Next Article in Journal
Solid Bitumen Occurrences in the Pyrenean Basin (Southern France): A Case Study across the Pliensbachian–Toarcian Boundary
Previous Article in Journal
Bio-Organic Mineral Fertilizer for Sustainable Agriculture: Current Trends and Future Perspectives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Combined Zircon/Apatite U-Pb and Fission-Track Dating by LA-ICP-MS and Its Geological Applications: An Example from the Egyptian Younger Granites

1
Geology Department, Faculty of Science, Port Said University, Port Said 42522, Egypt
2
Institute of Nature and Environmental Technology, Kanazawa University, Kanazawa 920-1192, Japan
3
Department of Food Science and Nutrition, College of Sciences, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
4
Department of Earth Sciences, Kanazawa University, Kanazawa 920-1192, Japan
*
Author to whom correspondence should be addressed.
Minerals 2021, 11(12), 1341; https://doi.org/10.3390/min11121341
Submission received: 3 October 2021 / Revised: 23 November 2021 / Accepted: 25 November 2021 / Published: 29 November 2021
(This article belongs to the Section Mineral Deposits)

Abstract

:
Laser Ablation-Inductively Coupled Plasma-Mass Spectrometry (LA-ICP-MS) is classically used in U-Pb dating to measure U and Pb isotopic concentrations. Recently, it has become frequently used in fission-track (FT) chronometry too. As an advantage, the U-Pb and FT double dating will enable efficiently determining the crystallization ages and the thermo-tectonic history concurrently as samples volume, analytical time, efforts, and cost will be greatly reduced. To demonstrate the validity of this approach, a Younger granite (Ediacaran age) sample from North Eastern Desert (NED), Egypt was analyzed for U-Pb and FT double dating. The integration of multiple geochronologic data yielded a zircon U-Pb crystallization age of 599 ± 30 Ma, after emplacement, the rock cooled /uplifted rapidly to depths of 9–14 km as response to the post-Pan African Orogeny erosional event as indicated by apatite U-Pb age of 474 ± 9 Ma. Afterwards, the area experienced a slow cooling/exhumation for a short period, most-likely as response to denudation effect. During the Devonian, the area was rapidly exhumed to reach depths of 1.5–3 km as response to the Hercynian tectonic event, as indicated by a zircon FT age of 347 ± 16 Ma. Then the studied sample has experienced a relatively long period of thermal stability between the Carboniferous and the Eocene. During the Oligocene-Miocene, the Gulf of Suez opening event affected the area by crustal uplift to its current elevation. This integration of Orogenic and thermo-tectonic information reveals the validity, efficiency, and importance of double dating of U-Pb and FT techniques using LA-ICP-MS methodology.

1. Introduction

The integration of different geochronological (e.g., U-Pb, K-Ar, Helium and FT dating) techniques provide useful information on the formation and development of the upper crust through geological time [1,2,3]. FTs are approximately linear damages, formed in the lattices of certain minerals and glasses by natural fission of heavy isotopes like U-238, U-235, Th-232 [4]. The preservation of FTs in minerals mainly depends on the temperature and cooling rate. For FTs in typical F-apatite, such as, for instance, Durango apatite, the temperature interval between 110 °C and 60 °C [5,6] is often referred to as the partial annealing zone (PAZ). So far, however, little is known about the mechanism and kinetics of FT annealing in Cl-rich apatite specimens. FTs in Cl-apatite are seemingly more resistant to thermal annealing if compared with tracks in F-apatite [7,8]. This feature can be especially observed during the dating of clastic samples which commonly contain partially reset to set apatite grains, factually Cl-rich ones, which may yield older FT ages as compared with F-apatite grains [9]. For example, for apatite with Cl of 3 wt.%, Donelick et al. [6] postulated that FTs begin to anneal at ca. 90 °C and experience total annealing at temperatures higher than 160 °C. This implies that the PAZ in Cl-rich apatite is wider and hotter if compared with the PAZ in F-apatite. For FTs in zircon, the effective closure temperature ranging 240 °C and 200 °C depending on cooling rates [10].
Partial annealing will reduce the apparent FT age and shorten the lengths of tracks, whereas total annealing will reset the FT age to zero removing all the spontaneous fission tracks. Therefore, FT lengths, combined with FT ages, can be used to reconstruct detailed thermal histories of rock samples [5,11].
The FT thermochronology is a powerful tool that may be applied to resolve several geological problems in a variety of geological settings [12,13,14,15]. The FT age is calculated by determining the number of spontaneous fission tracks (i.e., formed by daughter nuclides) and the number of U-238 atoms (i.e., parent nuclides) per weight of sample. In the conventional FT dating, the concentrations of U-238 are measured indirectly by irradiating of samples with thermal neutrons at a nuclear power reactor [3,16,17]. Currently, the FT ages may be determined rapidly and safely employing laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) technique for in-situ quantification of U-238 (ICP-MS: Yokogawa Analytical Systems, Japan; MicroLas: GeoLas Q-plus by MicroLas, Germany) [6,18]. The ability of LA-ICP-MS to measure rare earth elements simultaneously with U, Th, Sr, Y, Mn, Mg, Cl, and isotopic ratios—as done in many modern studies including this work—gives complementary chemical composition of apatite (or zircon, titanite) which may be used for simultaneous double dating of individual crystals by U–Pb and FT techniques as well as for petrogenetic, metallogenetic and provenance studies [19,20].
The U-Pb dating is based on decay of U and Th-232 isotopes to Pb isotopes. It is one of the most commonly used geochronological tools to obtain the protolith and magmatic ages of rocks [21]. The U-Pb system is extremely resistant to thermal overprinting (>900 °C; [22]) and is, therefore, unaffected by thermal events when the most suitable mineral, zircon, is analyzed. These advantages and improvements in the technology of LA-ICP-MS made it a rapid, precise and accurate method for U-Pb geochronology which has recently been extended [23,24,25,26,27,28,29,30,31,32]. The combination of U-Pb and FT methods could provide insights on history of the provenance that has undergone metamorphism. In this study, a Younger granite (Ediacaran age) sample from the North Eastern Desert (NED) of Egypt was analyzed to examine the possibility of achieving a simultaneous double dating of both U-Pb and FT by coeval measuring of all required isotopes using LA-ICP-MS with the final purpose of integrating the obtained results to investigate a comprehensive thermo-tectonic history of the studied sample since its crystallization. This sample is part of the Arabian-Nubian Shield (ANS) which represents the crust beneath north-eastern Africa, Arabia and SW Asia (Figure 1).

2. The Studied Sample

The ANS crops out in Egypt along the Red Sea coastline and in south Sinai (Figure 1). It was formed through multiple tectonic activities ended by accretion of island arcs during the Late Neoproterozoic [35,36,37,38]. Typical island-arc related granitoid rocks are common in the ANS [33,39,40,41,42,43]. Some previous studies proposed that these granitoids can be divided into two main groups; the older (Tonian-Cryogenian) calc-alkaline granitites and the younger (Ediacaran) alkaline granitoids [44,45,46,47]. However, more recent studies indicate that there is sometimes a coeval occurrence of the calc-alkaline and the alkalinegranitoids [48,49]. After the formation of the ANS, the Egyptian Nubian Shield (ENS) was eroded before the end of Cambrian-Silurian time and buried beneath fluvial and near shore marine sedimentary successions with Lower Cambrian-aged fossils [50,51,52]. After the Cambrian, a period of tectonic stability with the deposition of platform sediments was established [53]. This period of stability was interrupted by number of vertical movements and unconformities [54]. During the Devonian-Carboniferous, the Hercynian (Variscan) tectonic event occurred by the collision of Gondwana and Laurasia, causing significant uplift and erosion of the Lower Palaeozoic sedimentary successions by Late Carboniferous time [55,56,57,58]. During the Jurassic-Cretaceous, Gondwana began to break apart forming extensional basins with volcanism in the ENS [59,60]. Simultaneously, during the Mid-Jurassic time, the mid-Atlantic opening caused an eastward movement of Africa with respect to Eurasia producing a large sinistral shear, extensive alkaline magmatism, normal NW-SE and N-S trending faults, as well as the formation of the broad northeast trending folds of the Syrian Arc domal system [61,62,63,64]. During Oligocene to Miocene period, a significant period of uplift and erosion started along the flanks of the Red Sea-Gulf of Suez rift system, exposing the ANS terrane, active fault systems and basaltic dykes have emerged in the course of crustal uplift.
Numerous schemes classified the Egyptian granites; one of the simpler schemes divided the granites into two groups. This approach includes; the Grey and Gattarian granites [44]; the Older Grey and Younger Red granites [46]; or the syn-orogenic and Younger granites [65]. In Sabet, [65] schema, the granitic rocks became more silicic and potash rich throughout the East African Orogeny. The Older Grey granites include the assemblage of felsic plutonic rocks of essentially intermediate composition [66,67]. While, the Younger Red granite shows a pronounced shift from the dominantly trondhjemitic, tonalitic, granodioritic, and calc-alkalic plutonism to one characterized by predominance of more felsic, more alkalic, granodioritic and granitic rock types [50].
Formation of Younger granites could be resolved into two major events; the first magmatic event (630 Ma to 570 Ma) is called the Dokhan event, during which most the Younger granites of the Eastern Desert were emplaced. The second magmatic event (570 Ma to 530 Ma) is called the Katherina event which is restricted to northern ANS [68]. The studied sample was collected from the Younger granite of the Dokhan event at Gebel Somr El-Qaa area, western the Gulf of Suez, in the NED, Egypt (Figure 2). In this region, the outcrops of the ENS Precambrian rocks non-conformably overlain by the sandstone of the Araba and Naqus formations with Cambrian-Ordovician age and the consequent Cretaceous to Eocene marine carbonate rocks [69,70].
The Younger Granites are muscovite biotite granitoids, however, hornblende may occur in association with biotite in the less evolved varieties. Almandine-rich garnet is common in per-aluminous granites, while, arvedsonite and riebeckite are more common in peralkaline grantoids [71,72,73]. The Younger granites are characterized by SiO2 ranging between 70 wt.% and 80 wt.%, K2O > 3.8 wt.%, the K2O/Na2O ratios > 1, the FeO*/MgO ratios > 4, and the CaO ranges between 0.1 wt.% and 1.6 wt.%. They are enriched in the high field strength elements and total REEs and relatively depleted in Ba, Sr, and have high Rb/Sr and conversely low K/Rb ratios [74,75].

