You are on page 1of 22

PUBLICATIONS

Tectonics
RESEARCH ARTICLE Structural geology of the Rub’ Al-Khali Basin, Saudi Arabia
10.1002/2016TC004212
S. A. Stewart1
Key Points: 1
Eastern Area Exploration Department, Saudi Aramco, Dhahran, Saudi Arabia
• Cretaceous and Tertiary major
structures in Rub’ Al-Khali are
reactivations of Precambrian
basement domain boundaries Abstract The Rub’ Al-Khali basin lies below a Quaternary sand sea, and the structural evolution from the
• Evaporites present at several Late Precambrian to Neogene is known only from reflection seismic, gravity, and magnetic data, and wells.
stratigraphic levels have little Gravity and magnetic data show north-south and northwest-southeast trends, matching mapped
influence on structural style
Precambrian faults. The deepest structures imaged on reflection seismic data are undrilled Precambrian rifts
filled with layered strata at depths up to 13 km. The distribution of Ediacaran-Cambrian Ara/Hormuz mobile
salt is restricted to an embayment in the eastern Rub’ Al-Khali. The Precambrian rifts show local inversion
Correspondence to:
and were peneplained at base Phanerozoic. A broad crustal-scale fold (Qatar Arch) developed in the
S. A. Stewart,
simon.stewart@aramco.com Carboniferous and amplified in the Late Triassic, separating subbasins in the west and east Rub’ Al-Khali. A
phase of kilometer-scale folding occurred in the Late Cretaceous, coeval with thrusting and ophiolite
obduction in eastern Oman. These folds trend predominantly north-south, oblique to the northwesterly
Citation:
Stewart, S. A. (2016), Structural geology shortening direction, and occasionally have steep fault zones close to their axial surfaces. The trend and
of the Rub’ Al-Khali Basin, Saudi location of these folds closely matches the Precambrian lineaments identified in this study, demonstrating
Arabia, Tectonics, 35, 2417–2438,
preferential reactivation of basement structures. Compression along the Zagros suture reactivated these
doi:10.1002/2016TC004212.
folds in the Neogene, this time the result of highly oblique, north-northeast to south-southwest shortening.
Received 19 APR 2016 Cretaceous-Tertiary fold style is interpreted as transpression with minor strain partitioning. Permian, Jurassic,
Accepted 21 SEP 2016 and Eocene evaporite horizons played no role in the structural evolution of the basin, but the Eocene
Accepted article online 25 SEP 2016
evaporites caused widespread kilometer-scale dissolution collapse structures in the basin center.
Published online 19 OCT 2016

1. Introduction
The Rub’ Al-Khali basin occupies an area of over 300,000 km2 in southeast Saudi Arabia (Figure 1). The basin is
covered by the largest contiguous sand sea in the world with dunes 150 m or more in height [Holm, 1960;
Edgell, 2006]. Due to lack of bedrock outcrop (apart from escarpments on the west margin of the basin)
the tectonostratigraphic evolution is known only from geophysical data and wells. Relatively few publications
deal with the structural evolution of this basin, other than a small number of regional syntheses [Konert et al.,
2001; Ziegler, 2001; Cantrell et al., 2014] and case studies [e.g., Dyer and Husseini, 1991]. A relatively rich set of
publications address structural evolution in the surrounding regions, from the Arabian Shield and Yemen
[e.g., Whitehouse et al., 2001; As-Saruri et al., 2010; Stern and Johnson, 2010], through Oman, UAE, and Qatar
[e.g., Boote et al., 1990; Filbrandt et al., 2006; Zampetti et al., 2014a] to the Zagros collision zone [e.g.,
Mouthereau et al., 2012]. This paper aims to complement these publications by presenting a structural history
of the Rub’ Al-Khali basin from Late Precambrian to recent, integrating structural styles and trends observed
on reflection seismic and potential field data to identify the mechanostratigraphy of the basin, its basement
and its structural evolution in relation to the regional geology.
In spite of the remote and challenging nature of the terrain, a considerable amount of subsurface data have
been acquired in the Rub’ Al-Khali in the pursuit of hydrocarbons. Few analyses of the petroleum systems of
the basin have been published to date, [e.g., Craig et al., 2010; Stewart et al., 2016]. All drilled wells, reflection
seismic, and potential field data in the Rub’ Al-Khali were available for this study. Although details of this
database are not public, materials included some 150 exploration and delineation wells drilled to depths
of between 1 and 6 km, over 50,000 km2 of 3-D seismic, 100,000 line kilometers of 2-D seismic, and basin-wide
airborne gravity and magnetic data.

2. Regional Overview of Structural Evolution


A regional structural section through the Rub’ Al-Khali basin (Figure 2) shows a broad synform containing a
©2016. American Geophysical Union. relatively complete Phanerozoic succession resting unconformably on Precambrian basins and basement.
All Rights Reserved. Published estimates of crustal thickness are based on a limited set of earthquake seismic receiver stations

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2417


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

Figure 1. Regional geology of Saudi Arabia and environs, grid pattern shows the Rub’ Al-Khali basin. Black box is area of
maps presented in this paper, e.g., Figure 3. Map is simplified from unpublished Saudi Aramco mapping integrated with
Haghipour et al. [2009].

that give Moho depth estimates of about 40 km [Rodgers et al., 1999; Al-Damegh et al., 2005], although there
were no stations in the Rub’ Al-Khali itself, and the Moho is not imaged on the reflection seismic data
available to the present study. The basin has a thick lithospheric root relative to the rest of Arabia and
North Africa, according to interpretation of Rayleigh wave tomography [McKenzie et al., 2015].
The Precambrian of the Rub’ Al-Khali is represented by a number of local rifts and is otherwise undifferen-
tiated on reflection seismic data. The base Cambrian unconformity is by far the most significant break in
the stratigraphy of the basin as imaged on reflection seismic data, and the Precambrian and Phanerozoic
chapters in the evolution of the Rub’ Al-Khali basin separated by it are treated in turn here. There is no direct
evidence for the age of this unconformity in the Rub’ Al-Khali itself, but detailed work in the Arabian Shield
indicates the uplift and erosion is late Ediacaran [Al-Husseini, 2011], and the unconformity is termed

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2418


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

Figure 2. (a) Regional seismic section and (b) interpretation through the Rub’ Al-Khali basin, from Stewart et al. [2016].
Stratigraphy is constrained by 12 wells in the line of section in Saudi Arabia drilled to depths ranging from Jurassic to
Lower Paleozoic. Interpreted Omani tectonostratigraphy after Terken et al. [2001]. Vertical scales in depth, datum is mean
sea level. VE is approximate vertical exaggeration. Dotted red line is the surface mapped in Figure 3c. (c) Same section
without vertical exaggeration, including Earth curvature. Base of crust is shown at 40 km based on interpretation of
earthquake seismic data by Al-Damegh et al. [2005].

“Angudan” [Loosveld et al., 1996]. Regional affinity of this Late Precambrian event may be with the
“Pan-African” orogeny, a general term for oceanic closures in the assembly of Gondwana [Nehlig et al.,
2002; Kröner and Stern, 2004], or “Cadomian” orogeny, a more localized term for Late Precambrian collisions
on the north margin of Gondwana [Linnemann et al., 2008]. In any case, the position of the Arabian plate in
Gondwana during the Cambrian places it within the East African Orogen [Collins and Pisarevsky, 2005; Gray
et al., 2008; Fritz et al., 2013].

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2419


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

The Phanerozoic evolution of the Arabian plate was characterized by protracted periods of tectonic quies-
cence on the gradually subsiding, northern margin of Gondwana [Veevers, 2012]. Accommodation space
for sediment accumulation reflected the interplay between crustal subsidence and eustatic variation [Haq
and Al-Qahtani, 2005]. The mechanism of Phanerozoic crustal subsidence in Arabia is unclear and may be
some combination of passive margin subsidence (related to the north margin of Gondwana), dynamic
topography, and a response to the Late Precambrian events that formed the Rub’ Al-Khali basement
[Burgess et al., 1997; Flament et al., 2013; Holt et al., 2015]. In map view, the Rub’ Al-Khali basin is located in
an embayment where the Arabian margin swung from northwest-southeast to the northeast-southwest
trend that now forms the Musandam Peninsula (Figure 1).
Subsidence was punctuated by three prominent Phanerozoic tectonic events. First was a Carboniferous
regional unconformity associated with crustal-scale arch and basin structures that locally preserved
Devonian section, subcropping the Permian Unayzah and Khuff Formations [Faqira et al., 2009]. This event
is coeval with the Hercynian orogeny, which is recognized to be progressively more intense westward along
the north margin of Africa [Haddoum et al., 2001; Badalini et al., 2002; Coward and Ries, 2003; Bumby and
Guiraud, 2005]. The second prominent Phanerozoic event in the Rub’ Al-Khali is characterized by lateral
variations in Late Triassic isopachs and seismically imaged onlap of tilted Permo-Triassic sequences
[Stewart et al., 2016]. The structural style associated with this event in the Rub’ Al-Khali appears to be
restricted to basin and arch reactivation, though the prevailing strains in northern Gondwana in the mid to
Late Triassic were extensional, associated with Neotethys rifting [Bumby and Guiraud, 2005; Golonka, 2007].
Finally, from the mid-Cretaceous onward, pulses of fold growth affected the Rub’ Al-Khali basin, becoming
more pronounced from west to east. These compressional episodes record plate boundary effects in
Neotethys and the Arabian Sea, and ultimately final collision across the Zagros suture [Searle et al., 2004;
Filbrandt et al., 2006; Gaina et al., 2015].

