You are on page 1of 40

GeoArabia, vol. 14, no. 2, 2009, p.

101-140 Potential energy resources, northwest Jordan


Gulf PetroLink, Bahrain

Diagenetic and epigenetic alteration of Cretaceous to Paleogene


organic-rich sedimentary successions in northwestern Jordan,
typical of the western margin of the Arabian Plate
Harald G. Dill, Jolanta Kus, Abdulkader M. Abed, Reinhard F. Sachsenhofer
and Hani Abul Khair

ABSTRACT

Four reference sections through the calcareous-siliciclastic rocks of the Lower


Cretaceous Kurnub Group and the Amman and Muwaqqar formations of the
Upper Cretaceous to Paleogene Belqa Group in Jordan document the various
processes of accumulation and alteration of organic matter (OM). Sections at
Jerash, Sultani, Wadi Isal, and in the Kharazeh area were investigated by means
of sedimentary petrography/mineralogy, organic petrography, and organic
chemistry, and correlated with equivalent deposits in Syria and Egypt. The impacts
of oxidation potential (Eh), acidity/basicity (pH) and temperature variations during
the post-depositional alteration of these organic concentrations were assessed using
x-y plots. Syngenetic Aptian-Albian lignite-amber beds near Jerash developed
in a tide-dominated delta under marginally alkaline conditions and were altered
under slightly acidic conditions at temperatures of less than 100°C. Environmental
analysis focused on Maastrichtian-Paleocene (?), oil shales in the Sultani area that
were deposited in a small restricted basin on the continental shelf of the Neo-Tethys
Ocean. Abnormally high contents of vanadium (V), phosphorus (P), zinc (Zn),
and uranium (U) gave rise to yellow uranium ore minerals, phosphates and zinc
sulfides and so bridge the gap between OM concentrations and those of uranium.
The western part of the Arabian Peninsula is known for its uraniferous phosphorites
and U-bearing calcretes. Reducing conditions during deposition of syngenetic OM
in the oil shales may have shifted, in places, toward more oxidizing conditions in the
course of post-depositional alteration at temperatures well below 200°C. Epigenetic
fault-related concentration of OM was responsible for the Wadi Isal Aptian-Albian
tar sand deposit and the oil seepage system in calcareous wall rocks of the Kharazeh
Fault. The alteration of the tar sands (fluvial sandstones) is designated a high-
sulfidation type (introduced aluminum sulfate minerals), whereas the oil seepage
in Campanian shallow-marine carbonates is a low-sulfidation type (removal of
aluminum sulfate minerals). Both alteration patterns may have implications for
structure-bound base- and precious-metal deposits whose emplacement involves
hydrocarbons as a carrier of metals in the mineralizing fluids. The Jordanian oil
shales, tar sands and uranium mineralization are possible sources of energy. Recent
economic deals on the recovery and use of oil shales and uranium in Jordan are
reviewed.

ENERGY RESOURCES AT THE WESTERN MARGIN


OF THE ARABIAN PLATE

The Arabian Plate (Figure 1a) is bounded to the west by the well-known strike-slip Dead Sea Transform
Fault whose morphological expression is the Jordan Valley and the Gulf of Aqaba (Sharland et al.,
2001). Whereas the northeastern boundary of the Arabian Plate (Gulf Region) is marked by a cluster
of world-class gas and oil deposits elongated northwest, there are no energy deposits of similar size
along its western boundary. The energy resources of the Gulf Region have sparked numerous studies;
for example, Alsharhan and Nairn (1995), Al-Aswad (1997), Al-Siddiqi and Dawe (1999), Sadooni and
Alsharhan (2004), and Al-Jallal and Alsharhan, (2005).

In Jordan, the number and size of energy resources is manifestly less than those of the Arabian Gulf
region, but the diversity of potential energy resources is high and includes lignite, hydrocarbons, and
uranium mineralization. A small fraction of domestic energy requirements is met by the Hamzah

101
Dill et al.

a 30°E 40° 50° 60°

TURKMENISTAN
TURKEY
36° 38° Caspian
40°N Sea

CYPRUS
LEBANON
IRAN
SYRIA
Mediterranean
Sea IRAQ

JORDAN

Study Area
30°

BAHRAIN

ARABIAN PLATE QATAR


EGYPT
UNITED ARAB
EMIRATES

SAUDI OMAN
ARABIA

ne
Zo
Re

20°
d

re
Se

ctu
a

Fra
en
SUDAN YEMEN

Ow
ERITREA

Arabian
Sea

f Aden 0 500
ETHIOPIA Gulf O
km

Figure 1: Topographic and geological overview of the Kingdom of Jordan.


(a) The Arabian Plate showing the location of the study area and oil and gas fields of the Gulf
region.
(b) Geology of the Kingdom of Jordan and neighboring areas (simplified after Bender, 1974).
(c) Location of oil shales, tar sands, and uranium occurences; locations modified from Dyni
(2005) and Toukan (2008).

oil field near Azraq east of Amman, and the Risha gas field near the Jordan-Iraq border (Figure 1b)
(Abu-Ajamieh et al., 1988). Various aspects of the domestic oil shales have been studied, but the
depositional environment is still poorly known (Hufnagel et al., 1980; Abed and Amireh, 1983; Jaber
and Probert, 1997, 1999; and Hamarneh, 1998). The emphasis on the oil shales has eclipsed other
carbonaceous rocks as potential fossil fuels.

This paper presents the results of the investigation of a stratigraphic interval from the Upper Cretaceous
through the Paleogene that hosts hydrocarbon deposits, lignite seams and uranium mineralization
(Figures 1b, c). The organic-rich sedimentary rocks, locally containing uranium, are related to variety
of paleoenvironments that have been studied geologically, mineralogically and petrographically, and
their organic and inorganic chemical composition investigated.

102
Potential energy resources, northwest Jordan

b 35° 36° 37° 38° 39° Basalt

N
NO
Med. Cenozoic

BA
Sea
LE Lake SYRIA Mesozoic
33° Tiberias 33°
Paleozoic

IRAQ Precambrian
River Jordan Risha
Oil field
Jerash
Gas field

32° 32° Organic-rich study site


PALESTINE Amman
Hamzah River
Dead Sea

Kharazeh Area SAUDI ARABIA


Sultani Trench
31° 31°
Wadi Isal

JORDAN

30° 30°

N
0 100

km

Gulf of Aqaba
29° 29°
35° 36° 37° 38° 39°

c 35° 36° 37° 38° 39° International boundary


N
NO

1 Road
Med.
BA

Sea City/town
LE

SYRIA
33° 33° Oil shales
10 1: Ma'an
Ramtha IRAQ 2: Juref ed Darawish
3: El Hasa
3 Ruwayshid
4: Sultani
Jerash 5: Wadi Maghar
Zarqa 7 6: El Lajjun
Safawi
32° 32° 7: Attarat Umm Ghudran
PALESTINE Amman Azraq 8: Khan ez Zabib
9: Wadi Thamad
9 8 10: Yarmouk
8 11: Siwaga
11 6
6 Tar sands
Karak 7 SAUDI ARABIA
W

4 1: Hasbaya
ad

2
i

31° 5 31° 2: Wadi Aheimir


Si

3
rh
an

1 3: Wadi Dhira
Tafila 2
Uranium occurences
3 4
JORDAN
1: Wadi Araba, Dana
al-Jafr
1 2: Dubaideb, Mudawwara
5 3: Hamra, Hausha
Ma'an 4: Wadi Al Baheyya
30° 30°
5: Wadi Sahb Al Abiad
6: Central Jordan
N 7: South Rowaished
Aqaba 0 100 8: Qasr Al Kharraneh
2
km
Mudawwara
29° 29°
35° 36° 37° 38° 39°

103
Dill et al.

This study provides an overview of four sites that are rich in organic matter but differ in the types
of organic matter present. They are the Jerash amber-bearing coaly and organic-rich beds (Kurnub
Group), the Sultani oil shales (Muwaqqar Formation), the Wadi Isal tar sands (Kurnub Group), and
the oil seepage in the Kharazeh area (Amman Formation) (Figure 1b; Table 1). The types of organic
matter are typical of the western edge of the Arabian Plate and can be correlated with deposits in Syria
and Egypt. The main focus is placed on diagenetic and epigenetic alteration and the constraining
physico-chemical regime, an approach that had not been taken in Jordan although it had been used
elsewhere on the Arabian Plate (Dill et al., 2007). Papers published by, for example, Al-Rifaiy et al.
(1993) and Schulze et al. (2005) on the western parts of Jordan were based on common microfacies
analysis and exclusively centered on the calcareous platform sediments with no attention paid to the
chemical and mineralogical inventory.

Table 1
Age Group Formation Lithology

Eocene Shallaleh Chalk and marl


Paleogene
Paleocene- Rijam Chalk and chert
Maastrichtian Muwaqqar/less space Marl and oil shale
Belqa
Campanian Al-Hisa Phosphorite-chert-limestone
Amman Interbedded chert and limestone
Coniacian-Santonian Ghudran Chalk and chert
Late
Turonian Wadi Es Sir Limestone
Cretaceous
Shuaib Thin limestone alternating with marl
Ajlun Hummar Limestone
Cenomanian
Fuhais Marl
Naur Thick limestone alternating with marl
Early Fluvial sandstones with minor claystones;
Aptian-Albian Kurnub
Cretaceous marine interbeds in northern Jordan
Stratigraphy and lithology of Cretaceous and Paleogene stratigraphic units in northwestern Jordan.

GEOLOGICAL SETTING DURING THE CRETACEOUS AND PALEOGENE


IN NORTHWESTERN JORDAN

Cretaceous and Paleogene strata in Jordan are subdivided into three groups. From bottom to top they
are the Kurnub, Ajlun, and Belqa groups. Table 1 summarizes the groups and provides a stratigraphic
subdivision into formations and accompanying lithology.

Kurnub Group

The Kurnub Group is of Early Cretaceous (Aptian-Albian) age. During this time, the area that is
now Jordan was emergent and the locus for fluvial deposition (Abed, 1968; Amireh, 1993, 1997). The
Group is up to 300 m thick and consists of fluvial deposits dominated by varicolored quartz arenitic
sandstones with minor mud rocks arranged in fining-upward cycles (Abed, 1982; Amireh, 1997).
Terrestrial plants, especially ferns were widespread. Although there has been some attempt to divide
the Group into formations (for example, Amireh, 1999), Jordanian geologists prefer to keep it as a
group because of difficulties in identifying formations in the field. In northern Jordan, some marine
transgressions occurred during this time and marine deposits of dolomite and glauconitic sandstone
with some body and trace fossils are present, alternating with fluvial material.

Ajlun Group

This Late Cretaceous (Cenomanian-Turonian) Ajlun Group consists of five formations; from base
to top they are the Naur, Fuhais, Hummar, Shuaib, and Wadi Es Sir formations (Masri, 1963). Its
thickness increases from about 100 m in the south to more than 600 m in the north of the country.
The group consists exclusively of carbonate rocks, except in the extreme south where clastic rocks
make up a great part of the sequence. The group is rich in well-preserved body fossils, such as,

104
Potential energy resources, northwest Jordan

cephalopods, bivalves, gastropods, and echinoderms. The microfossils foraminifera, ostracodes, and
serpulids are also common. Lithologically, it consists of alternating limestones and dolostones and
some soft, interbedded yellowish-greenish marls to marly limestones. These intercalations are used
for subdividing the group into formations (Table 1).

In the Early Cenomanian, a major rise in sea level occurred and the Neo-Tethys Ocean transgressed
over large parts of the Eastern Mediterranean region, including present-day Jordan (Bender, 1974).
Jordan then became part of the southern epicontinental shelf. This paleogeographic setting favored
deposition of the carbonate series of the Ajlun Group. The group pinches out southward, and the
content of clastic rocks increases toward the Arabian-Nubian continent. During Turonian and
Senonian times, a compressional event known as the Syrian Arc Folding, affected the area (Bowen
and Jux, 1987). The northward movement of the African Plate triggered the event and the resulting
basin and swells account for the variations in thickness of the Turonian Wadi Es Sir Formation.

Belqa Group

The Coniacian-Santonian to Eocene Belqa Group consists of six formations; from the base upward
they are the Ghudran, Amman, Al-Hisa, Muwaqqar, Rijam, and Shallaleh formations (Table 1). As
with the Ajlun Group, the thickness of the Belqa Group decreases southward along with a relative
increase in clastic rocks. During deposition of the Belqa Group, the African Plate moved northward
and the basin-and-swell topography of the Syrian Arc became more pronounced. The resulting
paleotopography of the epicontinental shelf became an important factor in the depositional pattern
(Abed and Sadaqah, 1998).

During the Eocene, the Rijam and Shallaleh formations were deposited. The Rijam is composed of
chalk with alternating bedded chert and minor phosphorite horizons, indicative of strong upwelling.
There are nummulite buildups at certain places within the shallower parts of the platform.

By the end of the Eocene, the Neo-Tethys Ocean had completely receded from Jordan and most of
the Eastern Mediterranean interior, leaving Jordan emergent. Tectonic activity occurred during the
Oligocene and, as a result, conglomerates and sandstones were deposited in intermontane basins to
form the Dana Formation near the Dead Sea. During the Middle Miocene, movement along the Dead
Sea Transform Fault began and has continued into the present.

METHODOLOGY

For this overview, the four sites in northwestern Jordan (Jerash, Sultani, Wadi Isal, and the Kharazeh
area), each with different lithologies and organic matter type were sampled. The rock properties were
described visually and under the petrographic microscope using common schemes and comparison
charts: rock strength (Selby, 1987); grain morphology (Illenberger, 1991); sorting (Jerram, 2001);
fabric variations (Collinson and Thompson, 1982); and chemical data based on conventional X-ray
fluorescence analysis.

For X-ray diffraction, a Philips diffractometer PW 3710 (40 kV, 30 mA) with CuKα radiation and
equipped with a fixed divergence slit and a secondary graphite monochromator was used. Random
powder samples were scanned with a step size of 0.02° 2 theta and counting time of 1 second per step
over a measuring range of 2° to 65° 2 theta. Additional scans were performed from 17° to 28° with a
counting time of 30 second per step to evaluate the disordering of the kaolin group minerals.

An FEI QUANTA 600 FEG scanning electron microscope with an energy-dispersive system (SEM-
EDS), was used to assist in mineral identification and image analysis for morphological studies.

Dried coal samples were crushed to minus 1 mm and placed in a 3-cm-diameter mold and vacuum-
impregnated with epoxy resin. Following impregnation, particulate blocks were subjected to dry
grinding and polishing. For vitrinite reflectance, 50 vitrinite (see Glossary) particles were measured
from each polished particulate block in accordance with German Standard Guidelines (DIN 22020,
Part 5). A Leica DMRX microscope system was used for incident and fluorescent light microscopy

105
Dill et al.

using x 20, x 50, and x 100 oil immersion objectives and an MPV-2 photometer system. Digital images
were captured using Leica DC 300F and Image Manager (Leica, IM50) software.

The total carbon (TC), total organic carbon (TOC) and the total sulfur contents (TS) of the samples
were determined with a LECO CS 200 carbon-sulfur-analyzer by combustion in a high-frequency
furnace at between 1,800°C to 2,200°C. The carbon dioxide and sulfur dioxide contents generated
were measured with infrared detectors. For the TC and TS analysis, untreated rocks were used. The
TOC was determined after removal of carbonates with 2N HCl at 80°C.

Rock-Eval Pyrolysis was performed with a Rock-Eval 6 instrument according to standard methods
(Espitalié et al., 1977). About 100 mg of sample was pyrolized using nitrogen as a carrier gas. Pyrolysis
products were measured with a Flame Ionization Detector (FID), and the concurrent released carbon
dioxide was determined by infrared detection. The temperature program started with an isothermal
step for 3 minutes at 300°C followed by a heating step up to 650°C at a rate of 25°C/minute.
Hydrocarbons released during the isothermal and heating steps are quantified as S1 and S2 peaks,
respectively (both in milligram hydrocarbons per gram rock). Carbon dioxide generated up to 400°C
is summed up as S3 (in milligram carbon dioxide per gram rock).

