You are on page 1of 48

Accepted Manuscript

Organic geochemistry of the Silurian Tanezzuft Formation and crude oils, NC115
Concession, Murzuq Basin, southwest Libya

W. Sh. El Diasty, S.Y. El Beialy, T.A. Anwari, K.E. Peters, D.J. Batten

PII: S0264-8172(17)30209-X
DOI: 10.1016/j.marpetgeo.2017.06.002
Reference: JMPG 2936

To appear in: Marine and Petroleum Geology

Received Date: 26 November 2016


Revised Date: 3 May 2017
Accepted Date: 1 June 2017

Please cite this article as: El Diasty, W.S., El Beialy, S.Y., Anwari, T.A., Peters, K.E., Batten, D.J.,
Organic geochemistry of the Silurian Tanezzuft Formation and crude oils, NC115 Concession, Murzuq
Basin, southwest Libya, Marine and Petroleum Geology (2017), doi: 10.1016/j.marpetgeo.2017.06.002.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Organic geochemistry of the Silurian Tanezzuft Formation and crude oils, NC115
Concession, Murzuq Basin, southwest Libya

W.Sh. El Diasty *1, S.Y. El Beialy 1, T.A. Anwari 1, K.E. Peters 2, D.J. Batten 3
1
Geology Department, Faculty of Science, Mansoura University, Mansoura 35516, Egypt
2
Schlumberger, Mill Valley, CA 94941 & Geological Sciences Department, Stanford University, CA, USA

PT
3
School of Earth, Atmospheric and Environmental Sciences, University of Manchester, Oxford Road, Manchester M13
9PL, UK
*
Corresponding author: Dr. Waleed El Diasty; awaleed@mans.edu.eg, +20 0106 244 3492 – Mansoura – Egypt

RI
Abstract
Thirty-six Silurian cores and cuttings samples and 10 crude oil samples from Ordovician reservoirs in

SC
the NC115 Concession, Murzuq Basin, southwest Libya were studied by organic geochemical methods
to determine source rock organic facies, conditions of deposition, thermal maturity and genetic
relationships. The Lower Silurian Hot Shale at the base of the Tanezzuft Formation is a high-quality

U
oil/gas-prone source rock that is currently within the early oil maturity window. The overall average
AN
TOC content of the Hot Shale is 7.2 wt% with a maximum recorded value of 20.9 wt%. By contrast,
the overlying deposits of the Tanezzuft Formation have an average TOC of 0.6 wt% and a maximum
M

value of 1.1 wt%. The organic matter in the Hot Shale consists predominantly of mixed algal and
terrigenous Type-II/III kerogen, whereas the rest of the formation is dominated by terrigenous Type-III
D

organic matter with some Type II/III kerogen. Oils from the A, B and H oil fields in the NC115
Concession were almost certainly derived from marine shale source rocks that contained mixed algal
TE

and terrigenous organic input reflecting deposition under suboxic to anoxic conditions. The oils are
light and sweet, and despite being similar, were almost certainly derived from different facies and
EP

maturation levels within mature source rocks. The NC115-B oils were generated from slightly less
mature source rocks than the others. Based on hierarchical cluster analysis (HCA), principal component
analysis (PCA), selected source-related biomarkers and stable carbon isotope ratios, the NC115 oils can
C

be divided into two genetic families: Family-I oils from Ordovician Mamuniyat reservoirs were
AC

probably derived from older Palaeozoic source rocks, whereas Family-II oils from Ordovician
Mamuniyat–Hawaz reservoirs were probably charged from a younger Palaeozoic source of relatively
high maturity. A third family appears to be a mixture of the two, but is most similar to Family-II oils.
These oil families were derived from one proven mature source rock, the early Silurian, Rhuddanian
Hot Shale. There is a good correlation between the Family-II and -III oils and the Hot Shale based on
carbon isotope compositions. Saturated and aromatic maturity parameters indicate that these oils were
generated from a source rock of considerably higher maturity than the examined rock samples. The

1
ACCEPTED MANUSCRIPT

results imply that the oils originated from more mature source rocks outside the NC115 Concession and
migrated to their current positions after generation.

Keywords: Source rocks, Silurian, biomarkers, NC115 Concession, Murzuq Basin, Libya

1. Introduction

PT
With 44 billion barrels of oil and over 54 trillion cubic feet of natural gas, the largest proven reserves of

RI
hydrocarbons in Africa occur in Libya, an important founding member of the Organization of
Petroleum Exporting Countries (OPEC). From the exploration perspective, the country is divided into

SC
four major basins, three of which, Ghadames, Murzuq and Al Kufra, are essentially Palaeozoic basins.
The fourth, Sirte, is primarily a Mesozoic–Cenozoic basin (Fig. 1A). The Ghadames Basin continues
westwards into Tunisia and Algeria where it attains its greatest depth. By contrast, the Murzuq Basin

U
terminates against the Tihemboka Arch and does not extend into Algeria (Fig. 1B). It continues
AN
southwards into Niger, where it eventually pinches out onto basement.
Libya has 29 oil fields with greater than one billion barrels of oil originally in place; 22 of these
M

fields lie in the Sirte Basin, six in the Murzuq Basin (Hallett and Clark-Lowes, 2016) and one, Bouri, in
the Sabratah Basin, offshore northwest Libya. The reserves of the Murzuq Basin are now second only
to those of the Sirte Basin. A recent estimation of hydrocarbons in place for the Murzuq Basin is 6.0
D

billion barrels of oil and 1.0 trillion cubic metres of gas, which represents about 6.5% of the Libya's
TE

total production. Since oil exploration began in 1957, 62 wildcat wells resulted in the exploitation of 11
oil fields in the basin (Rusk, 2001), several of which, including El Shararah, and El Feel (Elephant), are
EP

giant fields. The focus of this paper is the NC115 Concession in the northwestern part of the basin,
some 800 km south of Tripoli. It covers approximately three-quarters of the area between latitudes
26°8' and 26°38'N and longitudes 11°30' and 12°30'E (Fig. 1C). Within the concession three major oil
C

fields, NC115-A, -B and -H, were developed in this part of the Sahara Desert. The NC115-B Field is
AC

the most southwesterly of these. It is approximately 50 km southwest of the NC115-A Field, which in
turn is approximately 10 km west of the NC115-H Field (Fig. 1D; RRI, 1998). REPSOL is the operator
of the NC115 exploration block, and its partners are OMV and TOTAL. Austria's OMV has been
present in Libya since 1975 and started production in 1985.
Previous studies of the Murzuq Basin mainly focused on the geochemical characteristics of
hydrocarbon fluids or on source rock evaluation, with only limited correlation between them (e.g.,
Lüning et al., 2000, 2003; Belaid et al., 2010; Hall et al., 2010; Hodairi and Philp, 2011, 2012; Butcher,

2
ACCEPTED MANUSCRIPT

2013; Loydell et al., 2013; Meinhold et al., 2013, and references therein). The main goals of this study
are to determine (1) environments of deposition, thermal maturity and hydrocarbon potential of the
Silurian Tanezzuft and its basal Hot Shale Member in the context of resource assessment, and (2) the
molecular and isotopic characteristics of a suite of crude oils and their genetic relationships.

PT
2. Structural/stratigraphic framework

RI
Oceanic anoxic events leading to oxygen depletion and deposits with high total organic carbon (TOC)
content occurred in parts of the world’s oceans several times in the past (Klemme and Ulmishek, 1991;

SC
Lüning et al., 2000, 2003, 2005). In North Africa and Arabia much of the Palaeozoic oil resources
resulted from the maturation and migration of hydrocarbons generated from Silurian organic-rich
marine shales (Klemme and Ulmishek, 1991; Boote et al., 1998; Lüning et al., 2000, 2003; Peters and

U
Creaney, 2004; Loydell et al., 2009; İnan et al., 2016). The intracratonic Murzuq Basin is no exception.
AN
Regional tectonic events and changes in relative sea level had a significant impact on
stratigraphic settings in this basin and the NC115 study area. Precambrian tectonic activity led to
vertical basement N–S trending faults that were counterbalanced by conjugated faults trending in a
M

NE–SW direction (Goudarzi, 1980; Echikh and Sola, 2000). Several subsequent compressional and
extensional tectonic episodes have generally been assigned to the Caledonian (Wenlockian, late Early
D

Silurian), Hercynian (Middle–Late Carboniferous) and Alpine (Bellini and Massa, 1980; Aziz, 2000;
TE

Echikh and Sola, 2000) orogenies.


The Murzuq Basin is situated between three major defining tectonic elements; the Al Qarqaf
High to the north, the Tibisti High to the east, and the Tihemboka High to the west (Fig. 1B). These
EP

prominent highs resulted from various tectonic events from mid-Palaeozoic to Tertiary times, but the
main periods of uplift took place during the mid-Cretaceous and Paleogene when the chief influence on
C

sedimentation was probably a NNW–SSE structural trend (RRI, 1998). This was disrupted by a number
AC

of later strike-slip faults of mainly dextral displacement that trend approximately perpendicular to the
main structures in an ENE–WSW orientation (Fig. 2A).
Within the NC115 Concession a strike-slip fault (Fig. 2A) runs from near the southwestern
corner to the northeastern corner, and all discovered oil fields are to the northwest of this fault. The
majority of the fields are located in basinal areas with the exception of the NC115-B field toward the
southwest of the concession, which is on a localized positive area bounded by two sub-parallel NNE–
SSW-oriented normal faults.

3
ACCEPTED MANUSCRIPT

The geology of the NC115 license area is generally typical of the Murzuq Basin as a whole
(Fig. 2B) in comprising Precambrian to Cretaceous rocks (Figs. 2B, 3), formation thicknesses being
broadly similar throughout the concession (RRI, 1998; Aziz, 2000; Hallett, 2002). The deposits consist
mainly of marine shale, siltstone, and sandstone unconformably overlying the Precambrian basement
complex. Their total thickness exceeds 3500 m in the central part of the basin. The hydrocarbons in

PT
Ordovician reservoirs in the basin were derived from Silurian source rocks (Boote et al., 1998;
Davidson et al., 2000; Echikh and Sola, 2000): our discussion of the stratigraphic setting of the Murzuq
Basin will, therefore, focus on these rock units.

RI
The Ordovician rocks of this area were first described by Massa and Collomb (1960). They are
widespread over large portions of the North African craton (Pierobon, 1991). Four formations represent

SC
the sequence, namely Ash Shabiyat (Lower Ordovician), Hawaz (Middle Ordovician), Melaz Shuqran
and Mamuniyat (Upper Ordovician) (Lüning et al., 2000). We focus here on the Hawaz and Mamuniyat

U
formations, which are important reservoirs (Fig. 3).
AN
The Hawaz Formation was first described and named after Jebel Hawaz on the Al Qarqaf High
by Massa and Collomb (1960). In the NC115 Concession, it is overlain by the Ordovician Melaz
Shuqran Formation (Fig. 3), where it reaches 150 m in thickness. Its probable mid-Ordovician age is
M

based on palynological evidence (Hallett, 2002). The formation consists of fine- to medium-grained
quartzitic sandstone that is moderately to well cemented by siliceous material. It is kaolinitic in parts,
D

has poor visual porosity, and contains thin streaks of moderately- to well-compacted, sub-fissile to
TE

fissile shale (Hallett, 2002; Ramos et al., 2006). These deposits reflect the onset of the first major
Palaeozoic marine transgression in the region (Echikh and Sola, 2000; Ramos et al., 2006; de Gibert et
al., 2011).
EP

The name Mamuniyat Formation is derived from the type section on the Al Qarqaf Arch. The
names of most stratigraphic units in the Lower Palaeozoic of Libya are based on locations either in
C

wadis (e.g., Tanezzuft Formation) or on arches (Mamuniyat Formation), as reported by Le Heron et al.
(2010). In common with the underlying units, the Upper Ordovician glacially-related strata occur on
AC

the Al Qarqaf Arch where two formations are formally recognized in both surface (Massa and Collomb
1960) and subsurface (El-ghali, 2005) mapping, namely the Melaz Shuqran Formation and the
overlying Mamuniyat Formation (Gundobin, 1985). The latter is predominantly a sandstone unit with
subordinate siltstone and shale interbeds (Aziz, 2000; Davidson et al., 2000). The sandstone is fine-
grained, off-white, medium hard to hard, and moderately to well cemented by siliceous material. It also
includes thin layers of fissile to sub-fissile grey shales. The lower and middle parts of the formation
represent glacial deposits associated with relative sea-level fall (El-ghali, 2005). The upper units
4
ACCEPTED MANUSCRIPT

comprise a prolific reservoir unit in the NC115 B oil field, and are in direct contact with the overlying
source rock, the Silurian (Llandovery–Wenlock) Tanezzuft Formation (Hallett, 2002). The sandstones
of the Mamuniyat Formation are the primary reservoir in the Murzuq Basin. The fluvial and shallow
marine sandstones of the Hawaz Formation are secondary in this respect. Oil discoveries in the NC115
Concession seem to follow a NW–SE trend. This regional trend is probably related to the early

PT
Palaeozoic structural configuration that controlled both the development of reservoir facies and source
rock richness. Oil discoveries in the Mamuniyat Formation have typically been at depths between 1500
and 2000 m, as in the A, B and H fields in the NC115 license (Echikh and Sola, 2000).

