic
Cretaceous Al Renk Shale Formation have been studied in this research (McHargue et al., 1992).
in the context of conventional petroleum resource investigation. As evidenced by geophysical and borehole data, the Melut Basin is
This study seeks to use a variety of geochemical techniques to filled with 11-km-thick sedimentary sequence that goes from the Lower
investigate the characteristics of organic matter in the Al Renk shales Cretaceous to the Quaternary (Dou et al., 2007), which is punctuated by
and the critical factors of the high thermal maturation that significantly several unconformities of different magnitudes and ages (Fig. 3). The
modifies the original organic matter and their relation to the potential Melut Basin is characterized by four significant erosion events, including
for thermogenic gas generation. Hence, the main objective of this Miocene, Eocene, Late Cretaceous, and Early Cretaceous, with eroded
research is to deepen our understanding of the richness of organic matter sections of 600, 200, 200, and 200 m, respectively (e.g., Dow, 1977; Katz
content, kerogen types, thermal alteration, petroleum generation po­ et al., 1988; Dou et al., 2007; Eisawi and Schrank, 2008; Mohamed et al.,
tential, as well as the depositional environment of these shale sediments 2016).
within the Al Renk Formation. Furthermore, we employed basin The Cretaceous succession in this basin includes four formations;
modelling approaches to simulate the generation and expulsion of pe­ from base to top, these are the Al-Gayger, Al Renk, Galhak, and Melut
troleum from the Al Renk shale source rock over time. This study could (Fig. 3). The older Cretaceous succession is primarily dominated by non-
also be used to advance the knowledge regarding the gas produced by marine clastics, such as lacustrine deposits (Dou et al., 2007), and is
the deep oil-source rocks and develop future strategy for exploring a new underlain by Precambrian rocks via an unconformity (Fig. 3). The
petroleum resource in the basin. Al-Gayger Formation is composed of basal fluvial–deltaic sandstone
with some shale horizons, followed by lacustrine shale (Fig. 3). It is
2. Geological background conformably overlain by the Lower Cretaceous Al Renk Shale Formation
(Fig. 3). The Al Renk Formation comprises mainly shale sediments
In Sudan, there are several rift systems in onshore regions (Fig. 1A) (Fig. 3), and is considered as a set of deep lacustrine sediments (Zhao
that are formed as a result of structural actions in the East African, et al., 2020). It is 1000 m thick in the basin’s deepest region and
Central, and Western Rift systems during the Mesozoic Era (i.e., Creta­ 200–500 m thick at the basin’s structure heights (Mohamed et al., 2016;
ceous ages), as reported in previous work by Browne et al. (1985) and Zhao et al., 2020). This formation is topped by the Upper Cretaceous
Schull (1988). Galhak and Melut formations (Fig. 3). The lithofacies of the Galhak
The Melut Basin is located in the southeast to south of Sudan, near Formation is similar to the overlain Al-Gayger Formation and consist
the Ethiopia and Kenya borders (Fig. 1A). It is an intercontinental rift primarily of sandstones with interlayered organic-rich shales and clay­
basin formed during the Mesozoic-Cenozoic (Genik, 1993), with a length stones (Fig. 3). A Cenomanian–Turonian border distinguishes the Gal­
of over 310 km and a maximum width of 100 km (Dou et al., 2007). It hak Formation from the Al-Gayger shales (Dou et al., 2007).
contains several sub-basin rift systems formed during the Early Creta­ The organic-rich shale and claystone sediments within the Al-Gayger
ceous (Guiraud and Maurin, 1992), namely, the southeastern, central, and Galhak formations represent important petroleum resources in the
western, eastern, and northern sub-basins (Fig. 1B). These sub-basins Melut Basin, South Sudan. Therefore, investigating the characteristics of

Fig. 2. Map view of major faults in the Melut Basin, including several cross-sections orientated NE-SW, showing the structural setting of hydrocarbons across
the Basin.

3
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

Fig. 3. Generalized stratigraphic column of the Cretaceous–Quaternary sequences of the Melut Basin, including the petroleum system elements (modified after Zhao
et al., 2020).

their organic matter inputs and evaluating the petroleum generation (Fig. 3). The Agor Formation, which was deposited during the Quater­
potential is necessary to understand their potential for conventional and nary, is comprised primarily of sandstones (Fig. 3), and unconformably
unconventional petroleum resources. These organic-rich shale sedi­ overlies the Late Miocene–Pliocene Daga Formation.
ments contain huge amounts of marine plankton and bacteria and minor
amounts of terrestrial plants that are accumulated under a reducing and 3. Datasets and methods
fresh lacustrine depositional environment (Dou et al., 2007; Hakimi
et al., 2021). 3.1. Samples and experimental section
The Melut Formation overlies the Galhak Formation conformably
and has a slightly similar lithology to that of the underlying Galhak Twenty seven cutting samples were taken from two exploratory well
Formation (Fig. 3). It is unconformably overlaid by a Tertiary mega­ sites (i.e., Adar-1 and Agordeed-1) in the Melut Basin (Eastern sub-
sequence, varying from the Paleocene to the Pliocene (Fig. 3). The basin), as depicted in Fig. 1B. These samples exhibit the Al Renk Shale
Tertiary megasequence includes seven formations, namely, the Daga, Formation’s organic-rich shale intervals (Fig. 3). Prior to analysis, the
Jimidi, Adar, Miadol, Samma, Lau, and Yabus formations, which consist cutting samples were rinsed with distilled cold water to eliminate im­
primarily of siltstones and sandstones with interlayered shales and purities such as borehole mud and some other drilling additives. The
claystones (Fig. 3). Most of the hydrocarbon exploration is confined to total organic carbon (TOC) analysis, programmed pyrolysis (Rock-Eval-
the sandstone reservoirs of the Paleocene Samma and Yabus formations II), bitumen extraction, gas chromatography (GC), and gas chromatog­
together with the Cretaceous Galhak and Melut sandstones, which are raphy–mass spectrometry (GC–MS) were employed to investigate the
regarded as the main petroleum reservoir rocks in the Melut Basin (Dou, organic geochemical properties of the cleaned samples.
2004). The overlying Adar Formation (Eocene) consists mainly of The total organic carbon (TOC) content of 100 mg of powdered shale
claystones and shales. This formation serves as an excellent regional seal was evaluated using a LECO CS-125 instrument, and the results were
for the hydrocarbon-bearing reservoir rocks (Cretaceous–Paleocene) primarily expressed as weight percentages (TOC wt. %). Additionally,

4
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

the 100 mg of powdered samples were subjected to programmed py­ utilizing ions such as m/z 191 and 217 mass fragmentograms (Table 2).
rolysis utilizing a Rock-Eval II instrument. The oven temperature was set Individual ion peak assignments in the aliphatic and hydrocarbon
between 300 and 600 ◦ C. Several parameters including free petroleum fractions were identified by matching the monitored ions’ retention
(S1, mg HC/g rock) and hydrocarbon production from kerogen cracking times and mass spectra with those from prior studies (Philp, 1985; Peters
(S2, mg HC/g rock) were determined. The S3 peak, which reflects the and Moldowan, 1993; Hakimi et al., 2012; Makeen et al., 2015).
quantity of CO2 generated by the pyrolysis of organic matter and is The vitrinite reflectance (% VRo) measurement was also done on
quantified in milligrams of CO2 per gram of rock (mg/g), was also twenty-nine (29) samples using polished block technique. The 29 sam­
produced during programmed pyrolysis. During the pyrolysis investi­ ples were collected from the Agordeed-1 well (Table 3), which included
gation, the maximum temperature (Tmax) of the S2 peak was also ob­ the Al Renk Shale Formation as a potential source rock. The whole shale
tained (Table 1). samples were made into polished blocks by mounting fine grades of
Other geochemical metrics, like the, hydrogen index (HI), oxygen crushed rock (2–3 mm in size) within Serrifix slow setting polyester resin
index (OI) and production index (PI), were determined using the TOC combined with a resin hardener, after which the samples were then
content and pyrolysis S1, S2, and S3 yields (Table 1), as outlined by allowed to harden. Mean vitrinite reflectance (% VRo) values were
Espitalie et al. (1977), Espitalie et al. (1985), and Peters and Cassa measured after calibration using a reflectance standard value of 0.20%
(1994). (glasses) and 5.10% (gemstones) under oil immersion, in a white plane-
Twenty shale samples were extracted using a mixture of dichloro­ polarized reflected light using a Zeiss microscope and a Leitz Orthoplan/
methane (DCM) and methanol (CH3OH) for 72 -hours. The recovered MPV photometric system.
bitumen from six shale rocks was further separated using liquid column
chromatography with silica gel yielding two hydrocarbon fractions,
aliphatic and aromatic, as well as polar (NSO) compounds. The aliphatic 3.2. PetroMod 1-D modelling dataset
fraction was further analysed using a Hewlett Packard 5890 gas chro­
matograph equipped with a flame ionizatıon detector. The temperature 3.2.1. Subsidence and burial history model of sedimentation frequency
of the furnace of the gas chromatograph was maintained at a rate of The basin modelling is important aspect in the oil and gas industry
1 ◦ C/min between 140 ◦ C and 211 ◦ C, and then kept at 300 ◦ C for 10 and it has a significant importance in petroleum system analysis, pore
min. pressure, and basin analysis (Mahdi et al., 2022; Radwan et al., 2020;
Two of these six samples had their aliphatic HC fractions analysed by Diab et al., 2022; Radwan, 2022). The basin modelling in this work used
GC–MS as well. The flame ionization detector was employed in the GC- the PetroMod 1-D modelling software (from Schlumberger; version 2010
MS experimental study on a HP–DB5–MS column with a distance of 30 SP1) to recreate the subsidence burial history through geological time
m, an internal diameter of 0.30 mm, and a film thickness of 0.25 m. The using geological data from a single well (Agordeed-1). The burial history
furnace was configured to maintain a temperature of 300 ◦ C for 25 min model was recreated using an unpublished compilation report that
at a rate of 3 ◦ C/min between 50 ◦ C and 180 ◦ C. Consequently, terpenes, included geological data inputs such as sedimentary rock types and
triterpenes, steranes, and diasteranes were thoroughly characterized by thickness (Table 3) as well as the age of sedimentary sequences based on
the stratigraphic data from Schull (1988), Kaska (1989), Mohamed et al.

