Cryptic Orogeny uplift of the at an alleged passive margin Reuben Johannes Hansman Academic dissertation for the Degree of Doctor of Philosophy in Geology at Stockholm University to be publicly defended on Friday 7 September 2018 at 10.00 in De Geersalen, Geovetenskapens hus, Svante Arrhenius väg 14.

Abstract Mountains evolve and grow because of the large forces that occur from the collision of tectonic plates. Plate boundaries change and move through time, and regions that were once stable, shallow-marine environments can be dragged into subduction zones and get transformed into vast mountain ranges. The Al Hajar Mountains in consist of carbonate rocks which show that during most of the Mesozoic (c. 268 Ma – 95 Ma) they had not yet formed but were flat and below sea level. Following this, in the Late Cretaceous (c. 95 Ma), a major tectonic event caused oceanic crust to be obducted onto this Mesozoic carbonate platform. Then after obduction a shallow marine environment resumed, and Paleogene sedimentary rocks were deposited. Currently, the central mountains are located on the Arabian Plate and are 200 km away from the convergent plate boundary with Eurasia. Here, Arabia is being subducted. Further towards the northwest Arabia and Eurasia are colliding, forming the Zagros Mountains which initiated no earlier than the Oligocene (c. 30 Ma). At this time the mountains were even further away from the plate boundary. The problem with the Al Hajar Mountains is that they record a collision, but are not in a collisional zone. To better understand the formation of the Al Hajar Mountains, a multidiscipline approach was used to investigate the timing at which they developed. This included applying low- temperature thermochronology, U-Pb dating of brittle structures, and balanced cross-sections. Results indicate that the orogeny began in the late Eocene and had concluded by the early Miocene (40 Ma – 15 Ma). Therefore, the uplift of the Al Hajar Mountains is not related to either the older Late Cretaceous ophiolite obduction or the younger Zagros collision, and a new tectonic model is proposed. This research shows that the Cenozoic tectonic history of northern Oman is more cryptic than what has been formerly presented.

Keywords: uplift, mountains, structural, low-temperature, thermochronology, dating, fission-track, (U-Th)/He, U-Pb, calcite, structure-from-motion, photogrammetry, UAV, trishear, fault-propagation, Hafit, Al Hajar Mountains, , Oman.

Stockholm 2018 http://urn.kb.se/resolve?urn=urn:nbn:se:su:diva-157506

ISBN 978-91-7797-338-6 ISBN 978-91-7797-339-3

Department of Geological Sciences

Stockholm University, 106 91 Stockholm

CRYPTIC OROGENY

Reuben Johannes Hansman

Cryptic Orogeny uplift of the Al Hajar Mountains at an alleged passive margin

Reuben Johannes Hansman ©Reuben Johannes Hansman, Stockholm University 2018

ISBN print 978-91-7797-338-6 ISBN PDF 978-91-7797-339-3

Cover: Photograph of the Al Hajar Mountains, Oman (colour correction by Mehran Hussain).

Printed in Sweden by Universitetsservice US-AB, Stockholm 2018 Distributor: Department of Geological Sciences, Stockholm University To my friend Graeme Blackburn 1984 – 2018

tetelestai

Sammanfattning

Bergskedjor bildas till följd av stora de stora krafter som inträffar vid en kontinentkollision. Kontinentala plattgränser förändras och flyttas över tiden och tidigare grundhavsområden kan omvandlas till långsträckta bergskedjor. Al Hajar-bergen i Oman består av karbonatbergarter som deponerades som horisontella lager under havsytan under mesozoikum (c. 268 – 95 Ma), vilket visar att bergskedjan måste ha bildats snare än så. Under senare delen av kritaperioden (c. 95 Ma) orsakade en storskalig tektonisk händelse upplyftning av oceanjordskorpa ovanpå denna mesozoiska karbonatplatform. Efter denna händelse återgick området till deponering av sediment i grundhavsmiljö, vilket givit upphov till paleogena sedimentära bergarter. I dagsläget befinner sig bergskedjans centrala delar inom den Arabiska kontinentalplattan, cirka 200 km från dess konvergenta plattgräns mot Eurasien under vilken den subduceras. Längre mot nordost sker istället aktiv kollision mellan dessa plattor, vilket givit upphov till bildandet av Zagrosbergen. Denna bergskedjeveckning initierades som tidigast under oligocen (c. 30 Ma). Vid denna tidpunkt befann sig Al Hajar-bergen ännu längre från plattgränsen jämfört idag. Den olösta gåtan om Al Hajar-bergen är alltså hur de kan uppvisa spår efter en kontinentkollision utan att befinna sig i en kollisionszon. En kombination av flera olika metoder användes för att undersöka tidpunkten för dess bildande och därmed öka förståelsen för hur de har bildats. Dessa metoder inkluderar lågtemperaturtermokronologi, U-Pb-datering av spröda strukturer samt studier av geologiska profiler. Resultaten indikerar att orogenesen inleddes i senare delen av eocen och var avslutad i den tidigare delen av miocen (40 – 15 Ma). Därmed är bildandet av Al Hajar-bergen varken relaterad till den äldre senkretaceiska ofiolitupplyftningen eller den yngre zagroskollisionen, varför en ny tektonisk modell måste föreslås. Detta visar att den Omans tektoniska historia under kenozoikum är mer komplex än vad tidigare rapporterats.

i

List of papers and authors’ contributions

Paper I Hansman, R. J., Ring, U., Thomson, S. N., den Brok, B., & Stübner, K. (2017). Late Eocene uplift of the Al Hajar Mountains, Oman, supported by stratigraphy and low-temperature thermochronology. Tectonics, 36, 3081–3109. https://doi.org/ 10.1002/2017TC004672

Paper II Hansman, R. J., Albert, R., Gerdes, A., & Ring, U. (2018). Absolute ages of multiple generations of brittle structures by U-Pb dating of calcite. Geology; 46 (3), 207–210. doi: https://doi.org/10.1130/G39822.1

Paper III Hansman, R. J., & Ring, U. (manuscript). Oligocene–Miocene trishear fault- propagation folding of the Jabal Hafit Anticline, supported by a three-dimensional geological model; and assessing structure-from-motion (SfM) photogrammetry of unmanned aerial vehicle (UAV) photographs for mapping.

Reuben Hansman carried out three field seasons (one for each paper) with the assistance of Uwe Ring. Hansman wrote the papers and created the figures and tables.

Paper I: Reuben Hansman carried out apatite and zircon mineral separation. For fission- track dating Reuben Hansman counted the tracks, and Stuart Thomson mounted, polished and etched the grains and coordinated the irradiation. For (U–Th)/He dating Hansman picked and packed the apatite grains and carried out the He-degassing and quadrupole mass spectrometer analysis. Uttam Chowdhury and Erin Able carried out the isotope dilution and solution high-resolution inductively coupled plasma-mass spectrometry. Zircon grains were sent to Konstanze Stübner who carried out the zircon (U–Th)/He dating. All co-authors helped with revisions.

Paper II: Reuben Hansman interpreted the data, Peter Späthe made thick-sections, Richard Albert and Axel Gerdes carried out the U–Pb dating, and all co-authors helped with revisions.