3. Materials and Methods

Apatite and zircon grains were concentrated using conventional mineral separation techniques such as rock crushing, sieving, Frantz separator (S. G. Frantz Co., Inc., City, PA, USA), and heavy liquids. Approximately 200 apatite grains extracted from the studied sample under a stereoscopic microscope (Nikon Solutions Co., Ltd., Tokyo, Japan) were mounted with EpoFix resin in a 2.5-cm-diameter plastic ring (most grains were mounted with their surfaces parallel to the crystallographic c-axis). Approximately 100 zircon crystals were mounted in a Teflon. Mounted crystals were polished to expose their internal surfaces (i.e., to 4π geometry). Polished apatite crystals were etched in 5.5 M HNO3 at 21 ± 1 °C for 20 s to reveal spontaneous fission tracks [6]. Zircon grains were etched in in a NaOH-KOH eutectic melt at 220 ± 5 °C [76] for 60–210 min.
FTs counting and horizontally confined tracks (CTs) length measurements were performed using a Nikon Eclipse 80i (Nikon Solutions Co., Ltd., Tokyo, Japan) upgraded with a digital camera, an image processing software, and dry objectives. For each apatite grain dated and CT measured, the etch pit values was determined as a kinetic parameter (Dpar; [5,6])
The U isotopic concentrations for both FT and U-Pb dating as well as the Pb isotopes for U-Pb dating were measured by the LA-ICP-MS at Kanazawa University, Kanazawa, Japan. Detailed operating conditions were reported by Morishita et al. [75] and are summarized in Table 1. For each single-spot analysis, time-resolved signals were monitored to ensure stable signal intervals; free from detectable inclusions, core–rim features, zones of high common Pb or strong evidence of fractionations. The isotopic concentrations were then calculated from background corrected intensities. In order to avoid mass bias and reduce elemental fractionation decoupling of U-Th from Pb during micro-sampling by laser ablation; (1) short wavelength of 193 nm was used [77,78], (2) stream of Ar and He mixture was used as a carrier gas to transport the ablated material, (3) short ablation time of 30 s was applied to avoid extreme heating of the ablation site and losing focusing state on the ablated surface [26,78,79], and (4) we measured repeatedly an external reference material (e.g., NIST glass) to correct the residual laser induced fractionation and instrument mass bias [26,79]. Thus, the U, Th and Pb signal intensities were calibrated against silicate glass reference SRM 610 [80]; the U-238 concentration of 456 ppm is based on the total amount of uranium and isotope ratio [81,82]; the total Pb concentration is 431 ± 15 ppm [82] and the concentration of each Pb isotope is estimated based on their isotopic ratios [83,84]. The results for measured isotopes using standard glasses were normalized using Ca-43 as an internal standard in apatite grains and taking the chemical data reported previously [8]. For zircon crystals, the results for measured masses were normalized using Si-29 as an internal standard and tacking the chemical composition from [85,86]. Zircon grains were imaged using cathodoluminescence (CL) probe facility of Kanazawa University.

4. Results and Discussion

U-Pb and FT (ZFT) dating was performed simultaneously in twelve zircon grains. Similarly, 49 apatite grains were analyzed synchronously for apatite U-Pb and apatite FT (AFT) dating. Concurrently, 28 measurements were done on the analytical standard NIST SRM 612 glass to monitor the precision of our measurement protocol.

4.1. Zircon (U-Th)/Pb

The separated zircon crystals are mostly transparent and euhedral. Grains with either visible inclusions or cracks were excluded from the U–Pb dating. Therefore, only twelve zircon grains were analyzed from the Younger granite sample (Figure 3). All the dated zircon grains have common Pb occurrence under our system detection limit with low average intensity of 7.6 cps, while the average detection limit is 72 cps. Thus, no common lead correction was performed. The U-238 content varies between 10.6 ppm and 269.8 ppm. The Th/U ratio ranges from 0.04 to 1.11 (Table 2), indicating the igneous origin for all the analyzed grains [87,88]. Zircon U-Pb ages were calculated and a Concordia diagram was plotted using the IsoplotR code [86]. Grain C7 was not used for calculations because of discordant percentage higher than 10%. This grain shows abnormal isotopic concentrations (207Pb/235U = 14.55 ppm) which recommend possible inclusions or distorted fractures (Table 2). All the other treated zircon grains are well clustered on the Concordia (Figure 3B).
Zircon grains A1 and A8 yielded concordant pre-Pan-African 206Pb/238U ages of 1730 ± 76 Ma and 1723 ± 68 Ma, respectively (Table 2 and Figure 3). Both grains are inherited from older crust, and therefore, they were not included in the weighted mean age calculations. Several previous studies already demonstrated that inherited zircon crystals are very common in the ENS [48,49,89,90,91,92,93]. Grain D3 has an age of 770 ± 36 Ma which is apparently derived from reworked Pan-African crust. Most grains dated define a uniform Pan-African U-Pb weighted mean age of 599 ± 30 Ma which can be interpreted as the crystallization age of the granitic sample studied. Cathodoluminescence (CL) imagery shows a variety of internal structures (Figure 4).

4.2. Apatite U-Pb

The apatite U-Pb ages were calculated for the same 49 grains of the FT method, where ages were calculated and isochrone diagram was plotted using the IsoplotR code [86]. These yielded apatite U-Pb weighted mean ages ranges between ca. 592 Ma and ca. 368 Ma with a weighted mean average age of ca. 440 ± 17 Ma (Table 3 and Figure 5). The effective closure temperature of apatite U-Pb varies from ca. 550 °C to ca. 450 °C [22,91,94]. The U-238 concentration ranges between ca. 22 ppm and ca. 129 ppm. Apatite grains B2, D6, H4, and I10 will not be used in age calculations because they have discordant percentage higher than 10% (Table 3). Grain B2 has a relatively higher common Pb concentration of 0.51 ppm and higher Th/U concentration of 0.84 which recommend unrecognizable inclusions. Grains D6, H4, and I10 show low 207Pb/235U concentration of ca. 0.18, 0.44, and 0.48, respectively (Table 3), which recommend a Pb-206 enrichment. All the other apatite grains define a post-Pan-African apatite U-Pb weighted mean age of 474 ± 9 Ma which is considered as a cooling age with tectonic significance. Although, this is the first apatite U-Pb study on the ENS, few studies presented high-temperature thermochronological data [54,95]. These, study produced Ar-Ar and Sphene FT ages which are comparative to our apatite U-Pb ages.

4.3. Zircon Fission Track

The ZFT ages (closure temperature of ca. 240–200 °C; [12,96]) were obtained for the same grains analyzed by the zircon U-Pb technique (Table 2), using the IsoplotR code [86]. The ZFT weighted mean ages from twelve grains range between 372 ± 69 Ma and 319 ± 39 Ma with a weighted mean age of 347 ± 16 Ma (Figure 6). The U-238 concentrations vary from ca. 93 ppm and ca. 372 ppm (Table 4). The analyzed grains pass the chi-square probability test, indicating that these zircon grains belong to a single cooling event. The lack of any discernible correlation between ZFT grains ages and uranium concentrations medicate that metamictisation effect is neglectable.
The weighted mean ZFT age of 347 ± 16 Ma is concordant with previously reported ZFT data from other parts of the ANS [54,97], indicating a tectonically driven uplift. These thermochronological results are consistent with the sedimentary hiatus in the region between the Cambrian and the Carboniferous [98,99].

4.4. Apatite Fission Track

The AFT ages (closure temperature ca. 110 °C; [100]) were calculated for the same 49 grains analyzed by the apatite U-Pb method (Table 3), using the IsoplotR code [86]. The AFT ages range between 331 ± 70 Ma and 235 ± 36 Ma with a weighted mean age of 283 ± 7 Ma (Figure 7). The U-238 concentrations vary from ca. 11 ppm to ca. 95 ppm (Table 5).
The Dpars (total 201), parallel to the c-axis of each measured grain were measured showing values range between 1.22 µm and 1.71 µm (Table 5). These Dpar values indicate a uniform typical near-end-member calcium-fluorapatites [7]. Concurrently, 185 CTs were measured.
These lengths show a wide distribution, tall of short tracks, a positive skewness, and mean track length of 10.33 µm (Table 5). Therefore, the AFT weighted mean age here indicates no special event as it represents a mixed (partially reset) age.