Hydrogen and oxygen indices (HI and OI) were S2 and S3 normalized to the TOC content. Tmax is
the temperature of the maximum generation of cracking products. About 6 g of the fine-grained
samples were extracted in a Dionex Accelerated Solvent Extractor 200 at 80°C and 1,200 psi with
dichloromethane/methanol (DCM/MeOH) 95:5 as the solvent. Sulfur was removed with activated
copper granulate. After a clean-up procedure using silicagel and hexane, the aliphatic fraction of
the extract was injected into a Hewlett Packard 5890 gas chromatograph, coupled to a 5972 Hewlett
Packard Mass Selective Detector. A 30 meters long nonpolar fused silica capillary column was used,
operated at a temperature program between 50°C and 300°C. In the Selected Ion Monitoring Mode,
mass fragments m/z 71 (n-alkanes), 191 (triterpanes), 217 (5α , 14α, 17α - steranes) and 218 (5α, 14β,
17β - steranes) were recorded. Total ion current traces and full spectra of selected compounds were
recorded using a 95 S Finnigan gas chromatograph/mass spectrometer system.

All samples containing organic matter were double-checked by coal petrographic and mineralogical
methods, for example, SEM-EDX (scanning electron microscope-energy dispersive using X-Rays),
to ensure that subaerial processes were of no impact on the samples taken. Degradation of organic
matter at outcrop can therefore be ruled out.

COALY AND ORGANIC-RICH AMBER DEPOSITS

An Overview of Coal on the Western Margin of the Arabian Plate

(See Glossary for coal/organic matter terminology.)

The western margin of the Arabian Plate is barren of economic coal accumulations. Coalified matter
is only of importance in this region in association with amber, a commodity that has brought Lebanon
to the attention of gemologists. Lower Cretaceous amber sites in Lebanon, Palestine and Jordan are
commonly considered to be of great scientific and jewelry interest (Azar, 2000; Poinar and Milki, 2001;
Kaddumi, 2005).

Egypt is the only Middle Eastern country to have viable deposits of coal. The Middle Jurassic Safa
Formation contains workable coal seams in the Maghara area of northern Sinai (Hassaan et al., 1992;
Mostafa and Younes, 2001). The extracts of the coal have unusual n-alkane distribution whereby
the gas chromatograms are dominated by unusual compounds eluting between n-C19 and n-C21.
These compounds suggest significant conifer contributions to the coal. Carbonaceous shales and
coal samples from these seams are characterized by organic material that could have been derived
from non-microbial organisms such as ferns. The Maghara coal was formed in a wet forest–swamp
environment situated in a marine-influenced lower delta plain (Mostafa and Younes, 2001). No
workable coal seams occur anywhere higher in the stratigraphic sequence; however, concentrations
of OM occur similar to those investigated in Jordan.

106
Potential energy resources, northwest Jordan

Formation\Group

Environment
Depositional

Stratigraphy
Thickness

Sequence
Period

(meter)
Limestone
Age
Era

Lithology
Marly limestone
B
B Bituminous marly limestone

Chert and limestone


Cenomanian

Marl
Naur

240 Subtidal
Sandstone
98 Ma
MFS Sandy dolomite
Carbona-
K120
ceous shale Carbonaceous shale
220 with lignite
Chalk
Chert
200 Sandstone
Tide- Plant debris
dominated
Sandy
delta to
dolomite
with trace
tidal flats Figure 2: Stratigraphy of the
180
fossils and Kurnub Formation in the Jerash
lignite
area, north of Amman (geologic
interpretation from Amireh, 1997
160 Sandstone
modified in this study). Vertical
Carbona-
ceous shale
bars denote stratigraphic sections
101 Ma
with lignite MFS studied. Sequence stratigraphic
140 and imprints
correlation: MFS = Maximum
Cretaceous

Meandering K110
of leaves
rivers
Mesozoic

fragments, flooding surface, SB = sequence


Aptian-Albian

abundant
amber and boundary. Ages refer to the
Kurnub

120 charcoal
chronological subdivision of
Sharland et al. (2001).

100

Braided to
Organic matter is also found in
Sandstone with
plant debris,
meandering Late Aptian to Early Cenomanian
80 rivers
unidirectional (mid-Cretaceous) sediments on
paleocurrent
the paleoshelf of Northern Sinai.
The organic facies is dominated by
60
hydrogen-depleted OM of either
terrestrial or degraded marine
106 Ma
Sandy dolomite Tide- MFS
origin (kerogen types III and IV)
40 with trace fossils
and lignite-coal
dominated K100 (Kim et al., 1999). The organic
delta
facies of mid-Cretaceous deposits
Sandstone with
plant debris, from northern Sinai indicates a
20 unidirectional proximal fluvio-deltaic or oxic shelf
paleocurrent Braided rivers
and thick environment. Random vitrinite
Kimmeridgian

sequence of
conglomerates
reflectance values ranges from
SB
Jurassic

0
0.6% to 0.7% Rr, equivalent to high
Mugha

Subtidal
volatile bituminous coalification
stages.

Coal- and organic-rich Amber Beds near Jerash

Stratigraphy and Geology


Near Jerash in northwestern Jordan, the Aptian–Albian Kurnub Group that is approximately
200 m thick, rests unconformably on the Kimmeridgian Mugha Formation and is overlain by the
Cenomanian Naur Formation of the Ajlun Group (Amireh, 1997) (Figures 2 and 3). In our study
area, the Kurnub Group consists of three regressive-transgressive depositional sequences, whereas
continental clastics predominate in central and southern Jordan.

107
a
Dill et al.

Limestone
and marl

Naur Formation
Sandstone

Glauconite-beds
MFS K 120

Fine-grained
sandstone
and claystone

Kurnub Formation
Fe-bearing

108
sandstone layer

Multicolored
sandstone, clay,
and marl

Figure 3: Lithology of the Aptian-Albian Kurnub b c d


Formation – Cenomanian Naur Formation transi-
tion in a road cut on the highway between Amman
and Irbid. Section stratigraphically overlies the
organic matter-bearing series in Figure 2.
(a) Cross-section through the Kurnub-Naur transi-
tion zone.
(b) Ferruginous sandstone underlain by multicol-
ored sandstones of the Kurnub Formation
(hammerhead 18 cm long).
(c) Glauconite beds (glaucony marker unit (GMU))
96.1 ± 1.1 Ma (Kurnub Formation). The glauco-
nite horizon is interpreted as MFS K120 (98 Ma) (pen 13 cm long).
(d) Biomicrite with randomly distributed articulate and disarticulate bivalve shells indicative of shallow-marine conditions (Naur Formation) (pen 13
cm long).
Potential energy resources, northwest Jordan

The lowermost part of the sequence in the Table 2


Jerash area is composed of sandstones and Sultani Wadi
Site Jerash Kharazeh
conglomerates containing abundant plant debris Trench Isal
that was deposited in a unidirectional stream SiO2 61.29 19.10 92.70 5.95
system with a paleocurrent directed toward the TiO2 1.21 0.07 0.07 0.05
north-northwest (Amireh, 1997). The sediments Al2O3 3.15 1.50 2.08 1.019
are generally poor in organic matter, and Fe2O3 19.26 1.09 0.07 0.47
sediments bearing organic matter traceable over
MnO 0.00 0.00 0.00 0.01
long distance are absent. However, the upper
MgO 0.09 0.54 0.12 14.70
half of the section contains abundant coaly and
CaO 0.05 26.86 0.06 30.73
organic-rich beds centimeters to decimeters thick,
Na2O 0.02 0.09 0.11 0.01
alternating with calcareous sandstones units and
oolitic ironstones. Organic-rich shales with leaf K2O 0.22 0.27 0.05 0.03
imprints, charcoal accumulations and amber P2O2 0.02 4.99 0.02 0.32
pebbles occur frequently (Figure 2; location Cl 0.01 0.04 0.01 0.41
indicated by the wide vertical bar between 120 F 0.05 0.42 0.05 0.05
and 150 m). Underlying the Cenomanian Naur As 7 26 4 4
Formation, glauconite beds and iron-bearing Ba 36 48 23 202
sandstones are interbedded with arenaceous and Ce 23 15 22 24
argillaceous bedsets (Figure 3). The glauconite Co 9 4 3 3
occurs in an arenaceous dolomite unit, referred Cr 22 429 8 79
to as the glauconite marker unit (GMU), in the Cu 6 140 21 111
upper part of the Kurnub Group that persists La 24 15 19 13
throughout Jordan (Figures 3a, c) (Amireh et Mo 3 227 3 5
al., 1998). The stratigraphic age of the Early
Nb 6 2 3 3
Cretaceous Kurnub Group is based on the K-
Nd 18 15 15 16
Ar dating of the glauconite. The apparent age,
Ni 26 200 11 31
constrained within the analytical uncertainty
Pb 14 3 6 4
limits and derived from the most evolved
glauconite, is 96.1 ± 1.1 Ma; this suggests that the Sm 19 13 13 14
GMU is of Early Cenomanian age (Figure 3c). Sr 41 867 123 173
Th 6 4 4 3

Petrography and Mineralogy U 4 31 3 5


Bandel and Haddadin (1979) recorded seams V 15 715 13 53
of low-quality, high-ash coal (lignite) and Y 7 22 2 7
associated amber along the Zarqa River Zn 3 1,659 2 137
northeast of Amman. The sequence shown in Zr 67 27 75 12
Figure 4 hosts a decimeter-thick bed of organic 13
C -22.9 -28.9 -28.9 -28.5
and coaly matter with pebbles of amber. The TOC 1.35 18.4 2.29 0.98
sedimentary section consists of centimeter-thick, TSC 18.2 5.19 0.38 2.75
thinly laminated layers of whitish-gray sand, Chemical composition of the Cretaceous OM-bearing sediments in
and dark-gray organic and coaly matter that northwestern Jordan referred to in the text. Data of major elements
from SiO2 (silica) to F (fluorine) are given in wt. %, trace elements from
are stained with sulfates. The lignite-amber bed As (arsenic) to Zr (zircon) are given in ppm and 13C data reported in
passes upward through bedded sandstones into the common per mil notation, TOC (total organic carbon content) and
massive amalgamated sandstone, and below TSC (total sulfur content) are listed in wt. %. The values represent
channel samples or mean values, respectively.
into gray organic-rich mudstones. Varicolored
sandstones containing abundant in goethite,
hematite and gypsum replace the lignite–sandstone couplet laterally. The fining-upward cyclicity
proposed for a major part of the Kurnub Group (see above) may locally turn into regressive cycles
as shown in Figure 4.

The host rocks of the organic-and coaly-rich sedimentary beds are mudshales with silty intercalations
composed of quartz, illite, kaolinite, and anatase, together with minor amounts of alkaline feldspar
and organic matter (Figure 5a, Table 2). The detrital grains are well sorted and subrounded to
rounded. The roundness parameter, however, is strongly affected by the replacement processes
along with cementation that has caused deterioration of the roundness of quartz clasts and destroyed
almost all labile constituents; for example, alkaline feldspar. Grains are freely floating within the

109
Dill et al.

Massive
sandstone

Coaly and organic-rich


amber beds

Bedded
sandstone

Gray mudstone
enriched in organic
matter, with interbedded
sandstones

Figure 4: Lithology and geology of


the regressive cycle hosting the coaly
and organic rich-amber beds near
Jerash at Zarka River crossing that
coincides with the MFS K110 marked
in Figure 2. The inset gives a close-up
view of the thinly laminated, “dirty”,
high-ash lignite-amber beds coated
in part with sulfates (hammer 25 cm
long).

Harald: What is the vertical scale?

- Please give specific location.


- Is the road cutting or a natural
exposure?

110
Potential energy resources, northwest Jordan

fine-grained argillaceous matrix and sulfide cement. The clean and mature sandstones are infiltrated
by goethite substituting for the argillaceous matrix of illite and kaolinite (Figure 5b). The sequence of
mineralization in the pore space of the sandstones is as follows: quartz-illite-OM ⇒ kaolinite-pyrite-
marcasite ⇒ goethite ⇒ Fe sulfate (Figure 5d). Eventually, the entire pore space of the sandstones and
mudshales is occluded by marcasite and pyrite and minor amounts of sphalerite (Figures 5c and d).

Sulfides are the dominant minerals (other than quartz) in some of the siliciclastic rocks (see TSC,
Table 2). Figure 5e shows the youngest generation of pyrite lining the pore spaces and infilling to
some extent the interstices. These pyrite crystals consist exclusively of octahedral crystals {111} with
their edges beveled and smoothed by dissolution. They are poor in trace elements such as Ni, Co
and As that may substitute for Fe in the pyrite structure (Vaughan and Craig, 1978; Duchesne et al.,
1983; Dill et al., 1993, 1997). The most recent sulfide precipitation involved pyrite oxidation into Fe-
Mg-sulfates that coat the lignite-amber beds (Figure 5f). Globular aggregates of botryogen (an Mg-Fe
hydrated sulfate) are responsible for the brown coatings visible with the naked eye, and needles of
ferroan epsomite have also been detected. Rhombs of native sulfur occur along the boundary between
quartz clasts and pyrite where kaolinite was almost totally destroyed and pyrite shows solutions pits
(Figure 5d inset).

Coal/Organic Petrography and Organic Chemistry of Amber and Coal


At the Jerash site, distinct amber fragments several centimeters in size were recovered. These are most
probably terpene-derived wound resins exuded at the plant surface (Taylor et al., 1998). Within large
amber fragments, plant remains and other small sized inclusions were recognizable in fluorescence
mode.

Some amber nodules have inclusions of subparallel aligned siliciclastic debris and the host nodules
are fractured and the resultant fissures filled with illite (Figure 6a). Figure 6b shows elongated dark-
yellow to dark-brown fluorescing streaks of possible plant origin embedded in a pale-green fluorescent
resinitic mass. In addition, small inclusions of variable fluorescence, color, and intensity are detectable,
but are of unknown origin. Similarly, numerous, vague and circular dark-brown fluorescent bodies
were observed (Figure 6c) that resemble to some degree, well-developed funginite. In Figure 6d, other
elongated bright yellow and dark-brown fluorescent bodies from 5 to 25 µm length are of unknown
origin, but their shape indicates a layered deposition.

In fine-grained sandstone to siltstone, collotelenite is porous and fissured (Figures 6e, f). Where the
fine-grained sandstone to siltstone is extremely rich in massive pyrite (grain size from 100 to 400 μm),
organic matter occurs in both the pore spaces and within clayey laminas. Solely porous collotelenite
as well as vitrodetrinite were observed. Liptinite macerals are represented by thin-walled cutinite,
microspores und isolated suberinites with a dark-yellow organic fluorescence (Figures 7a, b). In
Figures 7c and d, fissured huminitic tissue (ulminite) and suberinite (Figures 7e, f) are shown in
polished section in incident-light white-field mode and under blue-light excitation.

Collotelenite also occurs in medium-grained sandstone but is poorly preserved. It is porous, dissected
by fissures and strongly corroded giving an impression of naturally weathered vitrodetrinite. In
addition, suberinite and silty intercalations with liptodetrinite were also observed. Liptinite macerals
are represented by fine alginite, dinoflagellates (?), and isolated cutinite as well as numerous
microspores (Figures 8a, b) in samples from the Jerash section.

Coarse-grained sandstones are extremely rich in massive pyrite and frequently contain coaly particles.
The organic matter in these coarse siliciclastics shows mostly intensive yellow organic fluorescence
(vitrinite). The samples contain mainly ungelified to slightly gelified textinite.

In samples from the Jerash section, random vitrinite reflectance values range from 0.40 to 0.49% Rr
(Table 3) suggesting that the coaly material is sub-bituminous A coal (“Mattbraunkohle” according to
the German coal classification). The δ13C isotopic signature of huminite from Jerash is homogeneous
at about −23‰ (Table 2). The relatively heavy carbon isotopes probably reflect a high contribution of
gymnosperms to the biomass (see Bechtel et al., 2002, 2008 and references therein).