RI
The Tanezzuft Formation was named after Wadi Tanezzuft by Desio (1936). It overlies the
Mamuniyat Formation, as noted above, and underlies the (Ludlow–Pridoli) Akakus Formation. It is an

SC
important stratigraphic unit for petroleum exploration because it represents the most proven Palaeozoic
source rock in North Africa and northern Gondwana (Hallett, 2002). It reflects a widespread marine

U
transgression of the Silurian sea over the North African craton (Mamgain, 1980), although its areal
AN
extent was reduced by later mid-Devonian and post-Hercynian erosion.
Melting of the Late Ordovician ice sheets led to a major marine transgression from the north,
and culminated in a highstand with deposition of the Tanezzuft Formation, a consistently argillaceous
M

succession interbedded with very thin sandstone layers, the total thickness of which may reach more
than 700 m in some wells in the NC115 Concession. The lower part of the formation, which has been
D

dated as early Llandovery (Davidson et al., 2000), consists predominantly of grey to dark grey, fissile
TE

to sub-fissile, occasionally micaceous shales. The upper part begins with a light grey shale that is
usually soft to firm, fissile to sub-fissile, micaceous and silty.
The basal section of the Tanezzuft Formation is an early Llandovery, Rhuddanian, black shale
EP

known as the Hot Shale Member. This constitutes the most important source rock within the Murzuq
Basin, but its extent is limited as a result of Hercynian erosion (Lüning et al., 2000; Hallett, 2002). In
C

southern Libya the anoxic event that led to this deposit was confined to depressions in the eroded
Ordovician surface (Parrish, 1982; Finney and Berry, 1997; Lüning et al., 1999, 2000, 2003). It resulted
AC

from a favourable combination of marine upwelling associated with abundant organic productivity and
preservation under oxygen-deficient conditions (Waples, 1985; Finney and Berry, 1997). This
exceptionally organic-rich unit is absent from the eastern and southern parts of the basin. It is present
only in five (B42, H15, H28, H29, and H39) of the 15 wells in the NC115 Concession study area. Its
maximum thickness is about 45 m in the southeast (Aziz, 2000).
The Hot Shale deposits are commonly black, dark brown to dark grey, micaceous, and pyritic.
Isopach maps show that in North Africa they are thickest and most extensive in western Libya, Algeria
5
ACCEPTED MANUSCRIPT

and Tunisia, whereas in Arabia they are most prolific in Saudi Arabia, Oman, Jordan and Iraq. They are
absent from Egypt, which was a structurally high region at this time (Echikh and Sola, 2000; Lüning et
al., 2000).
The term ‘Hot Shale’ refers to the high level of natural radioactivity in certain shales owing to
an increase in authigenic uranium, thorium and potassium, and can be easily recognized on wireline

PT
logs (Lüning et al., 2000; Armstrong et al., 2005). High gamma ray values have also been recorded for
Hot Shales in Tunisia (Vecoli et al., 2009), Qusaiba in Saudi Arabia (McGillivray and Husseini, 1992;
İnan et al., 2016), and Jordan (Butcher, 2009).

RI
Most of the discoveries in the Murzuq Basin occur in structural traps. The multiphase evolution
of the area has produced a great variety of such traps of different ages. Mechanisms for the

SC
Mamuniyat/Hawaz play include fault block, drape anticlinal, faulted anticlinal, and buried hill traps.
Reverse thrust-faulted anticlines form traps in NC115, e.g. in the H and C oil fields. The faults

U
contribute to the seal on the eastern flanks of these fields.
AN
The degree of lateral continuity of the sealing surface of the Mamuniyat reservoir, in addition to
other factors such as lateral permeability and fault barriers, are the principal controls on the distribution
of migration pathways. In areas where the Tanezzuft Formation is absent or thin, the reservoir is
M

capped by different Middle–Late Devonian units with variable sealing capacities (Echikh and Sola,
2000; Hallett and Clark-Lowes, 2016).
D
TE

3. Material and methods

Ten core and 26 cuttings samples were collected from 11 exploratory wells (Table 1) from different
EP

depth intervals within the Tanezzuft Formation and the basal ‘Hot Shale' in Concession NC115 (A-, B-
and H-oil fields), southwest Libya (Fig. 1D; Table 1). In order to assess the hydrocarbon potential and
C

organic geochemical signature of this formation, data from analysis of TOC, Rock-Eval pyrolysis,
AC

bitumen extraction and stable carbon isotopic compositions (δ13C) were measured. In addition, 10 oil
samples were analyzed for bulk and elemental composition, stable carbon isotopic composition (δ13C),
gas chromatography (GC), and gas chromatography-mass spectrometry (GC-MS).
TOC and Rock-Eval pyrolysis are widely accepted as reliable methods for providing basic data
on the presence of contaminants, such as diesel oil or lignite from the drilling fluid, as well as for
determining organic richness, kerogen type, thermal maturity and petroleum generation potential
(S1+S2). The weight percentage of TOC was determined using a LECO C230 analyzer. The CO2

6
ACCEPTED MANUSCRIPT

generated by the combustion of organic matter in the sample was quantitatively measured using an
infrared detector. The precision of the TOC analyses is better than ±0.5% based on repeated analyses of
an internal standard. The petroleum potential of the rock samples was estimated using the Rock-Eval 6
instrument at StratoChem, New Maadi, Cairo.
For bitumen extraction, 50–100 g of powdered rock samples were extracted using Soxhlet

PT
apparatus for 48 h with a mixture of dichloromethane and methanol (93:7 v:v). The extracted bitumen
was then analyzed by a variety of techniques to compare source rocks and relate them to recovered oils.
The extracted bitumen and oil samples were subjected to asphaltene precipitation using excess

RI
n-hexane. The soluble maltene was fractionated into saturated and aromatic hydrocarbons and NSO
(nitrogen, sulfur, oxygen) polar compounds by column chromatography on activated silica gel using

SC
hexane, dichloromethane and dichloromethane/methanol (50:50) as eluents for the saturated, aromatic,
and polar fractions, respectively. The C15+ saturated hydrocarbon fraction was subjected to molecular

U
sieve filtration (Union Carbide S-115 powder) in order to concentrate the branched/cyclic biomarker
AN
fraction (West et al., 1990).
Stable carbon isotope compositions of the source rock extracts and crude oils were determined
using a Finnigan Delta E isotope ratio mass spectrometer with helium the carrier gas. Results of the
M

δ13C values are reported relative to the Pee Dee Belemnite (PDB) standard; uncertainty is ± 0.05‰.
Bitumens and crude oils were injected (split mode 70/1) on a 30 m × 0.32 mm J&W DB-5
D

column (0.2 µm film thickness) at temperatures programmed from 60–350°C at 12°C/min using an
TE

Agilent 7890A gas chromatograph. Helium was used as the carrier gas. Bitumen samples were not
subsequently analyzed for biomarkers as they were totally depleted during carbon isotope and routine
GC analyses.
EP

GC-MS analyses of C15+ branched/cyclic and aromatic hydrocarbon fractions (in order to
determine sterane, terpane and aromatic biomarker distributions and quantities) were carried out using
C

an Agilent 7890A GC System (split injection) interfaced to a 5975C mass spectrometer. Helium was
kept at a constant flow rate of 1 ml/min throughout the analysis. The J&W DB5 column (50 m × 0.2
AC

mm; 0.11 µm film thickness) was temperature programmed from 150–325°C at 2°C/min
(branched/cyclic) and 100–325°C at 3°C/min for aromatics. The mass spectrometer was operated in
electron ionization mode at 70 eV, 400 µA filament emission current and 230oC interface temperature.
Selected ion monitoring (SIM) was used to detect triterpanes (m/z 191), steranes (m/z 217), triaromatic
steroids (m/z 231), phenanthrenes (m/z 178, 192), and dibenzothiophenes (m/z 184, 198). Individual
sterane, hopane and aromatic components were identified by comparison of their retention times and

7
ACCEPTED MANUSCRIPT

mass spectra with published data (Peters et al., 2005; El Diasty et al., 2016). All biomarker ratios were
calculated by measuring peak areas in the GC-MS chromatograms.
Many multivariate statistical methods have been used to separate oils from different source
rocks into genetic families. Among the most common techniques for exploratory data analysis are
hierarchical cluster analysis (HCA) and principal component analysis (PCA). HCA and PCA of oil data
were determined using Pirouette® 2.03 chemometric software (Infometrix, Inc.; Bothell, Washington).

PT
They were completed using autoscale preprocessing, Euclidean metric distance and incremental linkage
in Pirouette. In autoscale pre-processing, all values for each parameter are normalized to the standard

RI
deviation for that parameter. This results in equal weights for each of the parameters in the assessment
of genetic affinity. The main purpose of PCA is to reduce the dimensionality of a geochemical data set

SC
from many variables to a few components that best explain the variation in the data, whereas the HCA
displays the degree of similarity in the geochemical characteristics between the oil samples (Peters et

U
al., 2005, 2016).
AN
4. Results and discussion

4.1. Geochemistry of source rocks


M

4.1.1. Organic matter richness, type and thermal maturity


The results of the TOC analysis and Rock-Eval pyrolysis on shale samples from the Tanezzuft
D

Formation are listed in Table 1 and displayed in Figs. 4–6. The TOC values from the Tanezzuft
TE

Formation range from 0.34–1.13 wt%, with an average of 0.62 wt% (Table 1), and the total pyrolytic
yield (S1+S2) ranges from 0.23–3.45 % (averaging 1.37 wt%). The Hot Shale samples display widely
variable but high TOC values and superior hydrocarbon generating potential compared to the rest of the
EP

Tanezzuft Formation, with TOC contents ranging from 1.7–20.9 wt% (averaging 7.2 wt%), and S1+S2
ranges from 4.15–66.89 mg HC/g rock with an average 21.5 wt% (Table 1).
C

The post-Rhuddanian rocks of the Tanezzuft Formation are, in general, organically-lean and do
AC

not make a significant contribution to petroleum generation (Lüning et al., 2000). This is consistent
with the plot of TOC versus S1+S2 (Fig. 4) in which most of the samples from the formation plot in the
poor to fair source rock region (organically lean), whereas those from the Hot Shale have good to very
good or excellent source rock characteristics (organically rich).
Kerogen types dispersed in sedimentary rocks are distinguished using the modified van
Krevelen diagram (Peters and Cassa, 1994). In the hydrogen index (HI) versus oxygen index (OI)
kerogen-type classification diagram here (Fig. 5), the Hot Shale contains mainly Type II/III kerogen

8
ACCEPTED MANUSCRIPT

whereas samples from the rest of Tanezzuft Formation yielded both Type III and Type II/III kerogen.
Variations in the latter reflect differences in the organic input, e.g., terrigenous plant debris as opposed
to aquatic, algal matter. The majority of Hot Shale samples are characterized by high HI and very low
OI, and thus indicate considerable potential for generating hydrocarbons, whereas samples from the
overlying beds commonly yield terrigenous plant debris that accumulated in a more oxidizing

PT
environment suggesting a tendency to generate mainly gas, although some oil might also be expected.
The current hydrocarbon potential of the Hot Shale is, however, less than expected from this proven
source rock.

RI
The ratio of S2/S3 suggests the quality of the organic matter. S2/S3 values of <1 indicate an
inability to generate any hydrocarbons, whereas S2/S3 values of >10 indicate an effective source rock

SC
for significant oil-generation potential (Peters and Cassa, 1994). The S2/S3 ratio for the Tanezzuft
Formation above the Hot Shale and Hot Shale Member is 0.28–30.89 (averaging 6.25) and 3.53–180.21

U
(averaging 44.75), respectively, which implies, not surprisingly, that the Hot Shale intervals have
AN
greater potential to generate liquid hydrocarbons than the rest of the formation (Table 1).
Rock-Eval Tmax data and vitrinite reflectance (Ro) are regarded as the most effective parameters
for assessing thermal maturity (Espitalié et al., 1977; Teichmüller and Durand, 1983; Peters and Cassa,
M

1994; Hunt, 1996). The maturity of Tanezzuft rocks cannot be determined using standard vitrinite
reflectance measurements because of the lack of vascular plant debris, so we use calculated vitrinite
D

reflectance Ro (Jarvie et al., 2001). This can be estimated from reliable Tmax data using the following
TE

calculation: Ro (%) = (0.0180*Tmax) – 7.16. Samples from the Tanezzuft Formation above the Hot
Shale have Tmax values between 433 and 439oC, with an average of 436oC. Those from the Hot Shale,
however, have relatively high Tmax values, ranging from 433 to 448oC and averaging 438oC (Table 1).
EP

The pyrolysis Tmax values and the calculated %Ro for the samples indicate early mature to mature
source rocks. This conclusion is consistent with the plot of Tmax versus production index (Fig. 6).
C

However, the Tanezzuft shales in the NC115 license area were probably not buried sufficiently for
petroleum generation during the Carboniferous Period (Echikh and Sola, 2000): this is likely to have
AC

taken place later, during the Mesozoic and Paleogene. In addition, uplift and erosion related to the
major Hercynian and Mesozoic tectonic phases particularly affected marginal parts of the present basin.

