Next Article in Journal
An Assessment of the Role of Combined Bulk Micro- and Nano-Bubbles in Quartz Flotation
Next Article in Special Issue
Types and Genesis of Siderite in the Coal-Bearing Beds of the Late Permian Xuanwei Formation in Eastern Yunnan, China
Previous Article in Journal
Machine Learning Methods for Quantifying Uncertainty in Prospectivity Mapping of Magmatic-Hydrothermal Gold Deposits: A Case Study from Juruena Mineral Province, Northern Mato Grosso, Brazil
Previous Article in Special Issue
The Kaolinite Crystallinity and Influence Factors of Coal-Measure Kaolinite Rock from Datong Coalfield, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tectono-Thermal Events of Coal-Bearing Basin in the Northern North China Craton: Evidence from Zircon–Apatite Fission Tracks and Vitrinite Reflectance

1
College of Mining Engineering, Taiyuan University of Technology, Taiyuan 030024, China
2
College of Geoscience and Surveying Engineering, China University of Mining and Technology, Beijing 100083, China
3
Wuxi Research Institute of Petroleum Geology, SINOPEC Petroleum Exploration & Production Research Institute, Wuxi 214126, China
*
Author to whom correspondence should be addressed.
Minerals 2022, 12(8), 942; https://doi.org/10.3390/min12080942
Submission received: 13 April 2022 / Revised: 20 July 2022 / Accepted: 21 July 2022 / Published: 26 July 2022
(This article belongs to the Special Issue Geochemistry and Mineralogy of Coal-Bearing Rocks)

Abstract

:
In order to further reveal the tectonic activity of the central and northern North China Craton (NCC) since late Paleozoic, the Datong coal-bearing basin was selected as the research object. The tectono-thermal events and uplifting cooling events of the basin were retrieved through zircon and apatite fission tracks and vitrinite reflectance measurements. The research shows that the Datong coal-bearing basin experienced three tectono-thermal events with ages of 245–207 Ma (middle–late Triassic), 179 ± 9 Ma (early Jurassic), and 140 Ma to 78 ± 11 Ma (middle–late Cretaceous), respectively. That just coincides with the lamprophyre activity, Kouquan fault activity, and Zuoyun basaltic andesite magmatic activity which surround the Datong coalfield. The basin also experienced three uplift events with the peak ages of 202 ± 18 Ma (late Triassic), 157 ± 7 Ma (late Jurassic), and 45 ± 3 Ma or 36 ± 3 Ma (middle Eocene), respectively. The Datong Permo-Carboniferous and Jurassic coal vitrinite reflectance proved that the average metamorphism temperature is 104–108 °C, even reaching 163–367 °C. The fission track results showed that the paleotemperature was even higher than 170–250 °C from 117 to 282 Ma and 80–120 °C from 20 to 68 Ma, in the Datong coal-bearing basin. The results show that the deep tectonic activities of the NCC were still active in the Mesozoic and even Cenozoic Paleogene.

1. Introduction

Fission-track is an effective thermochronological method which developed in the 1960s to obtain geochronological data. It can be used to determine geological ages by applying the decay rate of radioactive materials and tracking the quantities in a unit area produced during nuclear fission [1,2,3,4,5,6,7,8]. This technique has been widely used to analyze the uplift and cooling history of orogenic belts, to recover the thermal evolution and burial subsidence history of basins or provenance areas, to study the basin–mountain coupling relationship, fault activity time, hydrothermal metallogenic age, and other geological events. With the support of the research fruit, a new field of studying mineral-fission track annealing, using nuclear fission technology, has emerged [9,10,11,12,13,14,15,16]. Furthermore, the random huminite/vitrinite reflectance and apatite fission-track methods could be useful parameters for evaluating the paleotemperature of thermal and geodynamic evolution, and the burial history of coal-bearing sequences [17,18,19,20,21,22,23,24,25,26].
The Datong coalfield is located in the northern North China Craton (NCC) [27], between the west NCC and east NCC (Figure 1a,b). The Datong coalfield has a mining history of over 1500 years. It hosts Carboniferous–Permian and Jurassic coal-bearing strata.
In recent years, many researchers have studied the sedimentary, tectonic, and magmatic activities that have occurred since the late Paleozoic in the Datong area of the Shanxi Province [29,30,31,32,33]. They have confirmed that the Datong basin is still part of the NCC large sedimentary depression basin (i.e., the Ordos Basin) in the early–middle Triassic, although the NCC was subjected to the tectonic compression of the Yangtze and Siberian Plates in the late Triassic [34,35,36]. Because of the influence of the Siberian Plate, the northern Datong basin underwent a gradual uplift and denudation, which resulted in the late Paleozoic and Triassic strata becoming denuded in some areas [32,33,37,38]. During field explorations and coalfield drilling, numerous alkaline mafic–ultramafic lamprophyre sills and diabase dikes and a few carbonatite dikes were found in this basin, intruding the Taiyuan Formation or Shanxi Formation strata (Figure 1c and Figure 2) [39,40,41,42].
The aforementioned studies indicated multistage magmatic activity controlled by regional tectonic movements since the late Paleozoic, that has resulted in coal-bearing strata hiatus and magmatic intrusion. However, the specific characteristics and time intervals of these events are still unclear. Therefore, we used the fission track thermochronology of zircon and apatite to reconstruct the sedimentary–tectonic–magmatic events in this basin since the late Paleozoic. Furthermore, we attempted to elucidate the influences and restrictions of these events on the Carboniferous–Permian–Jurassic coalification and reveal the coalification evolution processes and coal metamorphism mechanisms. The results are significant for restoring the eastern boundary of the proto Triassic Ordos basin and helpful in understanding the surface response of the deep craton activity that was caused by the destruction of the NCC during the Mesozoic.

2. Geological Setting

As a part of the late Paleozoic giant coal-accumulating basin of the NCC, the Datong coal-bearing basin experienced weathering and denudation for up to 140Ma after the Ordovician, and was redeposited in the early late Carboniferous. The basement of the basin is Archaean Sanggan gneiss, and the Proterozoic, upper Ordovician, Silurian, Devonian, lower Carboniferous, Triassic, upper Jurassic, and upper Cretaceous strata are absent in the basin under the influence of multi-stage tectonic magmatic activities [43]. The lower Jurassic Yongdingzhuang Formation and the lower Cretaceous Zuoyun Formation are all in contact with the underlying strata in angular unconformity. The Permo-Carboniferous Taiyuan Formation, Shanxi Formation, and Jurassic Datong Formation are the main coal-bearing strata [44,45,46], in which, the Nos. 3, 5, and 8 (C8) coal seams are present in the Taiyuan Formation and the No. 4 (PS4) coal is in the Shanxi Formation; also, 15 other coal seams (Jd) in the Datong Formation (Figure 2). The dominant sedimentary environment changes from the coastal delta facies of the Taiyuan Formation to the river facies of the Shanxi Formation and the lake facies of the Datong Formation.
After Jurassic, the Datong Basin became the eastern margin of the wide and gentle folds of the Ordos Basin [47]. The basin boundary is marked by four tectonic features (Figure 1): the Hongtaoshan anticline (NW 50°) to the southeast; the Kouquan–Emaokou fault to the east; the Qingciyao reverse fault to the northeast; and the Maihutu–Weiyuanbao fault to the northwest. Only the gently plunging folds developed in the coalfield (Figure 1c).
Extensive alkaline mafic–ultramafic lamprophyre and carbonatite dikes intrude the Permo-Carboniferous coal strata (Figure 2 and Figure 3) [48]. The K–Ar and Rb–Sr geochronological ages of the lamprophyre biotite are 242 ± 53 Ma, 229 ± 11 Ma, and 224–198 Ma [38,48]. The intrusion age of the carbonatite dyke associated with the lamprophyre is 132.9 Ma [49], and the whole rock K–Ar age is 131–133 Ma. Additionally, a suite of early Cretaceous basaltic andesite covering an area of almost 3 km2 is emplaced in the Zuoyun Formation nearby the Zuoyun–Jiugaoshan district (Figure 1c), along the western side of the basin [31].