111
Dill et al.

a b
il/kao

goe

qtz
qtz
il/kao
0 µm 1,000 0 µm 200

c d py

Scale of inset?
s
qtz
qtz

0 µm ?
kao
py/mar

qtz py
0 µm 500 0 µm 200

e f

py
bo

0 µm 20 0 µm 10

Figure 5: Micrographs to illustrate the various host rocks of the coaly and organic-rich-amber beds
near Jerash.
(a) Mudshale composed of argillaceous layers of illite (il) and kaolinite (kao) and lenses of quartz
(qtz) (thin section; crossed Nicols).
[Harald: not seen in figure]

(b) Infiltration and replacement of illite and kaolinite by goethite (goe) (thin section; crossed Nicols).
(c) Sulfide-bearing fine-grained sandstones. The interstices between the quartz grains are filled with
Fe disulfides (pyrite (py)/marcasite (mar)) (SEM-EDX).
(d) Sequence of diagenetic mineralization quartz [illite absent] ⇒ kaolinite ⇒ pyrite (SEM-EDX).
The inset shows quartz with a rim of pyrite and well-crystallized rhombs of native sulfur (s)
(image width 100 µm).
(e) Octahedral pyrite crystals growing into the open pore space caused by dissolution (SEM-EDX).
(f) Globular aggregates of Fe-Mg sulfates composed of botryogen (bo) that is responsible for the
brown coatings visible with the naked eye, and needles of ferroan epsomite, both of which
originated from oxidized marcasite and pyrite (SEM-EDX).

112
Potential energy resources, northwest Jordan

a b

phyllosilicate
streaks

il

0 µm 200 0 µm 50

c d

0 µm 20 0 µm 50

e f

ct

ct

0 µm 50 0 µm 20

Figure 6: Amber and organic petrography of samples from the Jerash section.
(a) Amber nodule containing streaks of phyllosilicates arranged subparallel and fissures infilled
with illite (il) from the surrounding mudshales (thin section; crossed Nicols). Formation of amber
nodules occurred syn- to post-compaction; aging of the nodules produced shrinkage cracks later
filled with phyllosilicates.
(b) Elongated dark-yellow to dark-brown fluorescing streaks (plant origin?) embedded in a
pale-green fluorescent resinitic mass (polished section, fluorescent light).
(c) Circular dark-brown fluorescent bodies resembling sclerotinite (polished section, fluorescent
light).
(d) Yellow and dark-brown fluorescent bodies whose origin remains uncertain (phyllosilicates?).
(e) Fissured and porous collotelinite (ct) in pyrite-poor, fine-grained sandstone to siltstone (polished
section in incident light, white field mode).
(f) Extremely fractured “dirty” collotelenite (ct) in pyrite-poor fine-grained sandstone to siltstone
(polished section in incident light, white field mode).

113
Dill et al.

a b

cu cu

0 µm 50 0 µm 50

c d

ul ul

0 µm 50 0 µm 50

e f

su su

0 µm 50 0 µm 50

Figure 7: Coal petrography of coaly and organic rich-amber beds from the Jerash section.
(a) Thinly walled cutinite (cu) in fine-grained pyritiferous sandstone to siltstone (polished section
under blue light excitation).
(b) Thinly walled cutinite (cu) in fine-grained pyritiferous sandstone to siltstones (polished
section in incident light, white field mode).
(c) Fissured huminitic tissue (ulminite (ul)) in fine-grained pyritiferous sandstones to siltstones
(polished section in incident light, white field mode).
(d) Fissured huminitic tissue (ulminite (ul)) in fine-grained pyritiferous sandstone to siltstone
(polished section under blue light excitation).
(e) Suberinite (su) in fine-grained pyritiferous sandstones to siltstone (polished section in
incident light, white field mode).
(f) Suberinite (su) in fine-grained pyritiferous sandstone to siltstone (polished section under blue
light excitation).

114
Potential energy resources, northwest Jordan

a b

cu

cu

id

0 µm 20 0 µm 20

c d

fo

co
ch

0 µm 50 0 µm 50

e f

fo
dh

al
0 µm 50 0 µm 50

Figure 8: Petrography of coaly and organic rich-amber beds, oil shales and tar sands.
(a) Microspore and liptodetrinite in medium-grained sandstone, Jerash lignite-amber beds
(polished section under blue-light excitation).
(b) Cutinite (cu) in medium-grained sandstone, Jerash lignite-amber beds (polished section under
blue-light excitation).
(c) Foraminifers (fo) and char (ch) particle embedded in shaly laminae, Sultani oil shale (polished
section in incident light, white field mode).
(d) Collotelinite embedded in shaly laminae, Sultani oil shale (polished section in incident light,
white field mode).
(e) Alginite (al) streaks and tests of globular planctic foraminifers (fo) Sultani oil shale (polished
section under blue-light excitation).
(f) Solid bitumen between quartz grains, Wadi Isal tar sand (polished section in incident light,
white field mode).
dh = ?; id = ?
115
Dill et al.

Table 3
Location Lithology %Rr n s Q

Lignite-amber beds
Jerash 0.40 60 0.05 4
fine-grained sandstone to siltstone
Lignite-amber beds fine-grained
sandstone to siltstone, extremely rich in
Jerash 0.41 50 0.05 4
massive pyrite with grain size ranging
between 100 and 400 µm
Lignite-amber beds
Jerash 0.41 50 0.05 3-4
medium-grained sandstone
Lignite-amber beds
Jerash 0.49 50 0.06 3
medium- to coarse-grained sandstone
Sultani Trench Oil shale 0.32 33 0.07 3
Wadi Isal Tar sand 0.28 13 0.04 4
Wadi Isal Tar sand 0.23 1 n/a 0
Kharazeh Oil seep 0.68 4 0.01 5
Organic petrographic data of the organic matter-bearing sediments of the Kurnub Group: %Rr = random vitrinite reflectance; n = number of
measurements; s = standard deviation; Q = evaluation of the measured vitrinite particles based on size, texture and structure of the surface, as
well as the number of measurements and standard deviation: 1 = reliability very high; 2 = high; 3 = average (moderate); 4 = low; 5 = very low; 0
= not determinable or deducible.

Four samples were analyzed using organic geochemical techniques. The total organic content (TOC)
ranges from 0.5 to 2.3 percent. Three samples are very rich in sulphur (15−30%). The hydrogen
index (HI) is about 40 to 50 mgHC/gTOC indicating a prevalence of type III kerogen. The content of
liptinite, including resinite (amber), in these samples is low. Tmax values ranging from 396 to 411°C
indicate that the organic matter is thermally immature. The amount of soluble organic matter is low
(30–70 mg/gTOC), which agrees well with the low maturity and low HI values. Gas chromatograms

[Harald: K120 occurs in the Fuweis Shale of the Lower Aljun Group of the
Azraq basin (Sharland et al., 2001). Would it not be better to use the local
show a dominance of long-chain n-alkanes (see Figure 9). Pristane/phytane ratios range from 1.2 to
3.1. Figure 9 also shows high values of C28 steranes that are typical for coaly samples.

Depositional Environment
Amireh (1997) provided the first overview of the depositional environment of the Kurnub Group. A
more detailed interpretation is given in this paper for the Jerash area (see vertical bar between 120 m
and 150 m in Figure 2). The Kurnub Group consists of a 200-m-thick upward-fining succession that
is a transition from fluvial to tide-modified sedimentation and reflects an overall deepening of the
basin (Figures 2 and 3). Full marine sub-tidal conditions with glauconitic beds were attained during
deposition of the Cenomanian Naur Formation (Figure 3). This glauconite horizon dated at 96.1 ± 1.1
Ma has been correlated with the Maximum Flooding Surface (MFS) K120 (98 Ma) of Sharland et al.
(2001), in the mid-Cretaceous Natih Formation that crops out in Wadi Mu´aydin, Oman.
correlation rather than the Oman one?]
The conglomeratic lag deposits and coarse-grained siliciclastic sediments in the lower half of the
Kurnub Group record braided-alluvial processes with a limited preservation potential for OM. The
top strata of the various cycles within the Kurnub Group are best expressed in terms of sequence
stratigraphy by the MFS K100 and MFS K110 (Figure 2), which allow for a correlation with the Burgan
delta sediments in Kuwait and Iraq on a plate-wide scale (Al-Fares et al., 1998; Sharland et al., 2001).
Leaf imprints prevail over well-developed lignite seams. Within the section studied, mudstone drapes,
and vertically stacked sandstone-siltstone and mudstone laminations or tidal bundles typical of tidal
flats, were not observed. However, some large-scale coarsening-upward cycles can be recognized
within this overall fining-upward trend (Figure 4). The stacked cycles are attributed to a fluvial
drainage system changing from alluvial-braided into meandering to anastomosing drainage patterns
that, from time to time, were interrupted by marine inundation, probably related to absolute sea
level fluctuations in a tidal-dominated system. Point-source tide-dominated delta systems gradually
became replaced by linear shoreline-dominated tidal-flat systems toward younger series.

Estuarine series with peat-forming mangroves occur in the modern Fly River Delta of Papua
New Guinea and are recorded from Permian through Eocene beds (Coleman et al., 1970; Petersen
and Andsbjerg, 1996; Holz et al., 2002; Holz and Kalkreuth, 2003; Takano and Waseda, 2003). The

116
Potential energy resources, northwest Jordan

(a) (c)
27 29

Std.
30 31 Jerash Wadi Isal
28 Coal 30 Tar Sand

Please clarify the red marks and check if the positions of pristane, phytane etc. are correct.
33
32
34
Relative Intensity

Relative Intensity
25 26

Std.
31
mC18

35
mC17 pristane

24
mC20

23 36 28 29
mC19

32
22 37
phytane

21 38 39 27

phytane
40 23

pristane
22
25
20 24 26

mC17
19 21
18

Time Time

(?) (d)
Jerash Kharazeh
Std.

Std.
Coal Oil Seep
30
29
28

Relative Intensity
Relative Intensity

mC18
27 33
31 32 20

mC17
34 19 21 22
phytane 23 24 28
26 mC16 25 26 27 29 3031
pristene

25 35
phytane

32
pristane
mC18

24 3334 36
mC17

36
23 35 37 38
21 22
19 20 37 39 40
38 39

Time Time

(b)
29 (?) C28-Steranes
Sultani Trench
Std.

phytane
27 31 Oil Shale
18 17 19 23
21 28
16
Relative Intensity

20
30
22 25
24 32
26
mC15

33
34
Kharazeh
Oil Seep
35

36 38
37
39 40 Jerash
Sultani
Coal
Oil Shale

Time C27-Steranes C29-Steranes


Figure 9: Gas Chromatograph-Flame Ionization Detector (GC-FID) traces of extracted organic
matter from Jordanian organic matter-bearing rocks.
(a) Coal-amber beds at Jerash.
(b) Oil shales from the Sultani Trench.
(c) Tar sand in Wadi Isal
(d) Oil seepage near Kharazeh.

lignite-amber site studied in the Aptian-Albian Kurnub Group is almost identical in terms of age
and depositional environment with the lignite-amber site (“Burmite”) from the Hukawng Valley of
Myanmar (Cruickshank and Ko, 2003). In conclusion, the most appropriate site to concentrate organic
matter in the form of coaly and organic-rich sediments together with amber, is in a tide-dominated
delta overlying braided/meandering river deposits (Figure 2). The amber resin surrounded the
siliciclastic debris during or subsequent to the compaction of the host sediment when associated
sheet silicates were already parallel oriented {001}. Aging of the resin provoked shrinkage cracks that
became filled with phyllosilicates (Figure 6a).

117
Dill et al.

Harald: How do the captions relate to the figures and to the Eh/pH axes?

a b
1.0 1.0

FeSO4+
What is the significance of the blue and yellow areas?

0.5 Al+++ 0.5


Eh (volts)

Eh (volts)
FeSO4+
Kaolinite

Muscovite
Hematite

0 Maximum 0
Microline
Pyrite
Magnetite
-0.5 -0.5

25°C 25°C

0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
(pH) (pH)
Figure 10: Physico-chemical conditions during and after deposition of the coaly and organic-
rich-amber beds near Jerash.
(a) Stability fields in the pH-Eh diagram at 25°C and log a Al 3+ = -4, log a SiO2 = -4 log a K+ = -4, log a
SO4
2-
= -4.
(b) Stability fields in the pH-Eh diagram at 25°C and log a Fe 2+ = -4 and log a SO4 2- = -2.

The coaly and organic-rich sediments are characterized by porous and fissured collotelenite.
Deterioration of dispersed organic matter in siliciclastic sediments resulted from long-distance
transport leading to the mechanical rupture of organic particles (Taylor et al., 1998). Other causes of
corrosion of organic matter may include weathering associated with oxidation. These can generate
changes in texture and color, respectively, causing microfractures, “dirty” and inhomogeneous surfaces,
as well as dark oxidation rims, as detected in the examined samples. Such features can, however,
also result from recrystallisation by the influx of late-stage hydrothermal fluids. On thermodynamic
grounds, such hydrothermal alteration may be ruled out for the organic matter in the Jerash section
(Figure 10). Mineral assemblages in the Jerash region are stable under near-ambient conditions
and need no temperature increase. Random vitrinite reflectance as well as dark-yellow organic
fluorescence of land-derived liptinite point toward the immature stage of the examined sediments
in terms of thermal maturation of hydrocarbons. In some samples, the organic matter appears to be
purely terrestrial. In others, the presence of lamalginite (phytoplankton), with dinoflagellates among
the humic matter of land plants indicate, at least in part, deposition within a marine environment.
These coal petrographic results corroborate the interpretation of the depositional environment.

The data obtained from organic chemistry also contribute to the environment analysis. According to
Didyk et al. (1978), high pristane/phytane ratios (> 3.0) indicated aerobic conditions during diagenesis,
and values between 1.0 and 3.0 were interpreted as reflecting anaerobic environments. However,
pristane/phytane ratios are affected by maturation (Tissot and Welte, 1984) and by differences in the
precursors for acyclic isoprenoids (bacterial) origin (Goossens et al., 1984; Volkman and Maxwell,
1986; and ten Haven et al., 1987). Sterane patterns dominated by C29 steranes (80−86%) indicate a
dominance of land plants (Volkman, 1986), whereas very low 20S/(20S+20R) sterane isomer ratios
of the 5α(H),14α(H),17α(H)-C29 steranes (<0.04) result from low maturity (Mackenzie and Maxwell,
1981).

The coaly and organic-rich-amber beds near Jerash are rich in sulfur. There is a strong environmental
control on the sulfur content of coal seams (Casagrande et al., 1977; Diessel, 1992; Phillips and Bustin,
1996, and Chou, 1997). Studies of modern peat subjected to marine influences document that S
enrichment is attributable to sulfate-reducing bacteria promoting precipitation of pyrite in peat, and
that coals having marine roof rocks have higher sulfur contents than those with fresh- or brackish-
water–derived roof rocks. Similar, abnormally high S contents have also been reported from the
estuary of the Paleo-Naab River system in Germany. Here, at the passage from a Miocene braided-river
drainage system into a meandering fluvial system hosting coaly and organic-rich sediments, arenites
are cemented with pyrite (>7 wt.% S) (Dill et al., 1993). Carbonaceous matter was concentrated in both

118
Potential energy resources, northwest Jordan

environments under discussion in the low-gradient/low-energy near-shore fluvial drainage system,


whereas sulfides were concentrated beneath in the high-gradient/high-energy more landward part
of the river system. Dill et al. (1993) referred to this phenomenon of sulfur enrichment immediately
beneath coal seams as a “sulfur keel”.