4.1.2. Bulk stable carbon isotopic compositions


Stable carbon isotope compositions have long been used to provide a detailed summary of the
provenance and history of organic matter and depositional environments (Sofer, 1984; Waples, 1985;
Gratzer et al., 2011; El Diasty et al., 2016). Carbon isotope compositions of source rock extracts from
9
ACCEPTED MANUSCRIPT

the NC115 Concession are presented in Table 2. The measured values are isotopically light (more
negative) and broadly similar. The δ13C isotope values of the saturated hydrocarbon fractions in the
extracts range from –29.65‰ to –28.28‰, and for aromatics from –29.03‰ to –28.10‰ (Table 2).
The presence of 13C-enriched organic matter from the Hot Shale samples in the H15 well may
be attributed to the high productivity of algae typical of microbial/algal mats (Popp et al., 1998;

PT
Schouten et al., 2001). On the other hand, the H28 Hot Shale sample is isotopically lighter than all
other extracts with δ13C of –29.65‰ and –29.03‰ (Table 2). This variance in δ13C values may be
attributed to both differences in organic matter input and the effect of thermal maturity (Lewan, 1983).

RI
The plot of stable carbon isotope values of saturated versus aromatic hydrocarbon fractions
(Fig. 7) shows that all extracts plot within the field of marine organic matter (Sofer, 1984). This is

SC
consistent with values of the canonical variable (CV) for the extracts, which range from −3.42 to −0.16
(Table 2) indicating non-waxy organic matter of marine origin.

U
AN
4.1.3. n-Alkanes and isoprenoids
Typical examples of saturated hydrocarbon chromatograms of the Hot Shale source rocks and
overlying Tanezzuft Formation are shown in Fig. 8. The hydrocarbon compositions of the latter and the
M

Hot Shale are mostly characterized by short chain n-alkane distributions in the C15–C35 range that
typically maximize at n-C18 or n-C19 (Fig. 8). A unimodal distribution pattern and predominance of
D

light normal alkanes <n-C20 in all of the samples suggests high input of marine-derived organic matter
TE

or thermal maturation that has led to cracking of the long chain n-alkanes (Cranwell et al., 1987;
Meyers, 1997; Peters et al., 2005). Additionally, the dominance of short-chain n-C17 compared to long-
chain n-alkanes (n-C27) is considered to indicate the presence of algae and/or plankton (Peters et al.,
EP

2005). The elevated baselines for the rock extracts in Fig. 8 can be explained by the relatively low
thermal maturity (early window). This is common in low maturity rock extracts and also in
C

biodegraded oil.
The carbon preference index (CPI) is affected by kerogen type and thermal maturity. It is the
AC

ratio of the odd carbon-numbered to the odd plus even carbon-numbered n-alkanes in the range n-C23 to
n-C30 (Scalan and Smith, 1970; Peters et al., 2005). The values vary from 1.40 to 1.69 for the Tanezzuft
extracts from above the Hot Shale, indicating low thermal maturity and/or the presence of a large
amount of terrigenous organic matter. No odd/even predominance was observed. The calculated CPI
values for three of the four Hot Shale samples are close to unity, the exception being sample NC115-
A5 at depth 1537 m, in which there is a slight odd-over-even preference (CPI = 1.18; Table 2)
attributable to minor terrigenous organic input. The CPI values of ≈1.0 for the Hot Shale samples
10
ACCEPTED MANUSCRIPT

suggest that at least the early oil window of thermal maturity was reached at the time of the expulsion
(Peters et al., 2005).
The shale samples have pristane/phytane values that range from 0.91 to 1.51, suggesting
suboxic conditions. Pr/Ph values of >1.0 may either be the result of oxidation during the early stages of
chlorophyll degradation in the upper oxic water column or of alternating oxic-anoxic conditions during

PT
sedimentation (Didyk et al., 1978). Fig. 9 shows a plot of pristane/n-C17 versus phytane/n-C18,
suggesting that the samples analyzed are non-biodegraded hydrocarbons generated from mixed Type-
II/III kerogen and that the depositional environments ranged from suboxic to mildly anoxic

RI
(Shanmugam, 1985).

SC
4.2. Geochemistry of crude oils
4.2.1. Bulk and elemental compositions

U
The quality of crude oil largely determines producibility and ease of refining, and is an important
AN
consideration when planning the exploration of any basin. The quality of a crude oil is related to such
bulk physicochemical characteristics as American Petroleum Institute gravity (API gravity), pour point,
and sulfur, nitrogen, and metal contents (Hughes and Holba, 1988). Analytical data on bulk property
M

indicators of oil quality for a suite of 10 crude oil samples from the A, B and H oil fields in the NC115
Concession are summarized in Table 2. All of the oil samples have high API gravities ranging from
D

39.6 to 41.9° and very low sulfur contents in the range of 0.11 to 0.14 wt%, so they can be classified as
TE

light, sweet and highly mature crude oils.


Organometallic porphyrins account for vanadium and nickel concentrations and V/Ni ratios can
be used to indicate palaeo-redox depositional conditions and to classify and correlate crude oil (Lewan
EP

and Maynard, 1982; Peters et al., 2005). The potential for Ni and V enrichment depends upon the
amount of tetrapyrroles preserved and is highest when the organic matter was deposited under anoxic
C

conditions. In addition, higher concentrations of V than Ni in crude oils is characteristic of anoxic or


euxinic source rocks (Lewan and Maynard, 1982). The predominance of V over Ni (V/Ni=2.15–6.01)
AC

in the oil samples studied is a result of the greater relative stability of vanadium porphyrins under low
Eh conditions associated with sulfate reduction during anoxic sedimentation and diagenesis of marine
source rocks (cf. Lewan and Maynard, 1982). The low relative concentrations of both V and Ni in the
samples analyzed, however, may imply suboxic–anoxic sedimentation (cf. Moldowan et al., 1986).

11
ACCEPTED MANUSCRIPT

4.2.2. Bulk stable carbon isotopic composition


The stable carbon isotope data for the saturated and aromatic hydrocarbon fractions show that the A17,
A37i, H38, H40, H39i and H44i oil samples are more enriched in 13C (Table 2; Fig. 7) and do not
overlap the B21, B31, B41 and B42 isotopic ranges. The A- and H-oils exhibit slightly more 13C-
enriched values (–29.51 to –28.92) than the more 13C-depleted B-oils (–30.14 to –29.38; Table 2).

PT
These isotopic differences potentially allow the separation of the oil samples into genetic families. The
depletion of 13C in saturated hydrocarbon fractions may result from increasing thermal alteration
(Lewan, 1983; Schoell, 1984; Peters et al., 2005), but the isotopic composition of aromatic fractions

RI
seems to be less subject to thermal maturity (Murillo et al., 2016).
A canonical variable (Sofer, 1984) of >0.47 indicates terrigenous organic matter, but when it is

SC
<0.47 a marine source is indicated. The negative CV values that range from -0.40 to -1.33 (Table 2) for
the analyzed oil samples suggest non-waxy oil and a predominantly marine environment at the time of

U
deposition of the source sediments.
AN
4.2.3. n-Alkanes and isoprenoid distributions
Typical gas chromatograms of total oil and mass chromatograms of saturated hydrocarbons (terpanes
M

and steranes) for the Hawaz and Mamuniyat oils are presented in Fig. 10. The GC profiles show similar
unimodal n-alkane distributions in all samples (Fig. 10), suggesting a single source rock or very similar
D

source facies. The normal alkane distributions for the samples are typically dominated by short-chain
TE

alkanes and with Pr/Ph ratios ranging from 1.86–2.20 (Table 2), which suggests deposition in suboxic
conditions and substantial terrigenous input (Didyk et al., 1978; Murray and Boreham, 1992; Peters et
al., 2005). The GC fingerprints of the B-oils show a predominance of pristane and can be readily
EP

distinguished by high pristane/phytane ratios. Acyclic isoprenoids decrease in abundance with


increasing maturity. The consistently higher ratios of isoprenoids to normal alkanes in the oil from the
C

NC115-B field (Table 2) suggest that it is slightly less mature than the oils from the NC115-A and H
oil fields.
AC

The carbon preference indices (CPI) for all of the oil samples are equal or close to 1.0 with
almost no odd-over-even carbon preference in n-alkanes, indicating that thermally mature rock was the
source of the oils. The ratio of the amount of a lower carbon number (n-C17) versus the higher carbon
number n-alkane (n-C27) varies between 7.14 and 20.00 with a mean of 10.53 (Table 2). This ratio
increases with maturity, and these results indicate that the analyzed oils are mature. The oils from the B
oil field show the lowest values of n-C17/n-C27 suggesting that they are the least mature of the oils.

12
ACCEPTED MANUSCRIPT

The Pr/n-C17 and Ph/n-C18 ratios for all 10 oils show consistently high maturation, but the B
field oils are slightly less mature than the others (Fig. 9). This figure also shows that mixed Type-II/III
kerogen dominated the source rocks for the oils.
4.2.4. Terpenoids
Results of alkane fraction gas chromatography-mass spectrometry are graphically displayed in Figs. 10

PT
and 11, and biomarker ratios are presented in Tables 3 and 4. All of the oil samples show features
typical of Lower Palaeozoic oils, such as high tricyclic terpane contents and dominance of C29 over C27
and C28 steranes. Tricyclic terpane ratios (C19/C23TR, C22/C21TR, C24/C23TR and C26/C25TR) and the

RI
tetracyclic terpane ratio (Tet/C23 TR) from source rock extracts and oils are useful to assess organic
input, maturity, and correlation of crude oils (Didyk et al., 1978; Farrimond et al., 1998; Samuel et al.,

SC
2010). Relative abundances of tricyclic terpanes within a related sample set show a dramatic increase at
late mature levels owing to the greater relative thermal stability of tricyclic terpanes (Farrimond et al.,

U
1998, 1999; Samuel et al., 2010).
AN
As indicated in Fig. 10, the m/z 191 fragmentogram is characterized by C23 tricyclic terpane as
the dominant homolog, followed by, in order of magnitude, C24>C21> C20 tricyclic terpanes, these
being the next most dominant peaks (Table 3). C23 tricyclic terpane is often the main peak in freshwater
M

lacustrine or marine crude oils, while C19 and C20 tricyclic terpanes are more abundant in terrigenous
oils (Peters et al., 2005).
D

The tetracyclic to tricyclic terpane ratio (C24Tet /C23TR) is primarily source-related (Table 3).
TE

The oil samples are broadly similar: the C24Tet/C23TR ratios range from 0.17 to 0.23, the only
exception being the NC115-H44i oil sample, which has a higher ratio (0.34). Hence, the low levels of
tetracyclic terpanes in the oils examined are indicative of carbonate-poor source rocks (cf. Connan et
EP

al., 1986; Clark and Philp, 1989; Peters et al., 2005).


The C27Ts/Tm ratio (Table 3) is considerably lower in oil samples from the B-field (1.06–1.17)
C

than those from the other oil fields (values range from 1.60–3.53, mean = 2.73). The ratio of Ts/Tm
increases with increasing maturity, suggesting that the B-field oil is less mature than that in the A- and
AC

H-oil fields.
The presence of high concentrations of 17α(H)-28,30-bisnorhopane relative to C30 hopane in the
H- and A-oils is taken to indicate high productivity in marine upwelling conditions (Curiale et al.,
1985), and that the oils were generated from marine source rocks containing abundant aquatic, algal
organic matter deposited in suboxic–anoxic settings. On the other hand, the low bisnorhopane from the
m/z 191 chromatograms (Fig. 10; Table 3) supports the interpretation that the NC115-H44i- and

13
ACCEPTED MANUSCRIPT

NC115-B-oils were derived from more terrigenous source rocks deposited under suboxic conditions
(Noble et al., 1985).
The norhopane/hopane ratio (Table 3) is related to both source type and maturity. Within a
group of oils from the same source, the relative abundance of the C29 norhopane increases as a function
of maturity. The low C29/C30 hopane ratios for the samples (< 0.7) from the m/z 191 mass

PT
chromatograms suggest a very low carbonate content in the source rock (Peters and Moldowan, 1991).
Derivation from carbonate-poor source rocks is consistent with low C31R/H ratios ≤0.20 for the
samples (Table 3).