Table 1
Geochemical results of the analysed organic-rich shale samples within the Early Cretaceous Al Renk Shale Formation from two exploration well locations (Adar-1 and
Agordeed-1) in the eastern sub-basin of the Melut Basin, South Sudan, including TOC content and programmed pyrolysis (Rock-Eval-II).
Well Depth Bitumen TOC Wt. Rock-Eval pyrolysis data
(m) (ppm) %
S1-HC (mg/ S2-HC (mg/ S3–CO2 (mg/ HI (mg/ OI (mg/ S2/S3 (mg/ PI (mg/ Tmax
g) g) g) g) g) g) g) (OC)

Adar-1 4267 990 1.22 0.67 0.99 1.65 81 135 0.60 0.40 440
4359 760 1.39 0.26 0.76 1.20 55 86 0.64 0.25 466
4389 1750 1.98 0.50 1.74 1.31 88 66 1.33 0.22 459
4450 610 0.87 0.17 0.61 0.89 70 102 0.69 0.22 –
4481 750 1.08 0.21 0.75 0.66 69 61 1.13 0.22 473
4511 1330 2.04 0.43 1.33 0.65 65 32 2.03 0.24 473
4545 500 0.66 0.13 0.50 0.75 76 114 0.67 0.21 442
Agordeed- 3520 620 0.97 0.17 0.62 0.35 64 36 1.78 0.21 448
1 3540 300 0.74 0.10 0.30 0.33 41 45 0.91 0.25 –
3600 1300 1.37 0.57 1.30 0.73 95 53 1.79 0.30 443
3620 1.36
3630 1860 1.48 0.74 1.86 0.46 126 31 4.06 0.28 443
3640 1.25
3650 1.26
3660 1370 1.22 0.49 1.37 0.41 112 34 3.29 0.26 443
3670 1.21
3685 1.17
3700 1250 1.28 0.60 1.25 0.93 98 73 1.34 0.32 445
3710 1950 1.79 0.89 1.95 0.81 109 45 2.42 0.31 444
3720 1.97
3730 2130 1.90 0.84 2.13 0.74 112 39 2.87 0.28 444
3740 1.87
3750 2020 1.97 0.69 2.03 1.46 103 74 1.39 0.25 446
3760 2000 2.02 0.96 2.00 1.66 99 82 1.21 0.32 445
3770 2610 2.06 0.76 2.62 1.71 127 83 1.53 0.23 445
3780 2010 2.14 0.71 2.01 0.98 94 46 2.04 0.26 450
3794 1530 1.75 0.69 1.52 0.53 87 30 2.90 0.25 450

TOC = Total organic carbon; S1-peak = Free contents of hydrocarbon(mg HC/g rock); S2-peak = Remaining hydrocarbon potential (mg HC/g rock); S3 peak =
Produced carbon dioxide (mg CO2/g rock); HI=S2 × 100/TOC(mg HC/g rock); OI=S3 × 100/TOC (mg CO2/g TOC); Tmax = Maximum temperature at peak of S2(oC);
PI= Production index [S1/(S1+S2)].

5
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

(1999), Eisawi and Schrank (2008) and Mohamed et al. (2016).

Biomarker ratios of the representative organic-rich shale rocks of the Early Cretaceous Al Renk Shale Formation from one exploration well locations (Agordeed-1) in the eastern sub-basin of the Melut Basin, South Sudan,
Moreover, unconformities (erosion events, and hiatuses) that

M30/

0.12

0.11
occurred throughout the evolution of sedimentary succession should be

C30
considered when constructing conceptual models (Hakimi and Abdul­
lah, 2015; Hadad et al., 2017; Makeen et al., 2016; Hakimi et al., 2018).

ββ/(ββ+αα)
In the Melut Basin, there were significant erosion events in the Miocene,
Eocene, Late Cretaceous, and Early Cretaceous, with erosional thick­

0.56

0.57
C29 nesses of 600, 200, 200, and 200 m, respectively (Dow, 1977; Katz et al.,
20 S/(20 S + 20 1988; Dou et al., 2007; Eisawi and Schrank, 2008; Mohamed et al.,

Pr =Pristane; Ph = Phytane; CPI= Carbon preference index (1): {2(C23 + C25 + C27 + C29)/(C22 + 2[C24 + C26 + C28] + C30)}; Waxiness degree (WI) = Σ (n-C21- n-C31)/Σ (n-C15- n-C20).
Biomarker indicators thermal maturity

2016). These approximate estimations of eroded sediment were inte­


grated into the basin modelling work (Table 3) and made by utilizing a
significant effect on modelled maturity.
0.47

0.48
C29

R)

3.2.2. Estimation of paleo-heat flow and geothermal gradient


The thermal gradient and heat flow (HF, mW/m2) histories are
C32 22 S/(22 S +

needed as essential input parameters for modelling the basin’s thermal


backdrop, followed by the maturation of source rock (Lachenbruch,
1970; Allen and Allen, 1990; He and Middleton, 2002; Li et al., 2010;
22 R)

0.57

0.61

Welte et al., 2012; Shalaby et al., 2011, 2013).


The HF is believed to have formed in the mantle because of tectonic
M30/C30 = C30 moretane/C30 hopane; HCR31/HC30 = C31 regular homohopane/C30 hopane, G/C30 = gammacerane index (gammacerane/C30 hopane).

evolution and the heat produced by crustal and regional hydrodynamic


38.10

35.66
C29

fluxes (Allen and Allen, 1990 and 2005; Hantschel and Kauerauf, 2009).
Regular steranes (%)

Therefore, the HF is a more critical indicator of the thermal evolution in


20.18

18.97

the basin because the paleo HF was inferred based on major tectonic
C28

activities, particularly the extensional tectonics of the rifting phase


(Lachenbruch, 1970; Allen and Allen, 1990). High paleo-heat flow is
C29/C30 = C29 norhopane/C30 hopane; Ts= (C27 18α (H)-22, 29, 30-trisnorneohopane); Tm= (C27 17α (H)-22, 29, 30-trisnorhopane).
41.72

45.37

incorporated into episodes during the rifting phase, and an exponential


C27

reduction in heat flow is considered during the pre- and post-rift phases
(Mckenzie, 1978).
The Melut Basin is an intercontinental rift basin and has multiple
steranes
C27/C29
Regular

rifting episodes (McHargue et al., 1992). There are three rifting phases
1.09

1.27

that formed the Melut Basin including Early Cretaceous rifting, Late
Cretaceous rifting and Eocene to Miocene rifting (McHargue et al.,
1992). These three rifting phases incorporated in the heat flow model,
Biomarker indicators of source organic matter and depositional environment conditions

and made by utilizing a significant effect on modelled maturity.


HCR31/
HC30

Despite the heat flow is being an important parameter to evaluate the


0.23

0.20

thermal geo-history; it is difficult to define it for the geological past


(Allen and Allen, 1990; Lachenbruch, 1970). Thus, the paleo-heat-flow
illustrating source organic matter, depositional environment condition and thermal maturity.

can be estimated by using thermal calibration data, like vitrinite organic


0.97

0.83
C29/
C30

matter reflectance (% VRo) (Waples, 1980; Sweeney and Burnham,


1990). To data, the VRo data were often utilized to determine paleo-heat
0.12

0.09
C30

flow (Hakimi et al., 2010; Shalaby et al., 2011, 2013; Hakimi and
G/

Ahmed, 2016; Hadad et al., 2017; Makeen et al., 2016). However, two
aspects are believed to affect vitrinite reflectance (% VRo) measure­
0.82

1.13
Ts/
Tm

ments: i.e., maximum burial temperature and duration (geological


exposure time) (Waples, 1980; Sweeney and Burnham, 1990).
2.06
2.02
1.54
1.93
1.93
1.48

In this study, several thermal maturity models were developed uti­


WI

lizing Sweeney and Burnham’s (1990; Easy % Ro) method for thermal
1.09
1.09
1.05
1.05
1.06
1.07

modelling. The sensitivity of thermal maturity models is demonstrated


CPI

by a good match between measured calibration data (i.e., temperature


and %VR) as shown in Table 3 and calculated models (Easy % Ro) of
0.10
0.10
0.17
0.09
0.08
0.08

Sweeney and Burnham (1990) in the PetroMod basin modelling tool.


Ph/
C18

3.2.3. Hydrocarbon generation modelling


0.25
0.27
0.26
0.17
0.16
0.17
C17
Pr/

The timing of hydrocarbon generation was also simulated and


modelled using the initial total organic carbon (TOC) and initial
hydrogen index (HI) data. However, the low TOC and HI values
2.32
2.46
1.46
1.91
1.96
1.97
Pr/
Ph

extracted from geochemical results (Table 1) are the present-day


(measured) data. The alteration of the original organic matter (OM)
and decreasing TOC and HI values (conversion of kerogen to HCs) is
Depth

3630
3660
3710
3750
3770
3794

attributed to the high thermal maturation of the Al Renk shale source


(m)

rock in the studied well, as highlighted in the next subsections.