Post-depositional Alteration
Observations by Dill et al. (1997) on pyrites from various depositional environments show a close
link between the morphology of Fe disulfides (pyrite and marcasite) and processes related to the host
environments. Pyritization and the formation of marcasite are part of the post-depositional alteration
of sandstones of the Kurnub Group. Prior to the precipitation of pyrite, grain cement was leached
and the resultant pore space in the arenites became much larger than the amount of mineralizing
fluids available to fill the spaces. These well preserved pyrites show signs of exposure to fungal/
bacterial activities. Murowchick and Barnes (1987) found during their laboratory experiments that
any increase of temperature and/or degree of supersaturation sparks the formation of a sequence
from cube via octahedron to pyritohedron. These effects of temperature and degree of supersaturation
on morphology were investigated over a temperature range of 250°C to 500°C, which is beyond the
temperature regime to be expected in the environment under consideration. Marcasite forms from
aqueous solution under a restricted set of depositional conditions; for example, a pH less than 5,
temperature less than 240°C, and the presence of H2S (aq) (Murowchick, 1992).

The accumulation of organic matter and muscovite in the siliciclastic depositional environment
takes place under slightly alkaline conditions close to pH 8 (Figure 10a). Subsequent cementation
of siliciclastic minerals by kaolinite and marcasite suggest a lowering of the pH to 5 (Figure 10a).
Corrosion of sulfides and destruction of phyllosilicates attest to a hiatus between the reducing and
oxidation stages as marked by the precipitation of sulfur. Sulfur was derived from organic matter
concentrated in the interstices of the siliciclastics. According to the stability fields in the pH-Eh plot of
Figure 10b, a strong drop in pH and an increase in Eh may be assumed. Destruction of kaolinite and
formation of Fe(Mg)- sulfates are indicative of pH equal to or less than 4. An increase in the sulfur
fugacity and the temperature of formation rising to as much as 100°C would cause kaolinite to break
down, and alunite and diaspore to form instead. As these minerals were not observed in the samples
studied, we can exclude any hydrothermal effect on this lignite mineralization and the organic matter
(see previous paragraph).

Economic Considerations

Coal currently has no commercial value in Jordan. However, lignite is a marker for amber and
attention should be paid to amber as a raw material for handicrafts.

OIL SHALES

An Overview of Oil Shales along the Western Margin of the Arabian Plate

Oil shales have been discovered at various stratigraphic levels in Egypt (Troger, 1984; El Kammar,
1993; El Kammar et al., 1990). The oil shales at Al-Qusiema–Al- Kuntella and Al Thamad in East Sinai
are characterized by high contents of volatiles (9.98–8.98 wt%) and fixed carbon (9.02–21.92 wt %).
The evaluated reserves are 75 million tonnes per square kilometer. The oil shales are of Campanian
to Maastrichtian age and are somewhat older than the oil shales from Jordan discussed below. The
total volume of all organic-rich Cretaceous shale amounts to 4.5 billion barrels of oil in place in the
Safaga-Qusseir area, and 1.2 billion barrels of oil in place in the Abu Tartour area (Hassaan and Ezz-
Eldin, 2007).

Twenty oil shale deposits of Late Cretaceous age have been identified in Palestine with about 12
billion tonnes of oil-shale reserves identified (Minster, 1994). The average heating value of Palestinian
oil shales is 1,150 kcal/kg of rock with an average oil yield of 6 wt%. The organic content of these oil
shales is in the range 6 to 17 wt% with an oil yield of 60 to 71 l/t. The in-place oil shale resources are
4 billion barrels (Dyni, 2003).

119
Dill et al.

Marine oil shales of Late Cretaceous to Paleogene age occur in the Wadi Yarmouk basin of southern
Syria (Dyni, 2003). The Yarmouk deposit may prove to be an exceptionally large deposit that extends
into Jordan where, although the thickness of the oil shales diminishes, they remain of potential
economic size at several localities. As much as 90 percent of the oil shale in Jordan is amenable to
opencast mining (Dyni, 2003). The composition and size of the most important oil shale deposits are
listed in Table 4 and their locations shown in Figure 1c. The Sultani area has been selected for a more
detailed study of the diagenetic and epigenetic alteration processes, since they may also shed some
light on uranium redeposition, another source of energy that has had little publicity.

Table 4
Reserves

Jurf Ed- Attarat Umm


El-Lajjun Sultani Wadi Maghar
Darawish El-Ghudran
Area (km2) 20.4 24 150 226 --
Average thickness of oil shale (m) 29.6 31.6 63.8 45 40
Average thickness of OB (m) 25.8 69.3 47.3 53.2 40.5
Geological reserves (Mt) 1,196 11,30 8,000 11,300 31,600
Indicated reserves (Mt) 1,170 989 2,500 104,00 21,600
Chemical and Physical Properties

Jurf Ed- Attarat Umm


El-Lajjun Sultani Wadi Maghar
Darawish El-Ghudran
Average oil content (wt%) 10.5 7.5 5.7 11 6.8
Total organic matter (wt%) 22.1 21.5 18 -- --
Calorific value (kcal/kg) 1,590 1,210 864 -- --
CaCO3 (wt%) 54.3 46.96 69.11 -- --
SO3 (wt%) 4.8 4.4 4.3 -- --
Bulk density (g/cm ) 3
1.81 1.96 2.1 -- --
Moisture (wt%) 2.43 2.6 2.8 -- --
Oil shale deposits in Jordan (Source: Natural Resource Authority, Jordan). (a) Reserves. (b) Chemical and physical properties.

Oil Shales at Sultani

Stratigraphy and Geology


Oil shales of the Muwaqqar Formation form the lower part of the Belqa Group. The Muwaqqar
Formation is Maastrichtian to Paleocene in age and is well exposed in the Sultani Trench, an open pit
used for test mining (Figures 11a, b). The organic carbon content is as much as 37 percent of the total
volume in certain small basins such as El Lajjun about 20 km east of Karak (Yassini, 1979) (Figure 1c).
The sedimentary sequence at Sultani consists of bituminous marly limestones approximately 40-m-
thick, overlain by marls and marly limestones (Figure 11a). In the trench, bioclastic limestones and
biolithites are underlain by oil shales and unconformably overlain by polymictic conglomeratic wadi
sediments (Figure 11b). Limestone concretions with gastropods extend across the bottom of the trench
(Figure 11c) and are interbedded with limestone composed of disarticulated bivalve shells forming a
coquina (Figure 11d).

Petrography and Mineralogy


The oil shales at Sultani are well-laminated mudshales rich in calcite, illite, and calcareous biodetritus
predominantly of planctic foraminifers and the debris of unidentified bivalves, and also forams
that do not show any reworking or destruction of their tests (Figure 12a). Francolite (carbonate-
fluorapatite) debris is common to these mudshales (Figure 12b) as are sulfides. Zinc sulfides are most
common followed by pyrite and marcasite, as corroborated by the chemical analyses of the oil shales
that average 0.16 wt % Zn. As a result of scanning electron microscopic (SEM) observations, the ZnS
aggregates have been subdivided into cubic sphalerite (ZnS) and hexagonal wurtzite (Zn, FeS). Tiny
platelets among the ZnS aggregates suggest that in addition to sphalerite, wurtzite is present in the
oil shales (Figure 12c). The content of sphalerite/wurtzite ranges from 1.1 to 2.2 wt.% Fe.

120
Potential energy resources, northwest Jordan

(a)
Figure 11: The Muwaqqar For-

Rijam Formation/Group
mation in the Sultani oil-shale area.

Stratigraphy
Thickness

Sequence
Deposi-
(a) Columnar section through the
Period

(meter)
tional
Age
Era

Lithology
Environ- Muwaqqar Formation; the oil-
ment
shale section sampled during this
study is indicated by the vertical
bar in Figures 11a, b (modified
Eocene

90
after Abu Qudaira, 1996). A tenta-
85
tive position of for MFS K180 is
80 ? shown in the upper part of the
Shallow-marine MFS
75 with trace fossils, Muwaqqar Formation.
glauconite, and
K180
few phosphatic
(b) The trial mining oil shale opera-
70 particles tion at Sultani (depth of open pit
65 approximately 20 m).
60
(c) Calcareous concretions (25−30 cm
in size) abundant in gastropods.
Cretaceous-Paleogene

Maastrichtian-Paleocene
Mesozoic-Cenozoic

55
(d) Shell hash of randomly scattered
Small restricted
50 disarticulated bivalve shells
Muwaqqar

basins within the

45
continental shelf (“tempestite”) (pen 13 cm long).
40 B B
B B
B B
35 B B
B B
B B
30 B
B
B
B
B B
Small restricted
25 B
B
B
B
basins within the
B
B
B
B
continental shelf Limestone
20 B B with high pro-
B
B
B
B
ductivity due Marly limestone
15 B B to upwelling
B
B
B
B
B
currents B Bituminous marly limestone
10 B B
B B
Cherty marly limestone
Al-Hisa

B B
5
Shallow-marine Marl

b c

121
Dill et al.

a b

ap

fo
0 200 il 0 50
µm µm

c d

sp

wu

0 µm 50 0 µm 20

Figure 12: The mineralogy of oil shales viewed under the petrographic and scanning electron
microscope (SEM).
(a) Tests of globular planctic foraminifers (fo) in oil-bearing mudshales (thin section, crossed
Nicols).
(b) Francolite debris in illitic (il) matrix (SEM-EDX). (ap?)
(c) Aggregates of ZnS composed of isometric tetrahedral crystals of sphalerite (sp) and platelets of
wurtzite (wu) (SEM-EDX).
(d) Intergrowth of micaceous sheets of (meta)tyuyamunite (SEM-EDX).

Relatively high quantities of uranium in the oil shales at Sultani (Table 2) prompted a closer
examination of uranium minerals. The most common uranium mineral is (meta)-tyuyamunite
[Ca(UO2)2(VO4)2·6(H2O)] that precipitated from U- and V-bearing fluids percolating through the
oil shales (Figure 12d); the prefix (meta) has been added to indicate the effects of weathering. The
mineral is widespread in calcretes and in sediments enriched in organic matter (Dall’aglio et al.,
1974; Mann and Deutscher, 1978; Dill, 1983a). The presence of another yellow secondary uranium
ore mineral, wyartite [Ca3U(UO2)6(CO3)2(OH)18·3(H2O)] that is also present as tiny plates, cannot be
confirmed in this organic-rich environment by SEM analysis alone. No black uranium minerals such
as pitchblende, from which wyartite is derived, have been detected.

The oil shales also have anomalously high contents of Sr, Ni, Mo and Cr (Table 2) but at too low a
level to form minerals of their own. They are absorbed onto the organic matter or accommodated in
the structure of iron sulfides.

Organic Petrography and Organic Chemistry


The alginite and sporomorphs in dark-brown, fine-grained bituminous and calcareous shale show a
pale-yellow to intensive yellow organic fluorescence. The shale is characterized by well-developed
microlaminae and incompletely combusted particles of pyrolitic carbon and of amorphous bituminous

122
Potential energy resources, northwest Jordan

matter. The organic matter is composed of partially porous telohuminitic streaks with alginite and
sporomorphs (width 1−4 μm) (Figure 8e). Char particles and collotelenite are embedded into shaly
laminas (Figures 8c, d).

The oil shale is characterized by high TOC (18.4) and sulfur contents (5.2 %). A HI of 715 classifies the
organic matter as kerogen type I (to IIS) and agrees well with the high content of liptinite macerals. A
high percentage of C27 steranes (38 %; Figure 9) correlates with abundant alginate macerals and large
amounts of short-chain and long-chain n-alkanes. A vitrinite reflectance of 0.32 %Rr (Table 3), a Tmax
of 405 °C and the sterane isomer ratio of 0.05 show that the organic matter is immature (Peters, 1986;
Mackenzie and Maxwell, 1981).

The δ13C isotopic signature of organic matter in the oil shales of the Sultani Trench, the tar sand in Wadi

[Harald: How do GC-FID indicate these fossil communities?]


Isal, and the oil seepage near Kharazeh are almost identical and cluster around −29 ‰ compared with
–22.9 for the Jerash section (Table 2). The high proportion of lipid-rich organic material in both the oil
shale and the source rock of the migrated petroleum at Wadi Isal and Kharazeh explains the observed
depletion in δ13C relative to the Jerash coaly material.

Depositional Environment
The first description of the lithology of the Sultani oil shales was by Abu Qudaira (1996). The inorganic
and organic sedimentary features observed in the underlying and overlying rocks of the oil shales
of the Muwaqqar Formation attest to a shallow-marine depositional environment at the Cretaceous-
Paleogene boundary (Figure 11a). Strong productivity due to upwelling currents along the shelf edge
(Figure 11a) is documented by the GC-FID traces that indicate the presence of autochthonous and
allochthonous fossil communities (see Figure 9). A fairly high degree of physical and chemical stress as
a result of stenohaline conditions is a feature of this part of the Muwaqqar Formation. The stenohaline
conditions contributed to the small number of species. Bivalves are preserved disarticulated and
strong reworking of fossils was widespread. The shell hash of randomly distributed bivalve shells
reflects storm events that formed tempestites (Figure 11d). The storm surges caused ravining of the
underlying limestones. Gastropod colonies within the “ravines” attest to local tranquil conditions
that allowed mound-like built-ups to develop. The oil shale with its planctic forams was the result
of a moderate energy regime and deep-water conditions. The presence of glauconite and phosphate
particles indicates another upwelling event that occurred in the aftermath of the deposition of the oil
shales.

The upwelling currents that took cold water from the deep Neo-Tethys Ocean in the north and west
onto the epicontinental shelf were usually rich in Si, P and other elements necessary for the bloom of
planktonic organisms. This explains the completely different sedimentation regime of the Belqa Group
from that of the Ajlun Group. Abundant chert, phosphorite, porcellanite and oil shale deposits, none
of which are found in the Ajlun Group, characterize the Belqa Group. Throughout the Belqa Group,
organic-rich sediments are present in varying amounts. However, the richest and most extensive are
the oil shales of the Muwaqqar Formation (Yassini, 1979). Similar, but less important oil shales, are
present in the underlying Lower Maastrichtian Al-Hisa Formation (Table 1). The deposition of the
phosphorite and oil shale took place in small basins created by the Syrian Arc tectonic events that
were the locus of upwelling currents (Abed and Amireh, 1983; Abed et al., 2005). Within the small
basins, the lower part of the water column seems to have been anoxic with an overlying O2/H2S
interface, as in the present-day Black Sea. The basins had a shallow depth range of 40 to 50 m that may
explain the abundance of organic matter (Abed and Sadaqah, 1998).

These sedimentologic findings are supported by organic petrography. Numerous foraminifers,


bivalves, and gastropods and the presence of alginite among partially porous telohuminitic streaks
and sporomorphs, indicate short distance transportation in a proximal setting and deposition within
a lagoonal environment influenced by siliciclastic input into the small, restricted Sultani basin on the
continental shelf.

The organic chemistry supports the interpretation of the depositional environment. The high
percentage of C27 steranes (see Figure 9) correlates with large amounts of alginite derived from algae
and cyanobacteria. The rock extract is characterized by high concentrations of thiophenes. This is an
indication of enhanced organic sulfur incorporation. A very low pristane/phytane ratio of 0.29 argues
for anoxic conditions during deposition of the oil shale (Didyk et al., 1978).

123
Dill et al.

[Harald: I have added the blue-highlighted sentence. OK?]


a b
1.0 1.0

UO2H3PO+4+
0.5 0.5
Eh (volts)

Eh (volts)
VO++
UO2H2PO4+
Tyuyamunite
0
VO+++ 0 UO2(H2PO4)2 (UO2)3(PO4)2(C)

VOH++
-0.5 V2O3(C) -0.5 UO2(H2PO4)H3PO4+

25°C 25°C

0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
(pH) (pH)
Figure 13: Physico-chemical conditions during and after deposition of the oil shales in the Sultani
area. The mineralisation expressed by the stability diagrams refers to epigenetic processes
(a) Stability fields in the pH-Eh diagram at 25°C and log a Ca 2+ = -4 and log a V4 3+ = -4, log a U 4+ = -4.
(b) Stability fields in the pH-Eh diagram at 25°C and log a Ca 2+ = -4 and log a HPO4 2+ = -4, log a U 4+ =
-4.