RI
Diahopanes are thermally more stable than regular hopanes. The C30 diahopane/C30 hopane
ratio increases with increasing maturity. Again, the B-field oils (C30X/H = 0.13–0.14; Table 3) show

SC
lower maturity than the other oils. The A- and H-oils have high diahopane/hopane ratios, indicating
generation from a clay-rich, marine source rock reflecting deposition in a suboxic environment. They

U
contain few or no extended hopanes over C31 (Fig. 10; Table 3). By contrast, the B-field oils contain
AN
extended hopanes up to and including the C34 homolog. The lack of extended hopanes indicates
relatively higher maturity owing to preferential generation of shorter-chain hopanes, and again points to
the B-field oils being less mature than the other three analyzed oils. Because the C34 and C35
M

homohopanes were not detected in the m/z 191 mass chromatograms of the H- and A-oils, the C35/C34
homohopane ratio could not be calculated (Table 3). The absence of C34 and C35 homohopanes implies
D

that the organic matter was deposited under suboxic conditions (Peters and Moldowan, 1991;
TE

Sinninghe Damsté et al., 1995).


The extended hopanes exist as two isomers, 22R and 22S. The former is biologically inherited
and the latter forms by thermally-induced structural rearrangement (Peters et al., 2005). The ratio of the
EP

22S to 22R isomers therefore increases with increasing maturity from 0 to about 0.6 (0.57–
0.62=endpoint). Values available for the C32 extended hopanes show that the maturity of the B-field
C

oils (%C32 22S = 0.64) is similar to that of the A- and H-fields (%C32 22S = 0.64–0.69), but that all
have reached at least the beginning of the oil window (Mackenzie et al., 1982). This interpretation is
AC

further supported by the low moretane/hopane (M/H, Table 4) ratios (Table 4).
The biomarker gammacerane is commonly associated with high salinity and can be used as an
indicator of photic-zone anoxia and a stratified water column during the deposition of the sediments
that now comprise the source rock (Sinninghe Damsté et al., 1995; Peters et al., 2005). In this study, the
gammacerane/C30-hopane ratios are extremely low, ranging from 0.04 to 0.08 (Table 3). This suggests
a lack of water-column stratification and normal salinity. Low gammacerane indices are often
associated with relatively high Pr/Ph ratios (Table 2).
14
ACCEPTED MANUSCRIPT

4.2.5. Steroids
Fig. 10 shows m/z 217 mass fragmentograms for saturated hydrocarbon fractions from three
representative oil samples in the A-, B- and H-fields. The sterane baselines are slightly elevated. Rising
baselines (especially for the terpanes) are expected because these oils are highly mature and show
elevated API gravity (Table 2). Based on the close similarity of homologous distributions of steranes

PT
C27-C28-C29 (Fig. 10; Table 3), the oils could have originated from one source rock or similar source
rocks. The sterane mass chromatograms (m/z 217) are characterized by an abundance of C29 regular
steranes compared to the C27 and C28 homologues. The relative abundances of the three regular sterane

RI
groups is approximately 50% C29, 30% C27 and 20% C28 (Table 3), which is typical of many Lower
Palaeozoic oils (Grantham and Wakefield, 1988; Schwark and Empt, 2006).

SC
The ratio of C28 to C29 regular steranes is a function of geological time as a result of
evolutionary changes in photosynthetic organisms (Grantham and Wakefield, 1988). The increase in

U
C28/C29 sterane ratios is attributed to an episodic change in the green algal community from primitive
AN
C29-sterane-producing forms to C28-sterane-producing prasinophytes (Schwark and Empt, 2006). Table
3 shows distinct C28/C29 sterane ratios. Although similar on the Grantham and Wakefield (1988) plot,
the data suggest that the A- and H-oils (C28/C29 = 0.55–0.59) may be from a younger source rock than
M

the B-oils (C28/C29 ≈ 0.49), although both source rocks appear to be Palaeozoic. Where it could be
calculated for the NC115 oils, the C28/C29 regular sterane ratio is <0.55, which is typical of Lower
D

Palaeozoic marine source rocks containing green algae, with possibly some contribution from
TE

acritarchs (Schwark and Empt, 2006). Above the Devonian/Carboniferous transition the C28/C29-sterane
ratios are >0.55.
The C30 steranes, which indicate a marine depositional environment (Moldowan et al., 1985;
EP

Peters et al., 2005), could not be detected in any of the oil samples analyzed. However, consistent with
the other sterane distributions, the overwhelming dominance of diasteranes (2.25–2.83) also indicates
C

an origin from siliciclastic-rich source rocks (Mackenzie et al., 1982; Peters et al., 2005) and high
thermal maturity.
AC

The sterane/hopane ratio is a measure of eukaryotic versus prokaryotic input to the source rocks
(Peters et al., 2005). Steranes are more abundant in marine oils and less significant in terrigenous oils.
The sterane/hopane ratios for the oil samples from the NC115 Concession range from 2.10 to 4.83
(Table 3), suggesting significant contribution of planktonic and algal material relative to bacterial
organic matter, which is supported by the sterane/terpane ratios (Table 3).
The C29 sterane isomerization parameter reflects an increase in abundance of the thermally
more stable 20S configuration relative to the biologically derived 20R configuration. The ratios of
15
ACCEPTED MANUSCRIPT

20S/(S+R) for the C29 steranes in the studied oils are within the range of 0.46–0.47 (endpoint ~0.52–
0.55), and for ββ/(αα+ββ) the range is 0.53–0.55 (endpoint ~0.67–0.71: Table 4), suggesting that they
were generated from source rocks approaching the peak stage (Fig. 11A) of oil generation (Peters et al.,
2005). Visual inspection of the m/z 217 mass fragmentograms (Fig. 10; Table 4) reveals that the sterane
isomerization parameters have nearly identical ranges and are clearly seen as one cluster on Fig. 11A

PT
for all A-, B- and H-oils.

4.2.6. Aromatic biomarkers

RI
Results of aromatic fraction gas chromatography-mass spectrometry are presented in Tables 3 and 4.
The ratio of dibenzothiophene to phenanthrene (DBT/P) is a depositional environment indicator. All

SC
DBT/P ratios are similar, indicating a similar source for all of the oils, and very low values (0.02–0.07)
together with a very low sulfur content suggest an origin from clay-rich rather than carbonate source

U
rocks for the oils and marine/lacustrine depositional environments (Hughes et al., 1995).
AN
A number of methylphenanthrene-based maturation parameters have been proposed, together
with an empirical correlation against vitrinite reflectance. Consequently, different empirical
correlations of the methylphenanthrene ratio (MPI) are applicable, depending upon the maturity range
M

(Table 4). This allows an estimate of the equivalent reflectivity to be calculated from the ratios (Radke,
1988; Sofer et al., 1993). Two such ratios are available for the analyzed oils, and the equivalent
D

vitrinite reflectivities for the oils are indicated (%Rc1 and %Rc2). The values of MPI (1.00–1.07)
TE

together with maturity determined using the calculated vitrinite reflectance (%Rc1= 0.77–0.80%;
%Rc2=0.97–1.01%) indicate that the oils from the NC115 Concession have similar maturity levels and
were generated near the peak of the oil window. Based on the assumption of rapid burial during the
EP

Permian, Aziz (2000) suggested that oil generation in the concession area began during the
Carboniferous–Permian. Moreover, the relatively high thermal maturity in this area is believed to be a
C

result of a heat pulse caused by tectonism and volcanism during the Eocene Epoch (Echikh and Sola,
2000). Their present-day geothermal gradient map of the Murzuq Basin shows a high heat-flow trend
AC

extending southeastward from the NC115 region, which may indicate the area most affected by the
Eocene heating event. This is consistent with the methylphenanthrene distribution fractions F1 and F2
(Kvalheim et al., 1987) and the dibenzothiophene/C4 naphthalene (DBT/C4N) ratios (Table 4). Many
ratios involving naphthalene compounds are maturation dependent. The dibenzothiophene/C4
naphthalene (DBT/C4N) ratios increase with increasing maturity. All consistently show the B-field oil
to have the lowest ratio. Likewise, the MDR ratio compares the abundance of 4-
methyldibenzothiophene to 1-methyldibenzothiophene (Table 4), a parameter that increases with
16
ACCEPTED MANUSCRIPT

increasing maturity. The results show that the B-field oil samples have lower MDR ratios than the other
oil samples and are thus the least mature.
The relative abundance of the lower molecular-weight (C20 and C21) triaromatic steroids
increases with increasing maturity (Mackenzie et al., 1981) related to thermally induced scission and
loss of the side chain from the C26, C27 and C28 components. The relative abundances of the C26, C27

PT
and C28 triaromatic steroids cannot be calculated directly because of the coelution of C26 and C27
components (TA26R and TA27S). Since the relative abundances are a closed data set, the C26/C28 and
C27/C28 triaromatic steroid ratios (Table 4) indicate the variation within a sample set, which may be

RI
useful for correlation.

SC
4.2.7. Oil–oil correlation and cluster analyses of crude oils
For the chemometrics, we used 21 source-related biomarker and isotope ratios (Peters et al., 2005,

U
2016): C19/C23, C22/C21, C24/C23, C26/C25, Tet/C23, C28/H, C29/H, C30X/H, C31R/H, Ga/H, DiaS/Reg,
%C27, %C28, %C29, S/H, C27Ts/Tm, DBT/P, V/Ni, δ13Csat, δ13Caro, and CV (Tables 2 and 3). We
AN
omitted C35/C34 because there were five empty cells (0 or NA). Ster/Terp was also omitted because it is
similar to S/H (Table 3). We selected these 21 source-related parameters by looking at a large number
M

of such parameters (many are listed in Peters et al., 2005), but later removed ratios that showed little or
no difference between the samples.
D

Different source parameters are required, depending on the basin and facies of the source rock.
TE

The PCA output is shown in Fig. 12B, where the first three principal components account for 87.2% of
variance in the data (Factor 1 = 54.3%, Factor 2 = 24.8% and Factor 3 = 8.1%). Both the HCA
dendrogram (Fig. 12A) and PCA plot (Fig. 12B) identify two genetically distinct oil families with one
EP

outlier for the 10 oil samples from the NC115 Concession. Although the three oil families have roughly
similar biomarker characteristics, they may originate from different source rocks or from different
C

organic facies of the same source rock.


AC

4.3. Family-I oils


Four oil samples, B21, B31, B41 and B42, from the Ordovician Mamuniyat reservoirs in the NC115-B
field were assigned to Family-I (Tables 2–4; Figs. 12A, B, 13). Biomarker characteristics suggest that
they were generated from shale or marl-rich organofacies containing mixed Type-II/III kerogen
deposited under suboxic marine conditions.
B-field oils of this family have the lightest stable carbon isotopic ratios for saturates (−30.38 to
−30.14 ‰) and aromatics (−29.58 to −29.38 ‰) compared to the other oil families (Table 2; Fig. 7).
17
ACCEPTED MANUSCRIPT

Oils in Family-I are typified by the lowest C27Ts/Tm, diahopane/hopane, diasterane/regular steranes,
C31R/hopane, C22/C21, C24/C23, C19/C23 and C26/C25 tricyclic terpanes, %C28 steranes (<24%) and
steranes/hopanes, and the highest %C29 steranes (47.27–48.58%; Table 3; Fig. 13).
Fig. 13 is intriguing because the families show distinct C28/C29 sterane ratios. The data suggest
that the B-oils may be from an older source rock than the other oils, although both source rocks are

PT
apparently Palaeozoic (Grantham and Wakefield, 1988). Based on saturated and aromatic maturity
parameters, all ratios consistently show that the B-field samples are the least mature of the oils from the
NC115 fields.

RI
4.4. Family-II oils

SC
This family consists mainly of A- and H-oils, namely the A-17, H37i, H38, H39i and H40 samples
(Figs. 12, 13). They show differences in bulk carbon isotope composition and geochemical

U
characteristics, which suggest local variations in the age and quality of the source rock, but they have
AN
geochemical features that appear to be related to low-sulfur Type-II/III kerogen deposited in a suboxic–
anoxic marine depositional environment.
The stable carbon isotope data for the samples are isotopically light compared to other oils
M

worldwide (saturates −29.76 to −29.51‰, and aromatics −29.01 to −28.92‰), typical of oils generated
from lower Palaeozoic rocks (Chung et al., 1992). A plot of stable carbon isotope ratios for the
D

saturated versus aromatic hydrocarbon fractions of the oils shows a relatively tight cluster indicating a
TE

separate family and similar source material (Fig. 7). These oils are characterized by prominent 28,30-
bisnorhopane (Table 3; Fig. 13), very high diasterane/sterane ratios (>2.0) and high sterane/hopane
ratios (mostly >4.0) typical of clay-rich, marine source rocks that reflect deposition under suboxic–
EP

anoxic conditions (Curiale and Odermatt, 1989; Peters et al., 2005). Based on methyl-(C28) and ethyl-
(C29) cholestane relative abundances (Fig. 10; Grantham and Wakefield, 1988), these oils may originate
C

from Palaeozoic source rocks younger than those for Family-I.