4.5. Time-Temperature Modelling

For a better evaluation of the sample’s thermal history and representation of the event-corresponding age, the AFT data (ages, CTs and Dpars) were modelled with the “HeFTy” program (V. 1.8.3) [101]. Additionally, we used time-temperature constraints based on the obtained zircon U-Pb, apatite U-Pb, ZFT, and AFT ages. The fourth constraint represents the mixed AFT age; therefore, it is widened to represent wider age possibilities, the fifth represents the Oligocene-Miocene thermal event accompanied the Gulf of Suez formation (Figure 8). Thus, these boxes limiting possible solution to t-T model were added based on known and reasonable constraints.
The model paths (Figure 8) indicate a slow cooling after formation, flowed by a rapid cooling by the end of the Cambrian through the Devonian-Carboniferous until the block reaches the AFT PAZ equivalent depths during the Carboniferous. The ANS was affected by intense post-accretion erosional event before the Cambrian [50,102,103], which must have triggered isostatic rebound by rock exhumation. While, the Hercynian tectonic event affected the ENS by uplifting and erosion during the Devonian-Carboniferous [56,57,58,104]. Afterwards, the area experienced a relatively long period of thermal stability since the Carboniferous till the Eocene. This stability is not consistent with the tectonic active Cretaceous period [60,61,63,105], however, previous thermochronological studies on the ENS reported differential blocks movement [16,94,95,105,106,107]. Therefore, our sample might has been located within un-uplifted block during the Cretaceous. Finally, a rapid cooling event occurred as response to the Gulf of Suez initiation during the Oligocene-Miocene. The t-T model shows no thermal overprinting accompanying the Gulf event (Figure 8), that is supported by synchronous un-resetting of the AFT age.

5. Geological Interpretation

The obtained thermochronological data are consistent with the previously reported data; however, our data provides more robust and detailed information due to the simultaneous measurement of isotopic concentrations in zircon and apatite crystals. This double dating of U-Pb and FT techniques applied in this study enabled examination of the thermal history of the studies area through four different paleo-isotherms; zircon U-Pb (ca. 900 °C), apatite U-Pb (ca. 550 °C to 450 °C), ZFT (ca. 240 °C to 200 °C), and AFT (ca. 110 °C to 60 °C). We have integrated these thermochronological information and reconstructed the t-T history (Figure 8A). The zircon U–Pb results indicate that the Younger granite was formed during the Neoproterozoic (i.e., at 599 ± 30 Ma; Ediacaran). The magma source contained zircon grains with old ages of ca. 1.7 Ga which were inherited from a Pre-Pan-African crust. Another old grain yielded a Pan-African zircon U-Pb age of 770 ± 36 Ma. These older age grains indicate a local magmatic source. After the emplacement in the lower crust 599 ± 30 Ma, the granite cooled through the apatite U-Pb isotherm (ca. 550 °C to 450 °C) to reach depths of 22 ± 3 km at 474 ± 9 Ma (geothermal gradients of 21 °C/km; [109]). Assuming a homogeneous uplifting rate and starting at ca. 550 Ma, this event of uplifting was accompanied by ca. 0.3 km/Ma exhumation rate with a moderate cooling rate of ca. 6 °C/Ma. This exhumation event was initiated as response to the intensive erosional event inferred from the ANS being eroded before the end of the Cambrian followed by deposition of cycles of fluvial to near shore marine Lower Cambrian sediments [49,50,51]. Afterwards, a second exhumation event (or the first event was continued) at ca. 425 Ma (Figure 8A) was responsible for uplifting rocks from depths equivalent to the apatite U-Pb closure temperature (ca. 550 °C to 450 °C; ~22 ± 3 km) to those of the AFT (110°C to 60 °C; ca. 3 ± 2 km) through the ZFT (ca. 240 °C to 200 °C; ~10 ± 1 km) in ca. 100 Ma. This tectonic event occurred with an exhumation rate of ca. 0.2 km/Ma (i.e., assuming a homogeneous uplifting rate), and a cooling rate of ca. 3.4 °C/Ma. This exhumation event is previously reported in the region [53], which may indicate the continuity of the post-accretion erosional event till the Silurian. That was followed by a period of thermal stability and slow erosional uplifting within the apatite PAZ. This period of tectonic stability indicates the absence of any tectonic effect for the Cretaceous breaking apart of Gondwana on the studied block. The last uplifting stage accompanied by the Gulf of Suez formation which exhumed the rocks from depths equivalent to ca. 80 °C (ca. 2 km for geothermal gradient of 42 °C/km; [110]) to elevations close to the present (281 m.a.s.l.). This event accompanied by ca. 0.4 km/Ma exhumation rate and with an elevated cooling rate of ca. 12 °C/Ma, and a corresponding total rock uplift of ca. 2 km. The surprisingly small amount of uplift and the low thermal regime (ca. 80 °C) accompanied the Gulf of Suez formation may be explained by the development of the rift/basement bounding the listric normal faults at the first stage of rift formation. These listric faults kept the rift flanks unaffected by any of the rift development features. Otherwise, the later tectonic affects as well as the thermal overprinting were restricted to the main rift axis.

6. Conclusions

The use of LA-ICP-MS in U-Pb and FT chronometry has advantages for the measurement of a wide variety of elements and isotopes concurrently without any special preparation. Simultaneous calculation of U-Pb and FT ages for both zircon and apatite grains may, therefore, be provided. Based on this double dating approach, detailed insights on the thermo-tectonic history of a granitic sample from the ENS could be revealed. The granitic sample was crystallized at 599 ± 30 Ma (zircon U-Pb age). After emplacement, it was affected by three major exhumation events represent the influence of three tectonic events on the region: (1) Before 474 ± 9 Ma (apatite U-Pb age), the rock was uplifted as response to the ANS post-accretion erosional event. This event might have extended through the Silurian. (2) During the Devonian-Carboniferous, the granitic rock was exhumed to the AFT PAZ as response to the Hercynian tectonic event. (3) The Gulf of Suez opening caused exhumation of the sample to its current elevation during the Oligocene-Miocene.

Author Contributions

Conceptualization, S.M. and N.H.; methodology, N.H. and S.M.; validation, N.H. and S.M.; formal analysis, A.T. and S.M.; investigation, S.M.; resources, N.H., E.A., A.Y.E. and S.M.; data curation, S.M., E.A. and A.Y.E.; writing—original draft preparation, S.M.; writing—review and editing, N.H., E.A., A.Y.E. and S.M.; visualization, S.M.; supervision, N.H.; project administration, A.T.; funding acquisition, E.A., A.Y.E., S.M. and N.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data is contained within the article.