Post-depositional Alteration
Post-depositional diagenetic to epigenetic alteration is a feature of the oil shale. Heimbach and Rösch
(1982) recorded the presence of wolchonskoite (a chromium-bearing clay mineral) in the mottled
zone of oil shales near Khan ez Zabid southeast of Amman (see Figure 1c). Chromium in silicified
travertine in the same area is indicative of hydrothermal activity and may have been leached from
the underlying oil shales.

Data collected during our petrographic examinations of the oil shales do not support high
temperatures representative of sanidinite metamorphic facies. Iron values in the Zn sulfides are low
and similar to low-temperature hydrothermal vein-type deposits. In the case of the oil shales, they
may be explained by late diagenetic to low-temperature hydrothermal processes. Uranyl vanadates
are decisive in constraining the physical and chemical conditions during post-depositional alteration
of the oil shales in the Sultani area (Figure 13). Tyuyamunite (a yellow secondary calcium-uranium-
vanadium mineral) is stable over a wide pH range as illustrated in Figure 13a. Its presence explains
the anomalous U and V values in the oil shales. However, between 200° and 300°C it is unstable. Why
tyuyamunite developed instead of its phosphate analogue autunite is not known. It may be due to
a preponderance of VO4 3+ over PO4 3+ or the presence of more acidic pore fluids that did not allow
autunite to form (Figure 13b).

Economic Perspectives

Oil Shales
Oil, gas, and coal deposits are scarce in Jordan but there are large and high-grade oil shale deposits
that are potential future energy resources (Knutson et al., 1987; Willmon, 1992; Hamarneh, 1998;
and Jaber and Probert, 1997, 1999). Of the 18 known deposits, eight, namely El-Lajjun, Sultani, Jurf
Ed-Darawish, Attarat Um El-Ghudran, Wadi Maghar, Siwaga, Khan ez Zabib and El-Thamad, have
been studied most intensively. The reserves and chemical and physical properties of the five most
important locations are listed in Table 4.

In the sites investigated, the high TOC content, the kerogen type, and the amount of soluble organic
matter (142 mg/gTOC) characterize the oil shale as an excellent hydrocarbon source rock (Peters,
1986). According to MEES (2008a), the Jordanian government and Royal Dutch Shell were in the final
stages of concluding a concession agreement that will allow Shell to explore the country’s large oil
shale reserves. Shell is expected to invest more than US$20 billion in the oil shale project over a period

124
Potential energy resources, northwest Jordan

of 20 years. The concession extends from northern Jordan to Siwaga near Karak, southeast to al-Jafr,
east to Wadi Sirhan on the border with Saudi Arabia, and north to Azraq. According to the National
Resource Authority (NRA), Jordan has 40 to 60 billion tonnes of proven oil reserves in oil shale in the
18 known deposits, with a potential to produce up to 4 billion tonnes/year of crude oil.

The NRA signed a memorandum of understanding on September 10, 2008 with a consortium of
Petrobras from Brazil and Total from France for oil shale exploration in Wadi al-Maghar (see Figure 1c)
(MEES 2008b). The consortium will carry out a comprehensive one-year exploration program. If
successful, an additional three years would be needed before the start of commercial oil production
from the shales. Jordan had previously signed agreements with Shell and Eesti Energia (Estonian
Energy), and local companies to investigate oil shale development.

Uranium
The Belqa Group is not only important for its oil shales and phosphorites but, because of its primary
uranium content, it also bridges the gap between hydrocarbons and nuclear energy. The presence of
yellow secondary uranium minerals in the oil shales is a guide to primary uranium mineralization. The
world average uranium content in phosphate rock is 50 to 200 ppm. Marine phosphorite U deposits
are below average at between 6 and 120 ppm U, whereas organic phosphorite deposits contain up to
600 ppm U (Baturin, 2002) as U may substitute for Ca in the apatite lattice. Although world uranium
resources in phosphate rock are uncertain, marine phosphorites are low-grade, large-tonnage deposits
with millions of tonnes of uranium potentially available as a by-product of phosphate mining in
Morocco: (6.9), USA (1.2), Mexico (0.15), and Jordan (0.1).

In Jordan and the northern Negev and Judean Desert of Palestine, Senonian to Paleocene phosphorites
provided U that is contained in a varied mineral assemblage of gypsum, vanadate, meta-autunite,
(meta)-tyuyamunite, strelkinite and carnotite in calcretes and gypcretes of Quaternary age. Areas
abundant in uranium concentrations in Jordan are shown in Figure 1c. Uranium was leached from
the phosphorites and migrated toward the Dead Sea Rift Valley where it was concentrated in the
soil (Ilani and Strull, 1988; Minster et al., 2004). Pedogenic layers such as gypcrete acted as lithologic
traps and fixed uranium, particularly under reducing conditions. Reasonably assured resources plus
inferred resources of approximately 80,000 tonnes of recoverable U (at US $130/kg U) at Wadi Araba-
Dana, Dubaidel-Mudawwara, Hamra-Huasha and Wadi Al Baheyya in near-surface carbonates,
place Jordan currently twelfth in terms of uranium availability worldwide. In addition, Jordanian
phosphorite is a potential source of uranium. It is difficult to calculate the recoverable U from low-
grade, large tonnage deposits such as the Jordanian phosphorite, but they may contain as much as
100,000 tonnes of U (OECD NEA & IAEA 2007). According to Xinhua (2008), Jordan is currently
negotiating a cooperation deal with a French company to purchase a nuclear reactor to produce
electricity and enriched uranium for peaceful purposes. Its civil nuclear energy program, under
which a nuclear plant will be built by 2015, will enable nuclear power to make up 30 percent of its
energy production by 2030. Jordan is also negotiating with Canada, the USA, and China on nuclear
cooperation.

TAR SANDS

An Overview of Tar Sands on the Western Margin of the Arabian Plate

Currently, the largest single tar sand deposit in the world is the Athabasca deposit in northeastern
Alberta, Canada. There has been little modern research on this unconventional hydrocarbon-based
energy resource in the Middle East. However, in earlier times, bitumen was recovered from tar sands
in Mesopotamia and shipped to Ancient Egypt. Today, the most well-known tar sand deposit is at Al-
Bushri, west of Deir ez-Zour, in eastern Syria. Here, asphalt is found in Upper Cretaceous to Miocene
rocks (Alsharhan and Nairn, 1997), the source of which may lie in the Euphrates Depression. Asphalt
shows are recorded in the Middle and Upper Cretaceous rocks of the Palmyra Fold Belt. It is believed
that they represent oil generated as a result of deep burial during the Tertiary. The Hasbaya deposit in
Lebanon (Figure 1c) has been related to similar deposits along the Dead Sea Rift. Asphalt in Cretaceous
rocks in Lebanon resulted from the concentration of oil released by Tertiary erosion (Alsharhan and
Nairn, 1997). These processes may be relevant to Jordanian tar sands.

125
Dill et al.

(a) (b)

Formation/Group

Stratigraphy
Deposi-

Thickness

Sequence
Period

(meter)
Age tional
Era

Lithology
Environ-
ment
Cenomanian

98 Ma
Naur

Subtidal
240 MFS
K120

220
Meandering
rivers
200 Carbonaceous shale
Limestone
Carbonaceous MFS
180 shale with Tide-dominated K110 Marl
delta
lignite-coal +100
Sandstone
Meandering
160 rivers
Figure 14: Tar sands in the Wadi
Cretaceous

Isal area.
Mesozoic

140
(a) Section through the Kurnub
Formation. Vertical bar denotes
Aptian-Albian

120
Kurnub

stratigraphic section under


Sandstone with investigation with respect to
100 plant debris,
unidirectional
tar sands. For sequence strati-
paleocurrents
Braided rivers
graphic correlation see also
and thick
80 sequence of Figures 2 and 3.
conglomerates
Note: the tar (b) Tar sands at outcrop in the
60
sand occurs
in this unit
Wadi Isal section. Beds of
coarse-grained to fine-grained
well-sorted sandstones whose
40
foreset lamellae in the tabular
cross bedding are accentuated
20 (dark streaks) by the migrating
hydrocarbons. (Pen 13 cm
0 long).

Tar Sands of the Kurnub Group in the Wadi Isal Area

Stratigraphy and Geology


Tar sands have been found along the north-trending Al-Kharaze Fault in the Wadi Isal area
(Figure 14). The tar sands belong to the Kurnub Group and are stratigraphically equivalent to the
section shown in Figure 2. The tar sands in the Wadi Isal area are placed stratigraphically below the
coaly and organic-rich-amber beds in Figure 3. However, the Wadi Isal lithology is very different,
particularly with regard to the amount of syngenetic organic matter. In particular, siltstones are rich
in comminuted plant material and seams of lignite. They stratigraphically underlie the tar sands. In
general, the Kurnub Group is a clastic sequence made up of clean sandstones containing only minor
plant debris that alternate with conglomerates. The arenaceous rocks, up to 160 m thick, show tabular
cross bedding whose bedsets and foreset lamellae are highlighted by the hydrocarbons that migrated
into the sandstone (Figure 14b). Limestones and marls of the Naur Formation overlie the tar sands.

Petrography and Mineralogy


The mineral grains and lithoclasts of tar sands range from fine- to coarse-grained. In the various
bedsets, the sandstones are well sorted and moderately well rounded to angular. The mature or
high-silica sandstones are composed of quartz aggregates and quartz grains that are homoaxially

126
Potential energy resources, northwest Jordan

overgrown by a younger generation of quartz (Figure 15a). The predominance of quartz is also shown
by the high SiO2 contents in the chemical composition of the quartz arenites (Table 2). The only detrital
accessory mineral identified by SEM was zircon. Organic matter (OM) is concentrated intergranular
and intragranular. Hydrocarbons filled the pore spaces between quartz grains that show evidence of
corrosion, and solid bitumen is also present along the grain boundaries between the primary quartz
nuclei and their envelope of secondary quartz (Figures 15a and b). In addition to quartz and zircon,
kaolinite is the third silicate present in the quartz arenites (Figures 15c, d). Booklets of kaolinite
occur between the quartz grains and were transformed into organo-mineralic compounds by OM
migrating into the kaolinite aggregates. In places, the kaolinite platelets form concertina-like structures
(Figure 15d). Quartz adjacent to this intimate intergrowth is strongly corroded (Figure 15d). The SEM-
EDS analyses of these mineral assemblages revealed the presence of considerable amounts of sulfur.
Pyrite and alunite are responsible for the elevated S contents (Figure 15e). Due to the tiny grain size of
these pseudocubic mineral aggregates, alunite can only be described as a member of the aluminum-

[Only solid bitumen shown in Fig. 8f.]


phosphate sulfate supergroup and cannot be investigated in detail (Störr et al., 1991; Stoffregen, 1993;
Stoffregen et al., 1994; Jambor, 1999; and Dill, 2001). Another sulfur compound identified by means
of SEM-EDS is phosphoalunogen. The most recent post-depositional alteration, affecting these OM-
bearing sandstones, was the formation of salcretes that seal the sandstones (Figure 15f).

Petrography and Organic Chemistry


The organic matter of the tar sand is composed of solid bitumen In Figure 8f, porous detrohuminite
rather than solid bitumen fills the pore space between the quartz grains. The reflectance of the solid
bitumen is low (<0.3 %Rr, Table 3). Two samples of tar sands have been investigated geochemically
(Figure 9). These samples contain about 2.3 percent organic carbon and 0.4 percent sulfur. The amount
of soluble organic matter is very high (>1,000 mg/gTOC). Hydrogen index values in the range of 530
to 570 mgHC/gTOC are typical for migrated bitumen (for example, Peters, 1986). The Production
Index (S1/(S1+S2)) yields information on the ratio between hydrocarbons present within the rock
and those that can be generated during pyrolysis. In the presence of migrated bitumen, this ratio is
typically high, but very low (<<0.1) in the present case. This is probably a result of biodegradation
that resulted in the total removal of n-alkanes. According to the gas chromatogram of the Wadi Isal tar
sand shown in Figure 9, biodegradation reached stage 6 on the classification of Peters et al. (2005).

Depositional Environment
The environment of deposition in the Wadi Isal area is similar to that of the lignite-amber site in the
Jerash area, but more proximal to the sediment source area (Figures 2 and 14). Correlation of the two
sites by means of maximum flooding surfaces is difficult. Only MFS K 120 may be traced from the
Jerash area through to this section (Figure 14). The tar sand was laid down in a fluvial environment
(Figure 14). The presence of zircon as the only heavy mineral underscores the supermaturity of
these quartz arenites. The high ZTR (zircon-tourmaline-rutile) index (Hubert, 1962) suggests strong
redeposition and alteration of the primary siliciclastics. Inter- and intragranular organic matter attest
to a polystage migration of hydrocarbon into very clean sandstones. It is not known whether this
conspicuous depletion in labile mineral constituents observed in the tar sands is due to provenance
variation, or is a product of post-depositional alteration.

Post-depositional Alteration
At Wadi Isal, solid bitumen (migrabitumen) is present in the pore spaces between quartz grains. Based
on the reflectance of about 0.2% Rr and brown fluorescence, the detected solid bitumen is classified
as albertite (Jacob, 1989). Because of the severe biodegradation (see Figure 9), the significance of
biomarker ratios for the reconstruction of the depositional environment of the source rock is limited.

Hydrocarbon migration has been a polystage process prior to silicification and following kaolinization
(Figure 15). The presence of kaolinite and aluminum-phosphate sulfates in the Wadi Isal area indicate
a rather low pH for the mineralizing fluids percolating along the north-trending Al-Kharazeh Fault.
Judging by the mineral assemblage in the Jerash and Wadi Isal areas, the pH was much lower in the
tar sands than in the lignite-amber beds (Figures 10 and 16). An increase in sulfate activity to log a
SO4
2-
= -2 increases the stability field of alunite at the expense of kaolinite (Figures 10a and 16a). The
activites (log a) are used to calculate the stability diagrams. Fluid temperatures rising to more than

127
Dill et al.

a b
C

qtz qtz
C

0 µm 100 0 µm 100

c d
kao
kao qtz

qtz
C

C
0 100 0 50
qtz µm µm

e f

aps kao
hal
C

qtz
0 µm 50 0 µm 300

Figure 15: Mineralogy of tar sands viewed under petrographic and scanning electron (SEM)
microscopes.
(a) Mature quartz sandstone with migrated inter- and intragranular hydrocarbons (C). Quartz
(qtz) grains are homoaxially overgrown by quartz and corroded along the grain boundaries
(thin section, crossed Nicols).
(b) Clean sandstone “soaked” with intergranular hydrocarbons (black filling of interstices) (thin
section, crossed Nicols).
(c) Booklets of kaolinite (kao) between quartz (qtz) grains that are replaced along grain
boundaries (thin section, crossed Nicols).
(d) Concertina textures of kaolinite (kao) with hydrocarbon that migrated between the kaolinite
plates and into the pore space between quartz and kaolinite (SEM-EDX).
(e) Pseudocubic aggregate of an unidentified aluminium-phosphate sulfate (aps) between
hexagonal kaolinite (kao) platelets (SEM-EDX).
(f) Crusts of halite (hal) on quartz (qtz) grains and bitumen (SEM-EDX).

128
Potential energy resources, northwest Jordan

a b
1.0 1.0

0.5 AlSO4+ 0.5 AlSO4+

Alunite
Alunite
Eh (volts)

Eh (volts)
Muscovite
Kaolinite
Diaspore
0 Maximum 0
Microline
Al(OH)4-
-0.5 -0.5

25°C 200°C

0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
(pH) (pH)
Figure 16: Physico-chemical conditions prior to the immigration of hydrocarbons into the tar
sands in the Wadi Isal area.
(a) Stability fields in the pH-Eh diagram at 25°C and log a Al 3+ = -4, log a SiO2 = -4 log a K+ = -4, log a
SO4
2-
= -2.
(b) Stability fields in the pH-Eh diagram at 200°C and log a Al 3+ = -4, log a SiO2 = -4 log a K+ = -4, log
a SO4 2- = -2.

100°C cause further changes to the stability field and the destruction of kaolinite to form diaspore a
mineral not seen during examination of the samples (Figure 16b). The presence of alumino-sulfates
suggests an increase in temperature and another drop of the pH value toward pH 2. Such reactive
fluids could well explain the clean nature of the supermature sandstones and also the provision of
pore space or accommodation space to take up the hydrocarbons. However, because some kaolinite
is still present, we assume fluctuating pH values of between 5 and 2.