Following the next biomarker results, the Al Renk shales were
Agordeed-

deposited in a lacustrine under suboxic to relatively oxic environment


Table 2

Wells

settings with a mixture of aquatic organic matter (e.g., algal and mi­
1

crobial) and land plant inputs. The presence of mixed aquatic and

6
M.H. Hakimi et al.
Table 3
Basin model input data used to reconstruct the burial and thermal history one well location (Agordeed-1) in the eastern sub-basin of the Melut Basin, South Sudan as shown in Fig. 1B.
Formation Deposition Erosion ages Erosion Thickness (m) Agordeed-1 Well Calibration data from AGORDEED-1 well Formation
ages (Ma) (Ma)

From To From To Top (m) Bottom (m) Thickness (m) Vitrinite reflectance (VRo) Corrected bottom hole
temperatures

Depth (m) Values (%VRo) Number Standard deviation (SD) Depth (m) Values (OC)

Agor 0 8 0 148 148 520 0.31 50 0.067 Miadol


Daga 8 11 11 16 600 148 481 333 620 0.37 55 0.005 Jimidi
Miadol 16 22 481 590 109 700 0.30 55 0.006
Jimidi 22 34 34 36 200 590 734 144 760 0.43 11 0.056 Adar
Adar 36 55 734 966 232 1000 0.39 53 0.001 780 65
Yabus 55 60 966 1361 395 1160 0.33 35 0.005 1220 75 Yabus
Samaa 60 65 65 70 200 1361 1479 118 1200 0.43 55 0.002
Melut 70 91 1479 2236 757 1350 0.43 10 0.042
Galhak 91 97 97 103 200 2236 3260 1024
Al Renk 103 135 3260 3794 534 1525 0.42 5 0.231 Melut
7

Total depth 3794 1650 0.45 10 0.038


2230 0.60 55 0.003
2325 0.72 55 0.003
2335 0.73 55 0.003 Galhak
2515 0.76 55 0.001
2650 0.84 55 0.001 2600 115
2760 0.87 55 0.001
2835 0.96 55 0.003 2900 125
3065 0.97 55 0.005 3000 130
3215 0.99 55 0.001 3100 132
3445 1.10 56 0.002 3400 142 Al Renk
3520 1.00 55 0.003
3620 1.15 53 0.004
3670 1.20 55 0.005

Marine and Petroleum Geology 146 (2022) 105926


3700 1.30 55 0.001
3710 1.36 34 0.031
3740 1.36 20 0.020
3750 1.42 24 0.011
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

terrigenous organic matter indicates a mixture of Types II and III gaseous phase from Type II source rock.
kerogen. The presence of hydrogen-rich Type II in the Al Renk Shales is
also confirmed by high hydrogen index (HI) value of more than 400 mg 4. Results and interpretations
HC/g TOC as reported by recent work of Zhao et al. (2020). In this re­
gard, the TOC content and HI values of up to 3% and 590 mg HC/g TOC 4.1. TOC content and bitumen extraction
of the low mature Al Renk shale samples (early-mature of oil generation
window) in the Melut Basin (Zhao et al., 2020) were used as initial The TOC content is often used to evaluate the organic matter content
geochemical input data in the current modelling. and the organic matter’s ability to produce petroleum during maturity
In the used petroleum generation modelling, the kinetic model is (Peters, 1986; Jarvie, 1991; Peters and Cassa, 1994; Sohail et al., 2022).
another important input needed to estimate the kerogen conversion Th measured total organic carbon (TOC) content of all 27 Al Renk
ratio (TR) in terms of oil and gas generation. In this case, the kinetic data shale samples from the examined wells, is summarized in Table 1. The Al
from Burnham and Sweeney (1989)-T2 for Type II kerogen were used as Renk shale samples possess moderate to high levels of organic matter, as
the default to numerically simulate the petroleum generation from the indicated by TOC values ranging from 0.66 to 2.14 wt % (Table 1).
Al Renk Formation. The reaction aspects of the Type II kinetic were According to Bissada (1982) and Katz and Lin (2014), most of the
integrated with the thermal maturity and burial history to record the studied samples (n = 23) had a TOC level of more than 1% (1.08–2.14 wt
timing of the primary kerogen-oil cracking and retained oil cracking into %), suggesting that these shale sediments contain a substantial quantity

Fig. 4. Geochemical correlation between total organic carbon (TOC) content and bitumen in the analysed Al Renk shale samples from studied well locations in the
Melut Basin, showing (A) Good positive correction TOC and bitumen, and (B) Fair to very good petroleum generation potential.

8
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

of organic matter and are therefore regarded as favourable source rocks. are rich in bitumen with values > 1000 ppm, while other samples have
The TOC content of the other four shale samples varies between 0.66 and relatively lower EOM content in the range of 300–990 ppm (Table 1).
0.97 wt percent (Table 1), implying limited source rock potential. However, with a relative R2 value of 0.86 (Fig. 4A), there is a good
The extractable organic matter (EOM) content of the twenty ana­ relationship between EOM content and TOCs of the analysed samples,
lysed shale samples shows moderate to high values, ranging between implying that these shales are probable source rocks with fair to very
300 ppm and 2610 ppm (Table 1). Most of the analysed samples (n = 13) good potential for petroleum generation (Fig. 4B).

Fig. 5. Gas chromatograms show n-alkane and acyclic isoprenoid (e.g., pristane and phytane) distributions of the aliphatic hydrocarbon fraction in the analysed
shale samples.

9
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

4.2. Geochemistry of programmed pyrolysis (Rock-Eval-II)) Espitalie et al., 1985; Peters, 1986; Peters and Cassa, 1994; Waples,
1994).
Table 1 shows the geochemical parameters of twenty samples of the The S1, S2, and S3 yields of the analysed shale samples are 0.10–0.96,
studied Al Renk shales that were subjected to programmed pyrolysis 0.30–2.62, and 0.33–1.71 mg HC/g rock, respectively (Table 1). The S2
analysis. These pyrolysis parameters, coupled with TOC content, were and S3 yields are compatible with TOC content and are used to calculate
utilized to evaluate the amount and quality of organic matter in the Al HI and OI values according to Peters and Cassa (1994). The bulk kerogen
Renk shales, along with their thermal maturity (Espitalie et al., 1977; types are identified using the pyrolysis HI and OI results (e.g., Espitalie

Fig. 6. m/z 191 and 217 mass fragmentograms of the saturated hydrocarbon fraction for representative two shale samples of the Al Renk Shale Formation.

10
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

et al., 1985; Peters and Cassa, 1994; Mukhopadhyay et al., 1995). and homohopanes of C31–C35 (Fig. 6A). Because the abundance of C30
For the shale samples examined in this research, the estiamted HI hopanes in the extracted rock samples is higher than that of C29
and OI values are 41–127 mg HC/g TOC and 30–135 mg CO2/g TOC, norhopanes (Fig. 6A), the values (0.83–0.97) of C29/C30 hopane ratio are
respectively (Table 1). The majority of the samples (n = 14) have low HI <1, as shown in Table 2. Clay-rich facies are usually identified by a low
values < 100 mg HC/g TOC (41–99 mg HC/g TOC), while six samples C29/C30 hopane of <1 (Gürgey, 1999). C30-hopane is also more preva­
have relatively higher HI values > 100 mg HC/g TOC, ranging from 103 lent than moretane, gammacerane, and C31 22 R-homohopanes
to 127 mg HC/g TOC. In terms of OI values, the majority of the analysed (Fig. 6A), resulting in low C31 R/C30H, G/C30, and C30 M/C30H ratios of
samples (n = 21) have low OI values varying from 30 to 86 CO2/g TOC 0.20–0.23, 0.09–0.12, and 0.11–0.15, respectively (Table 2). The dis­
whilst the other three samples have relatively high OI values, ranging tribution of the homohopane series of extracted samples is observed on
from 102 to 135 CO2/g TOC. m/z 191 fragmentograms and shows that the C31 hopane dominates the
Tmax is the temperature in the S2 peak at which maximum hydro­ homohopane distribution with decreasing abundance as carbon number
carbon yield occurs. Tmax values with S2 yield close or higher than 1 mg/ increases (Fig. 6A). The biomarker maturity ratio of C32 22 S/(22 R + 22
g rock provide the most reliable results (Jarvie et al., 2001; Ahmed et al., S) ranges from 0.56 to 0.61 (Table 2). Furthermore, the m/z 191 mass
2020; Katz and Lin, 2021). In this case, the Tmax is chosen in this context fragmentograms show substantial amounts of 17 (H)- trisnorhopane
for most of the samples with S2 values between 0.99 and 2.62 mg HC/g (Tm) and 18 (H)-trisnorneohopane (Ts), with Ts/Tm values varying
TOC, (Table 1). Hence, most of the samples have Tmax results in the from 0.82 to 1.13 (Table 2).
range of 440–473 ◦ C (Table 1). Only six samples with lowest S2 values Additionally, sterane and diasterane are important biomarkers
(<1 mg/g rock) have Tmax values between 442 ◦ C and 473 ◦ C (Table 1). discovered in m/z 217 mass fragmentograms and identified in the
The production index (PI) values of shale samples were also obtained saturated HC fraction of extracted rocks (Fig. 6B). Mass fragmentograms
via Rock-Eval pyrolysis and ranged from 0.21 to 0.40 (Table 1). Many of of the m/z 217 ion show the relative abundance of standard regular
the analysed samples (n = 26) have PI values between 0.21 and 0.32, steranes of C27 to C29, which are distinguished by substantial abun­
with the highest recorded PI value of 0.40 in the last sample (Table 1). dances of C27 and C29 regular steranes when compared to C28 regular
steranes (Fig. 6B). The proportional amounts of regular steranes C27,
C28, and C29 are estimated to be 41.72–45.37%, 18.97–20.18%, and
4.3. Biomarker fingerprint of organic matter 35.66–38.10%, respectively (Table 3). Other sterane ratios of C29 ster­
ane, such as ββ/(ββ+αα), 20 S/(20 S + 20 R), and C27/C29 regular
The biomarker distributions of normal alkane, isoprenoid, sterane, sterane, were calculated and found to be between 0.47 and 0.48,
and terpane of the saturated HC fraction in the extracted shale samples, 0.56–0.57, and 1.09–1.27, respectively (Table 2).
as shown in gas chromatograms (Fig. 5) and mass fragmentograms of m/
z 191 and 217 ions (Fig. 6) were examined in this study.
In gas chromatograms, the normal alkanes and isoprenoids have a 4.4. Thermal history evolution and relation to maturity history of the
bimodal distribution, with the entire range of normal alkane compounds source rock
falling between C12 and C38 (Fig. 5). Short (C15–C20) to middle (C21–C25)
chain n-alkanes dominate the n-alkane patterns, with waxy alkanes The thermal evolution of the sedimentary basins affects the timing of
(>C25) also present (Fig. 5). The carbon preference index (CPI) values source rock maturation and therefore the petroleum generation over
were calculated using the following equation: CPI = [2(C23 + C25 + C27 geological history (He and Middleton, 2002; Li et al., 2010; Shalaby
+ C29)/(C22 + 2(C24 + C26 + C28] + C30)]; and range from 1.05 to 1.09 et al., 2011, 2013; Abeed et al., 2013; Hakimi and Ahmed, 2016; Hadad
(Table 2). The waxiness degree (WI) was also calculated using the et al., 2017; Makeen et al., 2016; Mohamed et al., 2016; Botor and
following equation: WI = Σ (n-C21-n-C31)/Σ (n-C15-n-C20) and shows that Bábek, 2019).
the waxiness of the analysed shale samples ranges from 1.48 to 2.06 Basinal thermal history is generally established based on sedimen­
(Table 2). These CPI and WI parameters can be employed to offer in­ tation, erosion history, burial temperature gradients, and paleo-heat
formation on organic matter inputs (e.g., Connan and Cassou, 1980; flow (HF) developments (Lachenbruch, 1970; Allen and Allen, 1990).
Johns, 1986; Zhang et al., 2015). However, the HF, rather than other factors (i.e., temperature gradients),
The chromatograms also show the presence of acyclic isoprenoids is a more accurate indicator of the basin’s thermal evolution since the
such as pristane (Pr) and phytane (Ph) in all extracted studied rocks paleo HF was inferred based on subsidence events, with emphasis on
(Fig. 5). For most of the samples, the Pr predominated over the Ph based major tectonic activities (Lachenbruch, 1970).
on these isoprenoids, resulting in a Pr/Ph ratio of 1.46–2.46 (Table 2). In this study, heat flow values in the studied basin were estimated
Two extracted shale samples have Pr/Ph ratios >2 (2.32–2.46), while from quantified %VRo and corrected BHT (available from the explora­
the other samples have Pr/Ph ratios between 1.46 and 1.97, as shown in tion well Agordeed-1; Table 3) using a 1D-basin simulation approach.
Table 2. Additionally, these isoprenoids are compared with n-alkane Herein, multi-scenario heat flow models are used in conjunction with
concentrations (C17–C18), resulting in Pr/n-C17 and Ph/n-C18 ratios of erosion events throughout the Cretaceous and Tertiary periods (Table 3)
0.16–0.27 and 0.08–0.17, respectively (Table 2). As a proxy for organic to reach a possible thermal history range. The observed reasonable fit of
matter input and redox conditions, such isoprenoid biomarkers and their modelled and measured vitrinite reflectance profiles is aided by incor­
ratios may provide insight into the paleodepositional environment (e.g. porating into the basin history of up to 1200 m of sediments that have
Didyk et al., 1978; Ten Haven et al., 1987; Chandra et al., 1994; Tserolas been eroded in the Cretaceous and Tertiary time (Eocene to Miocene)
et al., 2019). (Table 3; Fig. 7) as described by earlier studies (Dou et al., 2007; Eisawi
Other important biomarkers, such as hopanoids and steroids, are and Schrank, 2008; Mohamed et al., 2016) and further demonstrated by
detected in the saturated HC fraction of both analysed samples and are the profiles of burial history (Fig. 8). The best fit between measured data
often studied using GC-MS by examining the m/z 191 and m/z 217 ions, of % VRo and BHT in the studied well and model of Sweeney and
respectively (Fig. 6A and B). In the two extracted shale samples, mass Burnham’s maturation (EASY % Ro) and temperature gradient models
fragmentograms of the m/z 191 ion reveal a large amount of hopanes (Fig. 8A) was achieved by assigning paleo-heat flow values of 50–80
(Fig. 6A) with the majority of which are C30 hopanes, C29 norhopanes, mW/m2 (Fig. 8B). The adopted HF model also reveals that paleo-heat