Paper III: Reuben Hansman created the three-dimensional geological model, and the UAV– SfM mapping. Uwe Ring assisted with mapping and helped revise the paper.

iii

Acknowledgements

I am incredibly grateful to my supervisor Uwe Ring, who allowed me to go in my own direction but also guided me when I needed it. Also, the preparation, analysis, and interpretation of my samples would have been impossible without the help of Stuart Thomson, Richard Albert, Axel Gerdes, Peter Reiners, Erin Abel, Uttam Chowdhury, Konstanze Stübner, and Per-Olof Persson for which I am thankful. Additionally, I appreciate the time that Victoria Pease and Hemin Koyi gave to commenting on my licentiate thesis which helped to clarify and improve Paper I. I value the discussions I had with Bas den Brok, Alasdair Skelton, and Iain Pitcairn regarding my thesis and geology in general. I would also like to thank, Dan Zetterberg, Eve Arnold, Runa Jacobsson, Krister Junghahn, Malin Andersson, and Viktoria Arwinge for assisting me through the finer details of completing a PhD at Stockholm University. I equally thank Hagen Bender, Alexandre Peillod, Clifford Patten, Barbara Kleine, Alexan- der Lewerentz (additional thanks for the Sammanfattning translation), Emelie Axelsson, Xi- aojing Zhang, Elin Tollefsen, Fitsum Girum, and Ahmad Boskabadi for running the race with me. Also, cheers to Remi Vachon, Henrik Linnros and Hermes Pantazidis. Finally, but more importantly, I cannot thank my parents enough for their continual support. I also acknowledge and appreciate the funding from the Swedish Foundation for Inter- national Cooperation in Research and Higher Education (IB2015-6002), the Bolin Centre for Climate Research (RA1 and RA6), the Royal Swedish Academy of Sciences (GS2015-0002 and GS2017-0028), the Stiftelsen Lars Hiertas Minne (FO2015-0130 and FO2016-0159), the K & A Wallenberg Foundation 2016 Travel Grant, the Stiftelsen Anna-Greta och Holger Crafoords fond and Stockholm University.

v

Contents

Sammanfattning i

List of papers and authors’ contributions iii

Acknowledgements v

Summary–Kappa 1 1 Introduction...... 1 2 Thesis Aims...... 4 3 Methodology ...... 4 3.1 Paper I: Low-Temperature Thermochronology...... 5 3.2 Paper II: U-Pb Calcite Dating ...... 10 3.3 Paper III: Three-Dimensional Geological Modelling...... 12 4 Summary of Key Results...... 14 4.1 Paper I: AGU Journal Tectonics ...... 14 4.2 Paper II: GSA Journal Geology ...... 16 4.3 Paper III: Manuscript...... 17 5 Conclusion and Future Work...... 19

Tectonostratigraphic Summary of the Central Al Hajar Mountains 21

References 23

Appended Papers: I, II, and III 27

vii

Summary–Kappa

1 Introduction

The tectonic setting and history of north Oman (Fig.1a) appears, at first glance, to be rel- atively simple and well understood. However, there is currently no satisfactory explanation for the presence of the 3 km high mountain range which spans 700 km from Sur through to the (Fig1b). This mountain range, named the Al Hajar Mountains (translated as the Rocky Mountains), is currently located on continental crust at the northeast corner of the Arabian Plate. The central mountains consist of two large scale anticlines, the Jabal Akhdar and Saih Hatat culminations, where the older rocks are exposed in the eroded cores. The Arabian Plate is moving north at a rate of 2–3 cm yr–1, relative to Eurasia (ArRajehi et al., 2010). Most of this convergence boundary is a collisional one, where continental crust on both plates are colliding to form the Zagros Mountains. However, towards the southeast this boundary transitions into the Makran subduction zone. Here, oceanic crust on the Arabian Plate is subducting beneath the continental crust of Eurasia. This tectonic setting places the Al Hajar Mountains 200 km to the southeast of the Makran trench and accretionary wedge (Fig.2). The geology of the mountains (Fig1b) reveals a rich history that dates back to the Precambrian. For a detailed tectonostratigraphic column see p. 21. However, most of the older rocks are irrelevant to the uplift story of the mountains. This is because most of Oman contains middle Permian to Late Cretaceous (268 Ma – 93 Ma) carbonate platform rocks (Glennie et al., 1974), which demonstrates that the entire area was in a shallow marine environment with a flat terrain. During the Late Cretaceous (Fig.3), while deposition of the carbonate platform was ongoing, an oceanic crust was being formed in the north (Rioux et al., 2013). Soon after crystallization occurred at 95 Ma to 93 Ma (Hacker, 1994), this crust (Semail Ophiolite), together with deep-marine sediments (Sumeini and Hawasina nappes), began to be obducted on top of the carbonate platform. As the ophiolite was thrust towards the south, the northern edge of the carbonate platform in Saih Hatat (Fig.1b) was subducted to depths of c. 80 km to eclogite facies metamorphism by 79 Ma (Warren et al., 2005). The obduction process came to an end by c. 70 Ma. Also at this time eclogite was also exhumed to less than 10 km depth (Searle et al., 2004; Saddiqi et al., 2006). The ophiolite was subaerial as an erosional surface and laterite horizon formed on it, which was then covered by rudist-bearing Upper Maastrichtian limestones (Searle et al., 2004). This indicates that in most places, any positive relief that had formed during obduction was subsequently removed and the area returned once again into a shallow marine environment. Stable deposition of limestone continued from c. 65 Ma to 40 Ma (Nolan et al., 1990), and possibly blanketed the entire region.

1 2 Summary–Kappa

Figure 1. (a) Tectonic map of the Arabian Plate and location of the Al Hajar Mountains (adapted from Stern and Johnson, 2010). Note that the central mountains are currently not at a plate boundary but are c. 200 km south of the Makran subduction zone. (b) Simplified geological map of the Al Hajar Moutains (adapted from Forbes et al., 2010). Cryptic Orogeny: uplift of the Al Hajar Mountains at an alleged passive margin 3

What happened after 40 Ma is less known as all limestone in the central mountains was removed. However, on the flanks of the Jabal Akhdar and Saih Hatat culminations, there remain some late Eocene (c. 35 Ma) carbonate rocks. Even further from the culminations, in the foreland, at the Jabal Hafit anticline (Fig.1b), there are Oligocene and Miocene limestones (Warrak, 1996). At Jabal Hafit there are exposed early Eocene to early Miocene sedimentary rocks which are greatly deformed. The structure also has comparable kinematics to that of the Jabal Akhdar and Saih Hatat culminations. This evidence suggests that the uplift of the central mountains cannot be older than about 40 Ma, and that the foreland was deformed after the Oligocene or Miocene (depending if deposition was pre- or syn-tectonic). Therefore, it is unlikely that the Late Cretaceous ophiolite obduction generated the high relief of the modern day Al Hajar Mountains. The collision between Arabia and Eurasia may be responsible for the uplift. The ear- liest age for the onset of the Zagros collision is Oligocene (Gavillot et al., 2010), but peak deformation of the Zagros Mountains took place at the end of the Pliocene (Hessami et al., 2001). It is possible that there is a temporal correlation between the Zagros and Al Hajar Mountains. However, the central Al Hajar Mountains are presently 200 km from the Makran trench (Fig.2), and more than 400 km from the Zagros collision zone. The mountains were even further away at about 700 km in the Oligocene (McQuarrie and Van Hinsbergen, 2013), when the Zagros collision initiated. Therefore, it seems improbable that the Arabia and Eurasia Plate boundary directly generated the crustal thickening in northern Oman. However, to test this it is paramount to find out the time at which the Al Hajar Mountains formed. Following this, it will then be possible to discuss what may have caused the uplift and comprehend the full tectonic history of northern Oman.

Figure 2. Section A–A’ showing a simplified present-day view of the upper crust (for location see (Fig.1b). The oceanic crust of the Arabian Plate is subducting north under the Eurasian continental plate. Note that currently the Al Hajar Mountains are 200 km away from the Makran accretionary wedge. It is inconceivable that at an oceanic-continental convergence plate boundary there would be crustal thickening outboard of the subduction zone in the downgoing plate. Based on this configuration alone, the formation of the Al Hajar Mountains cannot have occurred at the collision boundary between the Arabian and Eurasian tectonic plates. It is unknown if the ophiolite on the carbonate platform of the Arabian Plate is connected with the oceanic crust in the (marked by ‘?’). 4 Summary–Kappa

Figure 3. A timeline illustrating the sedimentary and tectonic events of northern Oman, which can help to explain the evolution of the Al Hajar Mountains.