These findings have implications not only for hydrocarbon accumulations but also for fault-bound
mineral deposits containing uranium, base metals, mercury, and antimony. Solid hydrocarbons
are concentrated along fractures and quartz veins and often are associated with black uranium
ore minerals (Dill, 1983b). There are many different categories of native allochthonous bituminous
substances found as vein deposits and defined as mineral wax (ozocerite), asphaltite (gilsonite, glance
pitch, grahamite), and asphaltitic pyrobitumen (wurtzilite, albertite, impsonite) (Jacob, 1967,1989,
1993; Cardott et al., 1989). Mercury deposits such as New Almaden, USA, formed at less than 250°C in
mineralized fracture zones containing cinnabarite, native Hg, pyrite, stibnite, chalcopyrite, sphalerite,
galena, and bornite, accompanied by quartz and dolomite and hydrocarbons (Studemeister, 1984).
These fault-related hydrocarbon deposits may also assist in better understanding some world-class
gold-copper deposits designated as “Carlin-”, “High-sulfidation-” or “Alunite-kaolinite” types.

Economic Considerations of Jordanian Tar Sands

Khraisha (1999) investigated the extraction and pyrolysis of tar sand from Wadi Isal. He concluded
that mixing time, temperature, particle size and alkali concentration are the important parameters for
bitumen recovery. Kerosene extraction yielded a maximum bitumen recovery of about 43 percent at
80°C and 180 to 250 μm particle size, whereas hot water extraction is ineffective since only a small
amount of bitumen was obtained at the same temperature. As an unconventional energy resource,
tar sands may be of interest only provided there is sufficiently large volume amenable to open-pit
extraction. As a result, stratiform tar sands have eclipsed fault-related tar sands. Also, in view of the
importance of Jordanian stratiform oil shale reserves, the fault-related tar sands are at present of little
economic value, although the situation may change in the future. In addition to the tar sands in Wadi
Isal, there are occurrences in Wadi Aheimir and Wadi Dhira (Figure 1c).

129
Dill et al.

OIL SEEPS IN THE AMMAN FORMATION IN THE KHARAZEH AREA

Stratigraphy and Geology

The Campanian Amman Formation of the Belqa Group (Table 1) contains oil seepages that are
structurally related to the north-trending Al-Kharazeh Fault in the Kharazeh area. The section studied
is in the upper part of the Amman Formation that stratigraphically underlies the section exposed in
the Sultani Trench (Figures 11a and 17a). Chalky beds of the Ghudran Formation and limestones of
the Wadi Es Sir Formation (Aijun Group) underlie the Amman Formation. The Amman Formation
is composed of limestones interbedded with chert. The overlying units attributed to the Muwaqqar
Formation are lithologically different from those reported from the Sultani Trench as they are marly
limestones about 20 m thick. In the study area, the oil seepage is located in the footwall carbonates of
the Al-Kharazeh Fault, terminated above by a secondary fault dipping at a lower angle than the main
fault and in the opposite direction (Figure 17b).

Petrography and Mineralogy

The gray dolomitic and calcitic limestones at Kaharazeh have been impregnated by oil along grain
boundaries of the rhombohedral carbonate crystals (Figure 18a). As a result of dolomitization, calcite
has been replaced by dolomite to more than 60 vol. % in some parts of the limestones. Elevated sulfur
contents in samples (Table 2) are due to the presence of swallow-tail gypsum that marks the selvage
of the fault and also cements fragments of the mudshales within the fault gouge (Figure 18b). In
addition to the gypsum, numerous oxidized massive to framboidal pyrite grains were observed. As
in the Wadi Isal tar sand, halite is the youngest mineral, forming coats of salcrete on the calcareous
host rocks. A mineral species found in this oil seepage that cannot be identified precisely is a hydrated
chlorine-bearing Ca-(Mg) silicate, possibly amstallite (CaAl(Si,Al)4O8(OH)4·((H2O),Cl). This mineral
was first recorded from a graphite quarry in Amstall, Austria, and results from hydrothermal alteration
of metamorphosed carbonaceous rocks (Quint, 1987).

Petrography and Organic Chemistry

A few sub-microscopical agglomerates of amorphous bituminous substance in pore spaces represent


organic matter. Its reflectivity is high at 0.68 %Rr. A few intercalations of black, most probably organic
matter-rich laminae, were also identified.

The organic carbon content of the oil-bearing carbonate rock at Kharazeh is about 1 percent. A very
high oxygen index indicates weathering of the organic matter, but the presence of n-alkanes suggests
that biodegradation was not severe (Figure 9). A high 20S/(20S+20R) sterane isomer ratio (0.45)
argues for a mature source rock for the oil seepage (Mackenzie and Maxwell, 1981). The presence
of C27 and C28 steranes in considerable quantities suggests a high contribution of aquatic organisms
to the organic matter in the source rock (Volkman, 1986). This is also supported by a low δ13C ratio
(Table 2).

Interpretation

Depositional Environment
The Amman Formation provides an insight into depositional environments prevailing before oil
shale deposition. It may be correlated with the Jordanian oil shales as well as other hydrocarbon sites
on the Arabian Peninsula using sequence stratigraphic planar elements such as Maximum Flooding
Surfaces (MFS) (Figures 11 and 17). The evolution of the oil shale basins of the Muwaqqar Formation
was preceded by subsidence of shallow-marine basins (with local deeps) and succeeded by another
deepening of the basin. The entire lithological section illustrated in Figure 17a reflects the depth
variations on an extended shelf bordering the Neo-Tethys Ocean. With respect to the emplacement of
organic matter, this hydrocarbon trap is more akin to the Wadi Isal area with its fault-bounded OM.

130
Potential energy resources, northwest Jordan

(a) (b)

Formation/Group

Stratigraphy
Deposi-

Thickness

Sequence
Period

(meter)
Age tional
Era

Lithology
Environ-
ment

?
MFS
Open, relatively
deep-marine
Pg10
Eocene

Rijam

60 environment
within the
continental shelf
Paleogene
Cenozoic

C
Harald: Please clarify the extent of Al-Hisa Formation

Maastrichtian-Paleocene

50
Muwaqqar

Small restricted MFS


basins within the K180
continental shelf
(ex. top = ~35 m; bottom = ~32 m).

40

Limestone
Al-Hisa

Marly limestone

Chalk
30
Chert
Shallow-marine
Campanian

with occasional MFS Figure 17: Oil seepages in the Campanian


Amman
Cretaceous

deep-marine
Mesozoic

environment;
K170 Amman Formation of the Kharazeh area
contains
oil seepages along the north-trending Al-Kharazeh
Fault.
20
(a) Columnar section through the Creta-
ceous–Paleogene series with the oil-
seepage section marked by a vertical
bar. A plate-wide correlation may be
Coniacian

Deep restricted
Ghudran

basins MFS
within the K160 achieved using the MFS of Sharland
continental shelf
10 ? et al. (2001).
(b) Oil seepage (C) in the calcareous
MFS
K150 footwall part of the steeply dipping
Wadi Essir
Turonian

Shallow-marine Al-Kharazeh Fault. Note, the oil seep-


wide shelf
age is terminated by a smaller fault
dipping at low angle (accentuated by
the dashed line).

Post-depositional Alteration
Fine crystalline carbonate in the samples from the Kharazeh area display sub-microscopic agglomerates
of amorphous bituminous substance in pore spaces between sediment grains. The carbonate texture
is fully intact and there are no signs of strong corrosion by any fluid prior to hydrocarbon migration.
Given both the numerous massive iron oxide particles and oxidized massive to framboidal pyrite
grains, the activity of sulfate-reducing bacteria may account for the incorporation of sulfur into iron-
limited shallow-marine carbonates. Unlike the Wadi Isal area, this type of organic matter mineralization
is not the high-sulfidation or alunite-type of alteration that might create or at least enhance the pore
spaces. Pyrite, gypsum, and an uncertain chlorine-bearing Ca-Al-silicate, do not provide evidence as

131
Dill et al.

a b
gyp

dol

il

0 200 0 800
µm
C µm

Figure 18: Mineralogy of oil-impregnated limestone of the Al-Kharazeh Fault.


(a) Dolomite (dol) rhombs with intergranular pore filling of oil (thin section, plane-polarized
light).
(b) Fault gouge of illite (il) impregnated with hydrocarbons (C) (bottom right), and palisades of
swallow-tail crystals of gypsum (gyp) (thin section, crossed Nicols).

to the pH-Eh regime at this site. Nevertheless, strongly alkaline or acidic conditions may be ruled out,
and approximately neutral pH values are a plausible explanation for the fluids that gave rise to the
Kharazeh oil seepage system. [Harald: Is this OK?]

Economic Considerations

The seepage system may be ranked inferior to the fault-related tar sands with regard to the economic
potential. It may, however, shed some light on deeper resources. A deep-seated hydrocarbon source
rock in the form of Silurian organic-rich graptolitic shales is present in Jordan and has given rise to the
Risha gas field (Lüning et al., 2005). These “hot” (radioactive) shales occur in two horizons worldwide
and are renowned as “low-grade-large-tonnage uranium deposits” (Dill, 1986; Dill and Nielsen, 1986).
Some of these shales also contain lenses and nodules of phosphate, pyrite, marcasite, sphalerite,
and chalcopyrite (Dill, 1986; Dill and Nielsen, 1986). A closer look at the chemical composition of
the mineralization in the vicinity of the Kharazeh fault (Table 2) shows Zn, Ni, Cu, and Co to be
anomalously enriched, also known from the oil as well as hot shales.

CONCLUSIONS

The Cretaceous-Paleogene series in Jordan formed under variable conditions along the passive shelf
margin of the Neo-Tethys Ocean and in various grabens and troughs in what was the northeastern
margin of the African Plate. In Jordan, the marine to near-shore siliciclastic-calcareous series of
the Kurnub (Aptian-Albian), Amman (Campanian) and Muwaqqar formations (Maastrichtian to
Paleocene), host various types of organic matter and related uranium mineralization. Some of the
deposits are of economic interest.

The four types of energy resources that are the subject of this paper (lignite-amber near Jerash, oil
shales at Sultani, tar sands at Wadi Isal, and oil seepages near Kharazeh) are strongly facies controlled
and were subjected to hypogene and supergene post-depositional alteration, as follows:

• Syngenetic coaly and organic-rich amber-bearing beds developed in a paralic environment


(tide-dominated delta) of the Kurnub Group near Jerash, under slightly alkaline conditions
and were altered under slightly acidic conditions at temperature of below 100°C.

• Syngenetic concentrations of organic matter occurred in a small restricted basin on the


continental shelf and led to the formation of oil shales containing abnormally high contents of
V, P, Zn and U. In the course of post-depositional alteration at temperatures well below 200°C,
yellow secondary uranium minerals were formed. The close relationship of oil shales and

132
Potential energy resources, northwest Jordan

phosphorites as well as the contemporary strongly arid climatic conditions provided favorable
conditions for the emplacement of syn(dia)genetic and surficial uranium accumulations.
[Harald: OK?]

• Epigenetic fault-related concentrations of organic matter were responsible for the Wadi Isal tar
sands that are located in a similar environment of deposition to the Jerash lignite. Alteration of
the fluvial sandstones that host the tar sands is identified as a high-sulfidation type with the
introduction of alumino-sulfate minerals.

• In contrast to the Wadi Isal alteration, epigenetic fault-related oil seepage systems in shallow-
marine calcareous rocks along the Kharazeh fault are low-sulfidation types with alumino-
sulfate minerals migrating out as a result of post-depositional alteration.

• The two Wadi Isal types of alteration may have implications for stratiform base metal and
precious metal deposits elsewhere whose emplacement involves hydrocarbons as a carrier of
metals among the mineralizing fluids.

DEDICATION

The authors dedicate this paper to the memory of Professor Dr. Friedrich Bender who passed away
on May 27, 2008. Under his supervision, the senior author started his career as an economic geologist
in the German Federal Institute for Geosciences and Natural Resources (BGR). Friedrich was Head of
the Geological Mission of BGR to Jordan from January 1961 to May 1966. The Mission was involved
in geologic investigations and the mapping of the entire country, together with mineral exploration
and the training of Jordanian geoscientists. It assisted the Jordanian Government in establishing the
National Geological Survey in 1965, which in 1966 was renamed the Natural Resource Authority
(NRA). After 1966, Friedrich continued as an advisor to the NRA. He wrote the seminal work, “The
Geology of Jordan”, published in 1974.

ACKNOWLEDGMENTS

We are indebted to I. Bitz for her assistance during mineral separation and grain-size analysis, F. Korte
for the XRF chemical analyses, D. Klosa for SEM analyses, and D. Weck who performed the XRD
analyses. All investigations were carried out in the laboratories of the Federal Institute for Geosciences
and Natural Resources (BGR) in Hannover, Germany. C. Ostertag-Henning and G. Scheeder carried
out the organic chemistry analyses. Their contribution is kindly acknowledged.

We are grateful to two anonymous reviewers for their comments and to Moujahed Al-Husseini for
his encouragement to integrate the latest press releases on Jordanian energy deals into this paper.
We thank David Grainger for his editorial assistance and support, and GeoArabia designer Arnold
Egdane for preparing the final design.

The senior author would like to express his gratitude to the German Academic Exchange Service
(DAAD) who provided financial grants for his stays at Amman, Jordan. He would like to thank all his
friends from the University of Jordan Geological Department for their great support during his stay
as a visiting professor both in the field and in the classroom.

We extend our gratitude to Professor Dr. K. Toukan, Chairman of the Jordan Atomic Energy
Commission for his talk at the University of Jordan. [Harald: Please give date.]

GLOSSARY

Alginite – a maceral of sapropelic coal (e.g., boghead-type) and organic-rich sediments (e.g., oil
shales) within the liptinite group, derived from algal matter and particularly resistant. (Taylor
et al., 1998).
Alkane – any of the series of saturated hydrocarbons including methane, ethane, propane, and higher
members. [Harald: I’ve added alkane. OK?]