3. Samples and Methods

The random vitrinite reflectance (%Ro) method was applied to quantify the maximum paleotemperature experienced by the Permo-Carboniferous and Jurassic stratum. The original No. 8 coal of the Taiyuan Formation (C8) and No. 4 coal of the Shanxi Formation (PS4) samples were collected from the subsurface. However, there are not many Jurassic coal resources left, and only a part of the Jurassic coal samples was collected from the underground of the Jinhuagong mine (JJ), Qinciyao mine (JQ), Sitai mine (JS), Xinzhouyao (JX), and Yanya mine (JY) (Figure 3). Every coal maceral section was made using a cool forming method after the whole-rock samples were crushed into 0.1 to 1.0 mm and then polished with white fused alumina powder and aluminum oxide polishing slurry, following the coal petrography procedure standard [50]. Then, the %Ro values of the coal were measured, using an Olympus BX51 microscope equipped with a Y-3 spectrophotometer. For measuring the maximum reflectance on a particulate sample, 50× oiled objectives were used with a peak transmittance in the range of 546 ± 5 nm and a half-peak transmittance band of less than 30 nm. More than 80 measurements per sample were measured. The maximum reflectance was taken after the microscope stage was rotated 360° (ISO 7404-5:2009) [51]. The maceral classifications and terminology applied in the current study are based on ISO 7404-4: 2017 and the ICCP System, 1994 [50,52].
The random vitrinite reflectance (%Ro) is an important index for coal rank determination and the thermal history analysis of sedimentary basins, and is not affected by later tectonic uplift and exhumation [21,22,53,54,55]. Its value increases with increasing temperature and burial depth. Thus, it can record maximum paleotemperature [56,57]. Therefore, the changing trends of the %Ro values under unconformity surfaces can be used to restore the tectonic uplift history and stratum denudation thickness [23,55,58,59,60,61,62]. Many methods use %Ro to study burial paleotemperatures, such as the Hood diagram [63], Karweil diagram [64], Barker formula [54,65,66,67], vitrinite reflectance gradient method, fitting calculation TTI method, and an improved model based on the TTI method [68,69,70,71,72,73,74].
Barker and Pawlewicz [54,65] proposed that calculating the paleotemperatures using the mean %Ro is relatively simple, and it is now commonly used (Equation (1)) [65]. The advantage of this equation is that considering the heating duration variables is not necessary, which is suitable for the geological evolution of organic matter with maximum burial temperatures Tpeak between 25 °C and 325 °C and the %Ro of vitrinite between 0.2% and 4.0% [66]. Moreover, it has been proposed that, in the case of contact metamorphism, the accurate maximum paleotemperature could be obtained only when the coal seam is 0.3 times wider than the intrusive body or when the maximum paleotemperature is lower than 300 °C. In addition, Yao et al. [75] suggested that the paleotemperature value may be higher if this equation is used when the organic matter has achieved high maturity [75]:
Tpeak = [ln(Ro) + 1.19]/0.078
In this paper, %Ro,max was used to replace %Ro (Equation (2)), where %Ro,max is the mean maximum reflectance of organic matter and Tpeak is the maximum burial temperature:
Tpeak = [ln(Ro,max) + 1.19]/0.078
For the fission track analyses, three groups of sandstone samples were collected from the Xiaoliaotangou and Sigou outcrop sections in northern Kouquangou of the Datong coal-bearing basin (Figure 2). The samples included medium–coarse-grained sandstone (K2) (Figure 4a) from the base of the Taiyuan Formation (C2t) of the upper Carboniferous, medium–fine-grained sandstone (K3) (Figure 4b) from the bottom of the Shanxi Formation (P1s) of the lower Permian, medium–coarse-grained sandstone (K11) (Figure 4c) from the base of the Datong Formation of the middle Jurassic (J2d), and compact basaltic andesite samples (Figure 4d) from the lower Cretaceous Zuoyun Formation (K1z) in Jiugaoshan of Zuoyun county (Figure 1). The four group samples were collected from different field sites of each stratum to obtain reliable analysis results. The ZFT analysis results were obtained from three groups of sandstone samples. The AFT analysis results were obtained from the Datong Formation and Shanxi Formation sandstone samples and the basaltic andesite samples from the Zuoyun Formation.
The basic theory of fission tracks is founded on the annealing characteristics of mineral fission tracks. Fission tracks of natural minerals enriched in 238U can only be preserved below a certain critical temperature (closure temperature). Following the increase in temperature and heating time, the density and length of the tracks decrease and shorten, until they disappear completely, which is called annealing. The closure temperatures of the zircon fission track (ZFT) and the apatite fission track (AFT) are 210 ± 40 °C and 100 ± 20 °C, respectively [76]. These suggest that the temperature range of the zircon partial annealing zone is generally 170–250 °C, the temperature range of the apatite partial annealing zone is generally 80–120 °C, and the complete annealing temperatures of zircon and apatite are 250 °C and 120 °C, respectively. The fission-track ages of zircon and apatite record the time when the minerals cooled to below 210 °C and 100 °C, respectively, which are called the cooling ages [77,78]. Because the annealing processes of mineral fission tracks are only related to the annealing temperature and time and have no obvious relationship with annealing physical and chemical conditions, such as pressure, pH, and Eh, the annealing degree of the fission tracks can be regarded as a function of the mineral sealing temperature and time. The fission-track age measurement mainly adopted the direct testing method before the 1980s. More recently, it has used the zeta constant calibration method, which has a smaller measurement error.
The fission-track samples were analyzed at the State Key Laboratory of Geological Process and Resources, China University of Geosciences (Beijing, China). The Track-key [79] software was used to treat the data. After being observed under a Lecia DM4P microscope, the grains were crushed to 60 mesh, the samples were randomly selected and dried, and the grains were purified, using conventional magnetic separation and gravity separation to separate their zircon and apatite mineral grains. The ZFT testing process is as follows. First, zircon grains were placed on polytetrafluoroethylene propylene plastic slices, made into thin slices, and polished to reveal the inner surface of mineral grains. Second, the thin slices were etched with a high-temperature melt of KOH + NaOH for 20–35 h, to reveal spontaneous tracks at 210 °C. Subsequently, a low-uranium muscovite slice was used as an external detector to cover the optical slice, which contacts closely with the inner surface of the grains and is irradiated by thermal neutrons along with a CN2 standard uranium glass. Finally, the induced tracks were revealed by etching for 20 min in 40% hydrofluoric acid (HF) at 25 °C [10]. On another note, the AFT testing process is as follows. First, the apatite grains were placed on a glass slice, dripped with epoxy resin, and were ground and polished to expose the inner surface of the mineral. Second, the spontaneous track was revealed by etching 7% HNO3 at 25 °C for 30 s. Then, a low-uranium muscovite external detector was incorporated into the reactor for irradiation. After that, the induced tracks were revealed by 40% HF etching at 25 °C for 20 s. The neutron flux was calibrated using a CN5 uranium glass [10]. The age values were calculated, according to the N constant method recommended by the International Union of Geological Sciences and the standard fission-track age equation. The detailed experimental procedures and principles can be referred to in the literature [10,80].
In this study, the zeta constants of apatite and zircon are 392 ± 18.7 and 102.1 ± 3.8, respectively [81]. The results of the sample analysis were tested, using Chi-square statistics, and P2) was the test probability. P2) > 5% usually indicates that the ages of single mineral granules belong to the same age group, reflecting the same thermal history activity; thus, the pooled age is used. P2) < 5% suggests that a sample has not passed the test, indicating that the ages of the samples may not be of a single age group; thus, the central age is applied. In addition, the pooled age is still applied if the single grain age of the sample is 0 Ma [10].

4. Results

4.1. Random Vitrinite Reflectance and Paleotemperature

Considering the characteristics of coal metamorphism in the Datong coalfield, the %Ro,max of the Permo-Carboniferous coal seam is below 0.96, suggesting that Barker’s formula for calculating paleotemperatures is suitable in this study. The %Ro,max data were selected from the test results of the No. 8 coal of the Taiyuan Formation (C8 coal, 76 normal coal samples, and 12 contact metamorphism coal samples, a total of 88 coal samples) and the No. 4 coal of the Shanxi Formation (PS4 coal, 35 normal coal samples, and 11 contact metamorphism coal samples, a total of 46 coal samples). Figure 3 shows the sampling location.
The %Ro,max of the C8 normal coal is 0.57%–0.96% (mean = 0.74%) (Table 1), with a Tpeak of 80–147 °C (mean = 114 °C). The %Ro,max of the PS4 normal coal is 0.59%–0.83% (mean = 0.71%) (Table 2), with a Tpeak of 83–128 °C (mean = 108 °C). On the basis of the %Ro,max of the 12 contact metamorphism coal samples from the C8 and 11 samples from the PS4, the highest paleotemperatures recorded were approximately 255–362 °C (Table 3) in C8 and 163–367 °C in PS4. The highest Tpeak could cause the zircons to anneal and produce fission tracks, with the zircons even annealing completely until the fission tracks disappear. This indicates that the temperature could be higher than the annealing temperature of zircon.
We thoroughly collected the %Ro,max of the mining coal in the Jurassic Datong Formation, including 10 subsurface samples, one literature sample data [82], and 91 measurements from the exploratory geology report (Table 4). As can be inferred from Table 4, the %Ro,max of the Datong Formation coal ranges from 0.51% to 0.94%, with an average value of 0.71%. According to Barker’s formula for calculating Tpeak values, the maximum values recorded in the Datong Formation coal ranged from 66–144 °C (mean = 108 °C). The mean %Ro, max of the Datong Formation coal is slightly lower than that of C8, but similar to that of PS4. The similar %Ro,max of the Permo-Carboniferous and Jurassic coal seams in the Datong coalfield indicated that they experienced similar maximum paleotemperatures.

4.2. Zircon Fission Tracks (ZFT)

The samples from the Taiyuan Formation (KCZ) and the Shanxi Formation (KPZ) were analyzed, using Zircon Fission Tracks. All of the fission-track ages were lower than the stratigraphic age, indicating that the rocks experienced a higher paleotemperature and zircon had undergone annealing (Figure 5). Although some of the ZFT ages from the Datong Formation (KJZ) sandstone were older than the stratigraphic age, they can be used to analyze tectonic thermal events or to predict provenance.
The P2) test of the KJZ zircon samples was greater than 5%, indicating that they underwent the same thermal event. Therefore, the pooled age of the samples is presented as 157 ± 7 Ma (193–128 Ma) (Table 1, No. 3, and No. 6 in Figure 5a). According to the age analysis (Figure 5a), a total of 24 zircon detrital grains could be measured in the Datong Formation, with 20 grains younger than the host strata, with ages ranging from 128 to 174 Ma (No. 19 in Figure 5a), and four zircon grains older than the host strata, with ages ranging from 177 (No. 19 in Figure 5a) to 193 Ma.
The P2) tests for the KPZ and KCZ were less than 5% (Table 4), suggesting that these zircon grains have undergone more than two tectonic thermal events. Therefore, the central ages of the samples are presented as 202 ± 18 Ma (279–118 Ma) for the KPZ (Figure 5b) and 179 ± 9 Ma (282–117 Ma) for the KCZ (Figure 5c), with the approximate peak ages being 180–190 Ma (Figure 5f). Twenty-seven zircon detrital grains could be measured in the KPZ, with three particle fission-track ages greater than their host strata ages (836, 946, and 1028 Ma) (No. 27, No. 26, and No. 25 in Figure 5b). The KCZ fission-track result shows that their ages are less than those of the host strata, indicating that the samples were partially or completely annealed and that the strata paleotemperature was even higher than 180 °C [83].
According to the fission-track age diagram (Figure 5a–c), the heating ended at approximately 118–117 Ma. One zircon grain heating age from the KPZ is 279 Ma (No. 24 in Figure 5b), which is the earliest time from when the host stratum started experiencing high temperatures. Starting from 245 Ma (No. 20), the grains’ heating age frequency increased, with those of seven grains ranging from 245 to 207 Ma (No. 17) with the final cooling age of 118 Ma (No. 5). Conversely, the earliest thermal age in KCZ is about 282–277 Ma (No. 14 and No. 15 in Figure 5c); the grains’ heating age frequency increased from 232 Ma (No. 2), with five particle ages ranging from 232 to 209 Ma (No. 26) with the final cooling age of 117 Ma (No. 28). The age range of the zircon grains from the Shanxi Formation is similar to that of the grains from the Taiyuan Formation. The heating age of the zircon grains is substantially younger than that of the host strata, potentially reflecting the heating history of the former. The common heating age ranges of the ZFTs from the Shanxi and Taiyuan Formations indicate strong tectonic thermal events in the Datong coal-bearing basin during the middle to late Triassic.
Using fission-track analysis and the BINOMFIT [84] simulation software, an age-frequency distribution diagram of the ZFT is depicted (Figure 5d–f). We analyzed the uplift cooling history and tectonic thermal events of the Datong basin combined with the distribution and occurrence characteristics and sedimentary evolution history of its strata. The zircon ages of the three groups’ sandstone samples show that the primary age of the KCZ ranges from 166 (No. 27) to 232 Ma (No. 2), with a peak age of approximately 180 Ma, a weak peak age of approximately 210 Ma (Figure 5f), and another peak age of 151–133 Ma (No. 4 and 6) (Figure 5a–c). The main cooling age of the KPZ ranges from 162 (No. 2) to 245 Ma (No. 20), with an obvious peak age at approximately 180 Ma, a weak peak age of 210 Ma (Figure 5e), and another peak age at 140–128 Ma (No. 4 and 10) (Figure 5b–e).
In contrast with the KCZ and KPZ, the KJZ only shows one peak age at approximately 150–131 Ma (Figure 5d, No. 20, No. 23) and no peak age at 180 Ma. However, the KJZ began to experience thermal interactions after 193–182 Ma, according to the radar diagram (No. 4 and 10) (Figure 5b–e). Combining the analog curve and the radar diagram (Figure 5) shows that the KCZ, KPZ, and KJZ samples experienced significant thermal events at approximately 180 Ma (193–177 Ma) and 151–140 Ma. Other thermal events occurred at approximately 210 Ma and from 245–232 Ma in the KCZ and KPZ. Additionally, some of the grains retained the age information of detrital provenance in the KPZ, with a peak age of approximately 800 Ma.
The above results indicate that the KCZ and KPZ underwent at least three annealing processes, with one significant thermal event at approximately 180 Ma (179 ± 9 Ma), and two less distinct thermal events at approximately 210 Ma (202 ± 18 Ma) and 151–140 Ma. In contrast, the KJZ underwent the last annealing process at approximately 150 Ma (157 ± 7 Ma). Because the thermal age of the KJZ begins at approximately 193–182 Ma, which is older than the sedimentation age of the Datong Formation, its source may have undergone the same thermal events that the KCZ and KPZ underwent. Although the lacustrine conditions in the Datong Basin were common during the precursor peat accumulation, clastic influx into the depositional environment seems to be possible. Therefore, further provenance analysis should be completed in the future.