The oils in this family appear to have been generated at or past the peak of oil generation from
AC

organic matter of mixed origins based on API gravity, isoprenoids/normal alkanes, hopane
isomerization %C32 22S, absence of extended hopanes, very high Ts/Tm ratio, relative amounts of
moretane, high percentage of triaromatic steroids (Table 4; Fig. 11B), methyldibenzothiophene ratios,
methylphenanthrene index and estimated vitrinite reflectance maturities.

18
ACCEPTED MANUSCRIPT

4.5. Family-III oil


Only one oil sample, NC115-H44i, belongs to this family. The biomarker distribution has a mixed
character, but is similar to Family-II oils (Figs. 10, 12A, B, 13).

4.6. Oil–source rock correlation

PT
The relatively small number of available rock samples did not allow a reliable oil–source correlation.
Based on GC and carbon isotope data, the source rock extracts indicate the same kerogen type as the
oils, and the stable carbon isotope ratio of the H28 Hot Shale extract looks very similar to Family-II

RI
and -III oils (Figs. 7–9).
It is also notable that the oils from the A-, B- and H-fields are all relatively mature, although the

SC
B-field oil is slightly less mature than the others. The measured maturity level for the Hot Shale and
overlying Tanezzuft source rocks is lower than that for the oils based on the biomarker maturity

U
indicators and the maturity equivalent vitrinite reflectance. This means that oil generation must have
AN
occurred outside the NC115 Concession with oil accumulating there after migration from the mature
Hot Shale ‘kitchen’.
M

5. Conclusions
D

Organic geochemical characterization of lower Silurian source rocks and Ordovician-reservoired crude
TE

oils from the NC115 Concession in the Murzuq Basin in southwest Libya reveals clear differences in
source rock organic matter input, depositional environment and thermal maturity. The following
conclusions can be drawn.
EP

The Hot Shale Member of the Tanezzuft Formation is rich in organic matter whereas the
overlying deposits are commonly organic-lean. Variations in the composition of the organic matter,
C

redox conditions and organic productivity have had a significant impact on the hydrocarbon generation
potential of the samples analyzed. The Hot Shale is dominated by Type-II/III kerogen whereas the
AC

overlying deposits contain mainly Type-III to Type-II/III kerogen. For this reason, the Hot Shale has
considerable potential to produce liquid hydrocarbons, whereas the overlying deposits have the
potential to generate mostly gas, although a mixture of oil and gas cannot be ruled out.
Rock-Eval Tmax and PI maturity data show that all of the samples analyzed are sufficiently
mature for the generation of oil and gas. The stable carbon isotope signature of the rock extracts
suggests a marine environment and non-waxy organic matter. The distribution of n-alkanes and
isoprenoids in both the Hot Shale and overlying Tanezzuft extracts confirms the preservation of organic
19
ACCEPTED MANUSCRIPT

matter of mixed origins in the source rocks, and also suggests that mildly anoxic to suboxic conditions
prevailed at the time of deposition of the sediments.
Bulk and elemental compositions suggest that the NC115 A-, B- and H-oil fields contain sweet,
light crude oils. The depletion or enrichment of carbon isotope compositions easily differentiate these
oils. This isotope difference may be a result of both increasing thermal maturity and type of organic

PT
matter.
All terpane and sterane biomarker ratios point to marine shale source rocks that contain varying
proportions of mixed terrigenous and aquatic organic matter deposited under weakly reducing to

RI
suboxic conditions. Based on saturated and aromatic maturity parameters, the oil samples analyzed
reached an advanced level of thermal maturity. The B-oils are the least mature. Because of the

SC
difference in maturity between the oils and source rocks, it is reasonable to assume that the oils
originated from the Hot Shale and migrated up-dip to the NC115 Concession where at least two

U
genetically distinct oil families and one outlier are recognized on the basis of source-related
AN
geochemical parameters compared using HCA and PCA. Although the three families are similar on the
Grantham and Wakefield plot, the data suggest that the Family-II and -III oils may be from a younger
Palaeozoic source rock than the Family-I oils.
M

Acknowledgements
D
TE

We are indebted to the staff of National Oil Corporation (NOC) and Akakus Oil Operations (formerly
REPSOL oil operations) in Tripoli, Libya, for providing the rocks and crude oil samples used in this
study. We also thank the laboratories of GeoMark Research (Houston, Texas) and StratoChem Services
EP

(New Maadi, Cairo) for their continuing analytical support. We are grateful to the Associate Editor
Michael Abrams and two anonymous reviewers for constructive comments on an earlier version of the
C

manuscript.
AC

References

Armstrong, H.A., Turner, B.R., Makhlouf, I.M., Weedon, G.P., Williams, M., Al Smadi, A., Abu
Salah, A., 2005. Origin, sequence stratigraphy and depositional environment of an upper Ordovician
(Hirnantian) deglacial black shale, Jordan. Palaeogeography, Palaeoclimatology, Palaeoecology 220,
273–289.

20
ACCEPTED MANUSCRIPT

Aziz, A., 2000. Stratigraphy and hydrocarbon potential of the Lower Palaeozoic succession of license
NC-115, Murzuq Basin, SW Libya. In: Sola, M.A., Worsley, D. (Eds.), Geological Exploration in
Murzuq Basin. Elsevier, Amsterdam, pp. 349–368.
Belaid, A., Krooss, B.M., Littke, R., 2010. Thermal history and source rock characterization of a
Paleozoic section in the Awbari Trough, Murzuq Basin, SW Libya. Marine and Petroleum Geology

PT
27, 612–632.
Bellini, E., Massa, D., 1980. A stratigraphic contribution to the Palaeozoic of the southern basins of
Libya. In: Salem, M.J., Busrewil, M.T. (Eds.), The Geology of Libya. Academic Press, London, pp.

RI
3–56.
Boote, D.R.D., Clark-Lowes, D.D., Traut, M.W., 1998. Palaeozoic petroleum systems of North Africa.

SC
In: MacGregor, D.S., Moody, R.T.J., Clark-Lowes, D.D. (Eds.), Petroleum Geology of North
Africa. Geological Society, London, Special Publication 132, pp. 7–68.

U
Butcher, A., 2009. Early Llandovery chitinozoans from Jordan. Palaeontology 52, 593–629.
AN
Butcher, A., 2013. Chitinozoans from the middle Rhuddanian (lower Llandovery, Silurian) ‘hot’ shale
in the E1-NC174 core, Murzuq Basin, SW Libya. Review of Palaeobotany and Palynology 198, 62–
91.
M

Chung, H.M., Rooney, M.A., Toon, M.B., Claypool, G.E., 1992. Carbon isotope composition of
marine crude oils. American Association of Petroleum Geologists Bulletin 76, 1000–1007.
D

Clark, J.P., Philp, R.P., 1989. Geochemical characterization of evaporite and carbonate depositional
TE

environments and correlation of associated crude oils in the Black Creek Basin, Alberta. Canadian
Petroleum Geology Bulletin 37, 401–416.
Connan, J., Bouroullec, J., Dessort, D., Albrecht, P., 1986. The microbial input in carbonate-anhydrite
EP

facies of a sabkha palaeoenvironment from Guatemala: a molecular approach. Organic


Geochemistry 10, 29–50.
C

Cranwell, P., Eglinton, G., Robinson, N., 1987. Lipids of aquatic organisms as potential contributors to
lacustrine sediments – II. Organic Geochemistry 11, 513–527.
AC

Curiale, J.A., Cameron, D., Davis, D.V., 1985. Biological marker distribution and significance in oils
and rocks of the Monterey Formation, California. Geochimica et Cosmochimica Acta 49, 271–288.
Curiale, J.A., Odermatt, J.R., 1989. Short-term biomarker variability in the Monterey Formation, Santa
Maria Basin. Organic Geochemistry 14, 1–13.
Davidson, L., Beswetherick, S., Craig, J., Eales, M., Fisher, A., Himmali, A., Jhoon, J., Mejrab, B.,
Smart, J., 2000. The structure, stratigraphy and petroleum geology of the Murzuq Basin, southwest

21
ACCEPTED MANUSCRIPT

Libya. In: Sola, M.A., Worsley, D. (Eds.), Geological Exploration in Murzuq Basin. Elsevier,
Amsterdam, pp. 295–320.
de Gibert, J.M., Ramos, E., Marzo, M., 2011. Trace fossils and depositional environments in the Hawaz
Formation, Middle Ordovician, western Libya. Journal of African Earth Sciences 60, 28–37.
Desio, A., 1936. Prime notizie sulla presenza del Silurico fossilifero nel Fezzan. Bolletino della Società

PT
Geologica Italiana 55, 116–120.
Didyk, B.M., Simoneit, B.R.T., Brassell, S.C., Eglinton, G., 1978. Organic geochemical indicators of
palaeoenvironmental conditions of sedimentation. Nature 272, 216–222.

RI
Echikh, K., Sola, M.A., 2000. Geology and hydrocarbon occurrences in the Murzuq Basin, SW Libya.
In: Sola, M.A., Worsley, D. (Eds.), Geological Exploration in Murzuq Basin. Elsevier, Amsterdam,

SC
pp. 175–222.
El Diasty, W. Sh., El Beialy, S.Y., Littke, R., Farag., F.A., 2016. Source rock evaluation and nature of

U
hydrocarbons in the Khalda Concession, Shushan Basin, Egypt's Western Desert. International
AN
Journal of Coal Geology 162, 45–60.
El-ghali, M.A.K., 2005. Depositional environments and sequence stratigraphy of paralic glacial,
paraglacial and postglacial Upper Ordovician siliciclastic deposits in the Murzuq Basin, southwest
M

Libya. Sedimentary Geology 177, 145–173.


Espitalié, J., LaPorte, J.L., Madec, M., Marquis, F., LePlat, P., Paulet, J., Boutefeu, A., 1977. Rapid
D

method for source rocks characterization and for determination of petroleum potential and degree of
TE

evolution. Oil and Gas Science and Technology 32, 23–42.


Farrimond, P., Bevan, J.C., Bishop, A.N., 1999. Tricyclic terpane maturity parameters: response to
heating by an igneous intrusion. Organic Geochemistry 30, 1011–1019.
EP

Farrimond, P., Taylor, A., Telnaes, N., 1998. Biomarker maturity parameters: the role of generation
and thermal degradation. Organic Geochemistry 29, 1181–1197.
C

Finney, S.C., Berry, W.B.N., 1997. New perspectives on graptolite distributions and their use as
indicators of platform margin dynamics. Geology 25, 919–922.
AC

Gundobin, V.N., 1985. Geological Map of Libya Sheet NH-33 (Qararat al Marar) and Explanatory
Booklet. Industrial Research Centre, Tripoli.
Goudarzi, G.H., 1980. Structure–Libya. In: Salem, M.S., Busrewil, M.T. (Eds.), The Geology of Libya
3. Academic Press, London, pp. 879–892.
Grantham, P.J., Wakefield, L.L., 1988. Variations in the sterane carbon number distributions of marine
source rock derived crude oils through geological time. Organic Geochemistry 12, 61–73.

22
ACCEPTED MANUSCRIPT

Gratzer, R., Bechtel, A., Sachsenhofer, R.F., Linzer, H-G., Reischenbacher, D., Schulz, H-M., 2011.
Oil-oil and oil-source rock correlations in the Alpine Foreland Basin of Austria: insights from
biomarker and stable carbon isotope studies. Marine and Petroleum Geology 28, 1171–1186.
Hall, P.B., Bjøroy, M., Ferriday, I.L., Ismail, Y., 2010. Libyan Murzuq Basin source rocks. Search and
Discovery Article #10271, expanded abstract presentation at AAPG Annual Convention and

PT
Exhibition, New Orleans, Louisiana, April 11–14, pp. 1–24.
Hallett, D., 2002. Petroleum Geology of Libya. Elsevier, Amsterdam, 503 p.
Hallett, D., Clark-Lowes, D., 2016. Petroleum Geology of Libya. Second Edition, Elsevier,

RI
Amsterdam, 392 p.
Hodairi, T., Philp, P., 2011. Geochemical investigation of Tanezzuft Formation, Murzuq Basin, Libya.

SC
Search and Discovery Article #10344, expanded abstract presentation at AAPG Annual Convention
and Exhibition, Houston, Texas, USA, April 10–13, pp. 1–22.

U
Hodairi, T., Philp, P., 2012. Biomarker characteristics of crude oils from the Murzuq Basin, SW Libya.
AN
Journal of Petroleum Geology, 35, 255–272.
Hughes, W.B., Holba, A.G., 1988. Relationship between crude oil quality and biomarker patterns.
Organic Geochemistry 13, 15–30.
M

Hughes, W.B., Holba, A.G., Dzou, L.I., 1995. The ratios of dibenzothiophene to phenanthrenes and
pristane to phytane as indicators of depositional environment and lithology of petroleum source
D

rocks. Geochimica et Cosmochimica Acta 59, 3581–3598.