133
Dill et al.

Biolitites – An inclusive category for all organic limestone. It is usually applied to rocks made up of
organic structures in growth position and not to debris broken from the bioherm and forming
pocket fillings or talus slopes.
Collotelinite – a maceral of the vitrinite group, subgroup telovitrinite, with a homogenous, more-or-
less structurless appearance occurring in coals and organic rich sediments. (ICCP, 1998).
Cutinite – a maceral of coal and organic rich sediments within the liptinite group, derived from
cuticular layers and cuticles, formed within the outer walls of the epidermis of leaves, steams,
and other aerial parts of plants. (Taylor et al., 1998).
Detrohuminite –a subgroup of the maceral group huminite consisting of fine humic fragments (<10μm)
occurring in coals and organic rich sediments. It is composed of loosely packed cell fragments
of other humic plant debris. Depending on its gelification, detrohuminite is subdivided into
the macerals attrinite (not gelified) and densinite (gelified). (Sýkorová et al. 2005).
Dinoflagellate – a one-celled microscopic, flagellated organism, chiefly marine and usually solitary,
with resemblances to both animal and plant kingdoms.
Liptinite – a coal maceral group including sporinite, cutinite, alginite, resinite, and liptodetrinite, derived
from spores, cuticular matter, resins, and waxes (Taylor et al., 1998). It occurs also in organic
rich sediments.
Huminite – a group of macerals in brown coals, consisting of humic matter derived mainly from
lignin and cellulose occurring also in organic-rich sediments. It is the precursor of the vitrinite
group in bituminous coals. (Sýkorová et al., 2005).
Lamalginite – a maceral of the alginite type. It is derived from small, unicellular or thin-walled,
colonial planktonic or benthic algae with distinct lamellar form with little recognizable
structures in sections perpendicular to bedding. It occurs in coals and organic-rich sediments.
(Hutton et al., 1980)
Liptodetrinite – a maceral of coal and organic-rich sediments within the liptinite group, having no
recognisable structure and low reflectance and fluorescence. Because of its finely divided
condition it cannot be assigned with certainty to any of the other macerals of the group. (Taylor
et al., 1998).
Mudshale – a consolidated sediment consisting of no more than 10 percent sand and having a silt/
sand ratio of between 1:2 and 2:1; a fissile mudstone.
Phytane – a diterpenoid alkane.
Pristane – a natural saturated terpenoid alkane.
Resinite – a maceral of coal and organic-rich sediments within the liptinite group, consisting of resinous
compounds, often in elliptical or spindle-shaped bodies representing cell-filling matter or resin
rodlets. (Taylor et al., 1998).
Funginite – a maceral of coal and organic rich sediments within the inertinite group, consisting of
the sclerotia (hard, rigid fragments) of fungi or fungal spores, hyphae and mycelia (stromata,
mycorhiza). ICCP (2001).
Sporomorphs – originate from the outer cell walls of fossil pollen grains or spores. It occurs in coals
and organic-rich sediments. (Taylor et al., 1998).
Stenohaline – said of a narrow range of salinity.
Suberinite – a maceral of coal and organic-rich sediments derived from corkified cell walls that occur
mainly in barks, and also at the surface of roots, on stems and on fruits as a protection against
desiccation. (Taylor et al., 1998).
Telohuminitic – see telohuminite
Telohuminite – a subgroup of the maceral group huminite comprising macerals with preserved
intact variably visible botanical cell structures, and isolated cells. The subgroup consists of the
macerals textinite and ulminite. It occurs in coals and organic-rich sediments. (Sýkorová et al.,
2005).
Terpene – a large and varied class of hydrocarbons derived from a wide variety of plants, particularly
conifers. They are the main components of resins.
Textinite – a maceral of the huminite group, subgroup telohuminite, consisting of ungelified cell walls
either of isolated but intact individual cells or within tissues. It consists of humic substances
as well as of the remains of cellulose and lignin. Textinite derives from the cell walls of
parenchymatous and woody tissues of roots, stems and barks, rarely also from leaves. It
originated from both herbaceous and aborescent plants. It occurs in coals and organic-rich
sediments. (Sýkorová et al., 2005).

134
Potential energy resources, northwest Jordan

Thiophene – a hetroclycic colorless liquid that closely resembles benzene; the simplest sulphur-
containing compound. [Harald: Thiophene added. OK?]
Ulminite –a maceral of the huminite group, subgroup telohuminite that denotes the cell walls of more
or less gelified tissues. It consists of humic acids, humates and traces of lignin and cellulose. It is
derived from parenchymatous and woody tissues of roots, stems, barks, and leaves. (Sýkorová
et al., 2005).
Vitrodetrinite – a maceral of the subgroup detrovitrinite within the vitrinite group, occurring as
discrete small vitrinitic fragments of varying shape that become discernible when surrounded
by non-vitrinitic material. The particle size is less than 10 µm in the maximum dimension for
rounded grains and 10 µm in the minimum direction for thread-shaped fragments. It occurs in
coals and organic-rich sediments. (ICCP, 1998).
Vitrinite – a term introduced by Stopes (1935) to denote a microscopically recognizable constituent of
medium-rank coal. Vitrinite designates a group of grey macerals. It occurs in coal as relatively
pure layers or lenses ranging in thickness from several micrometers to several centimetres; as
the continuous phase of the coal`s groundmass binding other coal components; or as amorphous
fillings of cells, pores and fissures. It occurs in coals and organic-rich sediments (ICCP, 1998).
[Harald: Stopes (1935) is not in listed in References.]

REFERENCES
Abed, A.M. 1982. Depositional environments of the Early Cretaceous Kurnub (Hathira) sandstones, north
Jordan. Sedimentary Geology, v. 31, p. 267-279.
Abed, A.M. and B.S. Amireh 1983. Petrography and geochemistry of some Jordanian Oil Shales from North
Jordan. Journal of Petroleum Geology, v. 5, p. 261-274.
Abed, A.M. and R. Sadaqah 1998. Role of oyster bioherms in the deposition and accumulation of high-
grade phosphorites in central Jordan. Journal of Sedimentary Research, v. 68, p. 1009-1020.
Abed, A. M., K.H Arouri and C.J. Boreham 2005. Source rock potential of the Phosphorite-Bituminous
Chalk-Marl Sequence in Jordan. Marine and Petroleum Geology, v. 22, p. 413-425.
Abu-Ajameih, M.M., F.K. Bender, R.N. Eicher, K.K. El-Kaysi, F. Nimri, B.H. Qudah and K.H. Sheyyab 1988.
Natural Resources of Jordan. Natural Resources Authority, Amman, 224 p.
Abu Qudaira, M. 1986. The geology of the Wadi Attarat Umm Ghadran area. Geological Mapping Division,
Geological Directorate, National Resource Authority, Jordan, Bulletin 33, 31 p.
Al-Aswad, A.A. 1997. Stratigraphy, sedimentary environment and depositional evolution of the Khuff
Formation in South-Central Saudi Arabia. Journal of Petroleum Geology, v. 20, p. 307-326.
Al-Fares, A.A., M. Bouman and P. Jeans 1998. A new look at the Middle to Lower Cretaceous stratigraphy,
offshore Kuwait. GeoArabia, v. 3, no. 4, p. 543-560.
Al-Jallal, I.A. and A.S. Alsharhan 2005. Arabia and the Gulf. In, R. Selley, R. Cocks and I. Plimer (Eds.),
Encyclopedia of Geology, Elsevier, p. 140-152
Alsharhan, A.S. and A.E.M. Nairn 1995. Tertiary of the Arabian Gulf: sedimentology and hydrocarbon
potential. Palaeogeography, Palaeoclimatology, Palaeoecology, v.114, p.369-38.
Alsharhan, A.S. and A.E.M. Nairn 1997. Sedimentary basins and petroleum geology of the Middle East.
Elsevier, Amsterdam, 843 p.
Al-Rifaiy, I.A., O.H. Cherif and B.A. El-Bakri 1993. Upper Cretaceous foramineral biostratigraphy and
paleobathymetry of the Al-Baqa area, north of Amman (Jordan). Journal of African Earth Science, v. 17,
p. 343-357.
Al-Siddiqi, A. and R.A. Dawe 1999. Qatar’s oil and gas fields: A review. Journal of Petroleum Geology,
v. 22, p. 417-436
Amireh, B.S., 1993. Sedimentology and mineral composition of the Kurnub sandstone in Wadi Qsieb, SW
Jordan. Sedimentary Geology, v. 78, p. 267-283.
Amireh, B.S., 1997. Sedimentology and palaeogeography of the regressive-transgressive Kurnub Group
(early Cretaceous) of Jordan. Sedimentary Geology, v. 112, p. 69-88.
Amireh, B.S., 1999. The Early Cretaceous Kurnub Group of Jordan: Subdivision, characterization and
depositional environment development. Neues Jahrbuch für Geologie und Paläeontologie Monatschefte,
v. 1999, p. 289-300.
Amireh, B.S., G. Jarrar, F. Henjes-Kunst and W. Schneider 1998. K-Ar dating, X-Ray diffractometry, optical
and scanning electron microscopy of glauconites from the early Cretaceous Kurnub Group of Jordan.
Geological Journal, v. 33, p. 49-65.
Azar, D. 2000. Les Ambres Mésozoïques du Liban. Unpublished PhD thesis, Université de Paris XI, Orsay,
164 p.

135
Dill et al.

Bandel, K. and A. Haddadin 1979. The depositional environments of the amber-bearing rocks in Jordan.
Dirasat, v. 6, p. 39-65.
Baturin, G.N. 2002. Manganese and Molybdenum in Phosphorites from the Ocean. Lithology and Mineral
Resources, v. 37, p. 412-428
Bechtel, A., R.F. Sachsenhofer, R. Gratzer, A. Lücke and W. Püttmann 2002. Parameters determining the
carbon isotopic composition of coal and fossil wood in the Early Miocene Oberdorf lignite seam (Styrian
Basin, Austria). Organic Geochemistry, v. 33, p. 1001-1024.
Bechtel, A., R. Gratzer, R.F. Sachsenhofer, J. Gusterhuber, A. Lücke and W. Püttmann 2008. Biomarker
and carbon isotope variation in coal and fossil wood of Central Europe through the Cenozoic.
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 262, p. 166-175.
Bender, F. 1974. Geology of Jordan. Gebrueder Borntraeger, Stuttgart, 196 p.
Bender, F. 1994. Jordan. In, H. Kulke (Ed.), Regional Petroleum Geology of the World, P. I. Beiträge Regionale
Geologie der Erde v. 21, Gebrueder Borntraeger, Stuttgart, p. 487-493.
Bowen, R. and U. Jux 1987. Afro-Arabian Geology, 295 p. Chapman and Hall, London.
Cardott, B.J., N.H. Suneson and C.A. Ferguson 1989. The Page impsonite mine, Le Flore County, Oklahoma.
Oklahoma Geological Survey, Oklahoma Geology Notes, v. 49, p. 32-39.
Casagrande, D.J., K. Siefert, C. Berschinski and N. Sutton 1977. Sulphur in peat-forming systems of the
Okefenokee Swamp and Florida Everglades: origins of sulphur in coal. Geochimica et Cosmochimica
Acta, v. 44, p. 25-32.
Chou, L., 1997. Geological factors affecting the abundance, distribution and speciation of sulphur in coals.
In, Q. Yang (Ed.), Geology of Fossil Fuels – Coal. Proceedings 30th International Geological Congress
v. 18, pt. B, VSP, Utrecht, The Netherlands, p. 47–57.
Coleman, J.M., S.M. Gagliano and W.G. Smith 1970. Sedimentation in a Malaysian high tide tropical delta.
In, J.P. Morgan (Ed.), Deltaic sedimentation: modern and ancient. Society of Economic Paleontologists
and Mineralogists, Special Publication 15, p. 185-197.
Collinson, J.D. and D.B. Thompson 1982. Sedimentary Structures. Allen & Unwin, London.
Cruickshank, R.D. and K. Ko 2003. Geology of an amber locality in the Hukawng Valley, Northern Myanmar.
Journal of Asian Earth Sciences, v. 21, p. 441-455
Dall’aglio, M., R. Gragnani and E. Locardi 1974. Geochemical factors controlling the formation of the
secondary minerals of uranium. International Atomic Energy Authority, Symposium Proceedings,
Formation of Uranium Ore Deposits, Athens, May 6–10, 1974.
Didyk, B.M., B.R.T. Simoneit, S.C. Brassell and G. Eglinton 1978. Organic geochemical indicators of
paleoenvironmental conditions of sedimentation. Nature, v. 272, p. 216-222.
Diessel, C.F.K. 1992. Coal-Bearing Depositional Systems, Springer, New York 721 p.
Dill, H.G. 1983a. First occurrences of some uranium minerals and sulfides in the uranium exploration sites
of N. Bavaria, Germany . Neues Jahrbuch für Mineralogie Abhandlungen, v. 147, p. 184-190.
Dill, H G. 1983b. Vein- and metasedimentary-hosted carbonaceous matter and phosphorus from NE
Bavaria (F.R. Germany) and their implication on syngenetic and epigenetic uranium concentration.
Neues Jahrbuch für Mineralogie Abhandlungen, v. 148, p. 1-21.
Dill, H.G. 1986. Metallogenesis of the Early Paleozoic Graptolite Shales from the Graefenthal Horst
(Germany). Economic Geology, v. 81, p. 889-903.
Dill, H.G. 2001. The geology of aluminium phosphates and sulphates of the alunite supergoup: A review.
Earth Science Reviews, v. 53, p. 25-93.
Dill, H.G. and H. Nielsen 1986. Carbon-sulphur-iron-variations and sulphur isotope patterns of Silurian
Graptolite Shales (Germany). Sedimentology, v. 33, p. 745-755.
Dill, H.G., H. Wehner and N. Blum 1993. The origin of sulfide accumulation (“sulfur keel”) in arenaceous
rocks beneath carbonaceous horizons in fluvial depositions of late Paleozoic through Cenozoic age (SE-
Germany). Chemical Geology, v. 104, p. 159–173.
Dill, H.G., E. Eberhard and B. Hartmann 1997. Use of variations in unit cell length, reflectance and hardness
determining the origin of Fe disulphides in sedimentary rocks. Sedimentary Geology, v. 107, p. 281-
301.
Dill, H.G., H. Wehner, J. Kus, R. Botz, Z. Berner, D. Stüben and A. Al-Sayigh 2007. The Eocene Rusayl
Formation, Oman, carbonaceous rocks in calcareous shelf sediments: environment of deposition,
alteration and hydrocarbon potential. International Journal of Coal Geology, v. 72, p. 89-123.
Duchesne, J.C., A. Rouhant, C. Schoumacher and H. Dillen 1983. Thallium, nickel, cobalt and other trace
elements in iron sulphides from Belgian lead-zinc vein deposits. Mineralium Deposita, v. 18, p. 303-
313.
Dyni, J.R. 2003 Geology and resources of some world oil-shale deposits. Oil Shale, v. 20, p. 93-252.
Dyni, J.R. 2005. Geology and resources of some world oil-shale deposits. U.S. Geological Survey Scientific
Investigations Report 2005-5294, 42 p.