4.3. Apatite Fission Tracks (AFT)

In this study, we applied 80–120 °C as the temperature of the AFT annealing zone [78,85]. The AFT results show that the P2) test is significantly greater than 5% (Table 4), manifesting similar thermal events. Therefore, the pooled ages of the KJA and KPA samples are 45 ± 3 Ma and 36 ± 3 Ma, respectively, which are considered valid (the Paleogene Eocene). The AFT ages are much younger than those of the host strata. The samples underwent annealing and subsequent uplift cooling, which could be used for the uplift activity analysis. In addition, the apparent age and track length of the apatite from the Datong Formation (KJA) and Shanxi Formation (KPA) samples are comparable and are proven reliable. The average track lengths of the KJA and KPA are 12.1 ± 2.2 µm and 12.0 ± 2.2 µm, respectively, revealing a single peak pattern (Figure 6d–f), with more moderate track lengths and a wider track distribution. These correspond to the P2) values of 98.1% and 100%, respectively, and can represent a single thermal event. These results suggest that the maximum temperatures of the KJA and KPA samples were over 120 °C and that the fission track of the apatite underwent complete annealing, with a long residence time in the annealing zone.
According to the fission-track annealing theory, when the sedimentary diagenetic time is experienced earlier, the burial depth will be deeper with the higher maximum paleotemperature, and the annealing age will also be younger [38]. The age range of the KJA is 68–21 Ma (Figure 6a,d), whereas that of the KPA is 49–20 Ma (Figure 6b,e). The minimum thermal age of a single particle is the same in both the KJA and KPA. In contrast, the maximum age decreased considerably with greater depths, suggesting that the increase in burial depth and the temperatures in later periods of diagenesis may be the primary reasons for AFT annealing. However, the annealing ages of the different strata were completed simultaneously from 21 (No. 10 in Figure 6a) to 20 Ma (No. 21 in Figure 6b), indicating that both of the strata had undergone the same tectonic activity (uplifted to the near surface) during this time. The annealing age difference of the apatite at the start in the KPA and KJA is nearly 19 Ma, the pooled age difference is 9 Ma, and the annealing age difference at the end is only about 1 Ma, indicating that the uplift rate and the stratum burial depth changed during this period.
The basaltic andesite (KKA) from the Zuoyun Formation could not be used for the thermal history simulation because it contained fewer grains (Figure 6c,f), and displayed no statistical fission-track length. However, its corresponding P2) value is 98.9%, indicating that its age has reference significance. The age of the KKA ranges (Figure 6c) from 140 (No. 8) to 39 Ma (No. 11), with central and pooled ages of 78 ± 11 Ma. This suggests that the apatite grains annealed or completely annealed from the late Jurassic–early Cretaceous to the Eocene. The paleotemperature during this stage should be higher than the lowest limit of the annealing temperatures (80 °C), indicating that the tectono-thermal events occurred in the late Cretaceous, which is nearly the same age as that of the eruption of the Zhuozi basalt in Liangcheng, Inner Mongolia [86]. This means that the paleotemperature of the coal seams in the Jurassic Datong Formation and the Carboniferous–Permian Shanxi and Taiyuan Formations, which underlie the Cretaceous basaltic andesite, was at least 80 °C.
The AFT age of the KKA indicates that a tectonic thermal event may have occurred in the western coal-bearing basin during the late Cretaceous at 78 ± 11 Ma. The basin uplifted and eroded the Cretaceous strata. The AFT ages of KPA and KJA indicate that a tectonic thermal event occurred in the eastern coal-bearing basin at approximately 45–36 Ma during the Eocene. This age is close to the formation time of the Sanggan River fault depression and the Shanxi Graben system in northern Shanxi Province.

5. Discussion

According to the ZFT and AFT results, the following events were recorded: (1) a tectono-thermal event at 245–207 Ma; (2) a slight cooling event with a peak age of 210 Ma; (3) a thermal event with a peak age of 180 Ma; (4) a cooling event range from the late Jurassic to early Cretaceous (157 ± 7 Ma), during which the basin was uplifted and the earlier Cretaceous strata eroded; (5) a tectono-thermal event occurring in the western coal-bearing basin at 140–78 ± 11 Ma; and (6) an uplift event in the eastern basin at 45–36 Ma, indicated by the KPA and KJA samples’ AFT age data. In addition, the Tpeak values that were calculated using the vitrinite reflectance equation surpassed the ZFT and AFT annealing temperatures since the late Paleozoic. The following section discusses how these events are related to regional tectonic activity.

5.1. Tectonic–Magmatic Activity

As mentioned previously, the Datong coal-bearing basin experienced three tectonic–magmatic thermal events since the late Paleozoic [87]. The first event was related to lamprophyre magmatic activity, the second was linked to the Kouquan–Emaokou fault zone activity, and the third was related to a basaltic andesite magmatic activity.
Previous studies indicated that numerous alkaline basic–ultrabasic lamprophyre sills and carbonatite dikes intruded the Permo-Carboniferous coal-bearing strata in the Datong basin, even densely affecting the coal quality (Figure 2 and Figure 7b–f) [88]. The magmatic activity occurred at approximately 229 ± 11 Ma and cooled at approximately 224–119 Ma [38,48,49]. The results of the ZFT analysis show that the peak age of the first cooling interval was approximately 210 Ma, whereas 117–118 Ma was the earliest time that zircon could have reached the annealing temperature (120 °C). These sequences of heating–cooling stages coinciding with a large-scale lamprophyre magmatism indicate that the first tectono-thermal event since the late Paleozoic in the Datong area was the lamprophyre magmatism in the middle–late Triassic. According to the radar diagram of the ZFT (Figure 5a–c), the annealing temperatures were approached at approximately 245–232 Ma, which may be the earliest time recorded for magma initiation. It is believed that the magma was the product of the partial melting of the lithospheric mantle [48]. The researcher proposed that a deep fault formed in the early Indosinian period was the magma’s intrusion channel. The ZFT record of the Ningwu Basin sample [46] has a front peak age of 210 Ma, which reflects the Indosinian tectono-thermal events and coincides with the lamprophyre activity in the Datong basin. This means that the lamprophyre magmatism started at 245 Ma, and another event followed, with a peak age of 210 Ma. The latter event simultaneously developed in the Ningwu Basin. This inference indicates that the thermal event occurred earlier in the Datong basin than in the Ningwu basin, explains why the C8 and Jurassic coal have similar %Ro, and proves that the Triassic stratum were deposited around Datong.
During the Yanshan Movement Period, strong tectonic activity began in the Kouquan of Datong in the Early Jurassic, then forming the Kouquan–Emaokou overthrust fault belt [45]. The most obvious cooling age in the ZFT analysis was approximately 180 Ma, which coincides with the strong activity of the Kouquan–Emaokou fault, sufficiently proving that a tectono-thermal event had occurred in the Datong area since the Early Jurassic. The relative active time was from 193 to 173 Ma (No. 2 and 6 in Figure 5b), with a peak age of 179 ± 9 Ma, and persistent activity was observed from the early Jurassic to the middle–late Jurassic. Meanwhile, a strong intracontinental polydirectional compression was consistently recorded in the middle and late Jurassic (174–145 Ma) in the western Shanxi and the entire tectonic belt boundary of the Ordos basin [89]. Furthermore, the age of the pyroclastic rock in the lower Jurassic Yongdingzhuang Formation (J1y) is 179.2 ± 0.79 Ma [90], whereas the age of the tuffaceous carbonate rocks at the top of the middle Jurassic Yungang Formation (J2yg) is 160.6 ± 0.55 Ma [29,90]. These represent the initiation time of the Jurassic tectonic or magmatic activity (Yanshan Movement). These cooling ages record the tectono-thermal event of the reactivated Kouquan fault belt and are also consistent with the regional tectonic background, where a large-scale fault belt exists from Lishi along Ningwu, Huairen, and Datong to Tianzhen and Zhangjiakou region (Figure 7a).
The AFT age of the basaltic andesite reflects that thermal magmatic activity occurred around the early Cretaceous, beginning at 140 Ma and with a peak cooling age of 78 ± 11 Ma. The basaltic andesite is commonly distributed in the Jiugaoshan area of Zuoyun county along the western margin of the Datong basin, and occurs as stratiform or stratiform-like in the Cretaceous Zuoyun Formation. Consequently, the AFT of the Datong Formation and Shanxi Formation scarcely recorded this event. It is known that the basaltic andesite and the Bainvyangpan Formation volcanic rocks in Liangcheng–Zhuozi county (Figure 7), Inner Mongolia, are the volcanism result of the same period [86], characterized as 126.9-139.1 Ma in age from the K–Ar isotopic age data. The stratigraphic sequences and chronological data of the basaltic andesite indicate that its magmatic active age is the late Mesozoic (Cretaceous) of the Yanshanian, suggesting that magmatic activity remained active in the north–central part of the NCC. The latest zircon U–Pb isotope-sensitive high-resolution ion microscopy and laser-ablation inductively coupled plasma mass spectrometry dating results of the Zijinshan alkaline complex in the western Shanxi flexural fold belt are 125–138 Ma [91,92,93]. The intrusion age of the Huyanshan alkaline monzonite in the northwest Jiaocheng county, Jinzhong City, is 114–127 Ma [94]. The isotopic age of the quartz monzonite that intruded the middle Jurassic strata in the Nangufeng drilling in Qixian county (Figure 7) is 141 Ma [86], coinciding with that of the Datong basaltic andesite. The paleotemperature was restored using vitrinite reflectance, and fission-track analysis also showed that a tectono-thermal event occurred in the Ordos Basin and Qinshui Basin of the Shanxi Province from the late Jurassic to the early Cretaceous (150–140 Ma) [95,96].
The frequent tectonic magmatic activity in the study area during Mesozoic is related to the regional lithosphere thinning. This is also an important manifestation of the tectonic system transformation, from the extensional tectonic to the contractional tectonics of the Central and western NCC since late Jurassic to early Cretaceous. The AFT age of the KKA is 140 Ma in the Datong basin, indicating that it responded to this magmation during the early Cretaceous, but this ended in the late Cretaceous at 66 Ma (Figure 6c, No. 2).