TE

Hunt, J., 1996. Petroleum Geochemistry and Geology. 2nd edition, Freeman and Company, New York,
743 p.
İnan, S., Goodarzi, F., Mamm, A.S., Arouri, K., Qathami, S., Ardakani, O.H., İnan, T., Tuwailib, A.A.,
EP

2016. The Silurian Qusaiba Hot Shales of Saudi Arabia: an integrated assessment of thermal
maturity. International Journal of Coal Geology 159, 107–119.
C

Jarvie, D.M., Morelos, A., Han, Z., 2001. Detection of pay zones and pay quality, Gulf of Mexico:
application of geochemical techniques: Gulf Coast Association of Geological Societies,
AC

Transactions 51, 151–160.


Klemme, H.D., Ulmishek, G.F., 1991. Effective petroleum source rocks of the world: stratigraphic
distribution and controlling depositional factors. American Association of Petroleum Geologists
Bulletin 75, 1809–1851.
Kvalheim, O.M., Christy, A.A., Telnaes, N., Bjørseth, A., 1987. Maturity determination of organic
matter in coals using the methylphenanthrene distribution. Geochimica et Cosmochimica Acta 51,
1883–1888.
23
ACCEPTED MANUSCRIPT

Le Heron, D.P., Armstrong, H.A., Wilson, C., Howard, J.P., Gindre, L., 2010. Glaciation and
deglaciation of the Libyan Desert: the Late Ordovician record. Sedimentary Geology 223, 100–125.
Lewan, M.D., 1983. Effects of thermal maturation on stable organic carbon isotopes as determined by
hydrous pyrolysis of Woodford Shale. Geochimica et Cosmochimica Acta 47, 1471–1479.
Lewan, M.D., Maynard, J.B., 1982. Factors controlling enrichment of vanadium and nickel in the

PT
bitumen of organic sedimentary rocks. Geochimica et Cosmochimica Acta 46, 2547–2560.
Loydell, D.K., Butcher, A., Frýda, J., 2013. The middle Rhuddanian (lower Silurian) ‘hot’ shale of
North Africa and Arabia: an atypical hydrocarbon source rock. Palaeogeography,

RI
Palaeoclimatology, Palaeoecology 386, 233–256.
Loydell, D.K., Butcher, A., Frýda, J., Lüning, S., Fowler, M., 2009. Lower Silurian “hot shales” in

SC
Jordan: a new depositional model. Journal of Petroleum Geology 32, 261–270.
Lüning, S., Craig, J., Fitches, W., Mayouf, J., Busrewil, A., El Dieb, M., Gammudi, A., Loydell, D.,

U
McIlroy, D., 1999. Re-evaluation of the petroleum potential of the Kufra Basin (SE Libya, NE
AN
Chad): does the source rock barrier fall? Marine and Petroleum Geology 16, 693–718.
Lüning, S., Craig, J., Loydell, D.K., Štorch, P., Fitches, W., 2000. Lower Silurian ‘hot shales’ in North
Africa and Arabia: regional distribution and depositional model. Earth-Science Reviews 49, 121–
M

200.
Lüning, S., Kolonic, S., Loydell, D.K., Craig, J., 2003. Reconstruction of the original organic richness
D

in weathered Silurian shale outcrops (Murzuq and Kufra basins, southern Libya). GeoArabia 8, 299–
TE

308.
Lüning, S., Shahin, Y.M., Loydell, D., Al-Rabi, H.T., Masri, A., Tarawneh, B., Kolonic, S., 2005.
Anatomy of a world-class source rock: distribution and depositional model of Silurian organic-rich
EP

shales in Jordan and implications for hydrocarbon potential. American Association of Petroleum
Geologists Bulletin 89, 1397–1427.
C

Mackenzie, A.S., Brassell, S.C., Eglinton, G., Maxwell, J.R., 1982. Chemical fossils: the geological
fate of steroids. Science 217, 491–504.
AC

Mackenzie, A.S., Hoffmann, C.F., Maxwell, J.R., 1981. Molecular parameters of maturation in the
Toarcian shales, Paris Basin, France. III. Changes in aromatic steroid hydrocarbons. Geochimica et
Cosmochimica Acta 45, 1345–1355.
Mamgain, V.D., 1980. The pre-Mesozoic (Precambrian to Palaeozoic) stratigraphy of Libya: a
reappraisal. Industrial Research Centre Bulletin 14, 104 p.
Massa, D., Collomb, G.R., 1960. Observations nouvelles sur la région d’Aouinet Ouenine et du Djebel
Fezzan (Libye). 21st International Geological Congress, Copenhagen, Proceedings 12, 65–73.
24
ACCEPTED MANUSCRIPT

McGillivray, J.G., Husseini, M.I., 1992. The Paleozoic petroleum geology of Central Arabia. American
Association of Petroleum Geologists Bulletin 76, 1473–1490.
Meinhold, G., Whitham, A.G., Howard, J.P., Stewart, J.C., Abutarruma, Y., Thusu, B., 2013.
Hydrocarbon source rock potential of latest Ordovician–earliest Silurian Tanezzuft Formation shales
from the eastern Kufra Basin, SE Libya. Journal of Petroleum Geology 36, 105–115.

PT
Meyers, P.A., 1997. Organic geochemical proxies of paleoceanographic, paleolimnologic, and
paleoclimatic processes. Organic Geochemistry 27, 213–250.
Moldowan, J.M., Seifert, W.K., Gallegos, E.J., 1985. Relationship between petroleum composition and

RI
depositional environment of petroleum source rocks. American Association of Petroleum Geologists
Bulletin 69, 1255–1268.

SC
Moldowan, J.M., Sundararaman, P., Schoell, M., 1986. Sensitivity of biomarker properties to
depositional environment and/or source input in the Lower Toarcian of SW-Germany. Organic

U
Geochemistry 10, 915–926.
AN
Murillo, W.A., Vieth-Hillebrand, A., Horsfield, B., Wilkes, H., 2016. Petroleum source, maturity,
alteration and mixing in the southwestern Barents Sea: new insights from geochemical and isotope
data. Marine and Petroleum Geology 70, 119–143.
M

Murray, A.P., Boreham, C.J., 1992. Organic geochemistry in petroleum exploration. Australian
Geological Survey Organization, Canberra, 230 p.
D

Noble, R.A., Alexander, R., Kagi, R.I., Knox, J., 1985. Tetracyclic diterpenoid hydrocarbons in some
TE

Australian coals, sediments and crude oils. Geochimica et Cosmochimica Acta 49, 2141–2147.
Parrish, J.T., 1982. Upwelling and petroleum source beds, with reference to Paleozoic. American
Association of Petroleum Geologists Bulletin 66, 750–774.
EP

Peters, K.E., Cassa, M.R., 1994. Applied source rock geochemistry. In: Magoon, L.B., Dow, W.G.
(Eds.), The Petroleum System–from Source to Trap. American Association of Petroleum Geologists
C

Memoir 60, pp. 93–120.


Peters, K.E., Creaney, S., 2004. Geochemical differentiation of Silurian from Devonian crude oils in
AC

eastern Algeria. In: Geochemical Investigations in Earth and Space Science: A Tribute to Isaac R.
Kaplan. The Geochemical Society, Special Publication Series 9, 287–301.
Peters, K.E., Moldowan, J.M., 1991. Effects of source, thermal maturity, and biodegradation on the
distribution and isomerization of homohopanes in petroleum. Organic Geochemistry 17, 47–61.
Peters, K.E., Walters, C.C., Moldowan, J.M., 2005. The Biomarker Guide. Cambridge University
Press, Cambridge, 1155 p.

25
ACCEPTED MANUSCRIPT

Peters K.E., Wright, T.L., Ramos, L.S., Zumberge, J.E., 2016. Chemometric recognition of genetically
distinct oil families in the Los Angeles Basin, California. American Association of Petroleum
Geologists Bulletin 100, 115–135.
Pierobon, E.S.T., 1991. Contribution to the stratigraphy of the Murzuq Basin, SW Libya. In: Salem,
M.J. Belaid, M.M. (Eds.), The Geology of Libya. Elsevier, Amsterdam, pp. 1767–1783.

PT
Popp, B.N., Laws, E.A., Bidigare, R.R., Dore, J.E., Hanson, K.L., Wakeham, S.G., 1998. Effect of
phytoplankton cell geometry on carbon isotope fractionation. Geochimica et Cosmochimica Acta
62, 69–77.

RI
Radke, M., 1988. Application of aromatic compounds as maturity indicators in source rocks and crude
oils. Marine and Petroleum Geology 5, 224–236.

SC
Ramos, E., Marzo, M., de Gibert, J.M., Tawengi, K.S., Khoja, A.A., Bolatti, N.D., 2006. Stratigraphy
and sedimentology of the Middle Ordovician Hawaz Formation (Murzuq Basin, Libya). American

U
Association of Petroleum Geologists Bulletin 90, 1309–1336.
AN
RRI (Robertson Research International), 1998. Source rock geochemistry and burial history study,
NC115 Concession, Murzuq Basin, Libya. Project no. Ic/GN229, internal report 384 p.
Rusk, D.C., 2001. Libya: petroleum potential of the underexplored basin centers—a twenty-first-
M

century challenge. In: Downey, M.W., Threet, J.C., Morgan, W.A. (Eds.), Petroleum Provinces of
the Twenty-First Century. American Association of Petroleum Geologists Memoir 74, pp. 429–452.
D

Samuel, O.J., Kildahl-Andersen, G., Nytoft, H.P., Johansen, J.E., Jones, M., 2010. Novel tricyclic and
TE

tetracyclic terpanes in Tertiary deltaic oils: structural identification, origin and application to
petroleum correlation. Organic Geochemistry 41, 1326–1337.
Scalan, R.S., Smith, J.E., 1970. An improved measure of the odd-to-even predominance in the normal
EP

alkanes of sediment extracts and petroleum. Geochimica et Cosmochimica Acta 34, 611–620.
Schoell, M., 1984. Stable isotopes in petroleum research. In: Books, J., Welte, D.H. (Eds.), Advances
C

in Petroleum Geochemistry 1. Academic Press, London, pp. 215–245.


Schouten, S., Hartgers, W.A., Lopez, J.F., Grimalt, J.O., Sinninghe Damsté, J.S., 2001. A molecular
AC

isotopic study of 13C-enriched organic matter in evaporitic deposits: recognition of CO2-limited


ecosystems. Organic Geochemistry 32, 277–286.
Schwark, L., Empt, P., 2006. Sterane biomarkers as indicators of Palaeozoic algal evolution and
extinction events. Palaeogeography, Palaeoclimatology, Palaeoecology 240, 225–236.
Shanmugam, G., 1985. Significance of coniferous rain forests and related organic matter in generating
commercial quantities of oil, Gippsland Basin. American Association of Petroleum Geologists
Bulletin 69, 1241–1254.
26
ACCEPTED MANUSCRIPT

Sinninghe Damsté, J.S., van Duin, A.C.T., Hollander, D., Kohnen, M.E.L., de Leeuw, J.W., 1995.
Early diagenesis of bacteriohopanepolyol derivatives: formation of fossil homohopanoids.
Geochimica et Comsmochimica Acta 59, 5141–5157.
Sofer, Z., 1984. Stable carbon isotope compositions of crude oils: application to source depositional
environments and petroleum alteration. American Association of Petroleum Geologists Bulletin 68,

PT
31–49.
Sofer, Z., Regan, R.D., Muller, D.S., 1993. Sterane isomerization ratios of oils as maturity indicators
and their use as an exploration tool, Neuquén Basin, Argentina. XII Congreso de Geológico

RI
Argentino y II Congreso de Exploración de Hidrocarburos, Actas I, pp. 407–411.
Teichmüller, M., Durand, B., 1983. Fluorescence microscopical rank studies on liptinites and vitrinites

SC
in peat and coals, and comparison with results of the Rock-Eval pyrolysis. International Journal of
Coal Geology 2, 197–230.

U
Vecoli, M., Riboulleau, A., Versteegh, G.J.M., 2009. Palynology, organic geochemistry and carbon
AN
isotope analysis of a latest Ordovician through Silurian clastic succession from borehole Tt1,
Ghadamis Basin, southern Tunisia, North Africa: Palaeoenvironmental interpretation.
Palaeogeography, Palaeoclimatology, Palaeoecology 273, 378–394.
M

Waples, D.W., 1985. Geochemistry in Petroleum Exploration. International Human Resources


Development Corporation, Boston, 232 p.
D

West, N., Alexander, R., Kagi, R.I., 1990. The use of silicalite for rapid isolation of branched and
TE

cyclic alkane fractions of petroleum. Organic Geochemistry 15, 499–501.


C EP
AC

27
ACCEPTED MANUSCRIPT

Captions to figures

Fig. 1. A, topographic map showing the location of the Murzuq Basin in southwest Libya. B,
distribution of the Silurian and pre-Silurian deposits in the basin (modified after Lüning et al., 2003;
Meinhold et al., 2013). C, NC115 Concession map showing the distribution of the A-, B- and H-oil
fields studied. D, locations of the exploration wells on which this paper is based.