3. Precambrian Structure of the Rub’ Al-Khali


3.1. Precambrian Trends Interpreted From Seismic Reflection and Potential Field Data
West of the Rub’ Al-Khali, where the recent sand sea thins out and the Phanerozoic succession is eroded,
Precambrian basement is exposed and termed “Arabian Shield” (Figure 1). The Arabian Shield is composed
of Neoproterozoic oceanic island arcs and pre-Neoproterozoic continental crust slivers, accreted during the
closure of the Mozambique Ocean to become part of the East African Orogen [Nehlig et al., 2002; Stern and
Johnson, 2010; Johnson et al., 2011; Cox et al., 2012; Fritz et al., 2013]. These accreted Precambrian terranes
are separated from one another by shear zones with local ophiolites [Dilek and Ahmed, 2003] and broadly
trend northeast-southwest in the western portion of the Arabian Shield, and north-south to northwest-
southeast in the eastern portion [Nehlig et al., 2002]. The northwest-southeast trends are collectively
termed “Najd Fault System”; field relationships indicate these trends are the youngest Precambrian-
Cambrian structures in the Arabian Shield [Nehlig et al., 2002; Kusky and Matsah, 2003; Johnson, 2003;
Al-Husseini, 2015]. The Najd faults truncate northeast-southwest trends and contain fault-bound outliers
of Jibalah Group sediments that are Late Ediacaran to Cambrian in age [Al-Husseini, 2011, 2014]. The
Najd faults disappear southeastward under Phanerozoic cover, trending toward the Rub’ Al-Khali [Stern
and Johnson, 2010] (Figure 1). The basal Phanerozoic strata overlying the Najd fault trends at the southeast
limit of Arabian Shield exposure are Permian Unayzah and Khuff Formations, earlier Paleozoic strata having
been eroded from this part of the Arabian Shield by Hercynian uplift [Faqira et al., 2009]. The Permian
strata display no fault offsets or isopach variations where they overstep the Najd structures [Powers
et al., 1966; Vaslet et al., 2005].
The eastern portion of the Arabian Shield, consisting very broadly of north-south trending basement terranes
cut by the northwest-southeast trending Najd fault system, is considered to be representative of basement
below the westernmost Phanerozoic cover, principally on the basis of regional magnetic data [Doebrich
et al., 2007; Stern and Johnson, 2010]. The easternmost exposed basement is the Ediacaran Ar Rayn terrane
magmatic arc complex [Doebrich et al., 2007]. Speculations on basement configuration farther east in
the Rub’ Al-Khali include the presence of a continental microplate (on the basis of magnetic signature
[Johnson and Stewart, 1995]) and/or the youngest and principal suture zone between West and East
Gondwana (on the basis of general younging of terranes eastward [Cox et al., 2012]).

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2420


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

Regional magnetic and gravity data covering the Rub’ Al-Khali have been used in this study to detect base-
ment trends below the Phanerozoic basin (Figures 3a and 3b). Both data sets indicate a dominant north-
south fabric with subordinate northwest-southeast trends. In addition, the full database of reflection seismic
data has been interpreted to yield a “top basement” map (Figure 3c). In practice, this map represents the base
of seismically interpretable, layered strata—usually base Phanerozoic, but locally deeper where older layered
strata occur. In local, fault-bound basins, Precambrian packages occur between this surface and the base
Cambrian unconformity. The faults bounding these Precambrian basins are generally truncated by the base
Cambrian unconformity though the fault tips locally cross upward into the Cambro-Ordovician section. These
faults again follow the predominant pattern of north-south trends, with northwest-southeast trends in the
western Rub’ Al-Khali, and northeast-southwest trends in the eastern part of the basin (Figure 3c).
Reflection seismic data quality at these relatively deep levels is variable across the basin, but the quality
and seismic grid spacing is generally good enough to support a high level of confidence in the interpreted
age and trend of these faults. A confidence map based on reflection seismic data quality is shown in
Figure 3d. The resulting map of basement trends (Figure 3d) is a collation of directly observed
Precambrian structures from reflection seismic interpretation and those exposed in the shield in the extreme
west of the area of interest, together with trends interpreted from the gravity and magnetic data (Figures 3a
and 3b). In some cases the faults mapped on reflection seismic are collinear with trends interpreted from the
magnetic and gravity data, lending weight to the interpretation of basement domain boundaries on the
potential field data in areas where seismic data are unable to resolve Precambrian faults.
Considering the trend map (Figure 3d) simply in terms of candidate boundaries between basement structural
domains, there is a pattern of broadly north-south trends with a spacing of approximately 100 km across the
Rub’ Al-Khali; this pattern loses coherence east of 53°E longitude, in the vicinity of the Lekhwair High
(Figures 3c and 3d). The trend pattern alone does not indicate that the eastern Rub’ Al-Khali-Lekhwair area
is the eastern limit of amalgamated terranes, though the location coincides with speculative suture zone
of Cox et al. [2012]. The set of trends produced here adds local detail to the East African orogenic framework
(Figure 4), showing that the north-south trends in the Rub’ Al-Khali are broadly collinear with the dominant
East African orogenic trends. The basement trends are revisited later in this paper in the context of basement-
cover relationships during Cretaceous and Tertiary deformation.
3.2. Precambrian Well Penetrations in the Rub’ Al-Khali
Apart from the Arabian Shield, Precambrian basement has been described in Yemen [Windley et al., 1996;
Whitehouse et al., 2001; Heikal et al., 2014], Oman [Gass et al., 1990; Mercolli et al., 2006; Stern and Johnson,
2010], and UAE [Thomas et al., 2015]. There are some shallow well penetrations of basement just east of
the Arabian Shield (one within the present study area), but these wells simply confirm the depth of basement
that can be projected eastward from the shield and is clearly imaged on seismic in any case in these areas.
Within the Rub’ Al-Khali itself there are two possible deep well penetrations of Precambrian, one on the south
flank of the Lekhwair High, the other some distance west of this, in Shaybah Field, just south of the UAE
border (Figure 3d). The Lekhwair High well penetrated an undated volcanic and siliciclastic sequence—
otherwise unknown in the Phanerozoic stratigraphy of the basin—at a depth of approximately 4.3 km.
These undated sediments are overlain by Permian Unayzah and Khuff formations. The westerly of the two pos-
sible basement penetrations encountered silty metasediment at approximately 6 km depth. Core specimens
from the bottom of this well show greenschist facies metamorphism with bedding dips in the range 60°–70°.
These metamorphics are overlain by an undated siliciclastic sequence that may be Cambro-Ordovician Saq
Formation, and this is in turn overlain by Permian strata. There are additional Precambrian well penetrations
north of the study area in this paper (i.e., outside the area of mapping shown in Figure 3c), but assimilating
these wells into the wider story of basement evolution of the Arabian plate is beyond the scope of this study.
3.3. Precambrian Fault-Bound Basins
Mapping presented here identifies several Precambrian fault-bound basins subcropping the Rub’ Al-Khali
Phanerozoic succession (Figures 2b and 3c). The largest of these, the “Wajid Graben,” underlies a large area
in the western Rub’ Al-Khali and was previously reported by Dyer and Husseini [1991]. The new mapping
(Figure 3c) is based on a significantly tighter reflection seismic grid than that available to Dyer and Husseini
[1991]. Dyer and Husseini [1991] compare the style and areal extent of the Wajid Graben to the Mesozoic
graben of the North Sea, a comparison that underlines the scale and exploration potential of the Wajid

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2421


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

Graben, especially considering there are


no well penetrations. The new mapping
indicates that the Wajid Graben con-
tains up to 5 km of Precambrian layered
strata subcropping the Cambrian over
an area of approximately 35,000 km2.
Present-day depth to top Precambrian
of the Wajid Graben is nowhere shal-
lower than 5 km (Figure 3c).
In some places the Wajid Graben is a
symmetrical structure bound by clearly
imaged fault systems on either side,
elsewhere there is a single bounding
fault (e.g., Figure 5). The throw on these
bounding faults is up to 5 km. There
are also discrete structures visible on
reflection seismic in the basement
below the Wajid Graben layered
sequences (Figure 5). Along the line of
the available 2-D reflection seismic
sections, these fabrics dip at approxi-
mately 10°, parallel to the onlapped gra-
ben dip slope, are regularly spaced with
a vertical separation of approximately
4 km, and could be interpreted as

Figure 3. Basement structural trends in the


Rub’ Al-Khali. (a) Aeromagnetic data,
reduced to pole. (b) Bouguer gravity. UAE
Bouguer gravity from Ali et al. [2014], color
scale not matched to Rub’ Al-Khali data,
ranges from 10 mGal (red) to 80 (blue).
Rub’ Al-Khali potential field data are
proprietary of Saudi Aramco. (c) “Top
Basement” depth structure referenced to
mean sea level (MSL), contour interval
0.5 km. Map is based on interpretation of a
reflection seismic database mostly of 2-D
lines of various vintages with an average
spacing of approximately 5–10 km. Labels on
some prominent major structures are at top
basement level. This map represents the
surface (dotted red line) highlighted in
Figure 2b. Highlighted wells in the eastern
Rub’ Al-Khali are possible basement
penetrations on basis of lithology only, they
occur in the area of low confidence in the
seismic mapping and the map is tied to these
wells. (d) Basement trends, interpreted as
domain boundaries, based on direct
mapping of faults at top basement level or
outcrop (solid lines). Dashed lines are trends
interpreted from lineaments in the potential
field data shown in Figures 3a and 3b. Color
shading indicates areas of relatively high and
low confidence in the reflection seismic map
shown in Figure 3c, in terms of the seismic
data quality.

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2422


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

Figure 4. Gondwana circa 500 Ma after Gray et al. [2008] and Stern and Johnson [2010]. Showing Rub’ Al-Khali lineaments
from Figure 3d in the context of East African Orogen trends and the wider reference frame of Gondwanan orogenic belts
and plate margins.

westerly vergent thrust structures associated with terrane accretion in the East African Orogen (Figure 5).
Assuming these fabrics are parallel to the graben dip slope in 3-D, reactivation of these fabrics as low-angle
extensional faults may explain local asymmetry of the Wajid Graben, which is a structural style akin to the
Norwegian post-Caledonian basins [e.g., Serrane and Seguret, 1987; Fossen, 2010].
In the western part of the Rub’ Al-Khali basin, seismic sequences can be discerned only in the Wajid Graben—
elsewhere, with the exception of locally imaged dipping fabrics mentioned above, the basement is feature-
less on reflection seismic and interpreted as crystalline. Within the Wajid Graben, three sequences can be
separated on the basis of onlap and growth architectures (Figure 5). The oldest sequence, thickest of the
three at up to 4 km, is essentially parallel bedded. This sequence is folded and the thickest part of the over-
lying, second seismic sequence occurs in a syncline. The third sequence is conformable with the overlying
Cambrian section with no direct control on where the base Cambrian should be placed. In this study the base
Cambrian is assumed to be equivalent to the unconformity visible on the rift shoulders (Figure 5) and is inter-
preted within the graben on this basis. The base Cambrian could be interpreted deeper within the graben
than shown in Figure 5, or could be shallower, within the interval that onlaps the rift shoulders (Figure 5).
As interpreted in Figure 5, the reflective package near base Cambrian, which onlaps the rift shoulders, could
represent marginal facies of the Hormuz/Ara salt basins [e.g., Dyer and Husseini, 1991]. Alternatively, these
facies could be correlated with older strata known from the outliers on the Arabian Shield (the Jibalah
Group [Al-Husseini, 2011, 2014]). As interpreted in this study, the bounding faults are active as extensional
structures until late Cambrian to early Ordovician, though this could be driven to some extent by differential
compaction of the basin fill. Modeling of gravity profiles across the Wajid Graben indicates a best fit average
density of 2.66 g/cc for the graben fill, which could represent low-porosity siliciclastics or interbedded
siliciclastics and volcanics.
The other Precambrian basins are much smaller than the Wajid Graben (Figure 3c) and have been identified
in the eastern part of the Rub’ Al-Khali on the basis of fault-bound, dipping, layered sequences subcropping
the Phanerozoic with angular unconformity.