2 Thesis Aims

1. To constrain when the present day topography of the Al Hajar Mountains developed.

2. To temporally correlate the uplift of the mountains with a tectonic event that can plau- sibly explain the uplift. This could either include ophiolite obduction or the Zagros collision. However, if neither of these fit, then a new tectonic model will be proposed.

3 Methodology

Three different approaches were taken to answer the aims set out in this thesis. First, low temperature thermochronology was used to date the age of cooling. This, combined with stratigraphic evidence, was used to interpret the timing of erosion due to uplift of the Al Hajar Mountains (Paper I). Second, the age of the brittle structures associated with uplift in the central mountains were directly dated. This was carried out by U-Pb dating of calcite slickenfibers that formed on the structures during deformation (Paper II). Third, a balanced geological model of the Jabal Hafit anticline, located in the foreland of the Al Hajar Mountains, was restored. This anticline is important as it consists of the most complete succession of Cenozoic sedimentary rocks found in the Al Hajar Mountains. Therefore, restoration is well constrained and the syn-deformation sedimentary rocks can be used to constrain the timing of foreland deformation. If deformation of the mountains is in-sequence (foreland propagating), then deformation of Jabal Hafit should be subsequent to the uplift of the central Al Hajar Mountains, and therefore will provide a minimum age of the orogeny (Paper III). Cryptic Orogeny: uplift of the Al Hajar Mountains at an alleged passive margin 5

3.1 Paper I: Low-Temperature Thermochronology Four low temperature thermochronometers were used in the interpretation of the uplift history of the central Al Hajar Mountains. We carried out 15 apatite fission track, 10 apatite (U-Th)/He, and 4 zircon (U-Th)/He sample age analyses. These data were combined with the literature which included 16 apatite fission track and 17 zircon fission track ages. Apatite and zircon was first separated from 15 rock samples following the guidelines set out in Donelick et al. (2005). This involved crushing and milling the hand samples to a 300 µm grain size. Next, the samples were processed by a Wilfley table, sieving, Frantz magnetic separation, and two heavy liquids to separate out the apatite and zircon.

Apatite Fission Track Dating Spontaneous fission tracks occur naturally in uranium bearing minerals such as apatite and zircon. These tracks are the result of fission decay of 238U within a crystal. During fission a heavy 238U nucleus splits into two lighter nuclei that travel in opposite directions through the crystal lattice. This creates a damaged zone called a latent fission track. Spontaneous fission decay of 238U occurs at a constant rate in nature (Fleisher et al., 1975) and therefore the tracks produced in a crystal are a function of time and 238U concentration. For an apatite grain a radiometric age calculation can be made based on the concentration of 238U (parent nuclide), the number of fission tracks (product of daughter nuclides), and the decay constant of the parent nuclide (Tagami and O’Sullivan, 2005; Donelick et al., 2005). To measure the 238U content and the spontaneous fission tracks to calculate a radiometric age, the external detector method is applied (Gleadow, 1981). In this approach (Fig.4) the apatite grains are mounted into an epoxy which is then polished to expose the internal crystal surface. Because the latent fission tracks are too small to be seen with an optical microscope, the mounts are submerged into acid. This chemically etches and enlarges the spontaneous fission tracks, which dissolve faster than the undamaged host crystal lattice. Only the tracks that intersect the polished surface are exposed and will be etched. Next, the uranium concentration of the grain is measured by attaching a uranium-low/free muscovite sheet (the external detector) to the polished surface of the apatite crystal. The mount and detector are then irradiated in a nuclear reactor and are bombarded by thermal neutrons which artificially induces fission decay of 235U in the apatite grain. The new, induced fission tracks form in the apatite grain, and 235U decaying near the polished surface will also create tracks that penetrate into the external detector. The external detector is then chemically etched so that both the spontaneous and induced fission tracks (over a fixed area) can be counted using an optical microscope, to determine relative track densities (Fig.5). The 235U/238U is a known constant ratio in nature, thus the number (or density) of induced 235U-sourced fission tracks provides a measure of parent nuclide 238U concentration. 6 Summary–Kappa

Figure 4. Steps (a) through (f) schematically detailing the external detector method (adapted from Tagami and O’Sullivan, 2005).

Figure 5. (a) An optical microscope image of a polished apatite grain (from Oman) after etching of the spontaneous fission tracks. (b) The external detector (muscovite sheet) with etched induced tracks from grain shown in (a). Cryptic Orogeny: uplift of the Al Hajar Mountains at an alleged passive margin 7

An age is calculated from the ratio between the density of spontaneous versus induced fission tracks. However, this age is not a crystallization age, but a cooling age of the apatite grain. This is because spontaneous fission tracks that form within an apatite crystal become fully or partially annealed (Green et al., 1986). This is a process where new tracks which are typically 16 µm long (Gleadow et al., 1986) will progressively shorten and eventually disappear from the crystal lattice. Annealing of tracks is primarily controlled by temperature and time. Over geological time, if an apatite is at temperatures greater than c. 120◦C any fission tracks that form will be rapidly annealed, but when the apatite is below c. 70◦C the tracks will be retained (Ketcham et al., 1999, 2015; Laslett et al., 1987; Ketcham et al., 2007). The range from 70–120◦C is the partial annealing zone (PAZ), and for zircon the PAZ is estimated to be c. 230–350◦C (Bernet, 2009). Within the PAZ, tracks are gradually annealed, and the amount of track-shortening is dependent on how long the apatite remains in this temperature range. Because of this, the track lengths are also measured as they indicate the rate of cooling (Gleadow et al., 2002). This includes: (1) long tracks which indicate little to no annealing and therefore less time in the PAZ and rapid cooling, (2) short tracks which mean more time spent in the PAZ and slower cooling, and (3) mixed track lengths which reveal one or more reheating events and a multiphase thermal history. Only confined tracks can be used for track length analysis, and are only observable if they are also etched. This means the confined tracks must intersect a host track or a cleavage (Fig.6). Therefore, apatite fission track dating provides a cooling age based on the spontaneous and induced tracks, as well as information on the cooling rate from the track lengths distribution.

Figure 6. (a) Diagram of a polished and etched mineral showing how spontaneous confined tracks will become etched (reprinted from Tagami and O’Sullivan, 2005). (b) An optical microscope image of an apatite grain from Oman showing a confined track that is etched because it is intersected by a host track. 8 Summary–Kappa

(U-Th)/He Dating

This dating method utilizes alpha decay (α-decay) of 238U, 235U, 232Th, and 147Sm found in zircon and apatite (Farley, 2002; Harrison and Zeitler, 2005). Similar to fission, α-decay is a spontaneous event when a heavy U, Th, or Sm nucleus emit helium isotopes (4He or α- particles). This occurs naturally at an observable constant rate (Hourigan et al., 2005). When a zircon or apatite crystallizes, there is no 4He in its lattice. However, over geologic time as α- decay occurs, 4He will accumulate within the crystal structure. As with the PAZ for fission track dating, (U-Th)/He dating has a similar temperature control termed the partial retention zone (PRZ). At higher temperatures 4He will rapidly diffuse and escape from the crystal lattice. Conversely at low temperatures 4He will not diffuse and will be retained within the grain (Ketcham et al., 2011; Guenthner et al., 2013; Flowers et al., 2009). For apatite the PRZ is c. 55–80◦C (Farley, 2002), and for zircon it is c. 160–200◦C (Reiners et al., 2004). Therefore, an age calculation of a crystal can be made using the decay rate and the measurements of the parent isotopes (238U, 235U, 232Th, and 147Sm) and the daughter products (4He) from α-decay. This is a cooling age as it will date the time at which the crystal passed through the PRZ. To measure the parent and daughter isotopes a zircon or apatite grain is packed into a niobium tube (Fig.7) and is then heated using a laser. This forces diffusion of the 4He (daughter product) which will exit the grain by degassing. Degassing occurs in a vacuum chamber and the gases released from the sample are then spiked with 3He for measurement by isotope dilution with a quadrupole mass spectrometer. The measured isotope ratio is used to calculate the original 4He concentration of the crystal. After 4He degassing the parent isotope concentrations (238U, 235U, 232Th, and 147Sm) are also measured by isotope dilution. This is carried out by dissolving the crystal in acid and adding 232U– 229Th and 147Sm spikes. Isotope ratio measurements are carried out by high-resolution inductively coupled plasma– mass spectrometry (Reiners et al., 2004; Reiners, 2005).