11
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

Fig. 7. Left: burial history for Agordeed-1 well and Right: blue line is shown exclusively thickness of all formations, including erosion events in the Melut Basin. (For
interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

flow values are generally higher than those of today (Fig. 8B). The Accordingly, the stretching of the studied Melut Basin has a weaker
Cretaceous (Early-Late) and Tertiary (Eocene-Miocene) have high blanketing effect than the subsidence rate of the basin, as demonstrated
paleo-heat flow values varying between 80 and 70 mW/m2 (Fig. 8B). by the stretching factor of less than 2 for the lithosphere during the three
This is most likely due to the rifting systems that existed in Sudan’s rifting periods.
interior basins during these times (e.g. McHargue et al., 1992). This The maturity model, which was developed using Sweeney and
estimated heat flow model (50–80 mW/m2; Fig. 8B) for the current basin Burnham’s (1990) EASY % Ro maturation model and incorporated with
modelling is consistent with a number of burial geo-historical and the burial history in the studied well, shows that the Al Renk Shale
thermal history models tested and noted in the Melut rift Basin Formation (source rock) had a burial depth of more than 4700 m and
(Mohamed et al., 2016). reached the specific interest area of the oil and gas maturity windows
However, the Melut basin was largely formed during the Early during the Mid Cretaceous-Early Eocene (age 83-51 Ma) and Late
Cretaceous rifting phase (McHargue et al., 1992), as evidenced by the Eocene (age 40 Ma) to present day, respectively (Fig. 9A). The incor­
high paleo-heat flow value of 80 mW/m2 (Fig. 8B). With relatively low poration of burial temperatures in the studied well (Fig. 8) and the high
paleo-heat flow (70 mW/m2; Fig. 8B), the second and third phases of paleo-heat flow during the Cretaceous and Tertiary rift systems (Fig. 7B)
rifting during the Late Cretaceous and Tertiary contributed less. are most likely responsible for these differences in thermal maturity
In addition, we are also particularly interested in the blanketing ef­ (Fig. 9A).
fect in the studied rift-basin and study how the stretching of the crust According to the thermal history model, the Al Renk Shale source
and mantle might expect the blanketing effect to be noticeable (Wangen, rock entered the early oil-generation window of 0.55–0.70 Easy % Ro in
1995; Peshkov et al., 2021). The effect of thermal blanketing increases the Mid Cretaceous to Late Cretaceous (age 90-77 Ma). The peak oil
with the duration of the sedimentation periods, while the rapid sedi­ window of VR values between 0.70 and 1.00 Easy % Ro occurred in the
mentation in a short geological period lead to a minor thermal transient Late Cretaceous to the Early Eocene (age 77-51 Ma) (Fig. 9B). The
due to the short period (Wangen, 1995). thermal history model also shows that the Al Renk source rock reached
However, the blanketing leads to counteract the increased heat flow the late oil window (1.00–1.30 easy % Ro) during the Early to Late
that is caused by the stretching of the crust (lithosphere) and the up­ Eocene (age 51-40 Ma) (Fig. 9B). The base Al Renk Shale Formation in
welling of hot mantle (asthenosphere), as indicated in previous work by the studied well then reached gas maturity window around 40 Ma and
Wangen (1995). has continued to the present day (Fig. 9A), corresponding to an Easy %
Wangen (1995) has been broadly quoted to discuss blanketing effect Ro of greater than 1.30% (Fig. 9B).
based on the numerical calculations of Mackenzie (1978) and concluded
that the stretching factor of the lithosphere of greater than 2 indicates 5. Discussion
that the stretching of the basin has a stronger blanketing effect than the
subsidence of the basin. 5.1. Origin of organic matter input and the sedimentary environmental
In this study, a one-dimensional model was also carried out based on conditions
numerical calculations of Mackenzie (1978) stretching model, which is
implemented as the “Crustal Model” in the Petromod1D. This crustal The normal alkanes, isoprenoids, terpanes, and steranes molecular
model was used to obtain the stretching factors of the crust (lithosphere) organic biomarkers, as well as their ratios and parameters, were utilized
and mantle (asthenosphere) that were obtained in each rift phase in the to assess organic matter inputs in the source rocks, as well as their
Melut Basin, and then predict the necessary conditions for blanketing probable origin and paleoenvironmental conditions (Peters and Mol­
and evolution of the thermal transients during sediment deposition. In dowan, 1993; Peters et al., 2005; Hakimi and Abdullah, 2013; Makeen
this regard, the stretching factors of the crust and mantle during the et al., 2015; Hakimi et al., 2020).
three rifting phases in the basin were obtained and found to be between The characteristics of the organic matter, such as source of organic
1.8 and 3.20 for the Early Cretaceous rift phase (80 mW/m2; Figs. 8B), matter input, lithology, and sedimentary environmental conditions of
1.5 and 2.50 for the Late Cretaceous rift phase (70 mW/m2; Figs. 8B), 1.6 the Al Renk Shale source rock, were assessed in this study based on lipid
and 2.40 for the Eocene-Miocene rift phase (70 mW/m2; Fig. 8B). biomarker compositions in the saturated HC fraction and their ratios and

12
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

parameters described in the previous paragraphs.

Fig. 8. Basin models for the studied well, showing (A) Optimized fit of calibrated data i.e. bottom-hole temperatures (PHT) and measured vitrinite reflectance (%VR)] and models of EASY %Ro maturity and geothermal
Normal alkanes exhibit the highest abundance of short-to middle-
chain n-alkanes (n-C12–n-C23) in many of the analysed extracted rock
samples, with significant amounts of waxy alkanes (+n-C23), indicating
that the lacustrine sediments were home to mixed aquatic species along
with a volume of terrigenous OM, as proposed by Ekpo et al. (2005) and
Makeen et al. (2015). The CPI parameter and the waxiness degree both
support a mixed organic matter input (Fig. 10).
The isoprenoid ratios of Pr/Ph, Pr/n-C17, and Ph/n-C18 are widely
employed to infer organic matter inputs and their depositional condi­
tions (Didyk et al., 1978; Ten Haven et al., 1987; Chandra et al., 1994;
Tserolas et al., 2019). The ratio of Pr/Ph of the analysed shale samples
(1.46–2.46) implies that these shale sediments were most likely depos­
ited in typically suboxic to moderately oxic depositional conditions.
Pr/Ph ratios (2.32–2.46) > 2 indicate moderately oxic environmental
settings, while Pr/Ph ratios between 1.46 and 1.97 specify more
reducing (suboxic) settings during deposition. The Pr/n-C17 and
Ph/n-C18 ratios further indicate the large contribution of mixed organic
matter (OM) that was most likely deposited under suboxic environ­
mental conditions (Fig. 11A).
Additionally, the ratios of Pr/Ph and Pr/n-C17 can be combined to
distinguish between marine, lacustrine and terrestrial OM inputs
(Hughes et al., 1995; Hakimi et al., 2012; Alias et al., 2012; Sarki Yan­
doka et al., 2015). Marine OM has Pr/Ph ratios <1.5 and Pr/n-C17 ratios
<0.5 (Hadad et al., 2021; Barham et al., 2021), increasing contribution
of terrestrial OM source leads to higher ratios (Radke et al., 1980). The
relationship between the Pr/Ph and Pr/n-C17 ratios in this study suggests
a large contribution of a lacustrine OM inputs (Fig. 11B). This inter­
pretation is supported by the C31 homohopanes distribution, which
the studied well (Agordeed-1), and (B) The estimated heat flow range that achieved from calibration data (%VR and BHT).