3.4. Precambrian Salt Distribution and Influence on Structure


Late Ediacaran to Early Cambrian salt is well known in the Zagros foreland (Hormuz Series) and Oman (Ara
Group), e.g., Husseini and Husseini [1990], Allen [2007], and Smith [2012]. Various publications depict evaporite

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2423


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

Figure 5. Two-dimensional reflection seismic line showing the Neoproterozoic Wajid Graben, note the Vertical
Exaggeration factor (times 10). Vertical axis is in two-way time ranging from 0 to 5 s. The seismic datum approximates
the land surface elevation. Location is in Figure 6a. Neoproterozoic sequences are separated on basis of seismic architec-
ture but are undrilled and undated. Seismic interpretation from Tertiary to top Ordovician is tied to offset wells, the
Cambro-Ordovician is divided on the basis of relative thicknesses observed elsewhere in Saudi Arabia.

distribution into Saudi Arabia, but they do not distinguish between mobile evaporites (i.e., may become
involved in halokinesis) from immobile evaporite such as anhydrite. Many authors [e.g., Smith, 2012] follow
Konert et al. [2001], reproduced here in Figure 6a, showing a significant extent of evaporites continuing south
from the UAE into the eastern Rub’ Al-Khali and a separate evaporite subbasin in the western Rub’ Al-Khali
which corresponds to the Wajid Graben. The Hormuz subbasin nomenclature used here revises that used

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2424


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

in some previous studies—the area


called “East Rub’ Al-Khali salt basin” in
this study (because it occurs in the east
part of the Rub’ Al-Khali, Figure 6a) was
termed “West Rub’ Al-Khali salt basin”
by Allen [2007].
The extent of mobile salt has been
defined in this study on the basis of seis-
mic structural style, using all available
seismic data. In Oman, UAE, and the
Zagros foreland, Hormuz/Ara salt pro-
foundly influenced structural style, lead-
ing to diapirs [Jackson et al., 1990; Bosak
et al., 1998; Peters et al., 2003; Thomas
et al., 2015; Perotti et al., 2016], miniba-
sins [Heward, 1990; Al-Barwani and
McClay, 2008; Li et al., 2012], and detach-
ments [Alavi, 2007; Mouthereau et al.,
2012]. None of these styles have been
observed in the comprehensive reflec-
tion seismic database now available in
the Rub’ Al-Khali. Reflection seismic
lines across the Wajid Graben (e.g.,
Figure 5) do not show any evidence of
salt pillows or diapirs, in contrast to the
Oman salt basins [e.g., Peters et al.,
2003; Al-Barwani and McClay, 2008]. A
possible exception to this is an isolated
dome in the middle of the Wajid
Graben (not shown here; see Figure 8
of Stewart [2014]). While it is possible
to interpret this feature as a salt dome,
it can also be interpreted as an inversion

Figure 6. Hormuz salt in the Rub’ Al-Khali.


(a) Comparison of salt extent, mapped in
this study from seismic facies, with previously
published evaporite distribution [Konert et al.,
2001]. Salt distribution outside Saudi Arabia
was not reviewed in this study. Structural
contours are from basement map in Figure 3
c. (b) Three-dimensional reflection seismic
line over the East Rub’ Al-Khali salt basin,
location is in Figure 6a. Vertical axis is in
two-way time ranging from 0 to 5.75 s. The
seismic datum approximates the land surface
elevation. (c) Seismic interpretation of
Figure 6b. Pre-Triassic stratigraphy is tied to
distant wells. There are no worthwhile well
ties to the pre-Silurian section of this area.
(d) Enlargement of interpreted Ediacaran-
Cambrian sequence, showing the “corru-
gated” seismic facies that are interpreted as
layered salt. Note also the strong multiples
generated by seismic energy reflecting within
the Mesozoic section.

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2425


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

structure, e.g., Stewart [2014] or even an igneous pluton. The inversion interpretation is more in keeping with
the structural styles observed along trend within the graben, e.g., Figure 5, and in any case this is a singular
structure in an extensive basin otherwise devoid of this style. This general absence of evidence does not
definitively prove the absence of salt, but with the Wajid Graben fill typically containing coherent, layered
seismic reflectors (Figure 5) the only option for substantial thickness of mobile salt would be an interpretation
placing top salt on the deepest visible strong reflectors. However, that interpretation is not preferred because
there then are no reflectors available to place base salt, i.e., top basement. The interpretation presented here
essentially follows Dyer and Husseini [1991] in assigning a portion of the reflective fill in the Wajid Graben as
lateral, marginal carbonates and clastics that are stratigraphically equivalent to Hormuz/Ara salt. Since the
Hormuz/Ara represents the youngest part of the Precambrian megasequence [Allen, 2007] the correlatives
would necessarily be the shallower parts of the Wajid Graben fill.
The smaller Precambrian basins identified in this study in the vicinity of 51° 30′E 21°N show no sign of mobile
salt. On the other hand, the East Rub’ Al-Khali salt basin contains seismic facies that do suggest the presence
of mobile salt. This subbasin contains some 10 km of Phanerozoic stratigraphy, the thickest section in the
Rub’ Al-Khali (Figures 2 and 3c). The subbasin is overlain by an embayment of Holocene sabkha facies
extending south from the Gulf coastline (Sabkha Matti [Alsharhan and Kendall, 2003]), which may be localized
due to differential compaction of this exceptionally thick Phanerozoic section. Three-dimensional reflection
seismic data in the East Rub’ Al-Khali salt basin shows unusual, layer-bound “corrugated” seismic facies
(Figures 6b–6d) that are interpreted as folded intrasalt stringers, a structural style familiar from well-
constrained, layered evaporite sequences in Oman [Li et al., 2012] and the North Sea [Jenyon, 1989; Van
Gent et al., 2011; Strozyk et al., 2012]. While other interpretations of this undrilled facies are possible (e.g., bio-
herms or dunes), a deformed evaporite interpretation is preferred here on the basis that it is admissible
(sensu Elliott [1983]) and occurs in the approximate stratigraphic position expected for the Hormuz Series.
Correlation of the corrugated facies visible in the East Rub’ Al-Khali salt basin (Figure 6d) with the
Hormuz/Ara salt places a lower bound on a relatively undeformed, strongly reflective sequence up to 1 km
thick that is confined to this subbasin (Figure 6d) and also occurs in the Wajid Graben. This package could
represent interbedded immobile evaporites and carbonates in the upper part of the Hormuz/Ara, e.g.,
Schröder et al. [2003], or they could represent younger intervals, for example, early to mid Cambrian Siq
Formation siliciclastics and Burj Formation carbonates or their lateral equivalents in Oman, the Mahatta
Humaid Group [e.g., Al-Husseini, 2014].
In addition to the corrugated seismic facies, it is possible to interpret subtle detachments on the terraced east
margin of the East Rub’ Al-Khali salt basin, where individual subsalt faults have insufficient throw to offset the
salt and top salt reflectors, in contrast to the west flank of this subbasin where there is a single fault carrying
enough throw to offset the evaporite sequence (Figures 6b–6d). This “soft-linked” style is characteristic of
basement faults interacting with detachment horizons that are thick relative to the amount of fault throw
at base salt level [e.g., Harvey and Stewart, 1998]. Toward the border with UAE, the suprasalt sequence above
the east bounding fault system of the East Rub’ Al-Khali salt basin contains at least one small, low-amplitude
dome that may indicate a salt pillow associated with salt flow above the basement fault terraces.
The distribution of mobile salt mapped in the east Rub’ Al-Khali on this basis (Figure 6a) is less extensive
than that inferred by Konert et al. [2001] and essentially agrees with Dyer and Husseini [1991] and
Sharland et al. [2001]. Overall, the impact of this salt on structural style is very modest and restricted to
the East Rub’ Al-Khali salt basin.

4. Phanerozoic Structure of the Rub’ Al-Khali


4.1. Hercynian Structure and Unconformity
There is a major stratigraphic break on the Arabian plate below the Permian, subcropped by isolated
Carboniferous outliers and a northeast-southwest trending basin and arch structure of Cambrian to
Devonian strata (Figure 7a). Thickness of preserved Cambrian-Devonian sections in the Faydah-Jafurah and
East Rub’ Al-Khali basin trends indicate that the elevation of arches relative to basins was 2–4 km
[Faqira et al., 2009]. The structural wavelength is hundreds of kilometers, and the arch limbs dip at less
than 1° (Figures 7b and 7c). The timing of this event in Saudi Arabia is Serpukhovian to Moscovian [Cantrell
et al., 2014], so is contemporaneous with the Hercynian/Appalachian Orogeny [Nance et al., 2010], and could

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2426


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

be viewed as a distant, crustal-scale


response to it [e.g., Stephenson and
Cloetingh, 1991]. Hercynian effects
become progressively less pronounced
eastward along the north African mar-
gin away from the plate collision zone
between Gondwana and Laurentia
[Lecorche, 1983; Coward and Ries, 2003;
Guiraud et al., 2005]. In spite of the
distance from the plate collision and
orogenic belt, this stratigraphic break
and associated basin and arch deve-
lopment tends to be referred to as
“Hercynian” on the Arabian plate e.g.,
Faqira et al. [2009].
The “Hercynian subcrop” map can be
compared with the basement trends
interpreted in this study (Figure 7a).
There is no first-order relationship
between the two, the overall northeast-
southwest trend of Hercynian basin
and arch structures appears to passively
deform the basement lineaments and
domains between them. There are local,
relatively minor examples of basement
trends affecting the Hercynian subcrop
pattern, for example, on the north
margin of the Faydah-Jafurah Basin
(Figure 7a). Why basement faults such
as those illustrated in Figure 5 did not
reactivate more widely during the
Hercynian is unclear; the prevailing
stress field and strain rate were evi-
dently insufficient to reactivate these
trends in preference to more distributed
deformation. Also unknown is what
caused the Hercynian basin and arches
to trend northeast-southwest, see-
mingly crossing basement lineaments
and therefore passing from one base-
ment domain to another. Nonetheless,
the southwest-northeast trends mani-
Figure 7. Hercynian subcrop and structure. (a) Subcrop to Late fested at this time—Qatar Arch in
Carboniferous/Early Permian modified from Figure 10 of Konert et al. particular—subsequently proved to be
[2001] and Figure 2 of Faqira et al. [2009]. Gray dashed trends are
long-lasting features with multiple reac-
basement lineaments from Figure 3d. White dots are wells that penetrate
pre-Carboniferous strata. (b) Composite 2-D reflection seismic line tivation episodes in the Rub’ Al-Khali,
crossing west margin of Qatar Arch, location is in Figure 7a. Line is farther northeast in Qatar, and what is
flattened on an intra-Permian reflector (base Khuff Formation), and the now the Zagros foreland [Konert et al.,
shallowest 2–2.5 km of section has been clipped from the top of the 2001; Alavi, 2007; Perotti et al., 2011].
line to allow the pre-Carboniferous structure to be emphasized.
(c) Interpretation of Figure 7b constrained by wells to top Ordovician in 4.2. Triassic Structure
the line of section and tied from elsewhere. Top and base Cambrian is not and Unconformity
directly tied but has been inferred from comparison with distant well
penetrations. A Late Triassic (Norian) angular un-
conformity in the Rub’ Al-Khali is

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2427


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

documented by Stewart et al. [2016].