Figure 7. (a) Optical microscope image of an apatite grain from Oman used for (U-Th)/He dating. (b) A packed Nb tube with crimped ends ready for degassing. Cryptic Orogeny: uplift of the Al Hajar Mountains at an alleged passive margin 9

Cooling Ages and Orogeny Cooling ages from thermochronology can be used to indirectly estimate paleotopography and its temporal evolution. The apatite and zircon fission track and (U-Th)/He thermochronome- ters constrain the cooling history from c. 350◦C through to c. 55◦C. To cool a rock it needs to be exhumed to the surface. One way to achieve this is during mountain building (Reiners and Brandon, 2006; Reiners, 2007), where fold and thrust belts create high and unstable relief which is rapidly eroded and deposited into adjacent basins. Collecting a sample at the surface of a mountain and analyzing it with the four thermochronometers (mineral-pair method), will date the cooling history of that mountain (Fig.8) that can then be used to estimate its erosion, and hence uplift history. This can be done by converting the cooling history to a crustal depth using an assumed paleogeothermal gradient. For zircon fission track to apatite (U-Th)/He thermochronometers this is equivalent to a depth range from c. 15 to c. 1 km depth. With the aid of inverse thermal history modelling with the program HeFTy (Ketcham, 2005), a thermal history that best matches the data can be identified. Estimates of timing, rates and amounts of erosion can be supported by the regional stratigraphy where the age of such rocks is known, as well as from observing clasts eroded from the core of a mountain in sedimentary rocks in an adjacent basin. For a more detailed account of this methodology and its geological applications refer to Paper I, as well as Reiners and Ehlers(2005) and Reiners et al.(2017).

Figure 8. A basic schematic showing the link between low-temperature thermochronology cooling ages and exhumation that is driven by orogenic erosion over time. (a) A fold and thrust belt being eroded and exhuming a rock sample bearing apatite and zircon. (b) Time-temperature path defined by four thermochronometers from one sample (mineral-pair method). This gives the timing and rate of cooling which is interpreted to indicate the age of the orogenic event. 10 Summary–Kappa

3.2 Paper II: U-Pb Calcite Dating Calcite slickenfibers grow during faulting in the brittle crust. These fibers can be dated to constrain the timing of brittle deformation. Also, the direction at which the calcite fibers grow will indicate the kinematics of the fault. Reverse and strike-slip faults associated to NNE– SSW shortening was targeted for sampling, as they share the same kinematics as the large scale folds that form the Al Hajar Mountain range. Calcite slickenfibers (Fig.9a) on these faults were made into thick sections (Fig.9b) and were analyzed by laser ablation-inductively coupled plasma-mass spectrometry (LA-ICP-MS) U-Pb geochronology. A Thermo Scientific Element 2 Sector-Field ICP-MS was used with a detection limit for 206Pb and 238U at c. 0.2 and c. 0.03 ppb, respectively. From the Al Hajar Mountains 11 thick sections were made and 22 U-Pb ages were calculated. Ages from shortening structures were used to constrain the timing at which the central mountains were being deformed. A modified U-Pb dating method was used and is described in Gerdes and Zeh(2006, 2009). U-Pb is a robust geochronology technique as it relies on two independent decay chains with different constant decay rates: 238U to 206Pb and 235U to 207Pb. Analyzing a calcite crystal in a closed system will result in both decay schemes giving the same age. Due to the different decay rates and that the 235U/238U ratio is a known constant in terrestrial rocks, the 207Pb/206Pb ratio is a function of time. This is represented on a Tera-Wasserburg concordia diagram (Tera and Wasserburg, 1972) which plots the parent/daughter ratio ( 238U/206Pb) and the Pb isotopic ratio (207Pb/206Pb) on the x- and y-axes respectively (Fig.9c and d). On a Tera-Wasserburg diagram, in a closed system with no initial common Pb, radiogenic Pb will have distinct 207Pb/206Pb and 238U/206Pb ratios at any given point in time and will plot on the concordia curve. This provides a concordant age for the calcite crystallization. If there is common Pb in the calcite at crystallization, then repeated analyses throughout the sample (that have varying amounts of Pb isotope ratios) will plot on a straight line (a discordia line). The initial 207Pb/206Pb ratio can then be calculated where the discordia line intersects the y-intercept. The age of calcite crystallization is then given by the interception of the discordia line with the concordia curve (lower isochron intercept). In an open system, where a calcite sample has been disturbed by geological processes (e.g. diagenesis), daughter and/or parent isotopes may be added or lost from the calcite crystal. This will result in a discrepancy between the two decay chain ages and multiple analyses of a disturbed calcite crystal will not plot on the concordia curve (providing that it does not contain common Pb). Data from an open system will not define a discordia line, as the scatter will be too large. The mean square weighted deviation (MSWD) defines how well the data is aligned. A MSWD higher than c. 2.5 indicates that the scatter is due to open system behavior and cannot be accounted for by analytical error alone (Rasbury and Cole, 2009; Brooks et al., 1972). Therefore, the sample does not give any geologically meaningful ages, such as calcite crystallization, and should be discarded. Cryptic Orogeny: uplift of the Al Hajar Mountains at an alleged passive margin 11

Figure 9. (a) An example of calcite slickenfibers that were sampled from a fault surface from the Al Hajar Mountains in Oman. Growth of fibers indicate missing block moved left (white arrow) and are kinematically correlated to the formation of the Al Hajar Mountains. (b) A polished thick section of calcite fibers and location of LA-ICP-MS spot analyses. (c) Results of spot analyses plotted in a Tera- Wasserburg concordia diagram and age calculated as a lower isochron intercept. The MSWD is low, indicating that the scatter in the data is not due to a geological disturbance or a mixed age, and is only caused by analytical issues. This data can confidently be interpreted to indicate a crystallization age. (d) A labelled Tera-Wasserburg concordia diagram. The concordia curve applies to a closed system with no common Pb. Analyses of a closed system with initial Pb will plot on a discordia line, where the y-intercept defines the initial Pb isotope ratio and the lower isochrone intercept defines a crystallization age (where common Pb is 0%). 12 Summary–Kappa

3.3 Paper III: Three-Dimensional Geological Modelling The purpose of creating a three-dimensional geological model is to predict the unknown geometry of a geological structure (such as a blind thrust) based on the known geometry (such as the folded strata above a blind thrust). Chamberlin(1910) is considered the first study to do this, where the Appalachian lower detachment was estimated using restored cross- sections. However, it was Dahlstrom(1969) who introduced the term "balanced cross-section" and outlined a methodology of doing so. This involves maintaining a constant rock area or bed length between the deformed and restored cross-sections. To create balanced cross-sections field mapping was carried out to constrain the surface geology of Jabal Hafit, a large-scale anticline in the foreland of the Al Hajar Mountains. This an- ticline comprises carbonate rocks that were deposited in horizontal layers before deformation. Mapping included using an unmanned aerial vehicle to photograph the anticline. These im- ages were then processed with structure-from-motion software to create a three-dimensional terrain model from which structures (dip and dip-direction) could be directly measured from (Bemis et al., 2014). From these data, combined with a well log and 7 seismic profiles, 10 area balanced, two-dimensional cross-sections were made.