shows that the C30 hopane predominates over the C31R homohopane,
with a low C31 R/C30 hopane ratio of 0.20–0.23 (Table 2), implying a
lacustrine setting for the corresponding source rocks, as low values <
0.25 indicate lacustrine depositional environments (Makeen et al.,
2015). Furthermore, the combination of Pr/Ph and C31 R/C30H ratios is
a strong indicator of a lacustrine origin (Fig. 11C). The presence of minor
amounts of gammacerane in the analysed shale samples (Fig. 6A), with
low values of the gammacerane index (Table 2), further confirms a fresh
lacustrine setting.
The C27–C29 steranes distribution and their relative amounts are also
used to emphasize the importance of organic matter derived from the
aquatic region (Huang and Meinschein, 1979; Waples and Machihara,
1991; Peters et al., 2005). The predominant abundances of C27
compared with C28 and C29 regular steranes (Fig. 6B) indicate that the
analysed shale samples received high amounts of aquatic organic matter
(i.e., algal constituents) during deposition (e.g., Schwark and Empt,
2006). The proportions of the C27 to C29 regular steranes were plotted on
a ternary diagram developed by Huang and Meinschein (1979) and
shows a mixture of planktonic/algae and terrestrial plants (Fig. 12A).
Furthermore, the high aquatic organic matter contributions are consis­
tent with the high C27/C28 regular sterane ratios of more than 1 (Table 2;
Fig. 12B).

5.2. Thermal maturation of organic matter

In this study, the thermal alteration of organic matter within the


studied Al Renk shale samples from studied wells was investigated.
Numerous optical and geochemical indicators, including vitrinite
reflectance measurements, Rock-Eval Tmax and PI data, as well as spe­
cific biomarker maturity indicators, were employed to assess the ther­
mal maturation of organic matter in the examined Al Renk shale samples
(Tables 1–3).
The most precise indication of maturity is vitrinite reflectance (%
VRo) evaluated under a reflected light microscope, which offers vital
information about the stage of organic matter maturation and the pro­
gression of petroleum generation potential (Sweeney and Burnham,
1990; Waples, 1994). In this respect, the analysed Al Renk shale samples

13
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

Fig. 9. Burial overlap with thermal maturity and thermal gradient histories (colored areas) cross all rock units (left) and red and blue lines are shown exclusively for
computed vitrinite reflectance (right) of the base Al Renk Shale Formation in the studied Agordeed-1 Well. (For interpretation of the references to color in this figure
legend, the reader is referred to the Web version of this article.)

Fig. 10. Geochemical biomarker results of the CPI and waxiness degree, indicating that the analysed Al Renk shales contain a mixture organic matter during their
sedimentation time.

from the Agordeed-1 Well attained a high degree of thermal maturity, The state of thermal maturity is further supported by the PI values
which is congruent with late-mature oil and wet gas generation window obtained via the Rock Eval pyrolysis. Low PI value of <0.10 indicates
as revealed by high VRo varying from 1.10 to 1.42% VRo (Table 3). immature organic matter, whereas a PI value between 0.10 and 0.40
The Tmax, based on the S2 measurement values, can also be used to suggests mature source rock (Hazra et al., 2018; Peters, 1986). In
give information about organic maturity following several scientists agreement with the previously obtained VRo and Tmax maturity data, the
such as (Tissot and Welte, 1984; Peters, 1986; Tissot et al., 1987; Bor­ calculated PI values of more than 0.10 (0.21–0.40) as shown in Table 1,
denave, 1993). The Rock-Eval pyrolysis Tmax results for the analysed Al confirm that most of the studied shale samples are thermally mature
Renk samples in the studied wells is varying from 440 to 473 ◦ C source rocks and point to the attainment of the main petroleum gener­
(Table 1), indicating peak of oil generation window to gas generation ation phase (Fig. 14a).
window (Fig. 13B). However, the thermal maturity stages from the Tmax Further, multiple biomarker maturity ratios of the distribution of
values for the analysed samples in the Agordeed-1 well is slightly steranes and hopanes in the saturated hydrocarbon fraction of the
different than those obtained from the VRo measurements and this is analysed extracted rock samples can also be employed to assess the
probably attributed to the variation of the S2 measurements (Table 1). organic matter maturity (Seifert and Moldowan, 1986; Mackenzie et al.,

14
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

Fig. 11. Geochemical biomarker results of the analysed Al Renk shale samples from Agordeed-1 Well in the Melut Basin: showing (a) Pristane/phytane versus
pristane/n-C17, and (b) Pristane/n-C17 versus phytane/n-C18 (from Shanmugam, 1985) and (c) Pristane/phytane versus C31 regular homohopane/C30hopane
(HCR31/HC30), indicating that the Al Renk shale sediments contain a mixture organic matter and were deposited in a lacustrine environment under sub­
oxic conditions.

15
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

Fig. 12. (A) Ternary diagram of regular steranes (C27–C29)in the saturated HC fraction, and (B) cross-plot of pristane/phytane ratio versus C27/C29 regular steranes
ratio, of the representative two shale samples of the Al Renk Shale Formation from Agordeed-1 Well in the Melut Basin, indicating the relationship between sterane
compositions, concerning organic matter input (modified after Huang and Meinschein, 1979).

16
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

Fig. 13. Geochemical correlations between hydrogen index (HI), oxygen index (OI) and Tmax. The analysed shale samples are dominated by Type III kerogen.

1980). For example, the C32 22 S/(22 R + 22 S) homohopane ratio is conversion in the oil-source rock system within the Lower Cretaceous Al
commonly used to determine maturity (Seifert and Moldowan, 1986). In Renk Shale Formation based on the organic geochemical characteristics
this regard, the immature zone for the source rock is represented by a together with basin modelling results. The aforementioned integrated
C32 ratio of less than 0.50. In contrast, early to moderate mature source techniques results helped to characterize the current organic matter
rocks reach C32 ratios between 0.50 and 0.57, and an equilibrium point (kerogen) in the analysed Al Renk shales from exploration wells and
of more than 0.57 indicates the peak-oil window and higher maturity offer information on the high level of thermal maturity of the Cretaceous
stages (Seifert and Moldowan, 1986; Abdula, 2015). The C32 hopane and Tertiary rift systems in the basin, which resulted in the trans­
ratio for the analysed Al Renk shales ranges between 0.56 and 0.61 formation of primary organic matter into secondary organic matter
(Table 2), which implies an equilibrium phase of oil and indicates that (secondary cracking of oil), which then transformed the oil into ther­
the oil generation window has already been outstretched or passed mogenic gas.
(Seifert and Moldowan, 1986). In this study, the current kerogen type in the organic-rich analysed
The thermal maturity stage of the analysed shale samples is also shale samples of the Al Renk Formation from the studied well were
demonstrated by the C29 sterane 20 S/(20 S + 20 R) and ββ/(ββ + αα) assessed utilizing qualitative data from the programmed pyrolysis re­
ratios (Seifert and Moldowan, 1978, 1981, 1986). These ratios are sults (Table 1). Traditionally, qualitative kerogen classification was
greater than 0.40 and 0.50, respectively, and suggest a source rock interpreted using bulk pyrolysis data such as hydrogen index (HI) and
reaching oil window maturity and higher maturity stages (Hakimi et al., oxygen index (OI) (Espitalie et al., 1985; Mukhopadhyay et al., 1995;
2012; Gharib et al., 2021). Peters and Cassa, 1994). The HI and OI values of the samples under
In this study, the C29 sterane 20 S/(20 S + 20 R) and ββ/(ββ + αα) consideration were calculated utilizing the compatibility of pyrolysis S2
ratios for the analysed shale samples reached high values in the range of and S3 yields with TOC content (Table 1). As a consequence, the data
0.47–0.48 and 0.56–0.57, respectively (Table 3), indicating high mature and findings of the Rock-Eval pyrolysis (i.e., the HI and OI) imply that
stage (Fig. 14b). The C32 hopane ratios, along with the high C29 sterane the organic facies of the examined shale samples is currently comparable
ββ/(ββ+αα) ratio, pointing to mature source rock during the peak stage and mostly comprised of Type III kerogen, as shown by a modified van
of oil window and higher maturity stages (Fig. 14c). Krevelen diagram of HI vs. OI (Fig. 13a). The modified van-Krevelen
In addition, the C30 moretane/C30 hopane (C30 M/C30H) ratio was diagram of HI and Tmax values also supports the predominance of
used to evaluate the thermal maturation (Mackenzie et al., 1980). Ac­ Type III kerogen that is plotted on the late mature of oil window and wet
cording to Mackenzie et al. (1980), this ratio decreases with increasing gas maturity zones (Fig. 13b).
thermal maturity, with values of less than 0.15. The high level of As a result of these kerogen type findings, the analysed shale in­
maturity of the analysed samples is also coherent with low values of the tervals in the Al Renk Shale Formation are currently expected to
C30 moretane/C30 hopane (CM30/C30) ratio between 0.10 and 0.12 generate mainly gas HCs. The geochemical correlations between TOC
(Table 2). content and the pyrolysis parameters of HI and S2/S3 led to this
conclusion and confirm the probability for gas generation potential from
the analysed Al Renk shale samples in the studied wells (Fig. 15).
5.3. Thermogenic gas generation (in-source conversion of oil to gas)
However, the dominant presence of the Type III kerogen and its gas
generation potential is inconsistent with the lipid biomarkers’
This section discusses the thermogenic gas generation from oil-to-gas

17
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

Fig. 14. Geochemical cross-plot of the maturity indicators of the analysed shale samples pyrolysis (A) Tmax versus production index (PI), (B) C29 sterane 20 S/(20 S +
20 R) versus ββ/(ββ + αα) and (C) C32 hopane 22 S/(22 S + 22 R) versus C29 sterane ββ/(ββ + αα), showing the high thermal maturation of the analysed shale
samples, corresponding to main stage of petroleum generation (i.e., peak oil-generation window and higher).