PT
Fig. 2. A, fault map highlighting the oil fields studied (after RRI, 1998). B, schematic cross-section
through the Murzuq Basin showing major structures and sedimentary sequences (after Boote et al.,

RI
1998). Inset: map showing the extent and direction of the cross-section.

Fig. 3. General depositional model and chronostratigraphic chart for Upper Ordovician–Lower

SC
Devonian sediments in North Africa with emphasis on the lowermost Silurian Hot Shales (after Lüning
et al., 2000).

U
Fig. 4. Plot of TOC versus petroleum generation potential (S1+S2) showing the very good to excellent
AN
potential of the Hot Shale compared to the poor to fair potential of the rest of the Tanezzuff Formation.
In this figure and in Figs. 5–9, Tanezzuft Formation = deposits overlying the Hot Shale Member.
M

Fig. 5. Typical data set of source-rock hydrogen and oxygen indices plotted on a pseudo van-Krevelen
diagram for rock samples from the NC115 Concession. The Hot Shale contains mainly Type II/III
kerogen whereas samples from the rest of Tanezzuft Formation yielded both Type III and Type II/III
D

kerogen.
TE

Fig. 6. Plot of Rock-Eval Tmax versus production index (PI). This shows that the majority of the
samples are sufficiently mature for hydrocarbon generation.
EP

Fig. 7. Sofer diagram (Sofer, 1984) showing mostly marine organic matter and four groupings
according to the isotopic response. Stable carbon isotope ratios [‰ relative to Pee Dee Belemnite
C

(PDB)] for saturated versus aromatic hydrocarbons differ between extracts and oils from the NC115
AC

Concession.

Fig. 8. Selected gas chromatograms of aliphatic fractions showing the n-alkanes and acyclic
isoprenoids for rock samples from the basal Hot Shale and overlying deposits of the Tanezzuft
Formation, NC115 A-, B- and H-fields.

Fig. 9. Plot of pristane/n-C17 versus phytane/n-C18 indicating environment and redox conditions during
deposition. The samples analyzed were non-biodegraded hydrocarbons generated from mixed Type-
II/III kerogen; depositional environments range from suboxic to mildly anoxic.

28
ACCEPTED MANUSCRIPT

Fig. 10. Representative gas chromatograms, m/z 191 and 217 ion mass fragmentograms for Family-I, -
II and -III oils.

Fig. 11. A, sterane isomerization C29 20S/(20S+20R) versus C29 ββ/(ββ+αα), and B,
methylphenanthrene index (MPI) versus (C20+C21)/(C20–C28) triaromatic steranes for determining
thermal maturity of the oils from the NC115 A-, B- and H-fields. The sterane isomerization parameters

PT
have nearly identical ranges and are clearly seen as one cluster, suggesting that the oils were generated
from source rocks approaching the peak stage of oil generation.

RI
Fig. 12. NC115 Concession oils were separated into three genetic oil families based on chemometric
analysis of source-related biomarker and isotope ratios. A, hierarchical cluster analysis (HCA)

SC
dendrogram, and B, plot of principal component analysis (PCA) scores. The dashed repeatability line is
based on the assumption that samples B21 and B31 are representative of reproducibility because these
two oils are from similar depths (1448 and 1410 m, respectively) in the same pay zone (Mamuniyat
Formation) in wells ~1 Km apart.
U
AN
Fig. 13. Plots of biomarker-related parameters for oils from the NC115 A-, B- and H-fields. i, ratios of
diasteranes/regular steranes versus C27Ts/Tm; ii, C26/C25 tricyclic terpanes versus C31R/H; iii, %C29
M

versus %C28 regular steranes; iv, steranes/hopanes versus C28-bisnorhopane/hopane; v, C22/C21 tricyclic
terpanes versus C24/C23 tricyclic terpanes; and vi, C28-bisnorhopane/hopane versus C19/C23 tricyclic
D

terpanes.
TE

Captions to tables
EP

Table 1. TOC/Rock-Eval pyrolysis data for samples from wells in the NC115 Concession, Murzuq
Basin. In this table and Table 2, Tanezzuft Formation = deposits overlying the Hot Shale Member.
C
AC

Table 2. Bulk, elemental, stable carbon isotope and gas chromatographic results for oil and extract
samples from wells in the NC115 Concession.

Table 3. Biomarker data for oil samples from wells in the NC115 Concession.

Table 4. Maturity-related biomarker data for oil samples from wells in the NC115 Concession.

29
ACCEPTED MANUSCRIPT

Figure 1

MEDITERRANEAN SEA
o o o o
TUNISIA 10 E 12 14 16
28 E Al Qarqaf High
Cyrenaica
Sabha
Ghadames Platform NC115

PT
Basin
Sirte Basin
26

NC115

Murzuq Basin

RI
Murzuq L I B Y A Ghat
Basin
LIBYA
24
Al Kufra

SC
Basin Mourizidie Tibesti High
Horst
NIGER CHAD ALGERIA
0 400 km

U
NC115 Concession Djado Basin
A-field H-field
AN
NIGER
B-field
Silurian in subsurface
Silurian in outcrops
30 Km 0 150 Km pre-Silurian outcrops
M
D

B31 H29
H44i H28
A5
TE

B1
H39i
B41 B21

H15
EP

A17

H38
A37i

B42 H40
C

0 2 km 0 2 km 0 4 km
B-field A-field H-field
AC

30
ACCEPTED MANUSCRIPT

Figure 2

PT
RI
U SC
AN
M
D
TE
EP
C
AC

31
ACCEPTED MANUSCRIPT

Figure 3

A g e Formation Shale

PT
Pridoli
Ludfordian Libya
Ludlow
Gorstian

RI
Homerian
Wenlock
Sheinwoodian

Telychian

Llandovery Aeronian

SC
Rhuddanian Hot Shale

Mamuniyat
Hirnantian

Melaz
Sandbian Shuqran

U
Darriwilian
Middle Hawaz
Dapingian
AN
M
D
TE
C EP
AC

32
ACCEPTED MANUSCRIPT

Figure 4

100

Very good -

PT
excellent

10.0

RI
Good

SC
Fair

U
1.00
Very good
AN
Excellent
Poor
Good
Fair

0.10
0.10 1.00 10.0 100
D
TE
C EP
AC

33
ACCEPTED MANUSCRIPT

Figure 5
Hydrogen index (mg HC/g TOC)

PT
RI
U SC
AN
M
D
TE
EP
C
AC
34
ACCEPTED MANUSCRIPT

Figure 6

PT
RI
U SC
AN
M
D
TE
EP
C
AC

35
ACCEPTED MANUSCRIPT

Figure 7

A5 A5

PT
Terrigenous
Marine H15
A17

RI
H44i
A37i H39i
H38 H15

SC
H40
H28

B21
B41 B31

U
AN
B42
M
D
TE
EP
C
AC

36
ACCEPTED MANUSCRIPT

Figure 8

PT
RI
U SC
AN
M
D

110
TE
EP
C
AC

110

37
ACCEPTED MANUSCRIPT

Figure 9

PT
RI
U SC
AN
M
D
TE
EP
C
AC

38
ACCEPTED MANUSCRIPT

Figure 10
nC5

nC10

PT
nC15

RI
SC
Abundance

U
AN
M
D
TE
EP
C
AC
39
ACCEPTED MANUSCRIPT

Figure 11

PT
0.5

RI
U SC
AN
M
D
TE
C EP
AC

40
ACCEPTED MANUSCRIPT

Figure 12

H44i

H39i

PT
H40

RI
Similarity

SC
U
AN
M
D
TE
EP
C
AC

41
ACCEPTED MANUSCRIPT

Figure 13

A37i H40

H40

H39i
A37i H39i

A17

PT
H38 A17

B42 B21 H44i H38

RI
H44i

B21 B31 B41


B31 B41
B42

U SC
A37i
AN
M

B41

B21
D
TE
EP

A17
C
AC

42
ACCEPTED MANUSCRIPT

Table 1

Well Depth Sample TOC S1 S2 S3 Tmax S1/


No. Formation %Ro HI OI PI S2/S3 S1+S2
name (m) type wt% mg/g mg/g mg/g (oC) TOC
1 B42 Tanezzuft 1186 Cuttings 0.73 0.10 1.75 0.38 434 0.65 239 52 0.14 0.05 4.61 1.85

PT
2 B42 Tanezzuft 1286 Cuttings 0.42 0.08 0.67 0.32 436 0.69 161 77 0.19 0.11 2.09 0.75
3 B42 Tanezzuft 1396 Cuttings 0.34 0.09 0.14 0.50 433 0.63 42 149 0.27 0.39 0.28 0.23
4 B1 Tanezzuft 1338 Cuttings 1.13 0.50 1.50 0.68 437 0.71 133 60 0.44 0.25 2.20 2.00
5 H39i Tanezzuft 1213 Cuttings 0.40 0.05 0.60 0.51 436 0.69 152 129 0.13 0.08 1.18 0.65

RI
6 H39i Hot Shale 1449 Cuttings 9.70 2.58 26.63 0.98 433 0.63 275 10 0.27 0.09 27.17 29.21
7 H40 Tanezzuft 1225 Cuttings 0.35 0.05 0.40 0.67 438 0.72 114 191 0.14 0.11 0.60 0.45

SC
8 H40 Tanezzuft 1335 Cuttings 0.68 0.53 2.56 0.21 434 0.65 375 31 0.78 0.17 12.19 3.09
9 H40 Tanezzuft 1439 Cuttings 0.51 0.14 0.99 0.30 439 0.74 195 59 0.27 0.12 3.30 1.13
10 H15 Tanezzuft 1225 Cuttings 0.88 0.10 1.73 0.39 434 0.65 197 45 0.11 0.29 4.43 1.83
11 H15 Tanezzuft 1355 Cuttings 0.62 0.08 2.08 0.19 434 0.65 237 31 0.13 0.22 10.95 2.16

U
12 H15 Tanezzuft 1470 Cuttings 0.80 0.09 0.82 0.35 436 0.69 103 44 0.11 0.34 2.34 0.91
13 H15 Hot Shale 1475 Cuttings 2.18 1.12 3.03 0.35 437 0.71 139 16 0.51 0.32 8.66 4.15

AN
14 H15 Hot Shale 1480 Cuttings 8.19 1.16 16.70 0.65 440 0.76 204 8 0.14 0.20 25.69 17.86
15 H15 Hot Shale 1485 Cuttings 14.40 2.34 37.44 0.86 438 0.72 260 6 0.16 0.13 43.53 39.78
16 H15 Hot Shale 1489 Cuttings 14.30 3.56 34.18 0.71 439 0.74 239 5 0.25 0.16 48.14 37.74

M
17 H28 Tanezzuft 1225 Cuttings 0.53 0.14 0.90 0.41 433 0.63 170 77 0.26 0.10 2.19 1.04
18 H28 Tanezzuft 1307 Cuttings 0.59 0.15 0.59 0.35 438 0.72 100 59 0.25 0.08 1.68 0.74
19 H28 Tanezzuft 1426 Cuttings 0.59 0.12 1.12 0.22 438 0.72 190 37 0.20 0.18 5.09 1.24

D
20 H28 Hot Shale 1487 Cuttings 2.83 0.85 3.88 1.10 436 0.69 137 39 0.30 0.32 3.53 4.73
21 H28 Hot Shale 1489 Cuttings 2.56 0.90 3.40 0.66 435 0.67 133 26 0.35 0.05 5.15 4.30

TE
22 H28 Hot Shale 1490 Cuttings 5.27 1.25 10.06 1.05 435 0.67 191 20 0.24 0.18 9.58 11.31
23 H28 Hot Shale 1493 Cuttings 5.23 1.22 10.04 1.25 437 0.71 192 24 0.23 0.17 8.03 11.26
24 A37i Tanezzuft 1298 Cuttings 0.40 0.05 0.31 0.46 433 0.63 77 115 0.12 0.14 0.67 0.36
25 A37i Tanezzuft 1381 Cuttings 0.40 0.07 0.67 0.35 439 0.74 169 88 0.18 0.09 1.91 0.74
EP
26 A37i Tanezzuft 1463 Cuttings 0.48 0.08 0.55 0.39 438 0.72 114 81 0.17 0.13 1.41 0.63
27 B31 Tanezzuft 1409 Core 0.90 0.58 2.35 0.19 437 0.71 262 21 0.65 0.20 12.37 2.93
28 B31 Tanezzuft 1410 Core 1.08 0.67 2.78 0.09 439 0.74 257 8 0.62 0.19 30.89 3.45
C

29 A17 Tanezzuft 1479 Core 0.48 0.15 0.61 0.10 437 0.70 128 21 0.31 0.20 6.10 0.76
30 A17 Tanezzuft 1481 Core 0.57 0.21 0.79 0.13 438 0.72 138 23 0.37 0.21 6.08 1.00
AC