5.2. Uplift Cooling Events

Since the late Paleozoic, there have been three major uplift events in the Datong basin (Figure 8), represented by: (1) the absence of Triassic strata with an unconformity between the early Jurassic Yongdingzhuang Formation and its underlying Carboniferous–Permian rocks; (2) the absence of the Tianchihe Formation of the late Jurassic with an unconformity separating the early Cretaceous Zuoyun Formation or Quaternary from the Jurassic rocks; and (3) the absence of late Cretaceous and Paleogene rocks marked by an unconformity between the Neogene and underlying older deposits. Ren et al. [97] also thought that the southeastern Ordos basin experienced two denudation stages in the late Triassic–early Jurassic and the late Cretaceous [96].
Because of the tectonic–magmatic activity during the Mesozoic, the ZFT did not clearly show this uplift cooling event, which may be related to the large-scale intrusion of lamprophyres in the extensional tectonic setting. The partial uplift of the coal-bearing basin led to the erosion of the early Triassic strata. The uplift event could have lasted until the deposition of the Yongdingzhuang Formation in the early Jurassic. Considering that the Datong coalfield is located in the northeast Ordos basin and the eastern part of the north Lvliang Mountains, Liu et al. proposed that the Datong and Ordos coal-bearing basins should be considered as one large basin separated during the Lvliang Mountain uplift in the Triassic and Jurassic [81]. However, the thickness of the Triassic strata can reach more than a thousand meters in Ningwu, but only several hundred meters around the Datong area. On the basis of the ZFT analysis, the Datong area uplifted after a local magmatic activity at 202 ± 18 Ma.
Figure 6 shows that the age range of the AFT of KJA is 68–21 Ma, with a peak age of 45 ± 3 Ma. The age range of the KPA AFT is 20–49 Ma, with a peak age of 36 ± 3 Ma. The KJA and KPA apatite age data prove that an uplift cooling event occurred in the Datong area during the Eocene, with a time interval of 45 ± 3 Ma to 36 ± 3 Ma. The Eocene uplift cooling event is reflected specifically in the unconformity contact between the Neogene and underlying strata around the Datong basin. This result also demonstrates that a third uplift cooling event occurred around Datong since the late Paleozoic, characterized by a strong uplifting and rifting in the west and east Kouquan–Emaokou Fault zones, respectively (Figure 9). Subsequently, the Sanggan River downfaulted basin began to form, and its strong uplift occurred from 45 ± 3 Ma to 36 ± 3 Ma.
As proposed by Zhao [46], the AFT ages in the Ningwu–Jingle syncline are 62 ± 4 Ma and 70 ± 6 Ma, respectively. These ages are consistent with the transpression of Shanxi during the late Jurassic, when the NNE-trending anticline and syncline developed. A large-scale uplift had taken place in the Ordos basin and adjacent areas since the late Cretaceous, considering the regional sedimentary burial history. However, the early uplift (in the late Cretaceous) had manifested slowly until 70–62 Ma before the strata entered the partial annealing zone [46]. The ages of the two samples collected from the Liulin area are consistent with the cooling age of 23 Ma recorded by the AFT in the Lou No. 1 borehole from the western Shanxi Province [99], the Qian No. 1 borehole, and the Zhenchuan No. 1 borehole from the eastern slope of northern Shanxi Province. Because of the far-reaching impact of the tectonism that occurred at 22 ± 2 Ma in central and eastern Asia [100,101], the Ordos basin and its adjacent areas were also generally uplifted and denuded. The graben basins, such as the Hetao, Yinchuan, and Weihe basins, were eroded and accompanied by sedimentary discontinuity. The eastern Bohai Bay basins and the southwestern Qinghai–Tibet Plateau also recorded this obvious uplift and denudation event. Therefore, the fission-track ages of KPA and KZA likely record this tectonic uplift event that occurred in the western Shanxi fold belt and Lvliangshan area in the early Miocene. These events happened during the distinctive geological time limit of the unconformity formation since the late Paleozoic in the Datong coalfield.

6. Conclusions

The ZFT and AFT results of the Datong basin’s coal-bearing strata show that the zircon grains were annealed at 282–117 Ma, the apatite grains were annealed from 68 to 20 Ma. The high-temperature duration for the Datong Formation was at least 47 Ma, compared with at least 29 Ma for the Shanxi Formation. Moreover, on the basis of the vitrinite reflectance analysis, the Taiyuan Formation coal seam may undergo paleotemperatures higher than the zircon complete annealing temperature of 250 °C.
The AFT and ZFT data show that the Datong basin which lies in the northern NCC experienced three thermal events and three uplift events since the late Paleozoic. The age values of the three thermal events are 245–207 Ma, 179 ± 9 Ma, and 140 Ma to 78 ± 11 Ma, respectively. These coincide with the lamprophyre activity, Kouquan fault activity, and Zuoyun basaltic andesite magmatic activity in the Datong coalfield and surrounding areas, respectively. The peak ages of the three uplift cooling events are 210 Ma (202 ± 18 Ma), 157 ± 7 Ma, and 45 ± 3 Ma or 36 ± 3 Ma, corresponding to the late Triassic, late Jurassic, and Paleogene Eocene, respectively.
According to the combined local tectonic, sedimentary evolution, and vitrinite reflectance data, the ZFT and AFT ages recorded the absent strata events of the Triassic, upper Jurassic, upper Cretaceous, and Paleogene, also corresponding to the igneous intrusion during the middle Triassic and early Cretaceous and fault action during the early–middle Jurassic. The beginning age (245 Ma) recorded by ZFT and AFT is approximately equal to the magmatic dating age of 229 ± 11 Ma, with a cooling time limit of approximately 224–119 Ma [48]. These prove that the thermal events began at 245–232 Ma around the northern NCC and that the crust underwent uplift cooling at 202 ± 18 Ma in Datong. The results show that the deep tectonic activities of the NCC were still active in the Mesozoic and even the Cenozoic Paleogene.