Acknowledgments

We thank Taif University Researchers Supporting Project Number (TURSP-2020/32), Taif University, Taif, Saudi Arabia; Efforts of Mohamed Ahmed, Faculty of science, Suez Canal University, Egypt are greatly appreciated for their help during the field work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. De Grave, J.; Glorie, S.; Buslov, M.M.; Stockli, D.F.; McWilliams, M.O.; Batalev, V.Y.; Van den haute, P. Thermo-Tectonic History of the Issyk-Kul Basement (Kyrgyz Northern Tien Shan, Central Asia). Gondwana Res. 2013, 23, 998–1020. [Google Scholar] [CrossRef]
  2. Glorie, S.; Otasevic, A.; Gillespie, J.; Jepson, G.; Danišík, M.; Zhimulev, F.I.; Gurevich, D.; Zhang, Z.; Song, D.; Xiao, W. Thermo-Tectonic History of the Junggar Alatau within the Central Asian Orogenic Belt (SE Kazakhstan, NW China): Insights from Integrated Apatite U/Pb, Fission Track and (U–Th)/He Thermochronology. Geosci. Front. 2019, 10, 2153–2166. [Google Scholar] [CrossRef]
  3. Grujic, D.; Ashley, K.T.; Coble, M.A.; Coutand, I.; Kellett, D.A.; Larson, K.P.; Whipp, D.M., Jr.; Gao, M.; Whynot, N. Deformational Temperatures Across the Lesser Himalayan Sequence in Eastern Bhutan and Their Implications for the Deformation History of the Main Central Thrust. Tectonics 2020, 39, e2019TC005914. [Google Scholar] [CrossRef]
  4. Wagner, G.; van den Haute, P. Fission-Track Dating; Solid Earth Sciences Library; Springer: Dordrecht, The Netherlands, 1992; ISBN 978-94-010-5093-7. [Google Scholar]
  5. Gleadow, A.J.W.; Duddy, I.R.; Green, P.F.; Lovering, J.F. Confined Fission Track Lengths in Apatite: A Diagnostic Tool for Thermal History Analysis. Contrib. Mineral. Petrol. 1986, 94, 405–415. [Google Scholar] [CrossRef]
  6. Donelick, R.A.; O’Sullivan, P.B.; Ketcham, R.A. 3. Apatite Fission-Track Analysis. In Low-Temperature Thermochronology; Reiners, P.W., Ehlers, T.A., Eds.; De Gruyter: Berlin, Germany, 2005; pp. 49–94. ISBN 978-1-5015-0957-5. [Google Scholar]
  7. Carlson, W.D.; Donelick, R.A.; Ketcham, R.A. Variability of Apatite Fission-Track Annealing Kinetics; I, Experimental Results. Am. Mineral. 1999, 84, 1213–1223. [Google Scholar] [CrossRef]
  8. Barbarand, J.; Carter, A.; Wood, I.; Hurford, T. Compositional and Structural Control of Fission-Track Annealing in Apatite. Chem. Geol. 2003, 198, 107–137. [Google Scholar] [CrossRef]
  9. Ramírez-Calderón, M.; Bedoya, A.; Abdullin, F.; Martini, M.; Solari, L.; Ortega-Obregón, C. Triassic Breakup of Pangea in Southern Mexico: Thermochronological Evidence from the Tianguistengo Formation. Geochemistry 2021, 81, 125776. [Google Scholar] [CrossRef]
  10. Bernet, M.; Brandon, M.; Garver, J.; Balestieri, M.L.; Ventura, B.; Zattin, M. Exhuming the Alps through Time: Clues from Detrital Zircon Fission-Track Thermochronology. Basin Res. 2009, 21, 781–798. [Google Scholar] [CrossRef]
  11. Gallagher, K.; Hawkesworth, C.J.; Mantovani, M.S.M. The Denudation History of the Onshore Continental Margin of SE Brazil Inferred from Apatite Fission Track Data. J. Geophys. Res. Solid Earth 1994, 99, 18117–18145. [Google Scholar] [CrossRef]
  12. Hurford, A.J. Cooling and Uplift Patterns in the Lepontine Alps South Central Switzerland and an Age of Vertical Movement on the Insubric Fault Line. Contrib. Mineral. Petrol. 1986, 92, 413–427. [Google Scholar] [CrossRef]
  13. Cliff, R.A.; Yardley, B.W.D.; Bussy, F.R. U–Pb and Rb–Sr Geochronology of Magmatism and Metamorphism in the Dalradian of Connemara, Western Ireland. J. Geol. Soc. 1996, 153, 109–120. [Google Scholar] [CrossRef] [Green Version]
  14. Jess, S.; Koehn, D.; Fox, M.; Enkelmann, E.; Sachau, T.; Aanyu, K. Paleogene Initiation of the Western Branch of the East African Rift: The Uplift History of the Rwenzori Mountains, Western Uganda. Earth Planet. Sci. Lett. 2020, 552, 116593. [Google Scholar] [CrossRef]
  15. Barbarand, J.; Marques, F.O.; Hildenbrand, A.; Pinna-Jamme, R.; Nogueira, C.R. Thermal Evolution of Onshore West Iberia: A Better Understanding of the Ages of Breakup and Rift-to-Drift in the Iberia-Newfoundland Rift. Tectonophysics 2021, 813, 228926. [Google Scholar] [CrossRef]
  16. Omar, G.I.; Steckler, M.S.; Buck, W.R.; Kohn, B.P. Fission-Track Analysis of Basement Apatites at the Western Margin of the Gulf of Suez Rift, Egypt: Evidence for Synchroneity of Uplift and Subsidence. Earth Planet. Sci. Lett. 1989, 94, 316–328. [Google Scholar] [CrossRef]
  17. Gleadow, A.J.W.; Fitzgerald, P.G. Uplift History and Structure of the Transantarctic Mountains: New Evidence from Fission Track Dating of Basement Apatites in the Dry Valleys Area, Southern Victoria Land. Earth Planet. Sci. Lett. 1987, 82, 1–14. [Google Scholar] [CrossRef]
  18. Hasebe, N.; Barbarand, J.; Jarvis, K.; Carter, A.; Hurford, A.J. Apatite Fission-Track Chronometry Using Laser Ablation ICP-MS. Chem. Geol. 2004, 207, 135–145. [Google Scholar] [CrossRef]
  19. Khedr, M.Z.; El-Awady, A.; Arai, S.; Hauzenberger, C.; Tamura, A.; Stern, R.J.; Morishita, T. Petrogenesis of the ~740 Korab Kansi Mafic-Ultramafic Intrusion, South Eastern Desert of Egypt: Evidence of Ti-Rich Ferropicritic Magmatism. Gondwana Res. 2020, 82, 48–72. [Google Scholar] [CrossRef]
  20. Chew, D.; Drost, K.; Marsh, J.H.; Petrus, J.A. LA-ICP-MS Imaging in the Geosciences and Its Applications to Geochronology. Chem. Geol. 2021, 559, 119917. [Google Scholar] [CrossRef]
  21. Claesson, S. Isotopic Evidence for the Precambrian Provenance and Caledonian Metamorphism of High Grade Paragneisses from the Seve Nappes, Scandinavian Caledonides I. Conventional U-Pb Zircon and Sm-Nd Whole Rock Data. Contrib Mineral Petrol 1987, 97, 196–204. [Google Scholar] [CrossRef]
  22. Cherniak, D.J.; Watson, E.B. Pb Diffusion in Zircon. Chem. Geol. 2001, 172, 5–24. [Google Scholar] [CrossRef]
  23. Feng, R.; Machado, N.; Ludden, J. Lead Geochronology of Zircon by LaserProbe-Inductively Coupled Plasma Mass Spectrometry (LP-ICPMS). Geochim. Cosmochim. Acta 1993, 57, 3479–3486. [Google Scholar] [CrossRef]
  24. Kosler, J.; Tubrett, M.N.; Sylvester, P.J. Application of Laser Ablation ICP-MS to U-Th-Pb Dating of Monazite. Geostand. Geoanal. Res. 2001, 25, 375–386. [Google Scholar] [CrossRef]
  25. Kosler, J. Present Trends and the Future of Zircon in Geochronology: Laser Ablation ICPMS. Rev. Mineral. Geochem. 2003, 53, 243–275. [Google Scholar] [CrossRef]
  26. Jackson, S.E.; Pearson, N.J.; Griffin, W.L.; Belousova, E.A. The Application of Laser Ablation-Inductively Coupled Plasma-Mass Spectrometry to in Situ U–Pb Zircon Geochronology. Chem. Geol. 2004, 211, 47–69. [Google Scholar] [CrossRef]
  27. Xiao, L.; Zhang, H.-F.; Ni, P.-Z.; Xiang, H.; Liu, X.-M. LA-ICP-MS U–Pb Zircon Geochronology of Early Neoproterozoic Mafic-Intermediat Intrusions from NW Margin of the Yangtze Block, South China: Implication for Tectonic Evolution. Precambrian Res. 2007, 154, 221–235. [Google Scholar] [CrossRef]
  28. Cocherie, A.; Robert, M. Laser Ablation Coupled with ICP-MS Applied to U–Pb Zircon Geochronology: A Review of Recent Advances. Gondwana Res. 2008, 14, 597–608. [Google Scholar] [CrossRef]
  29. Carr, P.A.; Mercadier, J.; Harlaux, M.; Romer, R.L.; Moreira, E.; Legros, H.; Cuney, M.; Marignac, C.; Cauzid, J.; Salsi, L.; et al. U/Pb Geochronology of Wolframite by LA-ICP-MS; Mineralogical Constraints, Analytical Procedures, Data Interpretation, and Comparison with ID-TIMS. Chem. Geol. 2021, 584, 120511. [Google Scholar] [CrossRef]
  30. Chen, C.; Lv, X.; Gun, M.; Yang, J. Metallogenic Chronology and Tectonic Setting of the Erdaohe Pb–Zn–Ag Deposit in Inner Mongolia, NE China: Constraints from Sphalerite Rb–Sr Dating, Zircon U–Pb Dating, and Hf Isotope Analysis. Ore Geol. Rev. 2021, 134, 104067. [Google Scholar] [CrossRef]
  31. Hannon, J.S.; Dietsch, C.; Huff, W.D.; Garway, D. Tracking 40 Million Years of Migrating Magmatism across the Idaho Batholith Using Zircon U-Pb Ages and Hf Isotopes from Cretaceous Bentonites. Minerals 2021, 11, 1011. [Google Scholar] [CrossRef]
  32. Yang, Y.; Liang, C.; Zheng, C.; Xu, X.; Zhou, J.; Zhou, X.; Cao, C. Metamorphic Evolution of High-Grade Granulite-Facies Rocks of the Mashan Complex, Liumao Area, Eastern Heilongjiang Province, China: Evidence from Zircon U–Pb Geochronology, Geochemistry and Phase Equilibria Modelling. Precambrian Res. 2021, 355, 106095. [Google Scholar] [CrossRef]