31 H29 Hot Shale 1474 Core 4.86 1.96 11.04 0.21 440 0.76 227 4 0.40 0.15 52.57 13.00
32 H29 Hot Shale 1477 Core 8.61 2.98 22.58 0.23 444 0.83 262 3 0.35 0.12 98.17 25.56
33 H29 Hot Shale 1482 Core 20.90 7.42 59.47 0.33 438 0.72 285 2 0.36 0.11 180.21 66.89
34 H29 Hot Shale 1487 Core 1.71 0.84 3.90 0.17 438 0.72 228 10 0.49 0.18 22.94 4.74
35 A5 Hot Shale 1520 Core 2.46 1.49 6.30 0.22 440 0.61 256 9 0.60 0.19 28.64 7.79
36 A5 Hot Shale 1537 Core 5.21 3.29 13.28 0.88 448 0.71 255 17 0.63 0.20 15.09 16.57

43
ACCEPTED MANUSCRIPT

Table 2

Well Depth Formation Sample API S V Ni SAT ARO Pr/ Pr/ Ph/ nC17/
No. Age V/Ni CV CPI
name (m) / pay zone type (o) Wt% (ppm) (ppm) (δ‰) (δ‰) Ph n-C17 n-C18 nC27

1 B1 1338 Tanezzuft Silurian Extract NA NA NA NA NA NA NA NA 1.03 0.57 0.42 2.30 1.40

PT
2 A5 1520 Tanezzuft Silurian Extract NA NA NA NA NA -28.97 -28.11 -0.76 1.18 0.36 0.24 2.86 1.69

3 A5 1537 Hot Shale Silurian Extract NA NA NA NA NA -29.20 -28.10 -0.16 1.20 0.66 0.50 10.02 1.18

RI
4 H15 1475 Hot Shale Silurian Extract NA NA NA NA NA -28.28 -28.52 -3.42 0.91 0.66 0.43 2.51 1.01

5 H15 1485 Hot Shale Silurian Extract NA NA NA NA NA -28.60 -28.82 -3.27 1.51 0.48 0.33 8.10 0.99

SC
6 H28 1487 Hot Shale Silurian Extract NA NA NA NA NA -29.65 -29.03 -1.08 1.40 0.75 0.41 2.41 1.03

7 B31 1410 Mamuniyat Ordovician Oil 40.7 0.14 8.00 1.62 4.94 -30.26 -29.42 -0.40 2.02 0.56 0.35 10.00 1.01

U
8 B42 1415 Mamuniyat Ordovician Oil 41.0 0.14 6.00 1.84 3.26 -30.37 -29.58 -0.48 2.11 0.56 0.32 9.09 1.00

9 B41 1422 Mamuniyat Ordovician Oil 39.8 0.13 8.00 1.53 5.23 -30.38 -29.44 -0.15 2.03 0.58 0.33 7.14 1.01

AN
10 B21 1448 Mamuniyat Ordovician Oil 39.6 0.13 4.00 1.50 2.67 -30.14 -29.38 -0.62 2.20 0.59 0.32 7.14 0.99

11 A17 1481 Mamuniyat Ordovician Oil 41.9 0.11 4.00 1.47 2.72 -29.72 -28.92 -0.66 2.00 0.39 0.26 14.28 1.01

M
12 A37i 1531 Mamuniyat Ordovician Oil 41.1 0.13 8.00 1.57 5.09 -29.72 -28.98 -0.79 1.91 0.42 0.26 7.69 1.00

13 H38 1429 Hawaz Ordovician Oil 39.9 0.12 4.00 1.86 2.15 -29.76 -29.01 -0.76 2.15 0.39 0.25 20.0 1.02

D
14 H40 1440 Hawaz Ordovician Oil 39.7 0.13 9.00 1.57 5.73 -29.51 -28.98 -1.33 2.00 0.41 0.26 10.0 1.09

TE
15 H39i 1450 Hawaz Ordovician Oil 39.7 0.13 9.00 1.50 6.01 -29.65 -29.00 -1.02 1.97 0.43 0.27 10.0 0.99

16 H44i 1474 Hawaz Ordovician Oil 40.3 0.12 8.00 1.67 4.79 -29.68 -28.98 -0.90 1.86 0.40 0.26 10.0 0.99
EP
Notes: API: American Petroleum Institute gravity (60oF); sulfur content (S) expressed in wt%; V and Ni concentrations are expressed in ppm; SAT and ARO (δ‰):
Saturated and aromatic stable carbon isotope (δ‰); CV: Canonical variable=-2.53*δ13Csaturates+2.22*δ13Caromatics-11.65; Pr/Ph = pristane/phytane; Pr/n-C17=pristane to
C

n-C17 alkane; Ph/n-C18= phytane to n-C18 alkane; nC17/nC27= n-C17/n-C27 paraffin ratio, peak heights from whole-oil gas chromatogram; CPI= Carbon preference index
AC

(odd/even)=2*(n-C23+n-C25+n-C27+ n-C29)/[n-C22+2*(n-C24+ n-C26+ n-C28)+ n-C30]

44
ACCEPTED MANUSCRIPT

Table 3

Well Depth C19 C22 C24 C26 Tet/ C27Ts/ C28 C29 C31R C30X GA C35/ % % % DiaS/ Ster/ DBT/
No. Family S/H
name (m) /C23 /C21 /C23 /C25 C23 Tm /H /H /H /H /H C34 C27 C28 C29 Reg Terp P
1 B31 1410 1 0.06 0.22 0.60 1.14 0.18 1.14 0.08 0.63 0.16 0.14 0.04 0.65 29.25 23.48 47.27 2.25 3.32 3.32 0.03

PT
2 B42 1415 1 0.07 0.22 0.59 1.14 0.18 1.06 0.07 0.61 0.17 0.14 0.04 0.58 28.68 23.59 47.73 2.25 3.24 3.24 0.02
3 B41 1422 1 0.06 0.21 0.57 1.17 0.18 1.17 0.08 0.61 0.16 0.13 0.05 0.62 27.90 23.52 48.58 2.28 3.28 3.28 0.06

RI
4 B21 1448 1 0.08 0.21 0.57 1.19 0.19 1.15 0.07 0.63 0.17 0.13 0.04 0.52 28.82 23.41 47.76 2.22 3.20 3.20 0.05
5 A17 1481 2 0.09 0.25 0.71 1.31 0.17 2.60 0.17 0.60 0.18 0.47 0.05 0.75 28.94 25.31 45.75 2.83 4.83 4.83 0.03

SC
6 A37i 1531 2 0.12 0.22 0.67 1.29 0.18 3.53 0.16 0.59 0.19 0.40 0.05 0.00 29.21 25.39 45.41 2.49 4.61 4.61 0.07
7 H38 1429 2 0.11 0.25 0.70 1.31 0.23 2.29 0.12 0.60 0.17 0.30 0.06 0.00 29.06 26.17 44.77 2.70 3.77 4.60 0.04

U
8 H40 1440 2 0.10 0.25 0.74 1.45 0.19 3.31 0.16 0.55 0.20 0.41 0.06 NA 28.84 25.27 45.89 2.54 4.83 4.56 0.04
9 H39i 1450 2 0.09 0.26 0.69 1.31 0.19 3.04 0.16 0.58 0.19 0.42 0.08 NA 29.30 25.49 45.21 2.60 4.41 4.41 0.03

AN
10 H44i 1474 3 0.16 0.28 0.69 1.25 0.34 1.60 0.08 0.71 0.17 0.18 0.04 0.00 28.69 26.43 44.87 2.79 2.10 4.42 0.05

M
Notes: C19/C23: C19/C23 Tricyclic terpane peak areas from m/z 191; C22/C21: C22/C21 Tricyclic terpane peak areas from m/z 191; C24/C23: C22/C21 Tricyclic terpane peak

D
areas from m/z 191; C26/C25: C22/C21 Tricyclic terpane peak areas from m/z 191; Tet/C23: Tetracyclic C24 to tricyclic terpane C23 ratio peak areas from m/z 191; C27Ts/Tm
= 18α(H)-22,29,30 trisnorneohopane to 17α(H)-22,29,30 trisnorhopane; C28/H: Bisnorhopane/hopane; C29/H: Norhopane/hopane; C31R/H: Homohopane

TE
(22R)/hopane; C30X/H: C30 Diahopane/hopane; GA/H: C30 Gammacerane/hopane; C35/C34: C35 Extended hopane/C34 extended hopane (22S); %C27αααR:
%C27/(%C27+%C28+%C29); %C28 αααR: %C28/(%C27+%C28+%C29); %C29 αααR: %C29/(%C27+%C28+%C29); DiaS/Reg: 13β(H),17α(H)-diacholestanes (20S+20R) to 20S+20R isomers of
EP
C27 isosterane and C27 regular sterane; S/H: Steranes/hopanes; Ster/terp: Steranes/tricyclic terpanes; DBT/P = dibenzothiophene/phenanthene; NA: No analysis or
no data.
C
AC

45
ACCEPTED MANUSCRIPT

Table 4

Well %C32 DBT/


No. M/H %20S %ββR MPI %Rc1 %Rc2 F1 F2 MDR TAS1 TAS2 TAS3 TAS4 TAS5
name 22S C4N
1 B31 0.64 0.09 0.47 0.55 1.03 0.79 0.98 0.44 0.23 0.18 6.22 0.94 0.89 0.80 0.11 0.39

PT
2 B42 0.64 0.09 0.46 0.54 1.03 0.77 0.98 0.44 0.23 0.15 6.65 0.91 0.89 0.78 0.10 0.66
3 B41 0.64 0.09 0.46 0.53 1.01 0.77 0.97 0.44 0.24 0.17 6.10 0.94 0.87 0.78 0.10 0.33

RI
4 B21 0.64 0.08 0.46 0.55 1.00 0.77 0.97 0.45 0.24 0.13 7.25 0.91 0.89 0.79 0.08 0.64
5 A17 0.69 0.09 0.46 0.55 1.07 0.78 1.01 0.45 0.25 0.09 11.72 0.97 0.96 0.90 0.28 0.70

SC
6 A37i 0.67 0.09 0.46 0.55 1.02 0.78 0.97 0.45 0.24 0.23 8.32 0.98 0.96 0.90 0.22 0.32
7 H38 0.67 0.09 0.46 0.55 1.02 0.77 0.97 0.44 0.24 0.20 9.41 0.98 0.96 0.91 0.38 0.43

U
8 H40 0.67 0.08 0.46 0.54 1.02 0.78 0.97 0.44 0.24 0.15 9.44 1.00 0.96 0.91 0.25 0.50
9 H39i 0.65 0.10 0.47 0.55 1.02 0.80 0.97 0.44 0.24 0.18 10.10 0.99 0.96 0.91 0.34 0.24

AN
10 H44i 0.64 0.13 0.47 0.55 1.01 0.80 0.97 0.45 0.24 0.17 9.61 1.00 0.97 0.92 0.33 0.32

M
Notes: %C3222S: %C32αβ22S/(22S+22R) 17α(H),21β(H)-29-homohopane ratio (%); M/H: Moretane/hopnae; %20S: ααS/(ααS+ααR) 5α(H),14α(H),17α(H)-stigmastane 20S
and 20R ratio (%); %ββR: C29 ββR/(ααS+ββR); MPI = 1.5*(2MP + 3MP)/(P + 1MP + 9MP); MPI: Methylphenanthrene Index (1.5*[3MP+2MP]/[P+9MP+1MP]); %Rc1:

D
Inferred vitrinite reflectance of source rocks calculated as [0.5*(20S/20R C29 steranes)]+ 0.35 Sofer et al. (1993); %Rc2: 0.37 + 0.6*MPI-1 (Radke, 1988); F1:

TE
[3MP+2MP]/[3MP+2MP+9MP+1MP]; F2: 2MP/[3MP+2MP+9MP+1MP]; DBT/C4N: Dibenzothiophene/C4 naphthalene; MDR: Methyldibenzothiophene ratio =
4MDBT/1MDBT; TAS1: C20/[C20+C27] triaromatic sterane; TAS2: C21/[C21+C28] triaromatic sterane ; TAS3: [C20+C21]/[C20-C28] triaromatic sterane (cracking ratio); TAS4:
EP
C26S/C28S triaromatic sterane; TAS5: C27R/C28R triaromatic sterane; NA: No analysis or no data.
C
AC

46
ACCEPTED MANUSCRIPT
Highlights

1. Ordovician–Silurian oils and rocks from the Murzuq Basin have been investigated.

2. The Hot Shale is organically-rich with mixed algal-terrigenous, early mature Type-II/III kerogen.

3. The rest of the Tanezzuft Formation is organically-lean with abundant Type-III to II/III kerogen.

4. The NC115 oils studied were placed in two genetic families with one outlier.

PT
5. The analyzed oils record an advanced maturity level with the B-oils being the least mature.

RI
U SC
AN
M
D
TE
C EP
AC

You might also like