This unconformity is characterized by
local absence of the Minjur Formation
and a low-angle, onlap relationship
between the Minjur and underlying Jilh
Formation (Figure 8). The relatively iso-
pachous nature of overlying Jurassic
carbonates (Figures 8b and 8c) demon-
strates that there is a significant ele-
ment of onlap of tectonic architecture
rather than solely downlap of prograd-
ing Minjur fluviodeltaic systems over a
flat surface.
Minjur Formation distribution indicates
that some but not all of the Hercynian
basin and arch structures were
reactivated in this Late Triassic
event (Figure 8a). The Qatar and
Hadhramaut-Oman arches were active,
defining the intervening East Rub’
Al-Khali Basin, but the Al-Batin Arch
remained a low, continuous with the
Faydah-Jafurah Basin at this time
(compare Figures 7a and 8a). The
Minjur Formation onlaps the Qatar
Arch (Figures 8b and 8c). The relative
amplitude of arch and basin growth in
this event was subtle—in the order of
a few hundred meters, a tenth of the
vertical elevation developed in the
Hercynian event.
Late Triassic compression is recorded by
the Indosinian Orogeny in southeast
Asia [Carter and Clift, 2008; Metcalfe,
2013], but this area was separated
from North Gondwana by Neotethys at
the time [Stampfli and Borel, 2002] so
the events are presumably unrelated.
Permo-Triassic compressional events
are recorded in the Gondwanides from
South Africa to Patagonia [Pankhurst
Figure 8. Triassic isopach and structure, from Stewart et al. [2016]. et al., 2006; Tankard et al., 2009].
(a) Isopach of Minjur Formation (Norian-Pleinsbachian), contours in Local compressional structures within
meters. (b) Reflection seismic line flattened on top Marrat Formation,
Gondwana [Daly et al., 1992] and slight
showing onlap relationship of the Minjur Formation on to the Mid–Late
Triassic Jilh Formation on the east flank of the Qatar arch. Location is in uplifts, arching, and strike-slip faulting
Figure 8a. Seismic data flattened on top Marrat Formation (Toarcian) to on the North African margin [Guiraud
remove later structural distortion. Note the high vertical exaggeration of et al., 2005] have been correlated with
the display required to capture this relationship. (c) Interpreted version of these major events on the southern
Figure 8b.
continental margin but are Early
Triassic rather than the Late Triassic
deformation observed in the Rub’ Al-Khali. Late Triassic compression is described from Patagonia [Gregori
et al., 2016], which was very distant but demonstrates that local, compressional events occurred in
Gondwana about the time of basin and arch development on the Arabian plate. It is concluded that the

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2428


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

Late Triassic basin and arch develop-


ment in the Rub’ Al-Khali is a subtle,
local, crustal-scale event related to
some variance in plate motions but
unrelated to a major Gondwanan event.
See Cloetingh and Burov [2011] and
Wood [2015] for further discussion and
possible mechanisms.
4.3. Cretaceous Fold Timing
and Distribution
Late Cretaceous and Tertiary folds with
wavelengths of tens of kilometers and
amplitudes of about a kilometer are pro-
minent in the center and east Rub’ Al-
Khali. Some of these structures now
host giant hydrocarbon accumulations
[Mann et al., 2003; Alsharhan, 2014]. In
the eastern Rub’ Al-Khali fold growth
can be separated into two episodes
across a prominent angular uncon-
formity at base Tertiary. The Late
Cretaceous fold component is recorded
by an isopach map of Cretaceous
formations below the unconformity
(Figure 9a). The folds occur east of 50°E
and have a clear north-south trend,
switching to northeast-southwest trend
in the eastern Rub’ Al-Khali.

Figure 9. (a) Late Cretaceous structure—


isopach of Aruma, Wasia, and Shu’aiba
Formations (Maastrichtian-Aptian), contours
in meters. Red trends highlight Late
Cretaceous folds (isopach thins). Dashed
gray lines are basement lineaments
interpreted in this study (from Figure 3d).
Large transparent arrows are end-
Cretaceous shortening directions from
Johnson et al. [2005]. (b) Reflection seismic
line from a 3-D survey with a representative
example of Late Cretaceous and Tertiary
folding in the eastern Rub’ Al-Khali, location
is in Figure 9a. Vertical axis is in two-way time
ranging from 0 to 3.15 s. The seismic datum
approximates the land surface elevation.
(c) Annotated version of Figure 9b.
Figures 9b and 9c are slightly modified from
Stewart et al. [2016]. (d) Depth to base
Tertiary, contours in meters referenced to
mean sea level. Contour levels are negative
in the western Rub’ Al-Khali where the base
Tertiary is above mean sea level. Interpreted
fold and basement trend colors are as per
Figure 9a. Long arrows are recent plate
motions from Mouthereau et al. [2012].
Labeled lines (a) to (g) are locations of the
seismic examples shown in Figure 10.

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2429


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

Interpretation of reflection seismic data immediately below the Paleocene Umm Er Radhuma Formation is
hindered by strong seismic multiples that mask relatively steeply dipping Cretaceous reflectors, making the
angular unconformity occasionally appear stratigraphically deeper than it actually is (Figures 9b and 9c).
Where the Cretaceous primary seismic reflectors are weak, it becomes difficult to unambiguously interpret fold
timing and in places, older, intra-Cretaceous pulses of fold growth can be picked but this interpretation is
uncertain. On the basis of outcrop mapping northwest of the Rub’ Al-Khali, unconformities with local slight
angularity are reported by Powers et al. [1966] at base Aruma Formation (early Campanian) and base Wasia
Formation (early Cenomanian). East of the Rub’ Al-Khali, deformation in the foreland to the Oman thrust belt
is first recorded in the Turonian, reaching a climax relating to allochthon emplacement in the Santonian-
Campanian [Boote et al., 1990; Warburton et al., 1990; Cooper et al., 2014]. A second pulse of shortening occurred
in eastern Oman between the late Maastrichtian and Eocene (Batain Group and Masirah ophiolite [Marquer
et al., 1995; Schreurs and Immenhauser, 1999; Fournier et al., 2006]). Obduction gave way to burial and exten-
sional relaxation in Oman by the Paleocene [Fournier et al., 2006]. Compression is recorded northeast of the
Rub’ Al-Khali in the Zagros from Turonian onward [Alavi, 2004]. Compression is also recorded in the
Palmyrides and northern Egypt at end-Cretaceous [Guiraud and Bosworth, 1999]. These events in the surround-
ing area mean that documented correlatives can be found for possible structural events any time from the
Turonian to end-Cretaceous, but seismic reflection data and well control indicate that the most clearly defined
folding episode in the Rub’ Al-Khali occurred at or near end-Cretaceous (Figures 9b and 9c).

4.4. Tertiary Fold Timing and Distribution


The Paleocene-Eocene section (Umm Er Radhuma, Rus, and Dammam Formations) is folded, but nearly iso-
pachous, over the end-Cretaceous folds (Figures 9b and 9c). The near-constant thickness of the folded
Paleocene-Eocene section indicates that this folding represents a separate event reactivating the end-
Cretaceous folds rather than a continuous period of fold growth incorporating a period of erosion. Timing
of this event is difficult to constrain in the Rub’ Al-Khali due to the lack of reflection seismic and well data
in the post-Eocene section (the data are focused on deeper strata). Folds north and northeast of the Rub’
Al-Khali, such as the Dammam Dome and salt-cored highs in the Gulf, indicate that fold amplification
occurred from Neogene to Quaternary in an ongoing response to compression in the Zagros [Weijermars,
1999; Perotti et al., 2016]. Slight variations in the Paleocene-Eocene isopach are limited to the lower parts
of the Paleocene Umm Er Radhuma Formation which can be attributed to the last increments of the end-
Cretaceous folding or perhaps differential compaction of the Cretaceous section. Fold amplitudes of
Paleocene-Eocene markers are 40–50% of the total fold amplitude measured on Mesozoic markers, so just
under half of fold growth in the Rub’ Al-Khali occurred during Tertiary reactivation. As expected from the
reflection seismic evidence of fold reactivation, the map view pattern of Tertiary folding picked out by the
Tertiary isopach is similar to the mapped end-Cretaceous folds (Figures 9a and 9d).

4.5. Cretaceous-Tertiary Fold Origin and Style


Comparison of mapped end-Cretaceous Rub’ Al-Khali folds with the basement lineaments interpreted earlier
in this study shows a strong relation in trend and spacing, and in some cases the Cretaceous folds are
collinear with basement trends (Figure 9a). This relationship indicates that the end-Cretaceous folds are
controlled by Precambrian lineaments, a similar conclusion to that reached by As-Saruri et al. [2010] in
Yemen, Boote et al. [1990] in Oman, Zampetti et al. [2014a] in Qatar, and Kent [2010] in Iraq. The spatial and
temporal association of end-Cretaceous folds in the eastern Rub’ Al-Khali with the Cretaceous structural
development of Oman suggests a shared plate tectonic cause. In detail, the timing of end-Cretaceous folding
in the Rub’ Al-Khali correlates better with Batain Group thrusting and emplacement of the Masirah Ophiolite
in eastern Oman than the earlier Semail ophiolite in the main Oman thrust belt [e.g., Schreurs and
Immenhauser, 1999]. The wider tectonic context of ophiolite obduction in Oman and compressional tectonics
in general at this time is an outcome of relative motions of the Arabian and Indian plates, modeled in detail by
Johnson et al. [2005] as producing NW–SE compression in the UAE at end-Cretaceous (Figure 9a). This rather
surprising result, given the southwesterly directed shortening in the Oman thrust belt at the time, is corrobo-
rated in a synthesis of subsurface mapping in the Oman foreland [Filbrandt et al., 2006] and outcrop mapping
in the Batain Group and Masirah Ophiolite in eastern Oman [Schreurs and Immenhauser, 1999]. This shorten-
ing direction requires that end-Cretaceous fold growth in the Rub’ Al-Khali was associated with sinistral trans-
pression (Figure 9a).