18
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

Fig. 15. Geochemical correlations between TOC content and pyrolysis data (the HI and S2/S3 parameters), implying that the analysed shales are generally good
source rocks for mainly gas generation.

distribution in the saturated HC of the analysed extracted rock shale As above discussed by maturity indicators, the organic matter in the
samples. analysed Al Renk shale samples has reached a significantly high thermal
According to source and paleoenvironmental biomarker indicators, maturation stage to generate thermogenic gases. In this case, the high
the analysed Al Renk shale sediments contain a mixture of organic thermal maturity of the gas generation window is a crucial factor that
matter, with high contributions from aquatic organic matter (i.e., algal drastically altering the original organic matter (OM) and decreasing HI
constituents), and a significant contribution from land plants (Figs. 11 values (conversion of kerogen to hydrocarbons) as well as the major
and 12). The high amounts of mixture aquatic organic matter and land prevalence of gas-prone Type III kerogen. The high thermal maturation
plants are consistent with a mixture Types II and III kerogen as the of the shales can be explained by the high paleo-heat flow caused by the
original organic matter input of the Al Renk shale sediments during rifting actions in the East African, Central, and Western Rift systems
deposition, and thus they should generate both oil and gas. The presence (Browne et al., 1985; Schull, 1988) that occurred during the Cretaceous
of dominated Type II kerogen into the Al Renk shales during deposition and Tertiary Periods (Fig. 7B), which caused various deformation events
time is also supported by the high hydrogen index values between 580 and had an impact on the conversion of the oil into thermogenic gas in
and 600 mg HC/g TOC for low maturity samples of the Al Renk shale the Al Renk Shale source rock system. Apart from the high thermal heat
samples in the Melut Basin’s shallower parts as reported from the recent flow from tectonic rifting events that caused various deformation events,
work of Zhao et al. (2020). thick overburden rocks from the Late Cretaceous to Quaternary periods

19
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

also had an impact on the conversion of the retained oil in the Al Renk

Fig. 16. Model of cumulative hydrocarbon generation (left) and Evolution of the transformation ratio and computed Vitrinite reflectance (EASY % Ro) with age (right) for the base Al Renk Shale Formation in the
Shale source rock system, resulting in oil being converted into thermo­
genic gas. This finding backs up what the burial and thermal maturity
models predicted (see section 4.4).
The Al Renk Shale source rock in the known deeper location of
Agordeed-1 Well reached a high thermal maturity in oil- and gas-
windows, as demonstrated by measured VRo values of more than
1.0% VRo, according to burial and thermal models (Fig. 9). As a result,
the Al Renk Shale Formation is most likely an effective source rock
mostly for thermogenic gas that is formed due to secondary cracking of
retained oil in the Al Renk Shale source rock system.
This study also modelled the generation and expulsion of petroleum
from the Al Renk Shale source rock as well as the formation of ther­
mogenic gas from the oil retained in the source rock system over
geological time (Fig. 16). Oil and gas windows determined from VRo
values (Easy % Ro) were used to estimate the conversion ratios of Type-
II kerogen to oil and then to gas (Fig. 16). The main oil generation phase
occurred in the Late Cretaceous between 78 and 68 Ma (Fig. 16A), with a
corresponding transformation ratio (TR) of 10%–50% (Fig. 16B). The oil
expulsion was simulated up to a TR ratio of 50%, since a TR ratio greater
than 50% implies that the source has generated its maximum quantity of
oil and is saturated, forcing the oil to be expelled (e.g., Hantschel and
Kauerauf, 2009). Thus, the Al Renk Shale source rock started to expel oil
during the Late Cretaceous -Late Eocene (68–40 Ma; Fig. 16A) with TR
varying from 50% to 93% (Fig. 16B). Oil expelled from the Al Renk Shale
source rock system has been trapped or cracked into gas since the Late
Eocene (Fig. 16A), with TR values averaging 93–99% and estimated VRo
values of more than 1.3 Easy % Ro (Fig. 16B). As a result, this study can
be used as a foundation for further gas exploration in the Melut Basin,
particularly in the deeper parts of the basin, where these shales of the Al
Renk Formation are anticipated to reach high maturity levels of gas
generation window.

6. Conclusions

The organic-rich shale intervals of the Lower Cretaceous Al Renk


Shale Formation from two exploratory wells (i.e., Adar-1 and Agordeed-
1) in the eastern sub-basin of the Melut Basin, South Sudan, were sub­
jected to extensive geochemical analyses and a 1D basin modelling
process, to determine the input of organic matter, paleodepositional
environment, thermal maturity stages, and the potential for thermo­
genic gas generation from the conversion of oil. The following are the
findings of this research:

- The organic matter content of the analysed Al Renk shale samples is


moderate to high, as indicated by TOC content values ranging from
0.66 to 2.14 wt %, making them likely to be good source rocks.
- Organic geochemical results, especially Rock-Eval pyrolysis indicate
that the analysed shale samples from the studied wells are currently
have hydrogen index (HI) pyrolysis values ranging from 41 to 127
mg HC/g TOC and fall along the Type III kerogen pathway.
- Biomarkers show that lacustrine sediments contain a mixture of
aquatic organic matter (i.e. algae) and terrigenous land plant input.
Such organic matter inputs are mixed Type II and Type III kerogens.
This contradicts the low current Rock Eval HI of less than 150 mg
HC/g TOC (i.e., mainly Type III kerogen).
- The maturity-related indicators show that the Lower Cretaceous Al
Renk Shale Formation in the studied wells reached high thermal
maturity stages of late-mature oil and wet gas generation windows
and suggest they currently contribute to the generation of mainly
gas.
Agordeed-1 Well.

- Attainment the gas-window maturity was critical for converting the


initial Types II and Type III kerogen into a thermogenic gas gener­
ating in the Al Renk Shale source rock system.
- The 1-D basin models from this study imply that the primary gas
generation potential is thermogenic gas. The initial kerogen was