Author Contributions

Conceptualization, D.L. and A.Z.; methodology, F.Z.; software, J.L.; validation, D.L., F.Z. and A.Z.; formal analysis, R.Z.; investigation, Y.Z.; resources, D.L.; data curation, J.L and R.Z.; writing—original draft preparation, D.L.; writing—review and editing, F.Z.; visualization, J.L. and Y.Z.; supervision, A.Z.; project administration, D.L.; funding acquisition, F.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Dongna Liu grant number NSFC 41802191 and U1810202 and Fenghua Zhao grant number NSFC 41372164.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Thanks to the reviewers for their careful suggestions and academic editors for his sincere affirmation. Thanks to all those who helped with field sampling and experiments.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Tagami, T.; Dumitru, T.A. Provenance and thermal history of the Franciscan accretionary complex: Constraints from zircon fission track thermochronology. J. Geophys. Res. Solid Earth 1996, 101, 11353–11364. [Google Scholar] [CrossRef]
  2. Spiegel, C.; Kuhlemann, J.; Dunkl, I.; Frisch, W.; Von Eynatten, H.; Balogh, K. The erosion history of the Central Alps: Evidence from zircon fission track data of the foreland basin sediments. Terra Nova 2000, 12, 163–170. [Google Scholar] [CrossRef]
  3. Bernet, M.; Garver, J.I. Fission-track analysis of detrital zircon. Rev. Mineral. Geochem. 2005, 58, 205–237. [Google Scholar] [CrossRef]
  4. Cawood, P.A.; Hawkesworth, C.; Dhuime, B. Detrital zircon record and tectonic setting. Geology 2012, 40, 875–878. [Google Scholar] [CrossRef] [Green Version]
  5. Gao, C.; Liu, J.; Li, D.W.; Wang, F.; Liu, D.M. The constraints of the fission track thermochronology on active time of the Southern Tibetan Detachment System in Cho Oyu, Tibet. Earth Sci. Front. 2014, 21, 372–380. [Google Scholar]
  6. Chew, D.M.; Spikings, R.A. Geochronology and thermochronology using apatite: Time and temperature, lower crust to surface. Elements 2015, 11, 189–194. [Google Scholar] [CrossRef]
  7. Li, K.; Jolivet, M.; Zhang, Z.; Li, J.; Tang, W. Long-term exhumation history of the Inner Mongolian Plateau constrained by apatite fission track analysis. Tectonophysics 2016, 666, 121–133. [Google Scholar] [CrossRef]
  8. Glorie, S.; Agostino, K.; Dutch, R.; Pawley, M.; Hall, J.; Danišík, M.; Evans, N.J.; Collins, A.S. Thermal history and differential exhumation across the Eastern Musgrave Province, South Australia: Insights from low-temperature thermochronology. Tectonophysics 2017, 703, 23–41. [Google Scholar] [CrossRef]
  9. Gleadow, A.J.; Duddy, I.; Green, P.F.; Lovering, J. Confined fission track lengths in apatite: A diagnostic tool for thermal history analysis. Contrib. Mineral. Petrol. 1986, 94, 405–415. [Google Scholar] [CrossRef]
  10. Yuan, W.; Dong, J.; Bao, Z. Fission track evidences for tectonic activity in Altay Mountains, Xinjiang, northwestern China. Earth Sci. Front. 2004, 11, 461–468. [Google Scholar]
  11. Li, J.F.; Tang, W.H.; Liu, Z.; Zhang, Z.C. Apatite fission track analysis of Upper Jurassic Houcheng Formation at Qianjiadian area, Beijing and its geological significance. Chin. J. Geophys. 2010, 53, 2907–2917. [Google Scholar]
  12. Zou, B.; Wang, G.Z.; Deng, J.H. Evidence for apatite fission track of Pliocene rapid uplift of Zhongdian region on southeastern margin of Tibetan Plateau, China. J. Chengdu Univ. Technol. 2014, 41, 227–236. [Google Scholar]
  13. Glorie, S.; Alexandrov, I.; Nixon, A.; Jepson, G.; Gillespie, J.; Jahn, B.M. Thermal and exhumation history of Sakhalin Island (Russia) constrained by apatite U-Pb and fission track thermochronology. J. Asian Earth Sci. 2017, 143, 326–342. [Google Scholar] [CrossRef]
  14. Wang, S.C.; Jing, G.R.; Kang, T.S. Thermal history inversion based on apatite fission track data. Chin. Sci. Bull. 1994, 39, 816–819. [Google Scholar]
  15. Schito, A.; Andreucci, B.; Aldega, L.; Corrado, S.; Di Paolo, L.; Zattin, M.; Szaniawski, R.; Jankowski, L.; Mazzoli, S. Burial and exhumation of the western border of the Ukrainian Shield (Podolia): A multi-disciplinary approach. Basin Res. 2018, 30, 532–549. [Google Scholar] [CrossRef] [Green Version]
  16. Caricchi, C.; Aldega, L.; Barchi, M.; Corrado, S.; Grigo, D.; Mirabella, F.; Zattin, M. Exhumation patterns along shallow low-angle normal faults: An example from the Altotiberina active fault system (Northern Apennines, Italy). Terra Nova 2015, 27, 312–321. [Google Scholar] [CrossRef]
  17. Rantitsch, G.; Bhattacharyya, A.; Günbati, A.; Schulten, M.-A.; Schenk, J.; Letofsky-Papst, I.; Albering, J. Microstructural evolution of metallurgical coke: Evidence from Raman spectroscopy. Int. J. Coal Geol. 2020, 227, 103546. [Google Scholar] [CrossRef]
  18. Karayigit, A.I.; Atalay, M.; Oskay, R.G.; Córdoba, P.; Querol, X.; Bulut, Y. Variations in elemental and mineralogical compositions of Late Oligocene, Early and Middle Miocene coal seams in the Kale-Tavas Molasse sub-basin, SW Turkey. Int. J. Coal Geol. 2020, 218, 103366. [Google Scholar] [CrossRef]
  19. Bicca, M.M.; Kalkreuth, W.; da Silva, T.F.; de Oliveira, C.H.E.; Genezini, F.A. Thermal and depositional history of Early-Permian Rio Bonito Formation of southern Paraná Basin–Brazil. Int. J. Coal Geol. 2020, 228, 103554. [Google Scholar] [CrossRef]
  20. Yu, K.; Ju, Y.W.; Qian, J.; Qu, Z.H.; Shao, C.J.; Yu, K.L.; Shi, Y. Burial and thermal evolution of coal-bearing strata and its mechanisms in the southern North China Basin since the late Paleozoic. Int. J. Coal Geol. 2018, 198, 100–115. [Google Scholar] [CrossRef]
  21. Karayigit, A.I.; Mastalerz, M.; Oskay, R.G.; Gayer, R.A. Coal petrography, mineralogy, elemental compositions and palaeoenvironmental interpretation of Late Carboniferous coal seams in three wells from the Kozlu coalfield (Zonguldak Basin, NW Turkey). Int. J. Coal Geol. 2018, 187, 54–70. [Google Scholar] [CrossRef]
  22. Izart, A.; Barbarand, J.; Michels, R.; Privalov, V.A. Modelling of the thermal history of the Carboniferous Lorraine Coal Basin: Consequences for coal bed methane. Int. J. Coal Geol. 2016, 168, 253–274. [Google Scholar] [CrossRef]
  23. Littke, R.; Urai, J.L.; Uffmann, A.K.; Risvanis, F. Reflectance of dispersed vitrinite in Palaeozoic rocks with and without cleavage: Implications for burial and thermal history modeling in the Devonian of Rursee area, northern Rhenish Massif, Germany. Int. J. Coal Geol. 2012, 89, 41–50. [Google Scholar] [CrossRef]
  24. Sachsenhofer, R. Geodynamic Controls on Deposition and Maturation of Coal in the Eastern Alps. Mitt. Österr. Geol. Ges. 1999, 92, 185–194. [Google Scholar]
  25. Burkhard, M.; Kalkreuth, W. Coalification in the northern Wildhorn nappe and adjacent units, western Switzerland. Implications for tectonic burial histories. Int. J. Coal Geol. 1989, 11, 47–64. [Google Scholar] [CrossRef] [Green Version]
  26. Corrado, S.; Schito, A.; Romano, C.; Grigo, D.; Poe, B.; Aldega, L.; Caricchi, C.; Di Paolo, L.; Zattin, M. An integrated platform for thermal maturity assessment of polyphase, long-lasting sedimentary basins, from classical to brand-new thermal parameters and models: An example from the on-shore Baltic Basin (Poland). Mar. Pet. Geol. 2020, 122, 104547. [Google Scholar] [CrossRef]
  27. Zhai, M.G. Tectonic evolution and metallogenesis of North China Craton. Miner. Depos. 2010, 29, 24–36. [Google Scholar]
  28. Pan, G.T.; Xiao, Q.H.; Lu, S.N.; Deng, J.F.; Feng, Y.M.; Zhang, K.X.; Zhang, Z.Y.; Wang, F.G.; Xing, G.F.; Hao, G.J.; et al. Subdivision of tectonic units in China. Geol. China 2009, 36, 1–16+255+17–28. [Google Scholar]
  29. Li, Z.H.; Dong, S.W.; Qu, H.J. Timing of the initiation of the Jurassic Yanshan movement on the North China Craton: Evidence from sedimentary cycles, heavy minerals, geochemistry, and zircon U–Pb geochronology. Int. Geol. Rev. 2014, 56, 288–312. [Google Scholar] [CrossRef]
  30. Li, Z.H.; Qu, H.J.; Gong, W.B. Late Mesozoic basin development and tectonic setting of the northern North China Craton. J. Asian Earth Sci. 2015, 114, 115–139. [Google Scholar] [CrossRef]
  31. Li, Z.H.; Qu, H.J.; Yang, Y.H.; Gong, W.B. Late Mesozoic sedimentary-volcanic filling record in Yungang basin and its tectonic implications. Geol. China 2016, 43, 1481–1494. [Google Scholar]
  32. Zhou, R.; Liu, D.N.; Zhou, A.C.; Zou, Y. Provenance analyses of early Mesozoic sediments in the Ningwu basin: Implications for the tectonic–palaeogeographic evolution of the northcentral North China Craton. Int. Geol. Rev. 2019, 61, 86–108. [Google Scholar] [CrossRef]