Figure 10. (a) Retrodeformable cross-section of a fault-propagation fold (adapted from Suppe and Medwedeff, 1990). (b) A retrodeformed (restored) cross-section of (a), maintaining a constant area. This restores the stratigraphic layers to their original depositional orientation. The geometry of the anticline can be used to constrain the dip and dip-direction of a blind fault. When the bed thickness on the anticlines backlimb is constant through deformation it will have the same dip as the fault. The depth to the déollement can also be predicted, as well as the amount of horizontal shortening across the restored section. (c) Illustration of plane strain, where no rock mass has moved along the Y-axis, and all deformation can be examined on the X–Z plane. This is a suitable assumption for creating balanced cross-sections in fold and thrust belts. Cryptic Orogeny: uplift of the Al Hajar Mountains at an alleged passive margin 13

Balancing first involved drafting a section of the present deformed state of the anticline. The section was then retrodeformed using Midland Valley’s Move software, restoring the faults and folds to an undeformed state. Balancing was achieved by maintaining a constant area of the rock layers between the deformed and the restored sections. This requires that there are no gaps or overlaps of rock layers in the restored section (Fig. 10). The result is a geometrically plausible interpretation of the subsurface geology, but this does not guarantee a unique solution. The assumption with area balancing is that the rock area is maintained, meaning that there is no volume change (no change in density) and that there is no strain perpendicular to the section (plane strain, Fig. 10c). Therefore, this balancing method would not be appropriate to restore deformation across major strike-slip faults. However, it is well suited to restoring fold and thrust belts, such as observed in the Al Hajar Mountains. In a fold and thrust belt such as at Jabal Hafit, the geometry of the folds at the surface are controlled by the blind thrusts in the subsurface. Several models can explain the relationship between the fold and the thrust. Such as kink-band (Fig. 1a) fault-propagation folds (Suppe and Medwedeff, 1990) and trishear fault-propagation folds (Erslev, 1991). These structures form by gradual propagation of a thrust at depth, where strata above the fault tip are folded (Fig. 11). Based on these models the surface data can predict the depth and orientation of the thrust, and the amount of horizontal shortening that has occurred across the fault. In addition to predicting the geometry of a structure with balanced cross-sections, the timing of deformation can also be constrained. At Jabal Hafit there is a succession of syn- tectonic strata, where the older sedimentary rocks are more deformed than the younger rocks. As the carbonates are dated through fossil assemblages they can be used to understand the age of deformation. To help interpret the timing of deformation forward models were used. These are cross-sections starting with the undeformed horizontal carbonate rocks that were then progressively deformed by horizontal shortening with fault-propagation fold algorithms using Move (Midland Valley Ltd). After the balanced cross-sections were completed, they were connected to create a three- dimensional model of the Jabal Hafit anticline. This exposed new geometric problems that were then resolved in the cross-sections. The final model predicts the subsurface geometry of the anticline, the kinematics of the structures, and the timing of the thrust and fold. 14 Summary–Kappa

Figure 11. (a) An example of a small scale fault-propagation fold from the Al Hajar Mountains, and (b) an interpretation of the structure.

4 Summary of Key Results

4.1 Paper I: AGU Journal Tectonics

Title: Late Eocene Uplift of the Al Hajar Mountains, Oman, Supported by Stratigraphy and Low- Temperature Thermochronology.

Paper I represents the bulk work of this thesis, providing an age and causation for the Al Hajar Mountains. This research expanded a sparse dataset of apatite fission track ages from the central mountains and added new apatite and zircon (U-Th)/He ages. This paper also summarizes the stratigraphy of the area to help interpret the cooling ages. While the fission track results are comparable to previous work the interpretation of the data, with the new (U- Th)/He ages, is different. The low-temperature thermochronology data indicate a significantly different cooling history along the length of the mountain range, between the Jabal Akhdar and Saih Hatat culminations (Fig. 12). Cryptic Orogeny: uplift of the Al Hajar Mountains at an alleged passive margin 15

Figure 12. A concluding schematic map showing the propaga- tion of deformation into the foreland, based on low temperature thermochronological data and stratigraphic evidence. The en- tire event lasted from the late Eocene until the early–middle Miocene (adapted from Hansman et al., 2017).

The thermal history of the rocks at the Saih Hatat culmination is unique compared to the rest of the mountain range. This is because these rocks were subducted when obduction began at c. 95 Ma (Hacker, 1994). During this event the leading edge of the Arabian Plate subducted to depths of about 80 km by 79 Ma (Warren et al., 2003). These rocks were then exhumed to less than 3 km depth by c. 55 Ma, based on zircon fission track, zircon (U-Th)/He, and apatite fission track ages. This exhumation was buoyancy-driven because of slab break-off of the eclogitized root of the subducting slab crust (Searle et al., 2004). Rock moved back up the relic subduction zone as a tectonic wedge, which was bounded by a normal fault at the top and a reverse fault at the base. By 40 Ma, the Saih Hatat culmination had eroded level, with sedimentary rocks onlapping the flanks of the culmination. This thermal history did not occur at the Jabal Akhdar culmination, as it was further southwest and away from the subduction zone. Instead the rocks here were buried about 10 km by the advancing ophiolite that was being obducted until about 70 Ma (Searle et al., 2004). The topography formed by the obduction event was eroded, followed by sedimentation that onlapped onto the culmination from c. 65 Ma to c. 40 Ma. 16 Summary–Kappa

From 40 Ma to 30 Ma major cooling occurred at the Jabal Akhdar culmination, which is interpreted with sedimentary evidence to constrain the timing at which Jabal Akhdar was uplifted by 4–6 km, and Saih Hatat uplifted by 2 km. After 30 Ma, both culminations were not significantly uplifted as the deformation propagated to the southwest, where the foreland fold and thrust belts formed. Foreland deformation ceased by c. 15 Ma, after the Oligocene to early Miocene sediments that had deposited on the flanks of the culminations were uplifted. Finally, Pliocene marine terraces (Wyns et al., 1992) on the coast of Oman were uplifted c. 200 m above sea level. This uplift can be explained by a forebulge developing in the Arabian Plate, generated by the Makran subduction zone during the Pliocene and Pleistocene (Rodgers and Gunatilaka, 2002). The conclusion from Paper I is that the building of the Al Hajar Mountains began in the late Eocene and was completed by the early–middle Miocene. This timing indicates that the orogenic event was not associated to Late Cretaceous obduction or Miocene Zagros collision. These results reveal that the tectonic history of the northeast corner of the Arabian Plate is significantly more intricate than previously thought. Therefore, a new tectonic model is presented that attempts to explain the uplift of the Al Hajar Mountains. The model proposes that as the Arabian Plate moved north, subducting under Eurasia, the Makran subduction zone slowed but Arabia’s motion did not change. The convergence was then taken up by thrusting the Arabian continental crust over the oceanic crust in the Gulf of Oman. This caused folding and the formation of the Al Hajar Mountains above a backthrust propagating from a major décollement. As the crust thickened it progressively became mechanically stronger, and full convergence was then resumed at the Makran subduction zone.

4.2 Paper II: GSA Journal Geology Title: Absolute ages of multiple generations of brittle structures by U-Pb dating of calcite.

This paper builds on a previous structural study which documented nine events in the central mountains (Gomez-Rivas et al., 2014). The separate events were classified by Gomez-Rivas et al.(2014) based on cross-cutting relationships of structures with different kinematics that formed during each event. However, this only provided a relative timing for each event which was interpreted to have occurred from the Late Cretaceous through to the Neogene. In Paper II, calcite veins and slickenfibers from structures were sampled and dated with a modified U-Pb method. This is an innovative technique with considerable potential and was only first applied to directly date brittle structures by Ring and Gerdes(2016). The results from this paper provide 22 ages for diagenesis, ophiolite obduction, and several shortening events (Fig. 13). The results that are critical for the uplift phase of the Al Hajar Mountains are events 5 – 8. Calcite U-Pb ages from structures related to these events provide absolute constraints on the timing of horizontal shortening, which resulted in crustal thickening and the formation of the Al Hajar Mountains. These ages indicate that uplift initiated in the late Eocene and had finished by the early Miocene (c. 40 – 15 Ma). This new dataset helps to constrain the tectonic history of the Al Hajar Mountains in a way that has not been possible before. Cryptic Orogeny: uplift of the Al Hajar Mountains at an alleged passive margin 17

Figure 13. U-Pb calcite ages (n=22) from the central Al Hajar Mountains with 2-σ errors. These data provide absolute ages of 7 out of the 9 documented geological events, which includes diagenesis and brittle deformation. Structures related to events 5 to 8 formed due to horizontal shortening and crustal thickening and are used to constrain the uplift age of the Al Hajar Mountains (adapted from Hansman et al., 2018).

4.3 Paper III: Manuscript Title: Oligocene–Miocene trishear fault-propagation folding of the Jabal Hafit Anticline, supported by a three-dimensional geological model; and assessing structure-from-motion (SfM) photogram- metry of unmanned aerial vehicle (UAV) photographs for mapping.

A three-dimensional geological model based on 10 area-balanced cross-sections was created for the Jabal Hafit anticline. Prior to this study, only one published article using three balanced cross-sections over the anticline was carried out (Noweir, 2000). These three sections were located in the northern half of the anticline and available seismic profiles were not used in their construction. Paper III presents a new model that best fits all the available data, and includes a forward model used to constrain the timing of deformation (Fig. 14). Results from Paper III show that the geometry of the anticline can be described by a tris- hear fault-propagation fold. This was developed by a west-dipping backthrust that propagated 18 Summary–Kappa

from a major decollement, due to WSW-directed shortening. The Jabal Hafit anticline formed in the foreland of the Al Hajar Mountains, and therefore records the final stages of uplift in the mountain range (Fig. 15). Based on the geometry of pre- and syn-tectonic sedimentary rocks the formation of the anticline is interpreted to have initiated in the late Oligocene and had ceased by the early/middle Miocene (c. 28 – 15 Ma). These results constrain the timing of the uplift in the central mountains, which must be older than the late Oligocene.

Figure 14. A forward model of the Jabal Hafit anticline, constrained by the balanced geological model (adapted from Paper III). (a) End of deposition of pre-tectonic sedimentary rocks in the middle Oligocene. (b) Horizontal shortening and fault-propagation begins in the late Oligocene. (c) Deforma- tion continues, with deposition of syn-tectonic sedimentary rocks during the early and middle Miocene. The anticline developed from c. 28 Ma to 15 Ma, which provides a minimum age for the central Al Hajar Mountains which must be older than c. 28 Ma. Cryptic Orogeny: uplift of the Al Hajar Mountains at an alleged passive margin 19

Figure 15. A schematic of the Al Hajar Mountains showing the propagation of deforma- tion with a younging direction towards the WSW. The uplift of the central mountains must occur before the folding of the Jabal Hafit anticline.

5 Conclusion and Future Work

The work presented in this thesis aimed to better understand the age of the Al Hajar Mountains in northern Oman. This question was tackled using a multidisciplinary approach. In Paper I, low-temperature thermochronology dated the time of cooling, which occurred during uplift. Results indicate that in the central mountains a major event occurred during 40 Ma – 30 Ma. For Paper II, U-Pb dating of calcite slickenfibers shows that horizontal shortening took place from 40 Ma until 15 Ma. Finally, in Paper III the restoration of syn-tectonic sedimentary rocks in the foreland of the mountain range indicate that the fold and thrust belt was active from 28 Ma through to 15 Ma. This provides a minimum age for the central mountains which must have been deformed and uplifted before deformation in the foreland. These results establish that the age of the Al Hajar Mountains is late Eocene to early Miocene (Fig. 16). This temporally excludes ophiolite obduction or Zagros collision as the cause for the uplift. Therefore, in Paper I a new tectonic model is suggested, which can explain the uplift history of the Al Hajar Mountains. This thesis puts forward a completely new tectonic history for the Cenozoic of northern Oman. Consequently, the conclusions made in this work require future scrutiny by replicat- ing the results and applying new methodologies. The tectonic model proposes that there is currently no horizontal shortening occurring in the mountains, and this could be directly measured using GPS stations that record horizontal and vertical movement, which is work that we are currently conducting. Additionally, the sedimentary rocks in the Gulf of Oman should contain a complete record during the Cenozoic which has not been eroded. The majority of the cores of the Jabal Akhdar and Saih Hatat culminations have eroded to the north, into the Gulf. Seismic surveys and well data will provide the data required to further constrain the uplift of the mountains. Finally, the new tectonic model presented in Paper I predicts that the Semail Ophiolite and the oceanic crust in the Gulf of Oman are separated by a major structural boundary. This can be tested by deeper seismic data acquisition. 20 Summary–Kappa

Figure 16. A simplified timeline of the sedimentary and tectonic events and results from Papers I, II, and III. Data indicate that the formation of the present day topography of the Al Hajar Mountains is unrelated to obduction or Zagros collision. This means that the tectonic history of the mountains is more complex than what has previously been claimed. Therefore, a new tectonic model is required to explain the history of this orogeny, and is presented in Paper I.

Tectonostratigraphic Summary of the Central Al Hajar Mountains: Figure on previous page, based on Alsharhan and Nairn(1990); Bechennec et al.(1990); Bernoulli and Weissert (1987); Cherif et al.(1992); Clarke(1988); Cooper(1987); Forbes et al.(2010); Glennie et al.(1973, 1974); Graham(1980); Mann et al.(1990); Nolan et al.(1990); Robertson (1987); Robertson and Searle(1990); Robertson et al.(1990); Rollinson(2009); Rollinson et al.(2014); Skelton et al.(1990).

22 References

Alsharhan, A. and Nairn, A. (1990). A review of the Cretaceous formations in the and Gulf: Part I. Lower Cretaceous (Thamama Group) stratigraphy and paleogeography. Journal of Petroleum Geology, 9(4):365–391. ArRajehi, A., McClusky, S., Reilinger, R., Daoud, M., Alchalbi, A., Ergintav, S., Gomez, F., Sholan, J., Bou-Rabee, F., Ogubazghi, G., Haileab, B., Fisseha, S., Asfaw, L., Mahmoud, S., Rayan, A., Bendik, R., and Kogan, L. (2010). Geodetic constraints on present-day motion of the Arabian Plate: Implications for Red Sea and Gulf of Aden riftingArRajehi, A., McClusky, S., Reilinger, R., Daoud, M., Alchalbi, A., Ergintav, S., Gomez, F., Sholan, J., Bou-Rabee, F., Ogubazghi, G., Hailea. Tectonics, 29(3):1–10. Bechennec, F., Le Metour, J., Rabu, D., Bourdillon-de Grissac, C., de Wever, P., Beurrier, M., and Villey, M. (1990). The Hawasina Nappes: stratigraphy, palaeogeography and structural evolution of a fragment of the south-Tethyan passive continental margin. Geological Society, London, Special Publications, 49(1):213–223. Bemis, S. P., Micklethwaite, S., Turner, D., James, M. R., Akciz, S., T. Thiele, S., and Bangash, H. A. (2014). Ground-based and UAV-Based photogrammetry: A multi-scale, high-resolution mapping tool for structural geology and paleoseismology. Journal of Structural Geology, 69(PA):163–178. Bernet, M. (2009). A field-based estimate of the zircon fission-track closure temperature. Chemical Geology, 259(3-4):181–189. Bernoulli, D. and Weissert, H. (1987). The Upper Hawasina Nappes in the Central Oman Mountains: Stratigraphy, palinspastics and sequence of nappe emplacement. Geodinamica Acta, 1(1):47–58. Brooks, C., Hart, S. R., and Wendt, I. (1972). Realistic use of two-error regression treatments as applied to rubidium-strontium data. Reviews of Geophysics, 10(2):551–577. Chamberlin, R. T. (1910). The Appalachian Folds of Central Pennsylvania. The Journal of Geology, 18(3):228–251. Cherif, O. H., Al-Rifaiy, I. A., and El-Deeb, W. Z. M. (1992). “Post-Nappes” early Tertiary foraminiferal paleoecology of the northern Hafit area, south of Al-Ain City (United Arab Emirates). Micropaleon- tology, 38(1):37–56. Clarke, M. W. H. (1988). Stratigraphy and rock unit nomenclature in the oil-producing area of interior Oman. Journal of Petroleum Geology, 11(1):5–60. Cooper, D. J. W. (1987). Hamrat Duru Group: revised stratigraphy of a Mesozoic deep-water passive margin in the Oman Mountains. Geological Magazine, 124(02):157–164. Dahlstrom, C. D. A. (1969). Balanced cross sections. Canadian Journal of Earth Sciences, 6(4):743– 757. Donelick, R. A., O’Sullivan, P. B., and Ketcham, R. A. (2005). Apatite fission-track analysis. Reviews in Mineralogy and Geochemistry, 58(1):49–94. Erslev, E. A. (1991). Trishear fault-propagation folding. Geology, 19(6):617–620. Farley, K. A. (2002). (U-Th)/He dating: techniques, calibrations, and applications. Reviews in Mineralogy and Geochemistry, 47(1):819–844. Fleisher, R. L., Price, P. B., and Walker, R. M. (1975). Nuclear tracks in solids: principles and applications. Berkeley, Calif., University of California Press, 1975. 626 p., 1. Flowers, R. M., Ketcham, R. A., Shuster, D. L., and Farley, K. A. (2009). Apatite (U-Th)/He ther- mochronometry using a radiation damage accumulation and annealing model. Geochimica et Cos- mochimica Acta, 73(8):2347–2365. Forbes, G. A., Jansen, H. S. M., Schreurs, J., and Naft ’Uman, S. T. (2010). Lexicon of Oman: Subsurface Stratigraphy: Reference Guide to the Stratigraphy of Oman’s Hydrocarbon Basins. Gulf PetroLink.

23 24 Summary–Kappa

Gavillot, Y., Axen, G. J., Stockli, D. F., Horton, B. K., and Fakhari, M. D. (2010). Timing of thrust activity in the High Zagros Fold-Thrust Belt, Iran, from (U-Th)/He thermochronometry. Tectonics, 29(4). Gerdes, A. and Zeh, A. (2006). Combined U–Pb and Hf isotope LA-(MC-)ICP-MS analyses of detrital zircons: comparison with SHRIMP and new constraints for the provenance and age of an Armorican metasediment in Central Germany. Earth and Planetary Science Letters, 249(1):47–61. Gerdes, A. and Zeh, A. (2009). Zircon formation versus zircon alteration – new insights from combined U–Pb and Lu–Hf in-situ LA-ICP-MS analyses, and consequences for the interpretation of Archean zircon from the Central Zone of the Limpopo Belt. Chemical Geology, 261(3):230–243. Gleadow, A. J. W. (1981). Fission-track dating methods: what are the real alternatives? Nuclear Tracks, 5(1-2):3–14. Gleadow, A. J. W., Belton, D. X., Kohn, B. P., and Brown, R. W. (2002). Fission track dating of phosphate minerals and the thermochronology of apatite. Reviews in Mineralogy and Geochemistry, 48(1):579–630. Gleadow, A. J. W., Duddy, I. R., Green, P. F., and Lovering, J. F. (1986). Confined fission track lengths in apatite: a diagnostic tool for thermal history analysis. Contributions to Mineralogy and Petrology, 94(4):405–415. Glennie, K. W., Boeuf, M. G. A., Clarke, M. W. H., Moody-Stuart, M., Pilaar, W. F. H., and Reinhardt, B. M. (1973). Late Cretaceous nappes in Oman Mountains and their geologic evolution. AAPG Bulletin, 57(1):5–27. Glennie, K. W., Boeuf, M. G. A., Hughes-Clarke, M. W., Moody-Stuart, M., Pilaar, W., and Reinhardt, B. (1974). Geology of the Oman Mountains, Parts I, II, III, volume 31. Verhandelingen Koninklijk Nederlands Geologisch Mijnbouwkundidg Genootschap, Amsterdam. Gomez-Rivas, E., Bons, P. D., Koehn, D., Urai, J. L., Arndt, M., Virgo, S., Laurich, B., Zeeb, C., Stark, L., and Blum, P. (2014). The Jabal Akhdar dome in the Oman Mountains: evolution of a dynamic fracture system. American Journal of Science, 314(7):1104–1139. Graham, G. M. (1980). Structure and sedimentology of the Hawasina Window, Oman Mountains: Evolution of a passive continental margin, and emplacement of the Oman thrust belt. Phd, Open University, UK. Green, P. F., Duddy, I. R., Gleadow, A. J. W., Tingate, P. R., and Laslett, G. M. (1986). Thermal annealing of fission tracks in apatite 1. A qualitative description. Chemical Geology: Isotope Geoscience section, 59:237–253. Guenthner, W. R., Reiners, P. W., Ketcham, R. A., Nasdala, L., and Giester, G. (2013). Helium diffusion in natural zircon: radiation damage, anisotropy, and the interpretation of zircon (U-Th)/He thermochronology. American Journal of Science, 313(3):145–198. Hacker, B. R. (1994). Rapid emplacement of young oceanic lithosphere: argon geochronology of the Oman Ophiolite. Science (New York, N.Y.), 265:1563–1565. Hansman, R. J., Albert, R., Gerdes, A., and Ring, U. (2018). Absolute ages of multiple generations of brittle structures by U-Pb dating of calcite. Geology, 46(3):207–210. Hansman, R. J., Ring, U., Thomson, S. N., Brok, B., and Stübner, K. (2017). Late Eocene uplift of the Al Hajar Mountains, Oman, supported by stratigraphy and low-temperature thermochronology. Tectonics, 36(12):3081–3109. Harrison, T. M. and Zeitler, P. K. (2005). Fundamentals of noble gas thermochronometry. Reviews in mineralogy and geochemistry, 58(1):123–149. Hessami, K., Koyi, H., Talbot, C. J., Tabasi, H., and Shabanian, E. (2001). Progressive unconformities within an evolving foreland fold-thrust belt, Zagros Mountains. Journal of the Geological Society, 158:969–981. Hourigan, J. K., Reiners, P. W., and Brandon, M. T. (2005). U-Th zonation-dependent alpha-ejection in (U-Th)/He chronometry. Geochimica et Cosmochimica Acta, 69(13):3349–3365. Cryptic Orogeny: uplift of the Al Hajar Mountains at an alleged passive margin 25

Ketcham, R. A. (2005). Forward and inverse modeling of low-temperature thermochronometry data. Reviews in Mineralogy and Geochemistry, 58(1):275–314. Ketcham, R. A., Carter, A., Donelick, R. A., Barbarand, J., and Hurford, A. J. (2007). Improved modeling of fission-track annealing in apatite. American Mineralogist, 92(5-6):799–810. Ketcham, R. A., Carter, A., and Hurford, A. J. (2015). Inter-laboratory comparison of fission track confined length and etch figure measurements in apatite. American Mineralogist, 100(7):1452– 1468. Ketcham, R. A., Donelick, R. A., and Carlson, W. D. (1999). Variability of apatite fission-track annealing kinetics: III. Extrapolation to geological time scales. American Mineralogist, 84:1235–1255. Ketcham, R. A., Gautheron, C., and Tassan-Got, L. (2011). Accounting for long alpha-particle stop- ping distances in (U–Th–Sm)/He geochronology: refinement of the baseline case. Geochimica et Cosmochimica Acta, 75(24):7779–7791. Laslett, G., Green, P., Duddy, I., and Gleadow, A. (1987). Thermal annealing of fission tracks in apatite 2. A quantitative analysis. Chemical Geology: Isotope Geoscience section, 65(1):1–13. Mann, A., Hanna, S. S., and Nolan, S. C. (1990). The post-Campanian tectonic evolution of the Central Oman Mountains: Tertiary extension of the Eastern Arabian Margin. The Geology and Tectonics of the Oman Region, 49(1):549–563. McQuarrie, N. and Van Hinsbergen, D. J. J. (2013). Retrodeforming the Arabia-Eurasia collision zone: age of collision versus magnitude of continental subduction. Geology, 41(3):315–318. Nolan, S. C., Skelton, P. W., Clissold, B. P., and Smewing, J. D. (1990). Maastrichtian to Early Tertiary stratigraphy and palaeogeography of the Central and Northern Oman Mountains. Geological Society, London, Special Publications, 49(1):495–519. Noweir, M. A. (2000). Back-thrust origin of the Hafit structure, Northern Oman Mountain front, United Arab Emirates. GeoArabia, 5(2):215–228. Rasbury, E. T. and Cole, J. M. (2009). Directly dating geologic events: U-Pb dating of carbonates. Reviews of Geophysics, 47(3):1–27. Reiners, P. W. (2005). Zircon (U-Th)/He thermochronometry. Reviews in Mineralogy and Geochemistry, 58(1):151–179. Reiners, P. W. (2007). Thermochronologic approaches to paleotopography. Reviews in Mineralogy and Geochemistry, 66(1):243–267. Reiners, P. W. and Brandon, M. T. (2006). Using thermochronology to understand orogenic erosion. Annual Review of Earth and Planetary Sciences, 34(1):419–466. Reiners, P. W. and Ehlers, T. A. (2005). Low-tempertaure thermochronology: techniques, interpretations, and applications, volume 58. The Miner. Soc. of America Washington, Chantilly, VA. Reiners, P. W., Renne, P. R., Carlson, R. W., Cooper, K. M., Granger, D. E., McLean, N. M., and Schoene, B. (2017). Geochronology and thermochronology. John Wiley & Sons. Reiners, P. W., Spell, T. L., Nicolescu, S., and Zanetti, K. A. (2004). Zircon (U-Th)/He thermochronom- etry: He diffusion and comparisons with 40Ar/39Ar dating. Geochimica et Cosmochimica Acta, 68(8):1857–1887. Ring, U. and Gerdes, A. (2016). Kinematics of the Alpenrhein-Bodensee graben system in the Central Alps: Oligocene/Miocene transtension due to formation of the Western Alps arc. Tectonics, 35(6):1367–1391. Rioux, M., Bowring, S., Kelemen, P., Gordon, S., Miller, R., and Dudás, F. (2013). Tectonic development of the Samail ophiolite: high-precision U-Pb zircon geochronology and Sm-Nd isotopic constraints on crustal growth and emplacement. Journal of Geophysical Research: Solid Earth, 118(5):2085–2101. Robertson, A. (1987). Upper Cretaceous Muti Formation: transition of a Mesozoic nate platform to a foreland basin in the Oman Mountains. Sedimentology, 34(6):1123–1142. Robertson, A. H. F. and Searle, M. P. (1990). The Northern Oman Tethyan continental margin: stratigraphy, structure, concepts and controversies. Geological Society, London, Special Publications, 26 Summary–Kappa

49(1):3–25. Robertson, A. H. F., Searle, M. P., and Ries, A. C. (1990). The geology and tectonics of the Oman region. Geological Society of London special publication, 49(1):845. Rodgers, D. W. and Gunatilaka, A. (2002). Bajada formation by monsoonal erosion of a subaerial forebulge, Sultanate of Oman. Sedimentary Geology, 154(3-4):127–146. Rollinson, H. (2009). New models for the genesis of plagiogranites in the Oman ophiolite. Lithos, 112(3-4):603–614. Rollinson, H., Searle, M., Abbasi, I., Al-Lazki, A., and Al-Kindi, M., editors (2014). Tectonic evolution of the Oman Mountains. Geological Society, London, Special Publications, 392 edition. Saddiqi, O., Michard, A. N., Goffe, B. R., Poupeau, G. É., and Oberhänsli, R. O. (2006). Fission-track thermochronology of the Oman Mountains continental windows , and current problems of tectonic interpretation. Bull. Soc. Géol. Fr., pages 127–134. Searle, M. P., Warren, C. J., Waters, D. J., and Parrish, R. R. (2004). Structural evolution, meta- morphism and restoration of the arabian continental margin, Saih Hatat region, Oman Mountains. Journal of Structural Geology, 26(3):451–473. Skelton, P. W., Nolan, S. C., and Scott, R. W. (1990). The Maastrichtian transgression onto the northwestern flank of the Proto-Oman Mountains: sequences of rudist-bearing beach to open shelf facies. Geological Society, London, Special Publications, 49(1):521–547. Stern, R. J. and Johnson, P. (2010). Continental lithosphere of the Arabian Plate: a geologic, petrologic, and geophysical synthesis. Earth-Science Reviews, 101(1-2):29–67. Suppe, J. and Medwedeff, D. A. (1990). Geometry and kinematics of fault-propagation folding. Eclogae Geologicae Helvetiae, 83(3):409–454. Tagami, T. and O’Sullivan, P. B. (2005). Fundamentals of fission-track thermochronology. Reviews in Mineralogy and Geochemistry, 58(1):19–47. Tera, F. and Wasserburg, G. J. (1972). U-Th-Pb systematics in three Apollo 14 basalts and the problem of initial Pb in lunar rocks. Earth and Planetary Science Letters, 14(3):281–304. Warrak, M. (1996). Origin of the Hafit structure: implications for timing the Tertiary deformation in the Northern Oman Mountains. Journal of Structural Geology, 18(6):803–818. Warren, C. J., Parrish, R. R., Searle, M. P., and Waters, D. J. (2003). Dating the subduction of the Arabian continental margin beneath the Semail Ophiolite, Oman. Geology, 31(10):889–892. Warren, C. J., Parrish, R. R., Waters, D. J., and Searle, M. P. (2005). Dating the geologic history of Oman’s Semail Ophiolite: insights from U-Pb geochronology. Contributions to Mineralogy & Petrology, 150(4):403–422. Wyns, R., Béchennec, F., Le Métour, J., and Roger, J. (1992). Explanatory notes to the geological map of Tiwi, sheet NF40-8B. Sultanate of Oman, Ministry of Petroleum and Minerals, Muscat.