20
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

converted to oil and the retained oil was eventually cracked to form Allen, P.A., Allen, T.R., 1990. Basin Analysis: Principles and Applications. Blackwell
Scientific publications, Oxford.
thermogenic gas, which may have occurred due to the high paleo-heat flow
Allen, P.A., Allen, J.R., 2005. Basin Analysis: Principles and Applications. Blackwell,
from the rifting systems and high burial temperatures. Oxford, p. 549.
- Basin models also imply that the primary conversion of kerogen to oil Barham, A., Ismail, M.S., Hermana, M., Zainal Abidin, N.S., 2021. Biomarker
happened during the Late Cretaceous (78–68 Ma), with the oil characteristics of Montney source rock, British Columbia, Canada. Heliyon 7,
e08395.
window (0.62–0.85 Easy % Ro) and transformation ratios (TR) Bissada, K.K., 1982. Geochemical constraints on petroleum generation and migration—a
ranging from 10% to 50%. review. Proc ASCOPE Conf 81, 69–87.
Bordenave, M.L., 1993. Applied Petroleum Geochemistry. Editions Technip, Paris.
Botor, D., Bábek, O., 2019. Burial and thermal history modelling of the Upper
The retained oil was eventually cracked into thermogenic gas, which Carboniferous strata based on vitrinite reflectance data from Bzie-Dębina-60
may have occurred as a result of the gas window’s increased thermal borehole (Upper Silesian Coal Basin, southern Poland). Geologické výzkumy na
maturation (>1.30 Easy % Ro) during the Late Eocene and continuing to Moravě a ve Slezsku 26, 73–79.
Browne, S., Fairhead, J., Mohamed, I., 1985. Gravity study of the White Nile Rift, Sudan,
the present day. This research shows that the substantial thermogenic and its regional tectonic setting. Tectonophysics 113, 123–137.
gas potential from the Lower Cretaceous Al Renk Shale Formation for Burnham, A.K., Sweeney, J.J., 1989. A chemical kinetic model of vitrinite maturation
the prospects and exploration targets should be considered in the Melut and reflectance. Geochem. Cosmochim. Acta 53, 2649–2657.
Chandra, K., Mishra, C.S., Samanta, U., Gupta, A., Mehrotra, K.L., 1994. Correlation of
Basin’s deeper parts, where the Al Renk Shale Formation has reached different maturity parameters in the Ahmedabad–Mehsana block of the Cambay
deeper burial depths and gas-generation window. basin. Org. Geochem. 21, 313–321.
Cluff, R., Dickerson, D., 1982. Natural gas potential of the New Albany Shale Group
(Devonian-Mississippian) in Southeastern Illinois. Soc. Pet. Eng. J. 22 (2), 291–300.
Credit author statement Curtis, J.B., 2002. Fractured shale-gas systems. AAPG (Am. Assoc. Pet. Geol.) Bull. 86,
1921–1938.
Mohammed Hail Hakimi: Conceptualization, Data curation, Formal Diab, A.I., Sanuade, O., Radwan, A.E., 2022. An integrated source rock potential,
sequence stratigraphy, and petroleum geology of (Agbada-Akata) sediment
analysis, Investigation, Project administration, Writing – original draft. succession, Niger delta: application of well logs aided by 3D seismic and basin
Atif N. Abass: Conceptualization, Data curation, Formal analysis, modeling. J. Pet. Explor. Prod. Technol. 1–21 https://doi.org/10.1007/s13202-022-
Investigation. Aref Lashin: Funding acquisition, Writing – review & 01548-4.
Didyk, B.M., Simoneit, B.R.T., Brassell, S.C., Eglinton, G., 1978. Organic geochemical
editing. Ahmed E. Radwan: Writing – review & editing. Abbas F. indicators of palaeoenvironmental conditions of sedimentation. Nature 272,
Gharib: Software, Funding acquisition. Afikah Rahim: Funding acqui­ 216–222.
sition, Writing – review & editing. Adeeb Ahmed: Software, Writing – Dou, L., 2004. Formation style of petroleum accumulations in interior rift basins. Petrol.
Explor. Dev. 31, 29–31.
review & editing. Lanre Asiwaju: Writing – review & editing. Wafaa E
Dou, L., Xiao, K., Cheng, D., Shi, B., Li, Z., 2007. Petroleum geology of the Melut basin
Afify: Software, Writing – review & editing. and the great palogue field, Sudan. Mar. Petrol. Geol. 24, 129–144.
Dow, W.G., 1977. Kerogen studies and geological interpretations. J. Geochem. Explor. 7,
79–99.
Declaration of competing interest Eisawi, A.A., Schrank, E., 2008. Upper cretaceous to neogene palynology of the Melut
Basin, south Sudan. Palynology 32, 101–129.
Ekpo, B.O., Oyo-Ita, O.E., Wehner, H., 2005. Even-n-alkane/alkene predominances in
The authors declare that they have no known competing financia­ surface sediments from the Calabar River, SE Niger Delta, Nigeria.
linterestsor personal relationships that could have appeared to influence Naturwissenschaften 92, 341–346.
Erdmann, M., Horsfield, B., 2006. Enhanced late gas generation potential of petroleum
the work reported in this paper. source rocks via recombination reactions: evidence from the Norwegian North Sea.
Geochem. Cosmochim. Acta 70, 3943–3956.
Data availability Espitalie, J., Laporte, L., Madec, M., Marquis, F., Leplate, P., Pault, J., Boutefeu, A., 1977.
Methode rapid de caracterisation des rocks meres, deleur potential petrolier et
leurdegre devolution. Rev. Inst. Fr. Petrol 32, 23–42.
Data will be made available on request. Espitalie, J., Deroo, G., Marquis, F., 1985. La pyrolyse Rock-Eval et ses applications.
Partie 1. Revue de 1’Lnstitut Francois du Petrole 40, 563–579.
Gai, H., Tian, H., Xiao, X., 2018. Late gas generation potential for different types of shale
Acknowledgments source rocks: implications from pyrolysis experiments. Int. J. Coal Geol. 193, 16–29.
Genik, G.J., 1993. Petroleum geology of Cretaceous-Tertiary rift basins in Niger, Chad
The authors extend their admiration to the Sudan Ministry of Energy and Central African Republic. AAPG (Am. Assoc. Pet. Geol.) Bull. 77, 1405–1434.
Gharib, A.F., Özkan, A.M., Hakimi, M.H., Zainal Abidin, N., Lashin, A., 2021. Integrated
and Mining for assisting with this research and providing data. The geochemical characterization and geological modeling of organic matter-rich
authors wish to express their thanks to King Saud University-Deanship of limestones and oils from Ajeel Oilfield in Mesopotamian Basin, Northern Iraq. Mar.
Scientific Research (DSR) for supporting the work. Schlumberger is also Petrol. Geol. 126, 104930.
Guiraud, R., Maurin, J.C., 1992. Early Cretaceous rift of western and central africa: an
credited for making the free 1D PetroMod basin modelling software overview. Tectonophysics 213, 153–168.
available. We are also grateful to the Guest Editors board members Dr. Gürgey, K., 1999. Geochemical characteristics and thermal maturity of oils from the
Ahmed E. Radwan and Dr. David A. Wood and for organizing this special Thrace Basin (western Turkey) and western Turkmenistan. J. Petrol. Geol. 22,
167–189.
issue in the Journal of Marine and Petroleum Geology. The constructive Hadad, Y., Hakimi, M.H., Abdullah, W.H., Kinawy, M., ElMahdy, O., Lashin, A., 2021.
comments by Dr. Peshkov and other anonymous reviewers have Organic geochemical characteristics of Zeit source rock from Red Sea Basin and their
improved the original and revised manuscripts are gratefully acknowl­ contribution to organic matter enrichment and hydrocarbon generation potential.
J. Afr. Earth Sci. 177, 104151.
edged. Dr. Ahmed E. Radwan is thankful for the support provided by the
Hadad, Y.T., Hakimi, M.H., Abdullah, W.H., Makeen, Y.M., 2017. Basin modeling of the
Priority Research Area Anthropocene under the program “Excellence late Miocene Zeit source rock in the Sudanese portion of red sea basin: implication
Initiative—Research University” at the Jagiellonian University in for hydrocarbon generation and expulsion history. Mar. Petrol. Geol. 84, 311–322.
Kraków. Hakimi, M.H., Abdullah, W.H., 2015. Thermal maturity history and petroleum
generation modelling for the Upper Jurassic Madbi source rocks in the Marib-
Shabowah Basin, western Yemen. Mar. Petrol. Geol. 59, 202–216.
References Hakimi, M.H., Ahmed, A.F., 2016. Petroleum source rock characterization and
hydrocarbon generation modeling of the Cretaceous sediments in the Jiza sub-basin,
eastern Yemen. Mar. Petrol. Geol. 75, 356–373.
Abdula, R.A., 2015. Hydrocarbon potential of Sargelu formation and oil-source
Hakimi, M.H., Abdullah, W.H., Shalaby, M.R., 2010. Source rock characterization and oil
correlation, Iraqi Kurdistan. Arabian J. Geosci. 8, 5845–5868.
generating potential of the Jurassic Madbi Formation, onshore East Shabowah
Abeed, Q., Littke, R., Strozyk, F., Uffmann, A.K., 2013. The Upper Jurassic–Cretaceous
oilfields, Republic of Yemen. Org. Geochem. 41, 513–521.
petroleum system of southern Iraq: a 3-D basin modelling study. GeoArabia 18,
Hakimi, M.H., Abdullah, W.H., Shalaby, M.R., 2012. Molecular composition and organic
179–200.
petrographic characterization of Madbi source rocks from the Kharir Oilfield of the
Alias, F.L., Abdullah, W.H., Hakimi, M.H., Azhar, M.H., Kugler, R.L., 2012. Organic
Masila Basin (Yemen): palaeoenvironmental and maturity interpretation. Arabian J.
geochemical characteristics and depositional environment of the Tertiary Tanjong
Geosci. 5, 817–831.
Formation coals in the Pinangah area, onshore Sabah, Malaysia. Int. J. Coal Geol.
104, 9–21.