  33. Zhou, R.; Liu, D.N.; Zhou, A.C.; Zou, Y.; Xie, J. A synthesis of late Paleozoic and early Mesozoic sedimentary provenances and constraints on the tectonic evolution of the northern North China Craton. J. Asian Earth Sci. 2019, 185, 104029. [Google Scholar] [CrossRef]
  34. Li, H.Y. Destruction of North China Craton: Insights from temporal and spatial evolution of the proto-basins and magmatism. Sci. China Earth Sci. 2013, 56, 464–478. [Google Scholar] [CrossRef] [Green Version]
  35. Liu, S.F.; Su, S.; Zhang, G.W. Early Mesozoic basin development in North China: Indications of cratonic deformation. J. Asian Earth Sci. 2013, 62, 221–236. [Google Scholar] [CrossRef]
  36. Li, Z.H.; Xi, S.L.; Feng, S.B.; Liu, X.S. 150 Million year history of North China Craton disruption preserved in Mesozoic sediments of the Ordos Basin. Int. Geol. Rev. 2016, 58, 1417–1442. [Google Scholar] [CrossRef]
  37. Yang, M.H.; Liu, C.Y.; Zeng, P.; Bai, H.; Zhou, J. Prototypes of Late Triassic sedimentary basins of North China Craton (NCC) and deformation pattern of its early destruction. Geol. Rev. 2012, 58, 1–18. [Google Scholar]
  38. Liu, D.N. The Coupling Relationship of Coal Metamorphism and Sedimentary-Tectonic Magmatic Activity for Datong Double Period Coal-Bearing Basin. Ph.D. Thesis, Taiyuan University of Technology, Taiyuan, China, 2015. [Google Scholar]
  39. Lu, X.P.; Wu, F.Y.; Zhao, C.B.; Zhang, Y.B. Zircon U-Pb ages of Indochinese granites in the Tonghua area and their relationship with collisional orogeny in the Dabie-Sulu superhigh pressure belt. Chin. Sci. Bull. 2003, 48, 843–849. [Google Scholar] [CrossRef]
  40. Yang, J.H.; Chung, S.L.; Wilde, S.A.; Wu, F.Y.; Chu, M.F.; Lo, C.H.; Fan, H.R. Petrogenesis of post-orogenic syenites in the Sulu Orogenic Belt, East China: Geochronological, geochemical and Nd–Sr isotopic evidence. Chem. Geol. 2005, 214, 99–125. [Google Scholar] [CrossRef]
  41. Xu, W.L.; Yang, C.H.; Yang, D.B.; Pei, F.P.; Wang, Q.H.; Ji, W.Q. Mesozoic high-Mg diorites in eastern North China craton: Constraii on the mechanism of lithospheric thinning. Earth Sci. Front. 2006, 13, 120–129. [Google Scholar]
  42. Yang, J.H.; Wu, F.Y.; Wilde, S.A.; Liu, X.M. Petrogenesis of Late Triassic granitoids and their enclaves with implications for post-collisional lithospheric thinning of the Liaodong Peninsula, North China Craton. Chem. Geol. 2007, 242, 155–175. [Google Scholar] [CrossRef]
  43. Zhu, R.X.; Yang, J.H.; Wu, F.Y. Timing of destruction of the North China Craton. Lithos 2012, 149, 51–60. [Google Scholar] [CrossRef]
  44. Chang, J.L.; Ren, D.J.; Gao, Q. The Datong formation of Ningwu coalfield, Shanxi. J. Stratigr. 1996, 20, 262–270. [Google Scholar]
  45. Zhao, M.P. Studies on the relationship between Kouquan fault and Datong Jurassic Coal Field. North China Geol. Miner. Mag. 1996, 11, 93–98. [Google Scholar]
  46. Zhao, J.F.; Liu, C.Y.; Wang, X.M.; Ma, Y.P.; Huang, L. Uplifting and evolution characteristics in the Lüliang Mountain and its adjacent area during the Meso-Cenozoic. Geol. Rev. 2009, 55, 663–672. [Google Scholar]
  47. Zhao, J.F.; Liu, C.Y.; Wang, X.M.; Zhang, C. Migration of depocenters and accumulation centers and its indication of subsidence centers in the Mesozoic Ordos Basin. Acta Geol. Sin. 2009, 83, 278–294. [Google Scholar] [CrossRef]
  48. Shao, J.A.; Zhang, Y.B.; Zhang, L.Q.; Mu, B.L.; Wang, P.Y.; Guo, F. Early Mesozoic dike swarms of carbonatites and lamprophyres in Datong area. Acta Petrol. Sin. 2003, 19, 93–104. [Google Scholar]
  49. Liu, S.; Feng, C.X.; Hu, R.Z.; Lai, S.C.; Coulson, I.M.; Fung, G.Y.; Yang, Y.H. Geochemical, Sr-Nd isotope,and zircon U-Pb geochronological constraints on the origin of Early Cretaceous carbonatite dykes, northern Shanxi Province, China. Acta Petrol. Sin. 2014, 30, 350–360. [Google Scholar]
  50. International Committee for Coal and Organic Petrology (ICCP). The new vitrinite classification (ICCP System 1994). Fuel 1998, 77, 349–358. [Google Scholar] [CrossRef]
  51. International Organization for Standardization. Methods for the Petrographic Analysis of Coals. In Method of Determining Microscopically the Reflectance of Vitrinite; International Organization for Standardization: Geneva, Switzerland, 2009; ISBN 978-0-580-56505-2. [Google Scholar]
  52. International Committee for Coal and Organic Petrology (ICCP). The new inertinite classification (ICCP System 1994). Fuel 2001, 80, 459–471. [Google Scholar] [CrossRef]
  53. O’Keefe, J.M.K.; Bechtel, A.; Christanis, K.; Dai, S.; DiMichele, W.A.; Eble, C.F.; Esterle, J.S.; Mastalerz, M.; Raymond, A.L.; Valentim, B.V.; et al. On the fundamental difference between coal rank and coal type. Int. J. Coal Geol. 2013, 118, 58–87. [Google Scholar] [CrossRef]
  54. Barker, C.E.; Pawlewicz, M.J. An empirical determination of the minimum number of measurements needed to estimate the mean random vitrinite reflectance of disseminated organic matter. Org. Geochem. 1993, 20, 643–651. [Google Scholar] [CrossRef]
  55. Taylor, G.H.; Teichmüller, M.; Davis, A.; Diessel, C.; Littke, R.; Robert, P. Organic Petrology; Gebrüder Borntraeger: Berlin/Stuttgart, Germany, 1998; pp. 453–457. [Google Scholar]
  56. Balestra, M.; Corrado, S.; Aldega, L.; Gasparo Morticelli, M.; Sulli, A.; Rudkiewicz, J.-L.; Sassi, W. Thermal and structural modeling of the Scillato wedge-top basin source-to-sink system: Insights into the Sicilian fold-and-thrust belt evolution (Italy). GSA Bull. 2019, 131, 1763–1782. [Google Scholar] [CrossRef]
  57. Corrado, S.; Aldega, L.; Perri, F.; Critelli, S.; Muto, F.; Schito, A.; Tripodi, V. Detecting syn-orogenic extension and sediment provenance of the Cilento wedge top basin (southern Apennines, Italy): Mineralogy and geochemistry of fine-grained sediments and petrography of dispersed organic matter. Tectonophysics 2018, 750, 404–418. [Google Scholar] [CrossRef]
  58. Liu, B.; Teng, J.; Mastalerz, M.; Schieber, J. Assessing the thermal maturity of black shales using vitrinite reflectance: Insights from Devonian black shales in the eastern United States. Int. J. Coal Geol. 2020, 220, 103426. [Google Scholar] [CrossRef]
  59. Ferreiro Mählmann, R.; Le Bayon, R. Vitrinite and vitrinite like solid bitumen reflectance in thermal maturity studies: Correlations from diagenesis to incipient metamorphism in different geodynamic settings. Int. J. Coal Geol. 2016, 157, 52–73. [Google Scholar] [CrossRef]
  60. Fulton, P.M.; Harris, R.N. Thermal considerations in inferring frictional heating from vitrinite reflectance and implications for shallow coseismic slip within the Nankai Subduction Zone. Earth Planet. Sci. Lett. 2012, 335, 206–215. [Google Scholar] [CrossRef]
  61. Chen, Z.Z.; Liu, G.D.; Hao, S.S. A Corrected Method of Using Vitrinite Reflectance Data to Estimate the Thickness of Sediment Removed at an Unconformity. Acta Sedimentol. Sin. 1999, 17, 141–144. [Google Scholar]
  62. Teichmüller, M. The genesis of coal from the viewpoint of coal petrology. Int. J. Coal Geol. 1989, 12, 1–87. [Google Scholar] [CrossRef]
  63. Hood, A.; Gutjahr, C.C.M.; Heacock, R.L. Organic Metamorphism and the Generation of Petroleum1. AAPG Bull. 1975, 59, 986–996. [Google Scholar]
  64. Karweil, J. Die metamorphose der kohlen vom standpunkt der physikalischen chemie. Z. Dtsch. Geol. Ges. 1956, 107, 132–139. [Google Scholar]
  65. Barker, C.E.; Pawlewicz, M.J. Calculation of Vitrinite Reflectance from Thermal Histories and Peak Temperatures. In Vitrinite Reflectance as a Maturity Parameter; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 1994; pp. 216–229. [Google Scholar]
  66. Barker, C.E.; Bone, Y.; Lewan, M.D. Fluid inclusion and vitrinite-reflectance geothermometry compared to heat-flow models of maximum paleotemperature next to dikes, western onshore Gippsland Basin, Australia. Int. J. Coal Geol. 1998, 37, 73–111. [Google Scholar] [CrossRef]
  67. Barker, C.E.; Pawlewicz, M. The correlation of vitrinite reflectance with maximum temperature in humic organic matter. In Paleogeothermics; Springer: Berlin/Heidelberg, Germany, 1986; pp. 79–93. [Google Scholar]
  68. Teichmüller, M. Anwendung kohlenpetrographischer Methoden bei der Erdöl-und Erdgasprospektion. Erdöl Kohle 1971, 24, 69–76. [Google Scholar]
  69. Connan, J. Time-temperature relation in oil genesis: Geologic notes. AAPG Bull. 1974, 58, 2516–2521. [Google Scholar]
  70. Burnham, A.K.; Sweeney, J.J. A chemical kinetic model of vitrinite maturation and reflectance. Geochim. Cosmochim. Acta 1989, 53, 2649–2657. [Google Scholar] [CrossRef]
  71. Xiao, X.M.; Liu, Z.F.; Shen, J.G.; Liu, D.H. A new method for determining palaeo-geothermal gradients in hydrocarbon-bearing basins -vitrinite reflectance gradient method. Chin. Sci. Bull. 1998, 43, 2340–2343. [Google Scholar]