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2430


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

The nearly one-to-one relationship between Cretaceous folding and Tertiary reactivation leads to the same
conclusion regarding strong control of basement lineaments on Tertiary fold growth (Figures 9a and 9d).
Although the fold trends are identical between the end-Cretaceous and Tertiary phases, regional data indi-
cate that the Tertiary compression direction was markedly different. Plate motion syntheses indicate recent
northern movement of Arabia [DeMets et al., 1990; Vernant et al., 2004; Mouthereau et al., 2012] (Figure 9d),
and present-day maximum horizontal stress is oriented northeast-southwest in the Zagros and Oman
[Heidbach et al., 2010]. Reviews of tectonics in the Zagros and Indian Ocean indicate that these recent
shortening directions have been broadly consistent since the onset of final collision in the Eocene [Allen
et al., 2004; Gaina et al., 2015]. The acute angle between Tertiary shortening direction and the mapped fold
trends in the Rub’ Al-Khali underlines how important the Precambrian basement trends are in determining
late deformation episodes. The tight angular relationship also makes it difficult to infer the sense of transpres-
sion during this phase of deformation.
To determine the fold mechanism of these Cretaceous and Tertiary structures, several factors have to be con-
sidered: strong evidence for basement control (discussed above), general lack of change in marker elevations
from one side of the fold to another (Figure 10), and the relatively low amplitude of these structures when
viewed without vertical exaggeration (Figure 10). As discussed earlier, Hormuz salt is largely absent from
the Rub’ Al-Khali, so these folds are not salt-cored detachment folds. Indeed, there is no evidence in terms
of reflection seismic architecture for redistribution of any type of ductile material within these folds.
Fault-propagation folding can be considered unlikely because there is no evidence on reflection seismic data
for low-angle reverse faults associated with these structures within the seismically resolvable section
(Phanerozoic). The folds can be reproduced by forward models of listric fault hanging walls where the faults
have constant curvature (i.e., circular arc), detaching at approximately 20 km depth [Cantrell et al., 2014]. This
midcrustal detachment option requires that the modeled detachment be developed at the same depth in
potentially unrelated Precambrian terranes in the Rub’ Al-Khali basement. This option is also not fully
balanced because “downdip,” out-of-section accommodation of displacement on the putative midcrustal
detachment is undocumented [Cantrell et al., 2014]. The association of these folds with Precambrian linea-
ments points to two alternative fold mechanisms. One possibility is that they are inversion structures above
Precambrian rifts [e.g., Gillcrist et al., 1987]. The folds are more widespread than the mapped Precambrian
graben, so the folds cannot all be said to be inversion structures. A closer look at the geometry of these folds,
where 3-D reflection seismic data are available in the eastern Rub’ Al-Khali, shows that they anastomose
along trend (Figure 11a) and are internally partitioned with second-order folds and occasional faults
(Figure 11b). The steep, en echelon nature of these faults within the folds suggests that the faults are
strike-slip structures (Figure 11b). Taking these observations together with the basin-scale relationships
between the folds and basement lineaments, and the prevailing shortening directions during fold growth,
it is possible to conclude that these structures are transpressional folds.
Many strain distributions can occur within transpression zones [Dewey et al., 1998; Jones et al., 2004]. The most
relevant model to the Rub’ Al-Khali structures appears to be transpression with no boundary slip, minor
vertical expulsion and partitioning into secondary folds and faults (Figure 11c) [cf. Robin and Cruden, 1994;
Dutton, 1997]. Given the association of folds and basement lineaments in the Rub’ Al-Khali, the
Precambrian basement is assumed to be the controlling zone in the crust, represented in Figure 11c.
Relatively weak domains in the Precambrian basement expelled vertically when compressed between rela-
tively strong domains would passively fold the cover megasequence. Detailed imaging of the fault zones that
occasionally occur within the folds reveals occasional en echelon arrangement in plan view indicating dextral
sense of motion on the faults (Figures 11b and 11c), and by extrapolation, of transpression. Evidence for dex-
tral transpression reported here is difficult to reconcile with Late Cretaceous sinistral transpression in the Rub’
Al-Khali (Figure 9a), although it could be compatible with later, Tertiary, compression (Figure 9d).

5. Permian, Jurassic, and Tertiary Evaporite Influence on Structure


In addition to the Hormuz salt, several Phanerozoic evaporite layers are tectonically active in the Zagros
thrust system [Sherkati et al., 2005; Sepehr et al., 2006], but Phanerozoic evaporites, although present at sev-
eral stratigraphic levels, did not impact the tectonic evolution of the Rub’ Al-Khali. The Middle to Late Permian
Khuff Formation is 600–1000 m thick in the Rub’ Al-Khali, at present-day depths of approximately 5 km, and

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2431


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

contains 5–10% anhydrite in beds


individually up to 15 m thick
[Alsharhan, 1993, 2006]. The
Khuff anhydrites thin westward
and are absent at outcrop west
of the Rub’ Al-Khali [Vaslet et al.,
2005]. The Jurassic Hith and Arab
formations contain a significant
amount of evaporites in the Rub’
Al-Khali [Alsharhan and Kendall,
1994; Hughes and Naji, 2008].
These formations range up to
400 m in thickness, at depths of
approximately 3 km over much
of the basin, and contain up to
80–90% anhydrite. Within the
Hith and Arab anhydrites are thin
halite units (less than 3 m) that
constitute only a few percent
of the Jurassic evaporites. The
Jurassic evaporites thin but are
still present as far as outcrop west
of the Rub’ Al-Khali, where they
form a widespread zone of
dissolution collapse structures
[Powers et al., 1966; Sharief et al.,
1991]. Although the Permian
and Jurassic evaporite units are

Figure 10. Collection of reflection


seismic sections across the main fold
structures in the Rub’ Al-Khali. (a–g)
The seismic sections are shown with a
certain amount of vertical exaggera-
tion (VE) to clarify the fold structures.
All lines are oriented roughly west-
east, west on the left, line locations
shown in Figure 9. Vertical scale of all
lines 0–5.75 s two way time (TWT).
Colors and symbols are keyed in
Figures 10a and 10b. “Top basement”
corresponds to the surface mapped in
Figure 3c. Each seismic section is
shown alongside an equivalent section
with no Vertical Exaggeration (i.e.,
VE = 1). These depth sections include
nominal Moho depth of 40 km [Rodgers
et al., 1999; Al-Damegh et al., 2005] to
give a sense of scale. Also included in
the lower crustal parts of these sections
are the positions of basement linea-
ments (dotted white lines) interpreted
in this study (from Figure 3d). These
lineaments are represented as vertical
structures but there is no control on
their dip other than their having to be
steep enough to produce potential
field discontinuities.

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2432


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

Figure 11. Three-dimensional seismic imaging of Cretaceous-Tertiary folds, east Rub’ Al-Khali. (a) Large-scale oblique view encompassing all currently available 3-D
seismic data, colored surface is near top Hadriya Formation (Oxfordian). Color range on surface spans 1.4 (red) to 2.7 s (purple) TWT. The Hadriya surface shows the
anastomosing nature of the post-Jurassic folds, also local partitioning of strain into second-order folds and faults. (b) Close up of 3-D reflection seismic cube showing
internal structure of a fold in Figure 11a. Green arrow points north. Location indicated in Figure 11a. A steep fault zone is visible in the fold core and is en echelon in
map view. Inset block diagram is a kinematic interpretation. (c) Block diagram model of transpression zone with no boundary slip, vertical expulsion, and local strain
partition (fault). Modified from Figure 1b of Robin and Cruden [1994].

widespread, they have no discernible influence on structural style due primarily to low amplitude of the
Cretaceous and Tertiary folds (Figure 10), insufficient to trigger fold crest collapse mechanisms [cf. Stewart
and Coward, 1995, Figure 20b]. Second, the halite layers themselves were not thick enough to promote
pillowing or diapirism.
The other significant evaporite horizon in the Rub’ Al-Khali is the Eocene Rus Formation. This is an anhydrite
unit up to 250 m thick in the subsurface. The Rus Formation thins through dissolution where regional tilt or
folding has brought it to within a few hundred meters of the surface (Figure 12a). The upper parts of the

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2433


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

underlying Umm Er Radhuma carbo-


nates also thin in these areas
(Figures 12b and 12c)—this formation
is also known to be prone to dissolution
[Amin and Bankher, 1997; Zampetti et al.,
2014b]. At surface exposure outside the
Rub’ Al-Khali, less than 50 m of residual
material described as chalky limestone
and marl constitute the outcrop-based
Rus Formation type section [Powers
et al., 1966]. As with the Permian and
Jurassic evaporites, the folding that
affects the Eocene evaporites has been
insufficient to trigger secondary struc-
tures such as localized detachment sys-
tems. Dissolution of Rus Formation in
the central Rub’ Al-Khali has led to wide-
spread collapse structures up to 2 km in
diameter and 200 m in depth. These fea-
tures are imaged on 2-D reflection seis-
mic data in the shallow subsurface
(Figure 12d) and are similar in scale to
“giant” buried dissolution collapse struc-
tures found elsewhere [e.g., Lofi et al.,
2012]. The collapses are structurally
significant in their own right and are
geophysically important because their
chaotic fill affects reflection seismic ray-
paths leading to relatively poor imaging
of deeper strata and structures.

6. Conclusions
An extensive database of reflection seis-
mic, wells and potential field data has

Figure 12. Distribution of Eocene evaporites


(Rus Formation) and dissolution collapse
structures in these evaporites. (a) Rus
Formation distribution, fairly uniform in
thickness (100–200 m) in eastern Rub’
Al-Khali, thinning westward to residual thick-
ness (<50 m) and truncating at the base of
Quaternary sediment. Pattern of white circles
represents area of kilometer-scale dissolution
collapse structures; the circles are a pattern
fill and do not represent individual structures.
(b) Regional 2-D seismic line flattened on top
Aruma Formation, illustrating the Rus
Formation thinning westward, note the high
vertical exaggeration. (c) Annotated version
of Figure 12b, colored bands correspond to
the mapped zones in Figure 12a. (d) Three-
dimensional view of two intersecting 2-D
reflection seismic lines giving a closer look at
dissolution collapse structures (pits), in the
Rus Formation.