136
Potential energy resources, northwest Jordan

El Kammar, M.M. 1993, Organic and inorganic compounds of the Upper Cretaceous-Lower Tertiary black
shale from Egypt and their hydrocarbon potentialities. Unpublished Ph.D. Thesis, Faculty of Science,
Cairo University, 226 p.
El Kammar, A.M., M.M. Darwish, G. Philips and M.M. El-Kammar 1990. Composition and origin of black
shale from Quseir area, Red Sea Coast, Egypt. Journal of University of Kuwait (Science), v. 17, p. 177-
190.
Espitalié, J., G. Deroo and F. Marquis 1977. La pyrolyse Rock-Eval et ses applications, Partie I. Revue
l’Institute Français du Pétrole, v. 40/5, p. 563-579.
Goossens, H., J.W. de Leeuw, P.A. Schenck and S.C. Brassell 1984. Tocopherols as likely precursors of
pristane in ancient sediments and crude oils. Nature, v. 312, p. 440-442.
Hamarneh, Y. 1998. Oil Shale Resources Development in Jordan. National Resources Authority, Jordan,
82 p.
Hassaan, M.M. and M.A. Ezz-Eldin 2007. The black shale in Egypt: a promising tremendous resource of
organic matter – Let it Flow. Canadian Society of Petroleum Geologists/Canadian Society of Exploration
Geophysicists Convention, Cairo, Egypt, p. 664-665.
Hassaan, M., N.A. Abdel Hafez, A.A. Ahmed, A.A. Dardir and M.A. Erian 1992. Geological and geochemical
studies of the Safa and Masajid Formations and associated coal seam, Risan Eneiza, Northern Sinai,
Egypt. Egyptian Journal of Geology, v. 36, p. 233-252.
Haven ten, H.L., J.W. de Leeuw, J. Rullkötter and J.S. Sinninghe Damste 1987. Restricted utility of the
pristane/phytane ratio as a paleoenvironmental indicator. Nature, v. 330, p. 641-643.
Heimbach, W. and H. Rösch 1980. Die “Mottled Zone“ in Zentraljordanien. Geologisches Jahrbuch, v. B 40,
p. 3-17.
Holz, M. and W. Kalkreuth 2003. Sequence stratigraphy and coal petrology applied to the Early Permian
coal-bearing Rio Bonito Formation, Paraná Basin, Brazil. In, J.C. Pashin and R.A. Gastaldo (Eds.),
Sequence Stratigraphy, Paleoclimate and Tectonic of Coal-bearing Strata. 2004 51. American Association
of Petroleum Geologists Studies in Geology, p. 147-167.
Holz, M., W. Kalkreuth and I. Banerjee 2002. Sequence stratigraphy of paralic coal-bearing strata: an
overview. International Journal of Coal Geology, v. 48, 147-179.
Hubert, J.F. 1962. A zircon-tourmaline-rutile maturity index and the interdependence of the composition
of heavy-mineral assemblages with the gross composition and texture of sandstones. Journal of
Sedimentary Petrology, v. 32, p. 440-450.
Hufnagel, H., H. Schmitz and K. El-Kaysi 1980. Investigation of the Al-Lujjun Oil Shale Deposits. BGR,
Technical Cooperation Project No. 7821655. BGR, Hannover. [Harald: Please spell out BGR.]
Hutton, A.C. and A.C. Cook 1980. Influence of alginite on the reflectance of vitrinite from Joadja. NSW and
some other coals and oil shales containing alginite. Fuel 59: 711-714.
Ilani, S. and A. Strull 1988. Uranium mineralization in the Judean Desert and in the northern Negev, Israel.
Ore Geology Reviews, v. 4, p. 305-314.
Illenberger, W. 1991. Pebble shape (and size!). Journal of Sedimentary Petrology, v. 61, p. 756-767.
International Committee for Coal and Organic Petrology (ICCP) 1998. The new vitrinite classification
(ICCP System 1994). Fuel, v. 77, no. 5. p. 349-358.
International Committee for Coal and Organic Petrology (ICCP) 2001. The new inertinite classification
(ICCP System 1994). Fuel, v. 80, no. 4. p. 459-471.
Jaber, J.O. and S.D. Probert 1997. Exploitation of Jordanian oil shales. Applied Energy, v. 58, p.161-175.
Jaber, J.O. and S.D. Probert 1999. Predicted environmental and social impacts of the proposed oil shale
integrated tri-generation system. Oil Shale v. 16, p. 2-29.
Jacob, H. 1967. Petrologie von Asphaltiten und asphaltischen pyrobitumina. Erdöl und Kohle, Erdgas,
Petrochemie, v. 20, p. 393-400.
Jacob, H. 1989. Classification, structure, genesis and practical importance of natural solid oil bitumen
(“migrabitumen”). International Journal of Coal Geology, v. 11, p. 65-79.
Jacob, H. 1993. Nomenclature, classification, characterization, and genesis of natural solid bitumen
(migrabitumen). In, J. Parnell, H. Kucha, and P. Landais (Eds.), Bitumens in ore deposits. New York,
Springer-Verlag, p. 11-27.
Jambor, J.L. 1999. Nomenclature of the alunite supergroup. Canadian Mineralogist, v. 37, p. 1323-1341.
Jerram, D.A. 2001. Visual comparators for degree of grain-size sorting in two and three dimensions.
Computers and Geosciences, v. 27, p. 485-492.
Kaddumi, H.F. 2005. Amber of Jordan: the Oldest Prehistoric Insects in Fossilized Resin. Published by the
author, Amman, 168 p.
Khraisha, Y.H. 1999. Study of extraction and pyrolysis of Jordan tar sand. International Journal of Energy
Research, v. 23, p. 833-839.
Kim, J., T. Wagner, M. Bachmann and J. Kuss 1999. Organic facies and thermal maturity of Late Aptian to

137
Dill et al.

Early Cenomanian shelf deposits, Northern Sinai (Egypt). International Journal of Coal Geology v. 39,
p. 251-278.
Knutson, C.F., G.F. Dana and G. Solti 1987. Developments in oil shale. American Association of Petroleum
Geologists Bulletin, v. 71, p. 374-383.
Lüning, S., Y.M. Shahin, D. Loydell, H.T. Al-Rabi, A. Masri, B. Tarawneh and S. Kolonic 2005. Anatomy of
a world-class source rock: Distribution and depositional model of Silurian organic-rich shales in Jordan
and implications for hydrocarbon potential. American Association of Petroleum Geologists Bulletin,
v. 89, 1397-1427.
Mackenzie, A.S. and J.R. Maxwell 1981. Assessment of thermal maturation in sedimentary rocks by
molecular measurements. In, J. Brooks (Ed.), Organic maturation studies and fossil fuel exploration.
Academic Press, London, p. 239-254.
Mann, A.W. and R.L. Deutscher 1978. Genesis principles for the precipitation of carnotite in calcrete
drainages in Western Australia. Economic Geology, v. 73, p. 1724-1737.
Masri, M. 1963. The Geology of the Amman-Zerqa area. Unpublished Report, Central Water Authority,
Amman, Jordan, 74 p.
MEES 2008a. Jordan and Shell close to reaching an agreement for Oil Shale. Middle East Economic Survey,
v. 51, no. 29, p. 16, July 21, 2008, Nicosia, Cyprus.
MEES 2008b. NRA signs MOU with Petrobras and Total for oil shale exploration. Middle East Economic
Survey, 15 September 2008, Nicosia, Cyprus.
Minster, T. 1994. The role of oil shale in the Israeli energy balance. Energia. University of Kentucky, Center
for Applied Energy Research v. 5, p. 4-6.
Minster, T., S. Ilani, J. Kronfeld, O. Even and D. Godfrey-Smith 2004. Radium contamination in the Nizzana-
1 water well, Negev Desert, Israel. Journal of Environmental Radioactivity, v. 71, p. 261-273.
Mostafa, A.R. and M.A. Younes 2001. Significance of organic matter in recording paleoenvironmental
conditions of the Safa Formation coal sequence, Maghara Area, North Sinai, Egypt. International
Journal of Coal Geology, v. 47, p. 9-21.
Murowchick, J.B. 1992. Marcasite inversion and the petrographic determination of pyrite ancestry. Economic
Geology, v. 87, p.1141-1152.
Murowchick, J.B. and H.L. Barnes 1987. Effects of temperature and degree of supersaturation on pyrite
morphology. American Mineralogist, v. 72, p. 1241-1250.
OECD, NEA and IAEA 2007. Uranium 2007. Resources, Production and Demand (‘Red Book’). Organisation
for Economic Co-operation and Development, Nuclear Energy Agency and the International Atomic
Energy Authority.
Peters, K.E. 1986. Guidelines for evaluating petroleum source rock using programmed pyrolysis. American
Association of Petroleum Geologists Bulletin v. 70, p. 318-329.
Peters, K.E., C.C. Walters and J.M. Moldowan 2005. The Biomarker Guide. 2 vol., Cambridge University
Press, 1155 p.
Petersen, H.I. and J. Andsbjerg 1996. Organic facies development within Middle Jurassic coal seams,
Danish Central Graben, and evidence for relative sea-level control on peat accumulation in a coastal
plain environment. Sedimentary Geology, v. 106, p. 259-277.
Phillips, S. and M. Bustin 1996. Sulphur in the Changuinola peat deposit, Panama, as indicator of the
environments of deposition of peat and coal. Journal of Sedimentary Research, v. 66, p. 184-196.
Poinar, G.O. and R. Milki 2001 Lebanese amber, the oldest insect ecosystem in fossilized resin, Oregon State
University Press, Portland, 96 p.
Quint, R. 1987. Description and crystal structure of amstallite, CaAl(OH)2[Al0.8Si3.2O8(OH)2] .
[(H2O)0.8Cl0.2]; a new mineral from Amstall, Austria. Neues Jahrbuch Mineralogie Monatsheft, v. 1987,
p. 253-262.
Sadooni, F.N. and A.S. Alsharhan 2004. Stratigraphy, lithofacies distribution and petroleum potential of the
Triassic strata of the Northern Arabian plate. American Association of Petroleum Geologists Bulletin,
v. 88, p. 515-538.
Schulze, F., J. Kuss and A. Marzouk 2005. Platform configuration, microfacies and cyclicities of the upper
Albian to Turonian of west-central Jordan. Facies, v. 50, p. 505-527.
Selby, M.J. 1987. Rock slopes. In, M.G. Anderson and K.S. Richards (Eds.), Slope stability: Geotechnical
Engineering and Geomorphology, Wiley, Chichester, New York, p. 475-504.
Sharland, P.R., R. Archer, D.M. Casey, R.B. Davies, S.H. Hall, A.P. Heward, A.D. Horbury, and M.D.
Simmons 2001. Arabian Plate Sequence Stratigraphy. GeoArabia Special Publication 2, 371 p.
Störr, M., H.M. Köster, H. Kromer and M. Hilz 1991. Minerale der Crandallit-Reihe im Kaolin von Hirschau-
Schnaittenbach, Oberpfalz. Zeitschrift für geologische Wissenschaften, v. 19, p. 677-683.
Stoffregen, R.E. 1993. Stability relations of jarosite and natroalunite at 100-250°C. Geochimica et
Cosmochimica Acta, v. 58, p. 903-916.

138
Potential energy resources, northwest Jordan

Stoffregen, R.E., R.O. Rye and M.D. Wasserman 1994. Experimental studies of alunite: I 18O and D-H
fractionation factors between alunite and water at 250-450°C. Geochimica et Cosmochimica Acta, v. 58,
p. 903-916.
Stopes, 1935. [Harald: Please add reference -- new reference in Glossary (vitrinite)]
Studemeister, P.A. 1984. Mercury deposits of Western California: an Overview. Mineralium Deposita, v. 19,
p. 202-207.
Sýkorová, I., W. Pickel, K. Christianis, M. Wolf, G.H. Taylor and D. Flores 2005. Classification of huminite
– ICCP System, 1994. International Journal of Coal Geology, v. 62, p. 85-106.
Takano, O. and A. Waseda 2003. Sequence stratigraphic architecture of a differentially subsiding bay to
fluvial basin: the Eocene Ishikari Group, Ishikari Coal Field, Hokkaido, Japan. Sedimentary Geology,
v. 160, p. 131-158.
Taylor, G.H., M. Teichmüller, A. Davis, C.F.K. Diessel, R. Littke and P. Robert 1998. Organic Petrology.
Gebrüder Borntraeger, Berlin-Stuttgart, p.59-85.
Tissot, B.T. and D.H. Welte 1984. Petroleum Formation and Occurrences, 2nd Edition, Springer, Berlin.
Toukan, K. 2008. Oral presentation on nuclear energy in Jordan given to Faculty of Law, University of
Jordan, Amman. [Harald: please give date of presentation.]
Troger, U. 1984, Upper Cretaceous rocks in Quseir area. Bulletin Desert Institute of Egypt, v. 7, p. 35-53.
Vaughan, D.J. and J.R. Craig 1978. Mineral chemistry of metal sulphides. Transactions of the Institution of
Mining and Metallurgy, 493 p.
Volkman, J.K. 1986. A review of sterol markers for marine and terrigenous organic matter. Organic
Geochemistry, v. 9, p. 83-99.
Volkman, J.K. and J.R. Maxwell 1986. Acyclic isoprenoids as biological markers. In, R.B. Johns (Ed.),
Biological Markers in the Sedimentary Record. Elsevier, Amsterdam, p. 1-42.
Willmon, G.J. 1992. Oil shale and natural bitumen. In, M.G. Schomberg (Ed.), Survey of Energy Resources,
World Energy Council, p. 59-66.
Xinhua News Agency 2008. http://news.xinhuanet.com/english/2008-06/30/content_8459738.htm
Yassini, I. 1979. Maastrichtian-Lower Eocene biostratigraphy and the planktonic foraminiferal biozonation
in Jordan. Revista Espanola de Micropaleontologia, v. 11, p. 5-57.

ABOUT THE AUTHORS

Harald G. Dill received his MSc in Geology from Würzburg University


in 1975 followed by studies in economic geology at Technical University in
Aachen. In 1978, he graduated from Erlangen University with a PhD for
work on pyritiferous Pb-Cu-Zn deposits in Tuscany, Italy. Subsequently,
he had a one-year research post at Bayreuth University. Since 1979
he has worked for the Federal Institute for Geosciences and Natural
Resources (BGR), mainly involved in radiometric dating and in the study
of uranium deposits. From 1986 through 1991 he was a member of the
project management group of the Continental Deep Drilling Program of
the Federal Republic of Germany, being responsible for economic geology,
mineralogy and geochemistry. In 1982 he was made lecturer assistant
professor in applied geology at Mainz University, where he obtained his Dr. rer. nat habil. degree
in 1985 [Is this correct?]. In 1991 he was appointed associate professor at Hannover University,
where he lectures in economic geology, and in 2008 he was awarded an honorary professorship
at Mainz University. Through BGR’s technical cooperation schemes, he is involved in training
geologists from partner geological surveys in Asia and Africa and also teaches university courses
in applied sedimentology and economic geology. His main interests lie in the chemistry and
mineralogy of ancient and modern depositional systems and related fossil fuel, metallic and non-
metallic deposits. His work has led to 220 publications and the Quintino-Sella-Prize awarded at
the 32nd International Geological Congress (Florence, Italy). He was involved in the discovery
of a smectite-bearing clay deposit and a gypsum-celestite deposit. [Harald: Where are these
deposits? When were they discovered?]
dill@bgr.de
www.hgeodill.de

139
Dill et al.

Author: to send high-resolution color photo, at least 300 dpi, passport size in JPEG or TIFF.
Jolanta Kus is an internationally accredited coal and organic petrographer
at the Federal Institute for Geosciences and Natural Resources (BGR)
in Hannover (Germany). She recieved her BSc in geology (emphasis
palaeontology) and her MSc in exploration geology from the Imperial
College London (United Kingdom). From 1999 to 2000 Jolanta worked as
an exploration geologist at CalEnergy Gas Ltd. (London) where she was
involved in gas exploration in NW Poland. She joined BGR (Hannover) in
2000 and worked on the structural geology and hydrocarbon potential of
the Chilean accretionary complex. Since 2005 she is involved in the Sino-
German Research Initiative on coal fires in China. Her research interest
focuses on the evaluation of the hydrocarbon potential of sedimentary
basins world wide. She is a member of PESGB, ICCP and AKOP.
j.kus@bgr.de

Abdulkader M. Abed was awarded his PhD in sedimentology and

photo, at least 300 dpi, passport


send new high-resolution color
Author: poor photo quality; to
sedimentary geochemistry by Southampton University, UK, in 1972.
After graduation, he joined the King Abdul Aziz University in Jeddah,

size in JPEG or TIFF.


Saudi Arabia and later moved to the University of Jordan, Amman where
he became Professor of Geology in 1985. His research is concentrated on
the Upper Cretaceous phosphorites of Jordan and the overlying oil shales,
their mineralogy, geochemistry and source rocks. He has written more than
100 publications on various aspects of the geology of Jordan and writes
in Arabic on the geology of Jordan, Palestine, and the Dead Sea. More
recently he shifted interest to the paleoclimate of Jordan and adjacent areas.
He is a member of the SEPM, IAS and the Mineralogical Society and has
served on the scientific board of the IGCP from 1989 to 1995 and IUGS/UNESCO.
aabed@ju.edu.jo

Reinhard F. Sachsenhofer is head of Petroleum Geology at the University


of Leoben, Austria. He has a MSc in Economic Geology and a PhD in
Geology, (where?). Previously, he was Humboldt-Fellow at the Institute of
Petroleum and Organic Geochemistry Research Center, Jülich, Germany
and visiting professor at the Donetsk National Technical University,
Ukraine. His main research interests are in basin analysis, hydrocarbon
systems, source rocks, coal, and organic petrology. He has won awards
from the Austrian Geological Society, the Regional Government of Styria,
Austria, and the AAPG. He is on the editorial board of several journals
including the International Journal of Coal Geology.
reinhard.sachsenhofer@mu-leoben.

Hani Abdul Khair holds a BSc degree in Earth and Environmental


sciences, MSc degree in Sedimentary Geology, and PhD degree in
Petroleum Geology at the Uniniversity of Jordan, Amman Jordan. Over
12 years experience in petroleum exploration and in research assistance.
Specialized in sequence stratigraphy, seismic interpretation, petrophysics,
geohistory analysis, and petroleum evaluation of sedimentary basins.
Membership in the Society of Economic Paleontologists and Mineralogists
(SEPM), the American Association of Petroleum Geologists (AAPG), and
the International Association of Sedimentologists (IAS).
haaabulkhair@yahoo.com

Manuscript received June 24, 2008


Revised September 22, 2008
Accepted October 13, 2008
Press version proofread by authors mm/dd/yy

140

You might also like