  72. Jiang, G.H.; Hu, R.Z.; Fang, W.X. Progress in deduction of paleotemperature from vitrinite reflectance data. Geol.-Geochem. 2001, 29, 40–45. [Google Scholar]
  73. Wang, W.; Zhou, Z.; Yu, P. Relations between vitrinite reflectance, peak temperature and its neighboring thmperature variation rate: A comparison of methods. Chin. J. Geophys. 2005, 48, 1375–1386. [Google Scholar] [CrossRef]
  74. Ran, Q.G.; Hu, G.Y.; Chen, F.J. The inversion of thermal processes from vitrinite reflectance. Pet. Explor. Dev. 1998, 25, 29–32. [Google Scholar]
  75. Yao, H.; Ren, Y.; Shen, B. Reconstruction of the paleogeothermal gradient in the Zhongyuan area, Bohai Bay Basin, East China. Earth Sci. Front. 2006, 13, 135. [Google Scholar]
  76. Li, X. Study on Late Cretaceous-Cenozoic exhumation of the Yanji area, NE China: Insights from low-temperature thermochronology. Acta Geochim. 2019, 38, 815–833. [Google Scholar] [CrossRef]
  77. Jin, W.; Liu, Y.Q.; Wang, C.S. The application of the apatite fission track dating to the geothermal history in the northern piedmont zone of the Turpan depression, (Xinjiang). Sediment. Geol. Tethyan Geol. 2004, 24, 56–61. [Google Scholar]
  78. He, J.Q.; Ding, R.X.; Liang, S.Y.; Zhang, L.; Shan, X.J.; Jin, W.; Liu, Y.Q.; Wang, C.S. Study of thermal evolution of the North Yellow Sea basin based on apatite fission track data. Chin. J. Geophys. 2014, 57, 3347–3353. [Google Scholar]
  79. Dunkl, I. Trackkey: A Windows program for calculation and graphical presentation of fission track data. Comput. Geosci. 2002, 28, 3–12. [Google Scholar] [CrossRef] [Green Version]
  80. Tian, C.S.; Yuan, W.M.; Zeng, X.P.; Zhang, A.K.; Zhu, C.B.; Sun, F.F.; Ma, Z.Y.; Zhang, L.T.; Hao, N.N. Zircon fission track dating for mineralizing ages in IV ore-belt of Hariza polymetallic mining district, Eastern Kunlun Mountains, Qinghai-Tibet Plateau. Nucl. Tech. 2015, 38, 68–73. [Google Scholar]
  81. Yang, L.; Yuan, W.M.; Zhu, C.B.; Hong, S.J.; Li, S.Y.; Feng, Z.R.; Zhang, A.K. Mesozoic uplift exhumation history of East Kunlun. Acta Petrol. Sin. 2021, 37, 3781–3796. [Google Scholar]
  82. Bai, X.F.; Li, W.H. Coal characteristics and governed factors for coal properties of Datong Jassic 10-11# coal. Coal Convers. 2001, 24, 21–23+70. [Google Scholar]
  83. Qu, S.D.; Liu, C.Y.; Li, J.; Chen, S.Q. Fission track thermochronology evidence for Cenozoic uplifting of the southern Daheishan in Jilin, China. J. Cent. South Univ. 2014, 45, 3893–3899. [Google Scholar]
  84. Brandon, M. Decomposition of mixed grain age distributions using Binomfit. On Track 2002, 24, 13–18. [Google Scholar]
  85. Han, W.; Lu, J.C.; Zhang, Y.P.; Li, Y.H.; Wei, J.C.; Liu, X. Apatite fission track and its petroleum geological significance of the Ejina area and its vicinity, Western Inner Mongolia. Geotecton. Metallog. 2014, 38, 647–655. [Google Scholar]
  86. Yin, J.H.; Zhou, A.C. Petrological characteristics and tectonic significance of the Jiugaoshan volcanic rock in Datong, Shanxi Province. J. Taiyuan Univ. Technol. 2013, 44, 627–631+636. [Google Scholar]
  87. Chen, W.J.; Liu, R.X.; Sun, J.Z.; Li, D.M. Preliminary K-Ar age determination of the quaternary basalts in Datong volcanic area. Quat. Sci. 1986, 7, 96–102. [Google Scholar]
  88. Song, X.X.; Ma, H.T.; Li, K.J.; Liu, D.N.; Zhao, J.G.; Xue, D.S. Study on coal petrology characteristics of contact metamorphosed coal from Carboniferous-Permian in Datong Coalfield. Coal Sci. Technol. 2020, 48, 182–191. [Google Scholar]
  89. Zhang, Y.Q.; Liao, C.Z.; Shi, W. On the Jurassic deformation in and around the Ordos Basin, North China. Earth Sci. Front. 2007, 14, 182. [Google Scholar] [CrossRef]
  90. Li, Z.H.; Feng, S.B.; Yuan, X.Q.; Qu, H.J. Chronology and its significance of the Lower Jurassic tuff in Ordos Basin and its periphery. Oil Gas Geol. 2014, 35, 729–741. [Google Scholar]
  91. Yang, X.; Yang, Y.; Ji, L.; Su, C.; Zheng, M.; Zhao, L. Stages and characteristics of thermal actions in eastern part of Ordos Basin. Acta Geol. Sin. 2006, 80, 705–711. [Google Scholar]
  92. Xiao, Y.Y.; Ren, Z.L.; Qin, J.F.; Zeng, Z. Geochemistry and zircon LA-ICP MS U-Pb dating of the Zijinshan alkaline complex in the Linxian County, Shanxi Province: Geological implication. Geol. Rev. 2007, 53, 656–663. [Google Scholar]
  93. Ying, J.; Zhang, H.; Sun, M.; Tang, Y.; Zhou, X.; Liu, X. Petrology and geochemistry of Zijinshan alkaline intrusive complex in Shanxi Province, western North China Craton: Implication for magma mixing of different sources in an extensional regime. Lithos 2007, 98, 45–66. [Google Scholar] [CrossRef]
  94. Chang, Z.G.; Yang, H.; Zhou, A.C. LA-ICP-MS zircon U-Pb geochronology of the Huyanshan alkaline intrusive complex in the Lüliang Mountain of North China Craton and its geological significance. Geol. Bull. China 2017, 36, 372–380. [Google Scholar]
  95. Ren, Z.L.; Zhao, Z.Y.; Chen, G.; Wang, S.C. Tectonic thermal events of Late Mesozoic in Qinshui basin. Oil Gas Geol. 1999, 20, 48–50. [Google Scholar]
  96. Ren, Z.L.; Zhang, S.; Gao, S.L.; Cui, J.P.; Liu, X.S. Research on region of maturation anomaly and formation time in Ordos Basin. Acta Geol. Sin. 2006, 80, 674–684. [Google Scholar]
  97. Ren, Z.L.; De, M.X.; Chi, Y.L.; Ren, Y.G.; Liang, Y. Restoration of thermal history of the Permo-Carboniferous basement in the Songliao Basin. Oil Gas Geol. 2011, 32, 430–439. [Google Scholar]
  98. Guo, W.Z.; Cheng, Y.H.; Gao, Y.P.; Chen, Z.G.; Zhang, Z.W.; Wang, C. Structural characteristics and coal-bearing boundaries of Taiyuan Formation in Datong coalfield. Coal Geol. Explor 2015, 43, 1–7. [Google Scholar]
  99. Ren, Z.L. Thermal history of Ordos Basin assessed by apatite fission track analysis. Chin. J. Geophys. 1995, 38, 339–349. [Google Scholar]
  100. Liu, C.Y.; Zhao, H.G.; Gui, X.J.; Yue, L.P.; Zhao, J.F.; Wang, J.Q. Space-time coordinate of the evolution and reformation and mineralization response in Ordos basin. Acta Geol. Sin. 2006, 80, 617–638. [Google Scholar]
  101. Liu, C.Y.; Zhao, H.G.; Zhang, S.; Wang, J.Q. The important turning period of evolution in the Tibet-Himalayan tectonic domain. Earth Sci. Front. 2009, 16, 1–12. [Google Scholar]
Figure 1. (a) Inset of the North China Craton in China; (b) Schematic geological map of northern NCC (revised from Pan et al. [28]); (c) Detailed geological map of the Datong coalfield; Note for (b): II-2-Jin-Ji landmass; II-2-2-Zunhua-Wutai-Taihang mountain lava arc; II-2-5-Lvliang carbonate platform; II-3-Daqing Mountain-North Hebei paleo-arc basin system; II-3-1-Heng mountain-Chengde-Jianping lava arc; II-3-2-Daqingshan-Liangcheng continental marginal basin; II-4-1-Lang Mountain-Yin mountain landmass-Guyang-Xinghe ancient continent nucleus.
Figure 1. (a) Inset of the North China Craton in China; (b) Schematic geological map of northern NCC (revised from Pan et al. [28]); (c) Detailed geological map of the Datong coalfield; Note for (b): II-2-Jin-Ji landmass; II-2-2-Zunhua-Wutai-Taihang mountain lava arc; II-2-5-Lvliang carbonate platform; II-3-Daqing Mountain-North Hebei paleo-arc basin system; II-3-1-Heng mountain-Chengde-Jianping lava arc; II-3-2-Daqingshan-Liangcheng continental marginal basin; II-4-1-Lang Mountain-Yin mountain landmass-Guyang-Xinghe ancient continent nucleus.
Minerals 12 00942 g001
Figure 2. The stratigraphy of Kouquangou, Datong coalfield.
Figure 2. The stratigraphy of Kouquangou, Datong coalfield.
Minerals 12 00942 g002
Figure 3. The sample location and igneous rock distribution of the Datong coalfield.
Figure 3. The sample location and igneous rock distribution of the Datong coalfield.
Minerals 12 00942 g003
Figure 4. Representative field photographs and photomicrographs of the samples in the Datong basin. (a) sandstone from the Taiyuan Formation; (b) sandstone from the Shanxi Formation; (c) sandstone from the Datong Formation; (d) basaltic andesite from the Zuoyun Formation; (a-2,b-2,c-2,d-2) thin section under plane-polarized transmitted light (−); (a-3,b-3,c-3,d-3) and under orthogonality (+). Q—quartz; C.F.—lithoclast of chert; SSF.—lithoclast of sericite-rich schist; Ap.- apatite; Fel.—feldspar; Pl.—plagioclase; Aug.—pyroxene; Ol.—olivine.
Figure 4. Representative field photographs and photomicrographs of the samples in the Datong basin. (a) sandstone from the Taiyuan Formation; (b) sandstone from the Shanxi Formation; (c) sandstone from the Datong Formation; (d) basaltic andesite from the Zuoyun Formation; (a-2,b-2,c-2,d-2) thin section under plane-polarized transmitted light (−); (a-3,b-3,c-3,d-3) and under orthogonality (+). Q—quartz; C.F.—lithoclast of chert; SSF.—lithoclast of sericite-rich schist; Ap.- apatite; Fel.—feldspar; Pl.—plagioclase; Aug.—pyroxene; Ol.—olivine.