  33. Bentor, Y.K. The Crustal Evolution of the Arabo-Nubian Massif with Special Reference to the Sinai Peninsula. Precambrian Res. 1985, 28, 1–74. [Google Scholar] [CrossRef]
  34. Kennedy, W.Q. The Structural Differentiation of Africa in the Pan-African (±500 m.y.) Tectonic Episode. Leeds University Research Institute of African Geology and Department of Earth Sciences Annual Report on Scientific Results; Scientific Research: Leeds, UK, 1964. [Google Scholar]
  35. Kröner, A. Pan African Plate Tectonics and Its Repercussions on the Crust of Northeast Africa. Geol. Rundsch. 1979, 68, 565–583. [Google Scholar] [CrossRef]
  36. Engel, A.E.J.; Dixon, T.H.; Stern, R.J. Late Precambrian Evolution of Afro-Arabian Crust from Ocean Arc to Craton. Geol. Soc. Am. Bull. 1980, 91, 699. [Google Scholar] [CrossRef]
  37. Meert, J.G. A Synopsis of Events Related to the Assembly of Eastern Gondwana. Tectonophysics 2003, 362, 1–40. [Google Scholar] [CrossRef]
  38. Johnson, P.R.; Andresen, A.; Collins, A.S.; Fowler, A.R.; Fritz, H.; Ghebreab, W.; Kusky, T.; Stern, R.J. 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. 2011, 61, 167–232. [Google Scholar] [CrossRef]
  39. El-Gaby, S.; El-Nady, O.; Khudeir, A. Tectonic Evolution of the Basement Complex in the Central Eastern Desert of Egypt. Geol. Rundsch. 1984, 73, 1019–1036. [Google Scholar] [CrossRef]
  40. El Din, G.M.K.; Khudeir, A.A.; Greiling, R.O. Tectonic Evolution of a Pan-African Gneiss Culmination, Gabal El Sibai Area, Central Eastern Desert, Egypt. Zent Bl Geol Palaeont 1991, I11, 2637–2640. [Google Scholar]
  41. Rashwan, A.A. Petrography, Geochemistry and Petrogenesis of the Migif-Hafifit at Hafifit Mine Area, Egypt. 1991. Available online: https://www.iberlibro.com/9783893360390/Petrography-geochemistry-petrogenesis-MIGIF-HAFAFIT-gneisses-3893360395/plp (accessed on 25 November 2021).
  42. Gass, I.G. The Evolution of the Pan African Crystalline Basement in NE Africa and Arabia. J. Geol. Soc. 1977, 134, 129–138. [Google Scholar] [CrossRef]
  43. Nasseef, A.O.; Gass, I.G. Granitic and Metamorphic Rocks of the Taif Area, Western Saudi Arabia. Geol. Soc. Am. Bull. 1977, 88, 1721. [Google Scholar] [CrossRef]
  44. Hume, W.F. The Later Plutonic and Minor Intrusive Rocks. In Geology of Egypt; Government Press: Cairo, Egypt, 1935. [Google Scholar]
  45. Schurmann, H.M.E. The Precambrian of the Gulf of Suez Area. Int. Geol. Congr. Algiers CR 1953, 19, 115–135. [Google Scholar]
  46. El Ramly, M.F.; Akaad, M.K. The Basement Complex in the Central-Eastern Desert of Egypt between Lat. 24°30′ and 25°40′ N. Geol. Surv. Egypt Ann. 1960, 8, 1–35. [Google Scholar]
  47. El Bahariya, G.A. Geochemistry and Tectonic Setting of Neoproterozoic Rocks from the Arabian-Nubian Shield: Emphasis on the Eastern Desert of Egypt. In Applied Geochemistry with Case Studies on Geological Formations, Exploration Techniques and Environmental Issues; Felipe Mazadiego, L., De Miguel Garcia, E., Barrio-Parra, F., Izquierdo-Díaz, M., Eds.; IntechOpen: London, UK, 2020; ISBN 978-1-78985-884-6. [Google Scholar]
  48. Ali, B.H.; Wilde, S.A.; Gabr, M.M.A. Granitoid Evolution in Sinai, Egypt, Based on Precise SHRIMP U–Pb Zircon Geochronology. Gondwana Res. 2009, 15, 38–48. [Google Scholar] [CrossRef]
  49. Moreno, J.A.; Montero, P.; Abu Anbar, M.; Molina, J.F.; Scarrow, J.H.; Talavera, C.; Cambeses, A.; Bea, F. SHRIMP U–Pb Zircon Dating of the Katerina Ring Complex: Insights into the Temporal Sequence of Ediacaran Calc-Alkaline to Peralkaline Magmatism in Southern Sinai, Egypt. Gondwana Res. 2012, 21, 887–900. [Google Scholar] [CrossRef]
  50. Said, R. The Geology of Egypt, 2nd ed.; A.A. Balkema, Elsevier: Rotterdam, The Netherlands, 1990. [Google Scholar]
  51. Seilacher, A. Paleozoic Trace Fossils. In Geology of Egypt; A.A. Balkema, Elsevier: Rotterdam, The Netherlands, 1990; pp. 113–156. [Google Scholar]
  52. Bosworth, W.; Huchon, P.; McClay, K. The Red Sea and Gulf of Aden Basins. J. Afr. Earth Sci. 2005, 43, 334–378. [Google Scholar] [CrossRef]
  53. Alsharhan, A.S.; Nairn, A.E.M. Sedimentary Basins and Petroleum Geology of the Middle East; Elsevier: Amsterdam, The Netherlands, 1997; ISBN 978-0-444-82465-3. [Google Scholar]
  54. Gvirtzman, G.; Weissbrod, T. The Hercynian Geanticline of Helez and the Late Palaeozoic History of the Levant. Geol. Soc. Lond. Spec. Publ. 1984, 17, 177–186. [Google Scholar] [CrossRef]
  55. Stampfii, G.M.; von Raumer, J.F.; Borel, G.D. Paleozoic Evolution of Pre-Variscan Terranes: From Gondwana to the Variscan Collision. In Variscan-Appalachian Dynamics: The Building of the Late Paleozoic Basement; Geological Society of America: Boulder, CO, USA, 2002; ISBN 978-0-8137-2364-8. [Google Scholar]
  56. Craig, J.; Rizzi, C.; Said, F.; Thusu, B.; Luning, S.; Asbali, A.I.; Keeley, M.L.; Bell, J.F.; Durham, M.J.; Eales, M.H.; et al. Structural Styles and Prospectivity in the Precambrian and Palaeozoic Hydrocarbon Systems of North Africa. Geol. East Libya 2008, 4, 51–122. [Google Scholar]
  57. Dixon, R.J.; Moore, J.K.S.; Bourne, M.; Dunn, E.; Haig, D.B.; Hossack, J.; Roberts, N.; Parsons, T.; Simmons, C.J. Integrated Petroleum Systems and Play Fairway Analysis in a Complex Palaeozoic Basin: Ghadames-Illizi Basin, North Africa. Geol. Soc. Lond. Pet. Geol. Conf. Ser. 2010, 7, 735–760. [Google Scholar] [CrossRef]
  58. Hashad, A.H. Present Status of Geochronological Data on the Egyptian Basement Complex. Precambrian Res. 1978, 6, A24–A25. [Google Scholar] [CrossRef]
  59. Klitzsch, E. Plate Tectonics and Cratonal Geology in Northeast Africa (Egypt, Sudan). Geol. Rundsch. 1986, 75, 755–768. [Google Scholar] [CrossRef]
  60. Said, R. The Geology of Egypt, 1st ed.; Elsevier: Amsterdam, The Netherlands, 1962. [Google Scholar]
  61. Dewey, J.F.; Pitman, W.C.; Ryan, W.B.F.; Bonnin, J. Plate Tectonics and the Evolution of the Alpine System. Geol. Soc. Am. Bull. 1973, 84, 3137. [Google Scholar] [CrossRef]
  62. Awad, G.M. Habitat of Oil in Abu Gharadiq and Faiyum Basins, Western Desert, Egypt. AAPG Bull. 1984, 68, 564–573. [Google Scholar]
  63. Greiling, R.O.; Kriiner, A.; El Ramly, M.F.; Rashwan, A.A. Structural Relationships between the Southern and Central Parts of the Eastern Desert of Egypt: Details of a Fold and Thrust Belt. In The Pan-African of NE Africa and Adjacent Areas; El Gaby, S., Greiling, R.O., Eds.; Vieweg: Wiesbaden, Germany, 1988; pp. 121–145. [Google Scholar]
  64. Tawfik, H.A.; Ghandour, I.M.; Maejima, W.; Abdel-Hameed, A.T. Petrography and Geochemistry of the Lower Paleozoic Araba Formation, Northern Eastern Desert, Egypt: Implications for Provenance, Tectonic Setting and Weathering Signature. J. Geosci. Osaka City Univ. 2011, 54, 1–16. [Google Scholar]
  65. Sabet, A.H. An Example of Photo Interpretation of Crystalline Rocks. In; ITC Publication B.: Delft, The Netherlands, 1962; p. 34. [Google Scholar]
  66. El Shazly, E.M. On the Classification of the Precambrian and Other Rocks of Magmatic Affiliation in Egypt. In Proceedings of the 22nd International Geological Congress, New Delhi, India, 14 December 1964. [Google Scholar]
  67. Hassan, M.A.; Hashad, A.H. Precambrian of Egypt. In Geology of Egypt; Said, R., Ed.; Balkema Publications: Rotterdam, The Netherlands, 1990; p. 734. [Google Scholar]
  68. Abdallah, A.M.; Darwish, M.; El-Aref, M.; Helba, A.A. Lithostratigraphy of the Pre-Cenomanian Clastics of North Wadi Qena, Eastern Desert, Egypt. 1992. Available online: https://pascal-francis.inist.fr/vibad/index.php?action=getRecordDetail&idt=6498128 (accessed on 25 November 2021).
  69. Moghazi, A.M.; Hassanen, M.A.; Mohamed, F.H.; Ali, S. Late Neoproterozoic Strongly Peraluminous Leucogranites, South Eastern Desert, Egypt—Petrogenesis and Geodynamic Significance. Mineral. Petrol. 2004, 81, 19–41. [Google Scholar] [CrossRef]
  70. El-Bialy, M.Z.; Streck, M.J. Late Neoproterozoic Alkaline Magmatism in the Arabian–Nubian Shield: The Postcollisional A-Type Granite of Sahara–Umm Adawi Pluton, Sinai, Egypt. Arab. J. Geosci. 2009, 2, 151–174. [Google Scholar] [CrossRef]