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2434


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

been utilized in a review of the structural evolution of the Rub’ Al-Khali basin. The results of this study fill gaps
in the regional geology from Precambrian to Holocene and underpin more detailed structural analysis at the
scale of hydrocarbon fields. A Precambrian basement fabric is defined from gravity and magnetic data trends,
combined with direct mapping of reflection seismic data. The trends are consistent in orientation and spa-
cing with Precambrian terranes exposed in the Arabian Shield. Boundaries between these terranes are major
shear zones in the exposed shield, and the trends identified in the Rub’ Al-Khali are assumed to be associated
with similar zones of weakness. Hercynian and Triassic basin and arch development appears to be unrelated
to the basement trends. In the center and east Rub’ Al-Khali, major folds developed episodically in the Late
Cretaceous and Tertiary; these structures are economically significant because they host giant hydrocarbon
accumulations. These fold trends are closely associated in trend and location with basement lineaments. The
fold style is interpreted to be transpressional, driven by reactivation of basement lineaments, with local strain
partitioning into faults. Evaporites have been unimportant in the structural evolution of the Rub’ Al-Khali
basin. Hormuz/Ara evaporites are restricted to a single embayment in the northeast of the basin, and the
more widespread Permian, Jurassic, and Eocene evaporites are primarily composed of anhydrite and do
not form structural detachments, though the Eocene evaporites are locally replaced by kilometer-scale disso-
lution structures.

Acknowledgments References
The author thanks the Saudi Arabian
Ministry of Petroleum and Mineral Alavi, M. (2004), Regional stratigraphy of the Zagros fold-thrust belt of Iran and its proforeland evolution, Am. J. Sci., 304, 1–20.
Resources and the Saudi Arabian Oil Alavi, M. (2007), Structures of the Zagros fold-thrust belt in Iran, Am. J. Sci., 307, 1064–1095.
Company (Saudi Aramco) for granting Al-Barwani, B., and K. R. McClay (2008), Salt tectonics in the Thumrait area, in the southern part of the South Oman Salt Basin: Implications for
permission to publish this paper. mini-basin evolution, GeoArabia, 13, 77–108.
Release of data relating to this study is Al-Damegh, K., E. Sandvol, and M. Barazangi (2005), Crustal structure of the Arabian plate: new constraints from the analysis of teleseismic
limited to the materials presented in receiver functions, Earth Planet. Sci. Lett., 231, 177–196.
this paper. This paper reflects the Al-Husseini, M. (2011), Late Ediacaran to early Cambrian (Infracambrian) Jibalah Group of Saudi Arabia, GeoArabia, 16, 69–90.
work of many colleagues at Saudi Al-Husseini, M. (2014), Proposed correlation of Oman’s Abu Mahara Supergroup and Saudi Arabia’s Jibalah Group, GeoArabia, 19, 107–132.
Aramco, particularly C. Hofmann for Al-Husseini, M. (2015), Spatio-temporal position of the Ediacaran Thalbah Basin in the Najd Fault System, Arabian Shield, GeoArabia, 20,
contributions to mapping the Wajid 17–44.
Graben. The paper benefitted from Ali, M. Y., A. B. Watts, and A. Farid (2014), Gravity anomalies of the United Arab Emirates: Implications for basement structures and
review within Aramco by T. Barr, infra-Cambrian salt distribution, GeoArabia, 19, 85–112.
R. Hached, P. Nicholson, and X. Zhao, Allen, M., J. Jackson, and R. Walker (2004), Late Cenozoic reorganization of the Arabia-Eurasia collision and the comparison of short-term and
and external review by M. Allen, long-term deformation rates, Tectonics, 23, TC2008, doi:10.1029/2003TC001530.
W. Bosworth and an anonymous Allen, P. A. (2007), The Huqf Supergroup of Oman: Basin development and context for Neoproterozoic glaciation, Earth Sci. Rev., 84, 139–185.
referee. It has not been possible to Alsharhan, A. S. (1993), Facies and sedimentary environment of the Permian carbonates (Khuff Formation) in the United Arab Emirates,
reconcile input from all reviewers; Sediment. Geol., 84, 89–99.
opinions expressed throughout the Alsharhan, A. S. (2006), Sedimentological character and hydrocarbon parameters of the Middle Permian to Early Triassic Khuff Formation,
paper remain solely that of the author. United Arab Emirates, GeoArabia, 11, 121–158.
Alsharhan, A. S. (2014), Petroleum systems in the Middle East, in Tectonic Evolution of the Oman Mountains, Special Publications, vol. 392,
edited by H. R. Rollinson et al., pp. 361–408, Geol. Soc., London.
Alsharhan, A. S., and C. G. S. C. Kendall (1994), Depositional setting of the upper Jurassic Hith anhydrite of the Arabian Gulf: An analog to
Holocene evaporites of the United Arab Emirates and Lake MacLeod of Western Australia, Am. Assoc. Pet. Geol. Bull., 78, 1075–1096.
Alsharhan, A. S., and C. G. S. C. Kendall (2003), Holocene coastal carbonates and evaporites of the southern Arabian Gulf and their ancient
analogues, Earth Sci. Rev., 61, 191–243.
Amin, A. A., and K. A. Bankher (1997), Karst hazard assessment of eastern Saudi Arabia, Nat. Hazards, 15, 21–30.
As-Saruri, M. A., R. Sorkhabi, and R. Baraba (2010), Sedimentary basins of Yemen: their tectonic development and lithostratigraphic cover,
Arabian J. Geosci., 3, 515–527.
Badalini, G., J. Redfern, and I. D. Carr (2002), A synthesis of the current understanding of the structural evolution of North Africa, J. Pet. Geol.,
25, 249–258.
Boote, D. R. D., D. Mou, and R. I. Waite (1990), Structural evolution of the Suneinah Foreland, Central Oman Mountains, in The Geology and
Tectonics of the Oman Region, Special Publications, vol. 49, edited by A. H. F. Robertson, M. P. Searle, and A. C. Ries, pp. 397–418, Geol. Soc.,
London.
Bosak, P., J. Jaros, J. Spudil, P. Sulovsky, and V. Vaclavek (1998), Salt plugs in the eastern Zagros, Iran: Results of regional geological
reconnaissance, Geolines, 7, 3–174.
Bumby, A. J., and R. Guiraud (2005), The geodynamic setting of the Phanerozoic basins of Africa, J. Afr. Earth Sci., 43, 1–12.
Burgess, P. M., M. Gurnis, and L. Moresi (1997), Formation of sequences in the cratonic interior of North America by interaction between
mantle, eustatic, and stratigraphic processes, Geol. Soc. Am. Bull., 109, 1515–1535.
Cantrell, D. L., P. G. Nicholson, G. W. Hughes, M. A. Miller, A. G. Bhullar, S. T. Abdelbagi, and A. K. Norton (2014), Tethyan petroleum systems of
Saudi Arabia, in Petroleum Systems of the Tethyan Region, AAPG Memoir, vol. 106, edited by L. Marlow, C. C. G. Kendall, and L. A. Yose,
pp. 613–639, Am. Assoc. Pet. Geol., Tulsa, Okla.
Carter, A., and P. D. Clift (2008), Was the Indosinian orogeny a Triassic mountain building or a thermotectonic reactivation event?, C. R.
Geosci., 340, 83–93.
Cloetingh, S., and E. Burov (2011), Lithospheric folding and sedimentary basin evolution: A review and analysis of formation mechanisms,
Basin Res., 23, 257–290.
Collins, A. S., and S. A. Pisarevsky (2005), Amalgamating eastern Gondwana: The evolution of the Circum-Indian Orogens, Earth Sci. Rev., 71,
229–270.