Minerals 12 00942 g004
Figure 5. (ac) The grained zircon age radar diagram; (df) the fission track age frequency distribution, (a,d) sandstone from the Datong Formation; (b,e) sandstone from the Shanxi Formation; (c,f) sandstone from the Taiyuan Formation.
Figure 5. (ac) The grained zircon age radar diagram; (df) the fission track age frequency distribution, (a,d) sandstone from the Datong Formation; (b,e) sandstone from the Shanxi Formation; (c,f) sandstone from the Taiyuan Formation.
Minerals 12 00942 g005
Figure 6. (ac) The grained apatite age radar diagram; (df) the fission track age frequency distribution, (a,d) sandstone from the Datong Formation; (b,e) sandstone from the Shanxi Formation; (c,f) basaltic andesite from the Zuoyun Formation.
Figure 6. (ac) The grained apatite age radar diagram; (df) the fission track age frequency distribution, (a,d) sandstone from the Datong Formation; (b,e) sandstone from the Shanxi Formation; (c,f) basaltic andesite from the Zuoyun Formation.
Minerals 12 00942 g006
Figure 7. (a)—Internal tectonic sketch in northern Shanxi [48]; (bf) The samples of No. 8 coal from Taiyuan Formation of Wujiayao, Datong (reflected white light, oil immersion); (b,c)—The normal coal macerals; (df) The contact metamorphism coal macerals exhibit deformation, charking, and mineralization; Co—collotelinite; Col—collodetrinite; Cu—cutinite; Sp—sporinite; Fu—Fusinite; Sf—semifusinite; M—Minerals.
Figure 7. (a)—Internal tectonic sketch in northern Shanxi [48]; (bf) The samples of No. 8 coal from Taiyuan Formation of Wujiayao, Datong (reflected white light, oil immersion); (b,c)—The normal coal macerals; (df) The contact metamorphism coal macerals exhibit deformation, charking, and mineralization; Co—collotelinite; Col—collodetrinite; Cu—cutinite; Sp—sporinite; Fu—Fusinite; Sf—semifusinite; M—Minerals.
Minerals 12 00942 g007
Figure 8. Structural evolution in Datong coalfield (modified from Guo et al. [98]). The blue arrow represents horizontal motion and the gray arrow represents vertical motion.
Figure 8. Structural evolution in Datong coalfield (modified from Guo et al. [98]). The blue arrow represents horizontal motion and the gray arrow represents vertical motion.
Minerals 12 00942 g008
Figure 9. The outcrop of the Kouquan fault-fold zone (a) the fold transition in Benxi Formation C2b; (b) the small fold fault in Shanxi Formation P1s.
Figure 9. The outcrop of the Kouquan fault-fold zone (a) the fold transition in Benxi Formation C2b; (b) the small fold fault in Shanxi Formation P1s.
Minerals 12 00942 g009
Table 1. The maximum vitrinite reflectance (%Ro,max) and the paleotemperature (Tpeak/°C).
Table 1. The maximum vitrinite reflectance (%Ro,max) and the paleotemperature (Tpeak/°C).
No.C8 Ro,maxTpeakNo.C8 Ro,maxTpeakNo.C8 Ro,maxTpeakNo.C8 Ro,maxTpeakNo.PS4 Ro,maxTpeakNo.PS4 Ro,maxTpeak
10.96147 200.82127390.73112 580.67101 10.83128200.7106
20.96147 210.82127 400.72110 590.67100 20.81125 210.69105
30.94144 220.82126410.71109 600.669930.81125220.69105.
40.93143 230.8124 420.71108 610.659840.8124 230.67101
50.88135 240.8124 430.7107 620.659750.79121 240.67101
60.86134 250.79123 440.7107 630.649660.78120250.67101
70.86133 260.79123 450.7107 640.6496 70.76117 260.67101
80.86132 270.79123 460.69106 650.6495 80.75115 270.67101
90.85132 280.79122 470.69105 660.6494 90.75115 280.67101
100.85131290.79122 480.69105 670.6394100.75115 290.6699
110.85131300.78120 490.69105 680.6394110.75115 300.6598
120.85131310.78120500.69105 690.6393 120.75115 310.6495
130.85131 320.78120 510.69105 700.6393 130.74114 320.6495
140.84130 330.78120 520.69104 710.6292 140.74113 330.6292
150.84129340.77119 530.68103 720.6190 150.73112 340.6189
160.83128 350.77119 540.68103 730.6190160.73112 350.5984
170.83128 360.77119 550.68102740.6189 170.72110
180.83128 370.75116 560.67101 750.5985 180.72110
190.82127 380.73113 570.67101 760.5780 190.71108
Table 2. The contact metamorphism coal vitrinite reflectance (%Ro,max) and the calculated paleotemperature (Tpeak/°C).
Table 2. The contact metamorphism coal vitrinite reflectance (%Ro,max) and the calculated paleotemperature (Tpeak/°C).
No.C8 BoreholeRo,maxTpeak(°C)PS4 BoreholeRo,maxTpeak(°C)No.C8 BoreholeRo,maxTpeak(°C)PS4 BoreholeRo,maxTpeak(°C)
13035.16362W9075.36367.77053.70320 W72.34261
221054.96357W9014.88355819073.60316 25042.16251
384044.70350 W6034.58347921063.57315W51.92236
422034.41342 W34.233367 1084013.49312 W10061.15170
517064.13334 W133.85324 1123063.00293W141.09163
685183.75321W5063.21301 12T9012.24255
Table 3. The coal vitrinite reflectance (%Ro,max) and the calculated paleotemperature Tpeak(°C) for Jurassic Datong Formation.
Table 3. The coal vitrinite reflectance (%Ro,max) and the calculated paleotemperature Tpeak(°C) for Jurassic Datong Formation.
Coal seamRo,maxTpeakCoal seamRo,maxTpeakCoal seamRo,maxTpeakCoal seamRo,maxTpeak
J20.78120 J70.71108 J110.69105 J140.76117
J20.6597J70.67101 J110.73112J140.75115
J20.6393J70.75115J110.70107J140.6189
J20.6699J80.6189 J110.67101 J140.78120
J20.75115J80.6087 J110.69105J140.75115
J20.69105J80.74114 J110.73112J140.87134
J20.70107J80.78120J110.76117J140.72110
J20.70107 J90.67101J120.74114 J140.80124
J30.69105 J90.74114J120.76117J140.80124
J30.75115J90.5576 J120.70107 J140.70107
J30.75115 J90.6597J120.70107 J140.6495
J30.5576 J90.70107 J120.72110J140.87134
J30.75115 J100.6495 J120.6597 J150.89137
J30.6495 J100.6597 J120.69105 J100.81 *125
J30.6393 J100.72110J120.80124 JJ20.813126
J30.75115 J100.74114 J120.76117 JJ30.851136
J40.5473 J100.72110 J120.79122JJ70.872135
J50.5166 J100.6087 J120.77119 JS120.666100
J50.6291 J100.72110 J120.85131 JS140.734113
J50.6189 J100.72110 J120.76117 JX110.58784
J50.5269 J110.78120 J120.70107 JX150.746115
J70.72110 J110.74114 J140.93143JQ70.684104
J70.5985 J110.6495 J140.72110 JQ110.739114
J70.69105 J110.74114 J140.6393 JY140.57782
J70.94144 J110.6597 J140.68103
J70.69105 J110.67101 J140.76117
Note: 0.81 *—data from Bai and Li, 2001; J2~J14-No. 2~No.14 coal from Datong Fm.; JJ2-No. 2 coal from Jinhuagong mine; JQ7-No. 7 coal from Qinciyao mine; JX11-No. 11 coal from Xinzhouyao; JS12-No. 12 coal from Sitai mine, JY14-No. 14 coal from Yanya mine. For the sampling location, see Figure 3.
Table 4. The test results of fission track ages and lengths for the samples of Datong Coal-bearing basin.
Table 4. The test results of fission track ages and lengths for the samples of Datong Coal-bearing basin.
MineralSampleStrataStrata Age/Maρs/(105/cm2) (Ns)Ρi/(105/cm2) (Ni)Ρd/(105/cm2) (-)P2)/%Central Age/Ma (±1σ)Pooled Age
/Ma
(±1σ)
Length
/µm(n)
Grain
(N)
apatiteKKAK1z100.5
~145.0
2.736
(78)
12.383
(353)
18.079
(8116)
98.978 ± 1178 ± 11 12
KJAJ2d168.3 ± 1.3~174.1 ± 1.01.834
(330)
14.826
(2667)
18.52
(8116)
98.145 ± 345 ± 312.1 ± 2.2
(100)
28
KPAP1s283.5 ± 0.6~298.9 ± 0.151.674
(412)
16.741
(4119)
18.3
(8116)
10036 ± 336 ± 312.0 ± 2.2
(84)
28
zirconKJZJ2d168.3 ± 1.3~174.1 ± 1.0120.788
(5723)
50.882
(2408)
13.134
(10030)
61.6157 ± 7157 ± 7 24
KPZP1s283.5 ± 0.6~298.9 ± 0.15114.329
(7586)
36.758
(2439)
13.478
(10030)
0202 ± 18211 ± 10 27
KCZC3t298.9 ± 0.15~307.0 ± 0.178.346
(5545)
28.626
(2026)
12.905
(10030)
0.6179 ± 9178 ± 8 29
Note: ρs—spontaneous fission-track density in minerals; ρi—recorded mineral induction fission density by mica external detector; ρd—induction fission density by standard U glass monitor; (Ns), (Ni) and (-)—the fission quantity, respectively; P2) is Chi-sq inspection probability; N is the amount of actual mineral grain; n is the actual number of the fission-track branches.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, D.; Lin, J.; Zhou, A.; Zhao, F.; Zhou, R.; Zou, Y. Tectono-Thermal Events of Coal-Bearing Basin in the Northern North China Craton: Evidence from Zircon–Apatite Fission Tracks and Vitrinite Reflectance. Minerals 2022, 12, 942. https://doi.org/10.3390/min12080942

AMA Style

Liu D, Lin J, Zhou A, Zhao F, Zhou R, Zou Y. Tectono-Thermal Events of Coal-Bearing Basin in the Northern North China Craton: Evidence from Zircon–Apatite Fission Tracks and Vitrinite Reflectance. Minerals. 2022; 12(8):942. https://doi.org/10.3390/min12080942

Chicago/Turabian Style

Liu, Dongna, Junwei Lin, Anchao Zhou, Fenghua Zhao, Rui Zhou, and Yu Zou. 2022. "Tectono-Thermal Events of Coal-Bearing Basin in the Northern North China Craton: Evidence from Zircon–Apatite Fission Tracks and Vitrinite Reflectance" Minerals 12, no. 8: 942. https://doi.org/10.3390/min12080942

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