  71. Azer, M.K.; Samuel, M.D.; Ali, K.A.; Gahlan, H.A.; Stern, R.J.; Ren, M.; Moussa, H.E. Neoproterozoic Ophiolitic Peridotites along the Allaqi-Heiani Suture, South Eastern Desert, Egypt. Mineral. Petrol. 2013, 107, 829–848. [Google Scholar] [CrossRef]
  72. El-Bialy, M.Z. Precambrian Basement Complex of Egypt. In The Geology of Egypt; Hamimi, Z., El-Barkooky, A., Martínez Frías, J., Fritz, H., Abd El-Rahman, Y., Eds.; Regional Geology Reviews; Springer International Publishing: Cham, Switzerland, 2020; pp. 37–79. ISBN 978-3-030-15265-9. [Google Scholar]
  73. El-Bialy, M.Z. The Ediacaran Post-Collisional Dokhan Volcanics. In The Geology of the Egyptian Nubian Shield; Hamimi, Z., Arai, S., Fowler, A.-R., El-Bialy, M.Z., Eds.; Regional Geology Reviews; Springer International Publishing: Cham, Switzerland, 2021; pp. 267–294. ISBN 978-3-030-49771-2. [Google Scholar]
  74. Garver, J.I. Etching Zircon Age Standards for Fission-Track Analysis. Radiat. Meas. 2003, 37, 47–53. [Google Scholar] [CrossRef]
  75. Morishita, T.; Ishida, Y.; Arai, S.; Shirasaka, M. Determination of Multiple Trace Element Compositions in Thin (>30 µM) Layers of NIST SRM 614 and 616 Using Laser Ablation-Inductively Coupled Plasma-Mass Spectrometry (LA-ICP-MS). Geostand. Geoanal. Res. 2005, 29, 107–122. [Google Scholar] [CrossRef]
  76. Jeffries, T.E.; Pearce, N.J.G.; Perkins, W.T.; Raith, A. Chemical Fractionation during Infrared and Ultraviolet Laser Ablation Inductively Coupled Plasma Mass Spectrometry—Implications for Mineral Microanalysis. Anal. Commun. 1996, 33, 35–39. [Google Scholar] [CrossRef]
  77. Hirata, T.; Nesbitt, R.W. U-Pb Isotope Geochronology of Zircon: Evaluation of the Laser Probe-Inductively Coupled Plasma Mass Spectrometry Technique. Geochim. Cosmochim. Acta 1995, 59, 2491–2500. [Google Scholar] [CrossRef]
  78. Longerich, H.P.; Jackson, S.E.; Günther, D. Inter-Laboratory Note. Laser Ablation Inductively Coupled Plasma Mass Spectrometric Transient Signal Data Acquisition and Analyte Concentration Calculation. J. Anal. Spectrom. 1996, 11, 899–904. [Google Scholar] [CrossRef]
  79. Carpenter, B.S.; Reimer, G.M. Standard Reference Materials: Calibrated Glass Standards for Fission Track Use. 1974. Available online: https://www.osti.gov/biblio/4232790-standard-reference-materials-calibrated-glass-standards-fission-track-use-final-report (accessed on 25 November 2021).
  80. Pearce, N.J.G.; Perkins, W.T.; Westgate, J.A.; Gorton, M.P.; Jackson, S.E.; Neal, C.R.; Chenery, S.P. A Compilation of New and Published Major and Trace Element Data for NIST SRM 610 and NIST SRM 612 Glass Reference Materials. Geostand. Geoanalytical Res. 1997, 21, 115–144. [Google Scholar] [CrossRef]
  81. Walder, A.J.; Platzner, I.; Freedman, P.A. Isotope Ratio Measurement of Lead, Neodymium and Neodymium–Samarium Mixtures, Hafnium and Hafnium–Lutetium Mixtures with a Double Focusing Multiple Collector Inductively Coupled Plasma Mass Spectrometer. J. Anal. Spectrom. 1993, 8, 19–23. [Google Scholar] [CrossRef]
  82. Gagnevin, D.; Daly, J.S.; Waight, T.E.; Morgan, D.; Poli, G. Pb Isotopic Zoning of K-Feldspar Megacrysts Determined by Laser Ablation Multi-Collector ICP-MS: Insights into Granite Petrogenesis. Geochim. Cosmochim. Acta 2005, 69, 1899–1915. [Google Scholar] [CrossRef]
  83. Deer, W.A.; Howie, R.A.; Zussman, J. Rock Forming Minerals, 2nd ed.; Longman Scientific & Technical: Harlow, Essex, UK; New York, NY, USA, 1982; Volume 1A. [Google Scholar]
  84. Sokolov, V.A. Arc Furnace Assisted Carbothermal Decomposition of Zircon. Refract. Ind. Ceram. 2005, 46, 208–211. [Google Scholar] [CrossRef]
  85. Rubatto, D. Zircon: The Metamorphic Mineral. Rev. Mineral. Geochem. 2017, 83, 261–295. [Google Scholar] [CrossRef]
  86. Vermeesch, P. IsoplotR: A Free and Open Toolbox for Geochronology. Geosci. Front. 2018, 9, 1479–1493. [Google Scholar] [CrossRef]
  87. Kröner, A.; Eyal, M.; Eyal, Y. Early Pan-African Evolution of the Basement around Elat, Israel, and the Sinai Peninsula Revealed by Single-Zircon Evaporation Dating, and Implications for Crustal Accretion Rates. Geology 1990, 18, 545. [Google Scholar] [CrossRef]
  88. Kröner, A.; Kröner, J.; Rashwan, A.A.A. Age and Tectonic Setting of Granitoid Gneisses in the Eastern Desert of Egypt and South-West Sinai. Geol. Rundsch. 1994, 83, 502–513. [Google Scholar] [CrossRef]
  89. Moussa, E.M.M.; Stern, R.J.; Manton, W.I.; Ali, K.A. SHRIMP Zircon Dating and Sm/Nd Isotopic Investigations of Neoproterozoic Granitoids, Eastern Desert, Egypt. Precambrian Res. 2008, 160, 341–356. [Google Scholar] [CrossRef]
  90. Abu El-Enen, M.M.; Whitehouse, M.J. The Feiran–Solaf Metamorphic Complex, Sinai, Egypt: Geochronological and Geochemical Constraints on Its Evolution. Precambrian Res. 2013, 239, 106–125. [Google Scholar] [CrossRef]
  91. Eyal, M.; Be’eri-Shlevin, Y.; Eyal, Y.; Whitehouse, M.J.; Litvinovsky, B. Three Successive Proterozoic Island Arcs in the Northern Arabian–Nubian Shield: Evidence from SIMS U–Pb Dating of Zircon. Gondwana Res. 2014, 25, 338–357. [Google Scholar] [CrossRef]
  92. Chamberlain, K.R.; Bowring, S.A. Apatite–Feldspar U–Pb Thermochronometer: A Reliable, Mid-Range (∼450 °C), Diffusion-Controlled System. Chem. Geol. 2001, 172, 173–200. [Google Scholar] [CrossRef]
  93. Kohn, B.P.; Feinstein, S.; Foster, D.A.; Steckler, M.S.; Eyal, M. Thermal History of the Eastern Gulf of Suez, II. Reconstruction from Apatite Fission Track and K-Feldspar Measurements. Tectonophysics 1997, 283, 219–239. [Google Scholar] [CrossRef]
  94. Bojar, A.-V.; Fritz, H.; Kargl, S.; Unzog, W. Phanerozoic Tectonothermal History of the Arabian–Nubian Shield in the Eastern Desert of Egypt: Evidence from Fission Track and Paleostress Data. J. Afr. Earth Sci. 2002, 34, 191–202. [Google Scholar] [CrossRef]
  95. Vermeesch, P.; Avigad, D.; McWilliams, M.O. 500 m.y. of Thermal History Elucidated by Multi-Method Detrital Thermochronology of North Gondwana Cambrian Sandstone (Eilat Area, Israel). GSA Bull. 2009, 121, 1204–1216. [Google Scholar] [CrossRef] [Green Version]
  96. Yamada, R.; Tagami, T.; Nishimura, S.; Ito, H. Annealing Kinetics of Fission Tracks in Zircon: An Experimental Study. Chem. Geol. 1995, 122, 249–258. [Google Scholar] [CrossRef]
  97. Kohn, B.P.; Eyal, M.; Feinstein, S. A Major Late Devonian-Early Carboniferous (Hercynian) Thermotectonic Event at the NW Margin of the Arabian-Nubian Shield: Evidence from Zircon Fission Track Dating. Tectonics 1992, 11, 1018–1027. [Google Scholar] [CrossRef]
  98. Tewfik, N.; Harwood, C.; Deighton, I. The Miocene, Rudeis and Kareem Formations in the Gulf of Suez: Aspects of Sedimentology and Geochemistry. In Proceedings of the Egyptian General Petroleum Corporation Exploration Seminar, Cairo, Egypt, 7–10 February 1992. [Google Scholar]
  99. Darwish, M.; El Araby, A. Petrography and Diagenetic Aspects of Some Siliclastic Hydrocarbon Reservoirs in Relation to Rifting of the Gulf of Suez, Egypt. Geodynamics and Sedimentation of the Red Sea—Gulf of Aden Rift System. Egyptian J. Geol. 1994, 3, 1–25. [Google Scholar]
  100. Gleadow, A.J.W.; Duddy, I.R. A Natural Long-Term Track Annealing Experiment for Apatite. Nucl. Tracks 1981, 5, 169–174. [Google Scholar] [CrossRef]
  101. Ketcham, R.A. Forward and Inverse Modeling of Low-Temperature Thermochronometry Data. Rev. Mineral. Geochem. 2005, 58, 275–314. [Google Scholar] [CrossRef]
  102. Garfunkel, Z. Early Paleozoic Sediments of NE Africa and Arabia: Products of Continental-Scale Erosion, Sediment Transport, and Deposition. Isr. J. Earth Sci. 2002, 51, 135–156. [Google Scholar] [CrossRef]
  103. Kolodner, K.; Avigad, D.; McWILLIAMS, M.; Wooden, J.L.; Weissbrod, T.; Feinstein, S. Provenance of North Gondwana Cambrian–Ordovician Sandstone: U–Pb SHRIMP Dating of Detrital Zircons from Israel and Jordan. Geol. Mag. 2006, 143, 367–391. [Google Scholar] [CrossRef] [Green Version]
  104. Shehata, A.A.; El Fawal, F.M.; Ito, M.; Aboulmagd, M.A.; Brooks, H.L. Senonian Platform-to-Slope Evolution in the Tectonically-Influenced Syrian Arc Sedimentary Belt: Beni Suef Basin, Egypt. J. Afr. Earth Sci. 2020, 170. [Google Scholar] [CrossRef]
  105. Kohn, B.P.; Eyal, M. History of Uplift of the Crystalline Basement of Sinai and Its Relation to Opening of the Red Sea as Revealed by Fission Track Dating of Apatites. Earth Planet. Sci. Lett. 1981, 52, 129–141. [Google Scholar] [CrossRef]