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2435


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

Cooper, D. J. W., M. Y. Ali, and M. P. Searle (2014), Structure of the northern Oman Mountains from the Semail Ophiolite to the foreland basin,
in Tectonic Evolution of the Oman Mountains, Special Publications, vol. 392, edited by H. R. Rollinson et al., pp. 129–153, Geol. Soc., London.
Coward, M. P., and A. C. Ries (2003), Tectonic development of North African basins, in Petroleum Geology of Africa: New Themes and Developing
Technologies, Special Publications, vol. 207, edited by T. Arthur, D. S. MacGregor, and N. R. Cameron, pp. 61–83, Geol. Soc., London.
Cox, G. M., C. J. Lewis, A. S. Collins, G. P. Halverson, F. Jourdan, J. Foden, D. Nettle, and F. Kattan (2012), Ediacaran terrane accretion within the
Arabian–Nubian Shield, Gondwana Res., 21, 341–352.
Craig, J., D. Grigo, A. Rebora, G. Serafini, and E. Tebaldi (2010), From Neoproterozoic to Early Cenozoic: Exploring the potential of older and
deeper hydrocarbon plays across North Africa and the Middle East, in Petroleum Geology: From Mature Basins to New Frontiers—
Proceedings of the 7th Petroleum Geology Conference, edited by B. A. Vining and S. C. Pickering, pp. 673–705, Geol. Soc., London.
Daly, M. C., S. R. Lawrence, K. Diemu-Tshiband, and B. Matouana (1992), Tectonic evolution of the Cuvette Centrale, Zaire, J. Geol. Soc., 149,
539–546.
DeMets, C., R. G. Gordon, D. F. Argus, and S. Stein (1990), Current plate motions, Geophys. J. Int., 101, 425–478.
Dewey, J. F., R. E. Holdsworth, and R. A. Strachan (1998), Transpression and transtension zones, in Continental Transpressional and
Transtensional Tectonics, Special Publications, vol. 135, edited by R. E. Holdsworth, R. A. Strachan, and J. F. Dewey, pp. 1–14, Geol. Soc.,
London.
Dilek, Y., and Z. Ahmed (2003), Proterozoic ophiolites of the Arabian Shield and their significance in Precambrian tectonics, in Ophiolites in
Earth History, Special Publications, vol. 218, edited by Y. Dilek and R. T. Robinson, pp. 685–700, Geol. Soc., London.
Doebrich, J. L., A. M. Al-Jehani, A. A. Siddiqui, T. S. Hayes, J. L. Wooden, and P. R. Johnson (2007), Geology and metallogeny of the Ar Rayn
terrane, eastern Arabian shield: Evolution of a Neoproterozoic continental-margin arc during assembly of Gondwana within the East
African orogen, Precambrian Res., 158, 17–50.
Dutton, B. J. (1997), Finite strains in transpression zones with no boundary slip, J. Struct. Geol., 19, 1189–1200.
Dyer, R. A., and M. Husseini (1991), The Western Rub’ Al-Khali Infracambrian graben system, SPE, SPE-21396-MS, 8pp.
Edgell, H. S. (2006), Arabian Deserts: Nature, Origin and Evolution, pp. 664, Springer, Dordrecht, Netherlands.
Elliott, D. (1983), The construction of balanced cross-sections, J. Struct. Geol., 5, 101.
Faqira, M., M. Rademakers, and A. M. Afifi (2009), New insights into the Hercynian Orogeny, and their implications for the Paleozoic
hydrocarbon system in the Arabian Plate, GeoArabia, 14, 199–228.
Filbrandt, J. B., S. Al-Dhahab, A. Al-Habsy, K. Harris, J. Keating, S. Al-Mahruqi, S. I. Ozkaya, P. D. Richard, and T. Robertson (2006), Kinematic
interpretation and structural evolution of North Oman, Block 6, since the Late Cretaceous and implications for timing of hydrocarbon
migration into Cretaceous reservoirs, GeoArabia, 11, 97–140.
Flament, N., M. Gurnis, and R. D. Müller (2013), A review of observations and models of dynamic topography, Lithosphere, 5, 189–210.
Fossen, H. (2010), Extensional tectonics in the North Atlantic Caledonides: A regional view, in Continental Tectonics and Mountain Building:
The Legacy of Peach and Horne, Special Publications, vol. 335, edited by R. D. Law et al., pp. 767–793, Geol. Soc., London.
Fournier, M., C. Lepvrier, P. Razin, and L. Jolivet (2006), Late Cretaceous to Paleogene post-obduction extension and subsequent Neogene
compression in the Oman Mountains, GeoArabia, 11, 17–40.
Fritz, H., et al. (2013), Orogen styles in the East African Orogen: A review of the Neoproterozoic to Cambrian tectonic evolution, J. Afr. Earth
Sci., 86, 65–106.
Gaina, C., D. J. J. van Hinsbergen, and W. Spakman (2015), Tectonic interactions between India and Arabia since the Jurassic reconstructed
from marine geophysics, ophiolite geology, and seismic tomography, Tectonics, 34, 875–906, doi:10.1002/2014TC003780.
Gass, I. G., A. C. Ries, R. M. Shackleton, and J. D. Smewing (1990), Tectonics, geochronology and geochemistry of the Precambrian rocks of
Oman, in The Geology and Tectonics of the Oman region, Special Publications, vol. 49, edited by A. H. F. Robertson, M. P. Searle, and A. C. Ries,
pp. 585–599, Geol. Soc., London.
Gillcrist, R., M. Coward, and J.-L. Mugnier (1987), Structural inversion and its controls: Examples from the Alpine foreland and the French Alps,
Geodinamica Acta, 1, 5–34.
Golonka, J. (2007), Late Triassic and Early Jurassic palaeogeography of the world, Palaeogeogr. Palaeoclimatol. Palaeoecol., 244, 297–307.
Gray, D. R., D. A. Foster, J. G. Meert, B. D. Goscombe, R. Armstrong, R. A. J. Trouw, and C. W. Passchier (2008), A Damara orogen perspective on
the assembly of southwestern Gondwana, in West Gondwana: Pre-Cenozoic Correlations Across the South Atlantic Region, Special
Publications, vol. 294, edited by R. J. Pankhurst et al., pp. 257–278, Geol. Soc., London.
Gregori, D. A., B. Saini-Eidukat, L. Benedini, L. Strazzere, M. Barros, and J. Kostadinoff (2016), The Gondwana Orogeny in northern North
Patagonian Massif: Evidences from the Caita Có granite, La Seña and Pangaré mylonites, Argentina, Geosci. Front., 7, 621–638.
Guiraud, R., and W. Bosworth (1999), Phanerozoic geodynamic evolution of northeastern Africa and the northwestern Arabian platform,
Tectonophysics, 315, 73–104.
Guiraud, R., W. Bosworth, J. Thierry, and A. Delplanque (2005), Phanerozoic geological evolution of Northern and Central Africa: An overview,
J. Afr. Earth Sci., 43, 83–143.
Haddoum, H., R. Guiraud, and A. Moussine-Pouchkine (2001), Hercynian compressional deformations of the Ahnet-Mouydir Basin, Algerian
Saharan Platform: far-field stress effects of the Late Paleozoic orogeny, Terra Nova, 13, 220–226.
Haghipour, A., A. Saidi, A. Aganabati, A. Moosavi, A. Mohebi, M. Sadeghi, T. Delavar, S. Eskandari, F. Bagheri, and M. Zarei (2009), International
Geological Map of the Middle East, Commission for the Geological Map of the World (Subcommission for the Middle East).
Haq, B. U., and A. M. Al-Qahtani (2005), Phanerozoic cycles of sea-level change on the Arabian Platform, GeoArabia, 10, 127–160.
Harvey, M. J., and S. A. Stewart (1998), Influence of salt on the structural evolution of the Channel Basin, in Development, Evolution and
Petroleum Geology of the Wessex Basin, Special Publications, vol. 133, edited by J. R. Underhill, pp. 241–266, Geol. Soc., London.
Heidbach, O., M. Tingay, A. Barth, J. Reinecker, D. Kurfeß, and B. Müller (2010), Global crustal stress pattern based on the World Stress Map
database release 2008, Tectonophysics, 482, 3–15.
Heikal, M. T. S., S. A. Al-Khirbash, A. M. Hassan, A. M. Al-Kotbah, and K. M. Al-Selwi (2014), Lithostratigraphy, deformation history, and tectonic
evolution of the basement rocks, Republic of Yemen: An overview, Arab. J. Geosci., 7, 2007–2018.
Heward, A. P. (1990), Salt removal and sedimentation in Southern Oman, in The Geology and Tectonics of the Oman Region, Special
Publications, vol. 49, edited by A. H. F. Robertson, M. P. Searle, and A. C. Ries, pp. 637–652, Geol. Soc., London.
Holm, D. A. (1960), Desert geomorphology in the Arabian Peninsula, Science, 132, 1369–1379.
Holt, P. J., M. B. Allen, and J. van Hunen (2015), Basin formation by thermal subsidence of accretionary orogens, Tectonophysics, 639, 132–143.
Hughes, G., and N. Naji (2008), Sedimentological and micropalaeontological evidence to elucidate post-evaporitic carbonate
palaeoenvironments of the Saudi Arabian latest Jurassic, Volumina Jurassica, 6, 61–73.
Husseini, M. I., and S. I. Husseini (1990), Origin of the Infracambrian salt basins of the Middle East, in Classic Petroleum Provinces, Special
Publications, vol. 50, edited by J. Brook, pp. 279–292, Geol. Soc., London.

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2436


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

Jackson, M. P. A., R. R. Cornelius, C. H. Craig, A. Gansser, J. Stöcklin, and C. J. Talbot (1990), Salt Diapirs of the Great Kavir, Central Iran, pp. 139,
Geol. Soc. Am., Mem, Boulder, Colo.
Jenyon, M. K. (1989), Plastic flow and contraflow in superposed Zechstein salt sequences, J. Pet. Geol., 12, 477–486.
Johnson, C., T. Hauge, S. Al-Menhali, S. Bin Sumaidaa, B. Sabin, and B. West (2005), Structural styles and tectonic evolution of onshore and
offshore Abu Dhabi, UAE, in International Petroleum Technology Conference, 21-23 November, Doha, Qatar, edited, IPTC-10646-MS, 15 pp.
Johnson, P. R. (2003), Post-amalgamation basins of the NE Arabian shield and implications for Neoproterozoic III tectonism in the northern
East African orogen, Precambrian Res., 123, 321–337.
Johnson, P. R., and I. C. F. Stewart (1995), Magnetically inferred basement structure in central Saudi Arabia, Tectonophysics, 245, 37–52.
Johnson, P. R., A. Andresen, A. S. Collins, A. R. Fowler, H. Fritz, W. Ghebreab, T. Kusky, and R. J. Stern (2011), Late Cryogenian-Ediacaran history
of the Arabian-Nubian Shield: A review of depositional, plutonic, structural, and tectonic events in the closing stages of the northern East
African Orogen, J. Afr. Earth Sci., 61, 167–232.
Jones, R. R., R. E. Holdsworth, P. Clegg, K. McCaffrey, and E. Tavarnelli (2004), Inclined transpression, J. Struct. Geol., 26, 1531–1548.
Kent, W. N. (2010), Structures of the Kirkuk Embayment, northern Iraq: Foreland structures or Zagros Fold Belt structures?, GeoArabia, 15,
147–188.
Konert, G., A. M. Afifi, S. A. Al-Hajri, and H. J. Droste (2001), Paleozoic stratigraphy and hydrocarbon habitat of the Arabian plate, GeoArabia, 6,
407–442.
Kröner, A., and R. J. Stern (2004), Pan-African Orogeny, in Encyclopedia of Geology, edited by R. C. Selley, L. R. M. Cocks, and I. R. Plimer, pp. 12,
Elsevier, Amsterdam.
Kusky, T. M., and M. I. Matsah (2003), Neoproterozoic dextral faulting on the Najd Fault System, Saudi Arabia, preceded sinistral faulting and
escape tectonics related to closure of the Mozambique Ocean, in Proterozoic East Gondwana: Supercontinent Assembly and Breakup,
Special Publications, vol. 206, edited by M. Yoshida and B. F. Windley, pp. 327–361, Geol. Soc., London.
Lecorche, J.-P. (1983), Structure of the Mauritanides, in Regional Trends in the Geology of the Appalachian-Caledonian-Hercynian-Mauritanide
Orogen, vol. 116, edited by P. E. Schenk et al., pp. 347–353, Springer, Netherlands.
Li, S., S. Abe, L. Reuning, S. Becker, J. L. Urai, and P. A. Kukla (2012), Numerical modelling of the displacement and deformation of embedded
rock bodies during salt tectonics: A case study from the South Oman Salt Basin, in Salt Tectonics, Sediments and Prospectivity, Special
Publications, vol. 363, edited by G. I. Alsop et al., pp. 503–520, Geol. Soc., London.
Linnemann, U., R. D’Lemos, K. Drost, T. Jeffries, A. Gerdes, R. L. Romer, S. D. Samson, and R. D. Strachan (2008), Cadomian tectonics, in The
Geology of Central Europe Volume 1: Precambrian and Palaeozoic, edited by T. McCann, pp. 103–154, Geol. Soc., London.
Lofi, J., S. Berné, M. Tesson, M. Seranne, and P. Pezard (2012), Giant solution-subsidence structure in the Western Mediterranean related to
deep substratum dissolution, Terra Nova, 24, 181–188.
Loosveld, R. J. H., A. Bell, and J. J. M. Terken (1996), The tectonic evolution of interior Oman, GeoArabia, 1, 28–50.
Mann, P., L. Gahagan, and M. B. Gordon (2003), Tectonic setting of the world’s giant oil and gas fields, in Giant Oil and Gas Fields of the Decade
1990–1999, AAPG Memoir, vol. 78, edited by M. T. Halbouty, pp. 15–105, AAPG, Tulsa, Okla.
Marquer, D., T. J. Peters, and E. Gnos (1995), A new structural interpretation for the emplacement of the Masirah ophiolites (Oman): A main
Paleocene intra-oceanic thrust, Geodinamica Acta, 8, 13–19.
McKenzie, D., M. C. Daly, and K. Priestley (2015), The lithospheric structure of Pangea, Geology, 43, 783–786.
Mercolli, I., A. P. Briner, R. Frei, R. Schönberg, T. F. Nägler, J. Kramers, and T. Peters (2006), Lithostratigraphy and geochronology of the
Neoproterozoic crystalline basement of Salalah, Dhofar, Sultanate of Oman, Precambrian Res., 145, 182–206.
Metcalfe, I. (2013), Gondwana dispersion and Asian accretion: Tectonic and palaeogeographic evolution of eastern Tethys, J. Asian Earth Sci.,
66, 1–33.
Mouthereau, F., O. Lacombe, and J. Vergés (2012), Building the Zagros collisional orogen: Timing, strain distribution and the dynamics of
Arabia/Eurasia plate convergence, Tectonophysics, 532–535, 27–60.
Nance, R. D., G. Gutiérrez-Alonso, J. D. Keppie, U. Linnemann, J. B. Murphy, C. Quesada, R. A. Strachan, and N. H. Woodcock (2010), Evolution of
the Rheic Ocean, Gondwana Res., 17, 194–222.
Nehlig, P., A. Genna, and F. Asirfane (2002), A review of the Pan-African evolution of the Arabian Shield, GeoArabia, 7, 103–124.
Pankhurst, R. J., C. W. Rapela, C. M. Fanning, and M. Márquez (2006), Gondwanide continental collision and the origin of Patagonia, Earth Sci.
Rev., 76, 235–257.
Perotti, C. R., S. Carruba, M. Rinaldi, G. Bertozzi, L. Feltre, and M. Rahimi (2011), The Qatar–South Fars Arch development (Arabian Platform,
Persian Gulf): Insights from seismic interpretation and analogue modelling, in New Frontiers in Tectonic Research—At the Midst of Plate
Convergence, edited by U. Schattner, pp. 325–352, InTech.
Perotti, C., L. Chiariotti, I. Bresciani, L. Cattaneo, and G. Toscani (2016), Evolution and timing of salt diapirism in the Iranian sector of the
Persian Gulf, Tectonophysics, 679, 180–198.
Peters, J. M., J. B. Filbrandt, J. P. Grotzinger, M. J. Newall, M. W. Shuster, and H. A. Al-Siyabi (2003), Surface-piercing salt domes of interior North
Oman, and their significance for the Ara carbonate “stringer” hydrocarbon play, GeoArabia, 8, 231–270.
Powers, R. W., L. F. Ramirez, C. D. Redmond, and E. L. Elberg (1966), Geology of the Arabian Peninsula: Sedimentary geology of Saudi Arabia,
147 pp., U.S. Geological Survey Professional Paper 560-D.
Robin, P.-Y. F., and A. R. Cruden (1994), Strain and vorticity patterns in ideally ductile transpression zones, J. Struct. Geol., 16, 447–466.
Rodgers, A. J., W. R. Walter, R. J. Mellors, A. M. S. Al-Amri, and Y.-S. Zhang (1999), Lithospheric structure of the Arabian Shield and Platform
from complete regional waveform modelling and surface wave group velocities, Geophys. J. Int., 138, 871–878.
Schreurs, G., and A. Immenhauser (1999), West-northwest directed obduction of the Batain Group on the eastern Oman continental margin
at the Cretaceous-Tertiary boundary, Tectonics, 18, 148–160, doi:10.1029/1998TC900020.
Schröder, S., B. C. Schreiber, J. E. Amthor, and A. Matter (2003), A depositional model for the terminal Neoproterozoic-Early Cambrian Ara
Group evaporites in south Oman, Sedimentology, 50, 879–898.
Searle, M. P., C. J. Warren, D. J. Waters, and R. R. Parrish (2004), Structural evolution, metamorphism and restoration of the Arabian continental
margin, Saih Hatat region, Oman Mountains, J. Struct. Geol., 26, 451–473.
Sepehr, M., J. Cosgrove, and M. Moieni (2006), The impact of cover rock rheology on the style of folding in the Zagros fold-thrust belt,
Tectonophysics, 427, 265–281.
Serrane, M., and M. Seguret (1987), The Devonian basins of western Norway: tectonics and kinematics of an extending crust, in Continental
Extensional Tectonics, Special Publications, vol. 28, edited by M. P. Coward, J. F. Dewey, and P. L. Hancock, pp. 537–548, Geol. Soc.,
London.
Sharief, F. A., M. S. Khan, and K. Magara (1991), Outcrop-subcrop sequence and diagenesis of Upper Jurassic Arab-Hith formations, central
Saudi Arabia, J. King Abdulaziz Univ.: Earth Sci., 4, 105–136.

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2437


19449194, 2016, 10, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1002/2016TC004212 by Nat Prov Indonesia, Wiley Online Library on [20/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1002/2016TC004212

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, vol. 2, pp. 371, Gulf Petrolink, Manama.
Sherkati, S., M. Molinaro, D. Frizon de Lamotte, and J. Letouzey (2005), Detachment folding in the Central and Eastern Zagros fold-belt (Iran):
Salt mobility, multiple detachments and late basement control, J. Struct. Geol., 27, 1680–1696.
Smith, A. G. (2012), A review of the Ediacaran to Early Cambrian (“Infra-Cambrian”) evaporites and associated sediments of the Middle East,
in Geology and Hydrocarbon Potential of Neoproterozoic-Cambrian Basins in Asia, Special Publications, vol. 366, edited by G. M. Bhat et al.,
pp. 229–250, Geol. Soc., London.
Stampfli, G. M., and G. D. Borel (2002), A plate tectonic model for the Paleozoic and Mesozoic constrained by dynamic plate boundaries and
restored synthetic oceanic isochrons, Earth Planet. Sci. Lett., 196, 17–33.
Stephenson, R. A., and S. A. P. L. Cloetingh (1991), Some examples and mechanical aspects of continental lithospheric folding,
Tectonophysics, 188, 27–37.
Stern, R. J., and P. Johnson (2010), Continental lithosphere of the Arabian Plate: A geologic, petrologic, and geophysical synthesis, Earth Sci.
Rev., 101, 29–67.
Stewart, S. A. (2014), Detachment-controlled triangle zones in extension and inversion tectonics, Interpretation, 2, SM29–SM38.
Stewart, S. A., and M. P. Coward (1995), Synthesis of salt tectonics in the southern North Sea, UK, Mar. Pet. Geol., 12, 457–475.
Stewart, S. A., C. T. Reid, N. P. Hooker, and O. W. Kharouf (2016), Mesozoic siliciclastic reservoirs and petroleum system in the Rub’ Al-Khali
basin, Saudi Arabia, Am. Assoc. Pet. Geol. Bull., 100, 819–841.
Strozyk, F., H. Van Gent, J. L. Urai, and P. A. Kukla (2012), 3D seismic study of complex intra-salt deformation: An example from the Upper
Permian Zechstein 3 stringer, western Dutch offshore, in Salt Tectonics, Sediments and Prospectivity, Special Publications, vol. 363, edited by
G. I. Alsop et al., pp. 489–501, Geol. Soc., London.
Tankard, A., H. Welsink, P. Aukes, R. Newton, and E. Stettler (2009), Tectonic evolution of the Cape and Karoo basins of South Africa, Mar. Pet.
Geol., 26, 1379–1412.
Terken, J. M. J., N. L. Frewin, and S. L. Indrelid (2001), Petroleum systems of Oman: Charge timing and risks, Am. Assoc. Pet. Geol. Bull., 85,
1817–1845.
Thomas, R. J., R. A. Ellison, K. M. Goodenough, N. M. W. Roberts, and P. A. Allen (2015), Salt domes of the UAE and Oman: Probing eastern
Arabia, Precambrian Res., 256, 1–16.
Van Gent, H., J. L. Urai, and M. de Keijzer (2011), The internal geometry of salt structures—A first look using 3D seismic data from the
Zechstein of the Netherlands, J. Struct. Geol., 33, 292–311.
Vaslet, D., Y.-M. Le Nidre, D. Vachard, J. Broutin, S. Crasquin-Soleau, M. Berthelin, J. Gaillot, M. Halawani, and M. Al-Husseini (2005), The
Permian-Triassic Khuff Formation of central Saudi Arabia, GeoArabia, 10, 77–134.
Veevers, J. J. (2012), Reconstructions before rifting and drifting reveal the geological connections between Antarctica and its conjugates in
Gondwanaland, Earth Sci. Rev., 111, 249–318.
Vernant, P., et al. (2004), Present-day crustal deformation and plate kinematics in the Middle East constrained by GPS measurements in Iran
and northern Oman, Geophys. J. Int., 157, 381–398.
Warburton, J., T. J. Burnhill, R. H. Graham, and K. P. Isaac (1990), The evolution of the Oman Mountains Foreland Basin, in The Geology and
Tectonics of the Oman Region, Special Publications, vol. 330, edited by A. H. F. Robertson, M. P. Searle, and A. C. Ries, pp. 419–427, Geol. Soc.,
London.
Weijermars, R. (1999), Surface geology and Tertiary growth of the Dammam Dome, Saudi Arabia: A new field guide, GeoArabia, 4, 199–226.
Whitehouse, M. J., B. F. Windley, D. B. Stoeser, S. Al-Khirbash, M. A. O. Ba-Bttat, and A. Haider (2001), Precambrian basement character of
Yemen and correlations with Saudi Arabia and Somalia, Precambrian Res., 105, 357–369.
Windley, B. F., M. J. Whitehouse, and M. A. O. Ba-Bttat (1996), Early Precambrian gneiss terranes and Pan-African island arcs in Yemen: Crustal
accretion of the eastern Arabian Shield, Geology, 24, 131–134.
Wood, B. G. M. (2015), Rethinking post-Hercynian basin development: Eastern Mediterranean Region, GeoArabia, 20, 175–224.
Zampetti, V., L. Madsen, H. Cromie, G. Durance, M. Emang, N. Bounoua, and T. Gagigi (2014a), The role of regional basement fabric on
Cretaceous structural deformation; a case study from Al Shaheen Field, offshore Qatar, International Petroleum Technology Conference,
Doha, Qatar, 20–22 January 2014, pp. 11.
Zampetti, V., X. Marquez, S. Mukund, S. S. Bach, and M. I. Emang (2014b), 3D Seismic Characterization of UER Karst, Offshore Qatar, in
International Petroleum Technology Conference, Doha, Qatar, 20–22 January 2014, International Petroleum Technology Conference,
pp. 12.
Ziegler, M. A. (2001), Late Permian to Holocene paleofacies evolution of the Arabian plate and its hydrocarbon occurrences, GeoArabia, 6,
445–504.

STEWART STRUCTURAL GEOLOGY OF THE RUB’ AL-KHALI 2438

You might also like