21
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

Hakimi, M.H., Alaug, S.A., Ahmed, A.F., Yahya, M.M.A., El Nady, M.M., Ismail, I.M., Peters, K., Cassa, M., 1994. Applied source rock geochemistry. In: Magoon, L.B., Dow, W.
2018. Simulating the timing of petroleum generation and expulsion from deltaic G. (Eds.), The Petroleum System from Source to Trap, vol. 60. AAPG (Am. Assoc. Pet.
source rocks: implications for Late Cretaceous petroleum system in the offshore Jiza- Geol.) Memo, pp. 93–117.
Qamar Basin, Eastern Yemen. J. Petrol. Sci. Eng. 170, 620–642. Peters, K.E., Moldowan, J.M., 1993. The Biomarker Guide: Interpreting Molecular Fossils
Hakimi, M.H., Ahmed, A., Mogren, S., Ali Shah, S.B., Kinawy, M.M., Lashin, A.A., 2020. in Petroleum and Ancient Sediments. Prentice Hall, Englewood Cliffs, New Jersey,
Thermogenic gas generation from organic-rich shales in the southeastern Say’un- p. 363.
Masila Basin, Yemen as demonstrated by geochemistry, organic petrology, and basin Peters, K.E., Walters, C.C., Moldowan, J.M., 2005. The Biomarker Guide 2: Biomarkers
modeling. J. Petrol. Sci. Eng. 192, 107322. and Isotopes in Petroleum Exploration and Earth History, 2d ed. Cambridge
Hakimi, M.H., Abas, A.N., Hadad, Y.H., Al Faifi, H.J., Kinawy, M., Lashin, A., 2021. Bulk University Press, Cambridge, United Kingdom, p. 704.
pyrolysis and biomarker fingerprints of Late Cretaceous Galhak Shale Formation in Philp, R.P., 1985. Biological markers in fossil fuel production. Mass Spectrom. Rev. 4,
the northern Melut Basin, Sudan: implications on lacustrine oil-source rock. Arabian 1–54.
J. Geosci. 14, 520. Radke, M., Schaefer, R.G., Leythaeuser, D., Teichmuller, M., 1980. Composition of
Hantschel, T., Kauerauf, A.I., 2009. Fundamentals of Basin and Petroleum Systems soluble organic-matter in coals - relation to rank and liptinite fluorescence.
Modeling: Integrated Exploration Systems GmbH. Schlumberger Company, Springer- Geochem. Cosmochim. Acta 44, 1787–1800.
Verlag Berlin Heidelberg. Radwan, A.E., 2022. A multi-proxy approach to detect the pore pressure and the origin of
Hazra, B., Wood, D.A., Mani, D., Singh, P.K., Singh, A.K., 2018. Evaluation of Shale overpressure in sedimentary basins: an example from the Gulf of Suez rift basin.
Source Rocks and Reservoirs. Springer International Publishing, p. 142. Front. Earth Sci. 10, 967201 https://doi.org/10.3389/feart.2022.967201.
He, S., Middleton, M., 2002. Heat flow and thermal maturity modelling in the northern Radwan, A.E., Abudeif, A.M., Attia, M.M., Elkhawaga, M.A., Abdelghany, W.K.,
carnarvon basin, north west shelf, Australia. Mar. Petrol. Geol. 19, 1073–1088. Kasem, A.A., 2020. Geopressure evaluation using integrated basin modelling, well-
Huang, W.Y., Meinschein, W.G., 1979. Sterols as ecological indicators. Geochem. logging and reservoir data analysis in the northern part of the Badri oil field, Gulf of
Cosmochim. Acta 43, 739–745. Suez, Egypt. J. Afr. Earth Sci. 162, 103743. https://doi:10.1016/j.jafrearsci.2019.10
Hughes, W.B., Holba, A.G., Dzou, L.I.P., 1995. The ratios of dibenzothiophene to 3743.
phenanthrene and pristane to phytane as indicators of depositional environment and Radwan, A.E., Wood, D.A., Mahmoud, M., Tariq, Z., 2022. Gas adsorption and reserve
lithology of petroleum source rocks. Geochem. Cosmochim. Acta 59, 3581–3598. estimation for conventional and unconventional gas resources. In: Wood, D.A.,
Jarvie, D.M., 1991. Total organic carbon (TOC) analysis. In: Merrill, R.K. (Ed.), Treatise Cai, J. (Eds.), Sustainable Geosciences for Natural Gas Sub-surface Systems. Elsevier,
of Petroleum Geology: Handbook of Petroleum Geology, Source and Migration pp. 345–382 (Chapter 12).
Processes and Evaluation Techniques. AAPG. Am. Assoc. Pet. Geol.) Bull, Sarki Yandoka, B.M., Abdullah, W.H., Abubakar, M.B., Hakimi, M.H., Mustapha, K.A.,
pp. 113–118. Adegoke, A.K., 2015. Organic geochemical characteristics of cretaceous lamja
Jarvie, D.M., Claxton, B.L., Henk, F., Breyer, J.T., 2001. Oil and shale gas from the Formation from yola sub-basin, northern benue trough, NE Nigeria: implication for
Barnett shale, Fort Worth Basin, Texas. AAPG national convention, june 3-6, 2001, hydrocarbon-generating potential and paleodepositional setting. Arabian J. Geosci.
denver, CO. AAPG (Am. Assoc. Pet. Geol.) Bull. 85 (13), A100 (Supplement). 8, 7371–7386.
Jarvie, D.M., Hill, R.J., Ruble, T.E., Pollastro, R.M., 2007. Unconventional shale-gas- Schull, T.J., 1988. Rift basins of interior Sudan: petroleum exploration and discovery.
systems: the Mississippian Barnett Shale of north-central Texas as one model for Am. Assoc. Petrol. Geol. Bull. 72, 1128–1142.
thermogenic shale-gas assessment. AAPG (Am. Assoc. Pet. Geol.) Bull. 91, 475–499. Schwark, L., Empt, P., 2006. Sterane biomarkers as indicators of Palaeozoic algal
Kaska, H.V., 1989. A spore and pollen zonation of the Early Cretaceous to Cenozoic non- evolution and extinction events. Palaeogeogr. Palaeoclimatol. Palaeoecol. 240,
marine sediments of Central Sudan. Palynology 131, 79–90. 225–236.
Katz, B., Lin, F., 2014. Lacustrine basin unconventional resource plays: key differences. Seifert, W.K., Moldowan, J.M., 1978. Application of steranes, terpanes and
Mar. Petrol. Geol. 56, 255–265. monoaromatic to the maturation, migration and source of crude oils. Geochem.
Katz, B., Lin, F., 2021. Consideration of limitations of thermal maturity with respect to Cosmochim. Acta 42, 77–95.
vitrnite reflectance, Tmax and other proxies. AAPG (Am. Assoc. Pet. Geol.) Bull. 105, Seifert, W.K., Moldowan, J.M., 1981. Paleo-reconstruction by biological markers.
695–720. Geochem. Cosmochim. Acta 45, 783–794.
Katz, B.J., Pheifer, R.N., Schunk, D.J., 1988. Interpretation of discontinuous vitrinite Seifert, W.K., Moldowan, J.M., 1986. Use of biological markers in petroleum exploration.
reflectance profiles. AAPG (Am. Assoc. Pet. Geol.) Bull. 72, 926–931. In: Johns, R.B. (Ed.), 24, Methods in Geochemistry and Geophysics Book. Amsterdam
Lachenbruch, A., 1970. Crustal temperature and heat productivity: implications of the Series, pp. 261–290.
linear heat flow relation. J. Geophys. Res. 75, 3291–3300. Shalaby, M.R., Hakimi, M.H., Abdullah, W.H., 2011. Geochemical characteristics and
Li, M.J., Wang, T.G., Chen, J.F., He, F.Q., Yun, L., Akbar, S., Zhang, W.B., 2010. Paleo hydrocarbon generation modeling of the Jurassic source rocks in the Shoushan
heat flow evolution of the Tabei Uplift in Tarim basin, northwest China. J. Asian Basin, north Western Desert, Egypt. Mar. Petrol. Geol. 28, 1611–1624.
Earth Sci. 37, 52–66. Shalaby, M.R., Hakimi, M.H., Abdullah, W.H., 2013. Modeling of gas generation from the
Mackenzie, A.S., Patience, R.L., Maxwell, J.R., Vandenbroucke, M., Durand, B., 1980. Alam El-Bueib Formation in the Shoushan basin, northern western desert of Egypt.
Molecular parameters of maturation in the Toarcian shales, Paris Basin, France—I. Int. J. Earth Sci. 102, 319–332.
Changes in the configurations of acyclic isoprenoid alkanes, steranes and triterpanes. Shanmugam, G., 1985. Significance of coniferous rain forests and related oil , Gippsland
Geochem. Cosmochim. Acta 44, 1709–1721. Basin, Australia. Am. Assoc. Pet. Geol. Bull. 69, 1241–1254.
Mahdi, A.Q., Abdel-Fattah, M.I., Radwan, A.E., Hamdan, H.A., 2022. An integrated Sohail, G.M., Radwan, A.E., Mahmoud, M., 2022. A review of Pakistani shales for shale
geochemical analysis, basin modeling, and palynofacies analysis for characterizing gas exploration and comparison to North American shale plays. Energy Rep. 8,
mixed organic-rich carbonate and shale rocks in Mesopotamian Basin, Iraq: insights 6423–6442.
for multisource rocks evaluation. J. Petrol. Sci. Eng., 110832 https://doi.org/ Sweeney, J.J., Burnham, A.K., 1990. Evaluation of a simple model of vitrinite reflectance
10.1016/j.petrol.2022.110832. based on chemical kinetics AAPG. Am. Assoc. Petrol. Geol. Bull. 74, 1559–1570.
Makeen, Y.M., Abdullah, W.H., Hakimi, M.H., Mustapha, K.A., 2015. Source rock Ten Haven, H.L., de Leeuw, J.W., Rullkötter, J., Damsté, J.S.S., 1987. Restricted utility of
characteristics of the lower cretaceous Abu Gabra formation in the Muglad Basin, the pristane/phytane ratio as a palaeoenvironmental indicator. Nature 330,
Sudan, and its relevance to oil generation studies. Mar. Petrol. Geol. 59, 505–516. 641–643.
Makeen, Y.M., Abdullah, W.H., Pearson, M.J., Hakimi, M.H., Elhassan, O.M.A., Yousif, T. Tissot, B.P., Welte, D.H., 1984. Petroleum Formation and Occurrence, second ed.
A.H., 2016. Thermal maturity history and petroleum generation modelling for the Springer Verlag, Berlin, p. 699.
lower cretaceous Abu Gabra formation in the Fula sub-basin, Muglad Basin, Sudan. Tissot, B.P., Pelet, R., Ungerer, P.H., 1987. Thermal history of sedimentary basins,
Mar. Petrol. Geol. 75, 310–324. maturation indices and kinetics of oil and gas generation. Am. Assoc. Petrol. Geol.
Martini, A.M., Walter, L.M., Budai, J.M., Ku, T.C.W., Kaiser, C.J., Schoell, M., 1998. Bull. 71, 1445–1466.
Genetic and temporal relations between formation waters and biogenic methane: Tong, X., Dou, L., Tian, Z., Pan, X., Zhu, X., 2004. Geological mode and hydrocarbon
upper Devonian Antrim Shale, Michigan Basin, USA. Geochem. Cosmochim. Acta 62, accumulating mode in Muglad passive Rift Basin, Sudan. Acta Pet. Sin. 25, 19–24.
1699–1720. Tserolas, P., Maravelis, A.G., Tsochandaris, N., Pasadakis, N., Zelilidis, A., 2019. Organic
McHargue, T.R., Heidrick, J.L., Livingstone, J.E., 1992. Tectono-stratigraphic geochemistry of the upper miocene-lower Pliocene sedimentary rocks in the hellenic
development of the interior Sudan rifts, Central Africa. Tectonophysics 213, fold and thrust belt, NW corfu island, ionian sea, NW Greece. Marine Mar. Petrol.
187–202. Geol. 106, 17–29.
Mohamed, A.Y., Pearson, M.J., Ashcroft, W.A., Iliffe, W.A., Whiteman, A.J., 1999. Wangen, M., 1995. The blanketing effect in sedimentary basins. Basin Res. 7, 283–298.
Modelling petroleum generation in the southern muglad rift Basin Sudan. AAPG Waples, D.W., 1980. Time and temperature in petroleum formation: application of
(Am. Assoc. Pet. Geol.) Bull. 83, 1943–1964. Lopatin’s method to petroleum exploration. AAPG (Am. Assoc. Pet. Geol.) Bull. 4,
Mohamed, A.Y., Whiteman, A.J., Archer, S.G., Bowden, S.A., 2016. Thermal modelling of 916–926.
the Melut basin Sudan and South Sudan: implications for hydrocarbon generation Waples, D.W., 1994. Modeling of sedimentary basins and petroleum systems. In:
and migration. Mar. Petrol. Geol. 77, 746–762. Magoon, L.B., Dow, W.G. (Eds.), The Petroleum System from Source to Trap, vol. 60.
Mukhopadhyay, P.K., Wade, J.A., Kruge, M.A., 1995. Organic facies and maturation of AAPG (Am. Assoc. Pet. Geol.) Memo., pp. 307–322
Jurassic/Cretaceous rocks, and possible oil-source rock correlation based on Waples, D.W., Machihara, T., 1991. Biomarkers for geologists: a practical guide to the
pyrolysis of asphaltenes, Scotion Basin. Canada. Org. Geochem. 22, 85–104. application of steranes and triterpanes in petroleum geology. AAPG (Am. Assoc. Pet.
Peshkov, G.A., Chekhonin, E.M., Rüpke, L.H., Musikhin, K.A., Bogdanov, O.A., Geol.) Meth. Exp. 9, 91.
Myasnikov, A.V., 2021. Impact of differing heat flow solutions on hydrocarbon
generation predictions: a case study from West Siberian Basin. Mar. Petrol. Geol.
124, 104807.
Peters, K.E., 1986. Guidelines for evaluating petroleum source rock using programmed
pyrolysis. AAPG (Am. Assoc. Pet. Geol.) Bull. 70, 318–329.

22
M.H. Hakimi et al. Marine and Petroleum Geology 146 (2022) 105926

Welte, D.H., Horsfield, B., Baker, D.R., 2012. Petroleum and Basin Evolution: Insights Zhao, W., Shi, Z., Li, S., Chen, B., Luo, X., Pang, W., 2020. Petroleum geological
from Petroleum Geochemistry, Geology and Basin Modeling. Springer, Berlin characteristics and hydrocarbon accumulation patterns in the Melut Basin, South
Heidelberg. Sudan. Arabian J. Geosci. 13, 1149.
Zhang, S., Wu, T., Zhang, S., Cao, C., Ma, W., Shi, J., Sun, G., 2015. Organofacies and
paleoenvironment of lower carboniferous mudstones (dishuiquan formation) in
eastern junggar, NW China. Int. J. Coal Geol. 150–151, 7–18.

23

You might also like