  106. Omar, G.I.; Kohn, B.P.; Lutz, T.M.; Faul, H. The Cooling History of Silurian to Cretaceous Alkaline Ring Complexes, South Eastern Desert, Egypt, as Revealed by Fission-Track Analysis. Earth Planet. Sci. Lett. 1987, 83, 94–108. [Google Scholar] [CrossRef]
  107. Feinstein, S.; Eyal, M.; Kohn, B.P.; Steckler, M.S.; Ibrahim, K.M.; Moh’d, B.K.; Tian, Y. Uplift and Denudation History of the Eastern Dead Sea Rift Flank, SW Jordan: Evidence from Apatite Fission Track Thermochronometry. Tectonics 2013, 32, 1513–1528. [Google Scholar] [CrossRef]
  108. Ketcham, R.A.; Donelick, R.A.; Carlson, W.D. Variability of Apatite Fission-Track Annealing Kinetics; III, Extrapolation to Geological Time Scales. Am. Mineral. 1999, 84, 1235–1255. [Google Scholar] [CrossRef]
  109. Morgan, E.; Boulos, E.K.; Hennin, S.E.; EL-Sheriff, A.A.; Sayed, A.A.; Basta, N.Z.; Melek, Y.S. Heat flow in eastern Egypt: The thermal signature of a continental breakup. J. Geodyn. 1985, 4, 107–131. [Google Scholar] [CrossRef]
  110. Omar, G.I.; Steckler, M.S. Fission Track Evidence on the Initial Rifting of the Red Sea: Two Pulses, No Propagation. Science 1995, 270, 1341–1344. [Google Scholar] [CrossRef]
Figure 1. Location map for the ANS (the ENS; the Egyptian part of the Nubian Shield) in the frame of the African continent, and the studied sample (modified after [33,34]).
Figure 1. Location map for the ANS (the ENS; the Egyptian part of the Nubian Shield) in the frame of the African continent, and the studied sample (modified after [33,34]).
Minerals 11 01341 g001
Figure 2. Simplified Geologic Map for Gebel Somr El-Qaa area, NED, Egypt (modified after [70]) representing the studied sample location and the main lithology of the region; the ANS granite (750–530 Ma), Dokhan volcanics (630–570 Ma), Araba and Naqus Formations (Lower Palaeozoic age).
Figure 2. Simplified Geologic Map for Gebel Somr El-Qaa area, NED, Egypt (modified after [70]) representing the studied sample location and the main lithology of the region; the ANS granite (750–530 Ma), Dokhan volcanics (630–570 Ma), Araba and Naqus Formations (Lower Palaeozoic age).
Minerals 11 01341 g002
Figure 3. Concordia diagram [86]. (A) for all dated zircon U-Pb ages representing the two pre-Pan-African grains, (B) for all grains with Pan-African ages which gives weighted mean age of 599 ± 30 Ma (7 grains) if the older D3 grain was excluded.
Figure 3. Concordia diagram [86]. (A) for all dated zircon U-Pb ages representing the two pre-Pan-African grains, (B) for all grains with Pan-African ages which gives weighted mean age of 599 ± 30 Ma (7 grains) if the older D3 grain was excluded.
Minerals 11 01341 g003
Figure 4. Cathodoluminescence (CL) images showing the location of analysis spots and their 206Pb/238U weighted mean ages for some of the analyzed zircon grains.
Figure 4. Cathodoluminescence (CL) images showing the location of analysis spots and their 206Pb/238U weighted mean ages for some of the analyzed zircon grains.
Minerals 11 01341 g004
Figure 5. Tera–Wasserburg plots [86], displaying apatite U–Pb results obtained from the studied samples with excluding the discordant grains in Table 3.
Figure 5. Tera–Wasserburg plots [86], displaying apatite U–Pb results obtained from the studied samples with excluding the discordant grains in Table 3.
Minerals 11 01341 g005
Figure 6. Radial plot diagram [86] for all dated ZFT grains.
Figure 6. Radial plot diagram [86] for all dated ZFT grains.
Minerals 11 01341 g006
Figure 7. Radial plot diagram [89] for all dated AFT grains.
Figure 7. Radial plot diagram [89] for all dated AFT grains.
Minerals 11 01341 g007
Figure 8. (A): Thermal history model for the studied sample obtained using HeFTy [103]. Resulting t–T curves show four different reliability levels; green paths: acceptable fit (all t–T paths with a merit function value of at least 0.05), purple paths: good fit (all t–T paths with a merit function value of at least 0.5), black line: best fit and dashed blue line: the weighted mean path [101,108]. Four constrain boxes, based on zircon U-Pb, apatite U-Pb, ZFT, and AFT ages, and the Gulf of Suez initiation age were chosen to guide the randomness. P: number if inverse models’ iterations, A: acceptable fit models, G: good fit models, D: determined FT age ± 1-σ error, M: modelled FT age, G.O.F.: goodness of fit between the measured and the model age, N: number of single grains. (B): Confined track lengths (CTs) distribution for the c-axis projected tracks. D: determined CTs ± 1-σ error, M: modelled CTs ± 1-σ error, G.O.F.: goodness of fit between the measured and the model CTs, N: number of CTs. (C): comparison between the corresponding AFT details of AFT age, CTs lengths, and Dpar values.
Figure 8. (A): Thermal history model for the studied sample obtained using HeFTy [103]. Resulting t–T curves show four different reliability levels; green paths: acceptable fit (all t–T paths with a merit function value of at least 0.05), purple paths: good fit (all t–T paths with a merit function value of at least 0.5), black line: best fit and dashed blue line: the weighted mean path [101,108]. Four constrain boxes, based on zircon U-Pb, apatite U-Pb, ZFT, and AFT ages, and the Gulf of Suez initiation age were chosen to guide the randomness. P: number if inverse models’ iterations, A: acceptable fit models, G: good fit models, D: determined FT age ± 1-σ error, M: modelled FT age, G.O.F.: goodness of fit between the measured and the model age, N: number of single grains. (B): Confined track lengths (CTs) distribution for the c-axis projected tracks. D: determined CTs ± 1-σ error, M: modelled CTs ± 1-σ error, G.O.F.: goodness of fit between the measured and the model CTs, N: number of CTs. (C): comparison between the corresponding AFT details of AFT age, CTs lengths, and Dpar values.
Minerals 11 01341 g008
Table 1. Operating conditions for the LA-ICP-MS.
Table 1. Operating conditions for the LA-ICP-MS.
ICP-MSDetalis
ModelAgilent 7500 s
Forward power1200 W
Reflected power1 W
Carrier gas flow 1. 31/min−1 (Ar)
0.31/min−1 (He)
Auxiliary gas flow1.01/min−1
Plasma gas flow151/min−1
InterfaceNi sample cone
Ni skimmer cone
LaserDetails
ModelMicroLas GeoLas Q plus
Wavelength193 nm (Excimer ArF)
Repetition rate5 Hz
Pulse energy8 J/cm2
Pit diameter20 µm
Table 2. Zircon U-Th-Pb Data for the studied sample.
Table 2. Zircon U-Th-Pb Data for the studied sample.
Grain No.Intensities (cps)Conc. (ppm)Isotopic Ratios and 1σ ErrorsAge (Ma) and 1σ Errors% dis-c.
204Pb238UTh/U±1σ206Pb/238U±1σ207Pb/235U±1σ208Pb/232Th±1σ206Pb/238U±1σ207Pb/235U±1σ208Pb/232Th±1σ238U/235U
A1 *−17953.54.51.030.0390.3080.0054.9020.1570.0930.0021730761810601440394.0
A8 *−1513690.89.81.110.0060.3060.0014.9260.0100.1000.001172376183660608244.6
B19.1119135.312.40.840.0020.1030.0010.9750.0100.0360.0016303068835791258.8
B6 *53.560266.110.10.290.0020.0960.0030.8680.4330.0346.586591286343338532366.8
C7 *−37310.68.70.040.0020.2200.00114.5500.0088.1400.0011283582851686981954.0
D3 *0.36456.216.10.430.0020.1270.0011.1790.0140.0460.001770368023950642.6
D82.096269.815.20.730.0060.0910.0020.7750.0230.0320.0015602760332425153.9
E713.276101.211.30.370.0090.0980.0010.8200.0060.0330.0016012962633796131.1
F16.9132141.76.70.390.0160.0980.0010.8490.0060.0330.0016022963533596153.6
G418.4122102.86.81.080.0060.0920.0010.7640.0080.0310.0015672759632489221.7
G51.17170.98.51.100.0080.0990.0010.7900.0220.0340.00161029617334226−3.2
H75.570144.59.90.260.0050.1030.0010.8520.0210.0330.00163430643345409−1.3
A1 is symbol of analyzed grain, grains with * was not used in the weighted mean age, Intensities represent background corrected signal intensities, 3σ calculated for signal background to representing the detection limit of our system, Conc. = Isotopic concentrations, and isotopic ratios area background corrected values, 204Pb negative concentrations indicate being the 204Pb concentration in the mineral grain is lower than that in the background of gas flow, ±1σ error was calculated for both the isotopic ratios and ages.
Table 3. Apatite U-Th-Pb Data for the studied sample.
Table 3. Apatite U-Th-Pb Data for the studied sample.
Gr.Intensities (cps)Conc. (ppm)Isotopic Ratios and 1σ ErrorsAge (Ma) and 1σ Errors% dis-c.
204Pb238UTh/U±1σ206Pb/238U±1σ207Pb/235U±1σ208Pb/232Th±1σ206Pb/238U±1σ207Pb/235U±1σ207Pb/206Pb±1σ238U/235U
A20.0917.933.53.30.320.020.070.0090.4880.0080.0210.009407164031242611−1.0
A30.1145.927.62.80.100.010.080.0100.6020.0100.0250.010489234791850613−2.2
A60.0312.419.62.00.200.010.1030.0120.7030.0120.0270.011570315402353213−5.6
A70.0971.234.23.30.060.010.070.0090.5980.0100.0230.0094391847617469127.7
B10.107.019.72.00.160.010.090.0120.7630.0130.0270.0115442857627533135.4
B2 *0.5112.321.40.80.840.040.080.0101.3010.0210.0240.01046821846674801244.7
B30.1015.437.71.40.080.010.060.0080.4070.0070.0200.00836813346939610−6.1
B40.089.064.46.30.040.010.060.0080.4140.0070.0190.00837013352938810−5.1
B50.088.611.50.40.220.010.090.0120.6730.0110.0290.012558305222157614−6.9
B60.0821.318.10.50.160.010.090.0120.7690.0130.0290.0125763257927581150.5
B80.084.031.53.10.080.010.070.0090.5530.0090.0210.009452194471542611−1.0
B90.081.830.63.10.090.010.080.0100.5530.0090.0260.010465204471550913−4.0
C60.097.618.21.90.150.010.0840.0110.7010.0120.0260.0105182654023518133.9
C90.0833.729.92.90.110.010.060.0080.4710.0080.0210.008401153921142311−2.3
D10.084.926.72.80.110.010.080.0100.6340.0100.0240.0104812249919489123.4
D6 *0.0617.994.82.40.010.010.080.0110.1900.0030.0760.030498231772147637−181.9
E30.0845.927.61.00.180.010.080.0100.6350.0110.0220.0094862249919444112.7
E60.5112.446.03.10.560.030.090.0110.7840.0130.0260.0105372758828519138.7
E80.0971.249.81.10.060.010.090.0110.8040.0130.0290.0125402859929576149.8
E100.097.017.00.50.660.030.100.0130.7210.0120.0300.012592335512459415−7.3
F50.1212.325.91.00.300.010.070.0090.5880.0100.0210.0084331847017421117.8
F70.1015.411.12.10.110.010.070.0090.4970.0080.0220.0094081641012445110.5
F90.109.025.30.80.490.020.070.0100.5560.0090.0240.010453194491547812−0.9
F100.088.624.30.70.080.010.080.0110.6640.0110.0280.0114992451721551143.6
G20.1021.313.92.00.130.010.080.0100.6020.0100.0240.0104632047918479123.2
G30.124.037.13.70.120.010.070.0090.4830.0080.0240.009438184001247212−9.4
G40.081.845.33.70.350.020.080.0100.6260.0100.0230.0094672149319466125.4
G50.117.648.11.80.230.010.070.0090.5220.0090.0200.0084191742614390101.6
G60.4133.728.93.00.490.020.070.0090.5790.0100.0230.0094301746416454117.3
G70.084.930.11.10.070.010.070.0090.5450.0090.0240.0104401844215484120.5
G80.1217.949.71.50.630.030.080.0100.5550.0090.0250.010485224481550313−8.2
G90.0845.931.03.00.080.010.070.0100.5320.0090.0250.010452194331449412−4.4
G100.1012.456.84.60.070.010.070.0100.6100.0100.0240.0104511948318482126.7
H20.0871.228.82.90.080.010.080.0100.5880.0100.0250.010476214691750713−1.3
H30.077.013.81.50.150.010.100.0130.8790.0150.0300.0125893364034590158.0
H4 *0.0912.346.43.30.090.010.090.0110.4470.0070.0240.009540283751047412−43.9
H60.0815.424.62.60.200.010.080.0100.5670.0090.0230.009488234561645812−6.9
H90.109.035.51.30.120.010.070.0090.5250.0090.0220.0094151642914435113.2
I10.098.669.31.90.070.010.070.0100.6260.0100.0230.0094622049419456116.5
I20.0821.310.91.10.200.010.090.0120.7890.0130.0290.0115673159128572144.0
I30.584.022.02.20.870.040.070.0090.4720.0080.0220.009425173931143011−8.3
I40.081.832.23.00.110.010.070.0090.5920.0100.0230.0094381847217449117.2
I90.087.612.41.50.280.010.090.0120.7620.0130.0300.012576325752759515−0.2
I10 *0.09337.037.33.30.150.010.080.0100.4830.0080.0160.00647021400123148−17.7
J20.104.919.72.00.140.010.080.0110.6310.0100.0290.012497234971958215−0.2
J50.127.068.82.10.060.010.070.0090.5910.0100.0240.0104311747117477128.6
J60.0812.316.00.70.210.010.090.0110.6840.0110.0230.009534275292245211−0.9
J90.0915.441.64.40.180.010.060.0080.4300.0070.0190.008374133631038010−3.0
A20.089.022.60.80.230.010.070.0090.5990.0100.0220.0094421847717442117.3
Gr = Grains, A1 grain symbol, and B2 * refers to grains with %Discordance ˃10 which were excluded from age calculations, Intensities represent background corrected signal intensities, 3σ calculated for signal background to representing the detection limit of our system, Conc. = concentration by µg/g, ±1σ error for 238U was estimated as 5%, ±1σ error was calculated for both the isotopic ratios and ages, % dis-cord. = % discordance between 206Pb/238U and 207Pb/235U ages.
Table 4. Zircon fission-track ages.
Table 4. Zircon fission-track ages.
Gr.238UNsAρsAge
[ppm](×103 Track/cm2)[Ma]
A192.74.5113814.631939
A8209.49.8105618.136964
B1266.112.4117430.233042
B6217.410.1118524.434640
C7187.58.7132622.737269
D3345.516.1139720.535063
D8326.415.297425.133469
E7243.211.391331.334947
F1144.56.777326.535421
G4144.96.8193449.835115
G5182.28.571418.335154
H7212.59.9124816.033526
Weighted Mean ZFT Age34716
Gr = Grains, A1 grain symbol, U; Uranium concentration in ppm, 1σ = uncertainty of 1-sigma, Ns; number of spontaneous tracks, A; Area in square microns, ρs; density of spontaneous tracks (103 tr/cm2), W.M. age; weighted mean age in million years calculated using IsoplotR [86].
Table 5. Apatite fission-track ages.
Table 5. Apatite fission-track ages.
Gr.238UNsAρsCTLSDθDparSDAge
[µg/g](×103 Track/cm2)[µm][µm]°[µm][Ma]
A233.53.359125.1 10.20.6481.50.0229048
A327.62.83184.0 9.12.1631.50.2027757
A619.62.01863.1 10.31.4401.50.1630277
A734.23.33584.5 10.81.9401.50.1125350
B119.72.01562.6 10.72.8601.60.0825169
B221.40.81243.1 9.80.4601.70.1927785
B337.71.43065.2 10.43.2471.50.0426255
B464.46.3111129.6 7.31.5651.60.0628339
B511.50.41061.7 8.60.5631.60.1928795
B618.10.51142.8 9.40.4451.60.1829995
B831.53.1322655.4 9.63.2751.50.0929263
B930.63.1102205.3 10.30.4541.60.1432737
C618.21.928122.4 10.80.7501.50.1225454
C929.92.93293.7 9.21.5671.50.1723648
D126.72.82364.0 9.50.3551.60.1028466
D694.82.42221812.7 10.10.5661.50.0725720
E327.61.03894.4 9.30.9571.40.0630258
E646.03.12646.7 10.44.2751.50.1027961
E849.81.16288.0 8.92.5521.40.0530750
E1017.00.51882.3 10.00.3411.70.1326167
F525.91.02893.2 9.71.5231.50.0723851
F711.12.1741.8 12.12.5501.50.05310121
F925.30.83493.9 10.90.7441.50.0229458
F1024.30.71243.1 9.11.2671.60.1424475
G213.92.01362.2 9.40.2631.50.0730790
G337.13.744104.5 9.61.5741.40.1523536
G445.33.798166.3 9.41.3551.60.1526730
G548.11.86597.5 9.93.1751.50.1429638
G628.93.051124.4 9.72.6781.50.1729050
G730.11.150124.3 9.90.8471.40.0427347
G849.71.5165208.5 11.62.6501.70.0732731
G931.03.04194.7 11.41.8571.50.0829054
G1056.84.6116158.0 10.41.1561.30.0526933
H228.82.92965.0 9.92.1491.40.0433070
H313.81.532122.2 9.81.2681.50.1230454
H446.43.35386.1 10.00.8541.60.0725143
H624.62.63763.8 10.72.6731.20.2729657
H935.51.331155.3 10.22.1561.30.0928759
I169.31.95599.5 11.02.4611.70.3626144
I210.91.114101.6 9.80.5541.60.1628180
I322.02.21162.8 11.81.2471.50.2924779
I432.23.04865.5 10.72.3471.60.0932657
I912.41.51792.0 10.12.5731.50.1830179
I1037.33.32045.2 12.62.6551.40.0526565
J219.72.02493.1 9.60.7631.40.2230164
J568.82.17199.2 10.71.1671.40.0725540
J616.00.73742.5 10.41.7471.40.1230350
J941.64.45785.9 12.01.7631.50.0927139
Weighted Mean AFT Age2837
Gr = Grains, A1 grain symbol, U; Uranium concentration in ppm, 1σ = uncertainty of 1-sigma, Ns; number of spontaneous tracks, A; Area in square microns, ρs; density of spontaneous tracks (103 tr/cm2), CTL; confined track lengths, SD; Standard deviation of CTLs, θ; Angle from the c-axis in degrees, W.M. age; weighted mean age in million years calculated according to the equation in [18].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mansour, S.; Hasebe, N.; Azab, E.; Elnaggar, A.Y.; Tamura, A. Combined Zircon/Apatite U-Pb and Fission-Track Dating by LA-ICP-MS and Its Geological Applications: An Example from the Egyptian Younger Granites. Minerals 2021, 11, 1341. https://doi.org/10.3390/min11121341

AMA Style

Mansour S, Hasebe N, Azab E, Elnaggar AY, Tamura A. Combined Zircon/Apatite U-Pb and Fission-Track Dating by LA-ICP-MS and Its Geological Applications: An Example from the Egyptian Younger Granites. Minerals. 2021; 11(12):1341. https://doi.org/10.3390/min11121341

Chicago/Turabian Style

Mansour, Sherif, Noriko Hasebe, Ehab Azab, Ashraf Y. Elnaggar, and Akihiro Tamura. 2021. "Combined Zircon/Apatite U-Pb and Fission-Track Dating by LA-ICP-MS and Its Geological Applications: An Example from the Egyptian Younger Granites" Minerals 11, no. 12: 1341. https://doi.org/10.3390/min11121341

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop