Next Article in Journal
Mafic VMS Mineralization in the Mesozoic Metavolcanic Rocks of the Evros Ophiolite, Xylagani Area, Greece
Previous Article in Journal
Mafic Enclaves Reveal Multi-Magma Storage and Feeding of Shangri-La Lavas at the Nevados de Chillán Volcanic Complex
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dolomitization Facilitated by Clay Minerals on Mixed Siliciclastic-Carbonate Shoals of Carboniferous Age in the Tarim Basin, China: Constraints on Element Mobility and Isotope Geochemistry

1
Petroleum Exploration and Production Research Institute, SINOPEC, Beijing 102206, China
2
State Energy Key Laboratory for Carbonate Oil and Gas, Beijing 102206, China
3
College of Energy, Chengdu University of Technology, Chengdu 610059, China
4
Department of Earth Science, Durham University, Durham DH1 3LE, UK
*
Authors to whom correspondence should be addressed.
Minerals 2025, 15(4), 419; https://doi.org/10.3390/min15040419
Submission received: 11 February 2025 / Revised: 10 April 2025 / Accepted: 11 April 2025 / Published: 17 April 2025
(This article belongs to the Section Clays and Engineered Mineral Materials)

Abstract

:
In the western Tarim Basin, Carboniferous granular dolostones deposited on a carbonate platform contain a small amount of terrigenous materials of sand-size fraction, agglomerated clay minerals, or similar phases. However, the role of terrigenous materials on dolomitization is still unclear. The aim of this study was to reveal the dolomitization mechanism. The granular dolomites have small crystal size, earthy yellow color, and fabric-retentive texture, with relatively good order. These features indicate dolomites precipitated during early diagenesis. The ratio of rare earth elements (RREs) abundance of the stable isotopes 87Sr/86Sr relative to Post-Archean Australian Shale (PAAS) normalized patterns was used to study the source of the dolomitizing fluids. The composition of REEs is characterized by heavy rare earth (HREE) enrichment (average NdSN/YbSN = 0.83). There is a positive (La/La*)SN anomaly and slightly positive (Gd/Gd*)SN and (Y/Y*)SN anomaly; δ18O of seawater in fractionation equilibrium with granular dolostones was from −2.8‰ to 1.7‰ PDB, implying the dolomitizing fluid was contemporary, slightly evaporated seawater. The granular dolostones on the relatively thick shoals were subject to subaerial exposure before pervasive dolomitization, with evidence that the input of detrital kaolinite predated the formation of dolomites. Higher 87Sr/86Sr values and ∑REE in granular dolostones than the values in equivalent limestones indicate that dolomitization was related to terrigenous materials. Within the terrigenous materials, the negative-charged clay minerals may have catalyzed the dolomitization, resulting in dramatically decreased induction time for precipitation of proto-dolomites. A greater amount of terrigenous materials occurred on the shoals at the sea level fall, resulting from enhanced river entrenchment and downcutting. As a result, after subaerial exposure, the penesaline water flow through the limy allochems sediments lead to dolomitization, with the catalysis of illite on relatively thick shoals.

1. Introduction

Stratabound dolomites occur throughout the geological record, from Pre-Cambrian to the Late Pliocene [1,2,3]. The origins of these dolomites have been the subject of much research and significant debate [4,5,6,7,8,9,10,11]. In many cases, the petrographic features; δ18O, δ13C; and 87Sr/86Sr ratio of dolomites show the dolomitizing fluid could be seawater or slightly modified seawater [4,12,13], mesohaline brines [6,8,14], hypersaline brines [15,16], and hydrothermal fluids [9,13]. The mechanism of dolomitizing fluid through carbonate sediments after deposition could be the storm-driven flooding of the near-coastal supratidal flats [17,18], reflux of seawater/slightly evaporated seawater [19], evaporative pumping [15,20], and thermal convection [21,22] or geothermal convection along fault [23,24]. Accordingly, several conceptual models for dolomitization have been proposed to reveal the origin of stratabound dolostones in the last two centuries [3,6,7,8,9,14,16,22,23,24,25,26,27,28,29,30]. In particular, the dolomitization of island dolostones (e.g., Cenozoic Bahamas Bank) at considerable depth is performed by the seawater model [3,30].
The nucleation and growth of dolomite has been restricted by many kinetic inhibitors, e.g., temperatures (below about 50 °C), Ca2+/Mg2+ ratio, and salinity [31,32,33,34]. Up to now, microbes have been regarded as effective catalysts that promote non-stoichiometric, highly disordered and metastable proto-dolomite [25,35,36,37]. However, it would be difficult for abiotic dolomitization to occur under low temperature as a result of the kinetic barrier [33], unless occurring at elevated temperatures, with most temperatures being >175 °C in simulation experiments [34,38,39]. However, in some island dolostones, dolomitization occurs in the seawater mixed with meteoric water under low temperature below 50 °C without microbial catalysts, [3,30], e.g., Little Bahamas Bank, Xisha Island, Cayman Islands. Catalysts overcoming the kinetic barrier to promote abiotic dolomite formation at low temperature might exist in the above environment. Clay minerals are ubiquitous and may contribute to dolomitization as the catalyst, rather than only providing Mg2+. A simulation experiment has confirmed that highly negative-charged clay minerals can facilitate the precipitation of abiotic proto-dolomite under ambient conditions [10].
In this study, granular dolostones from the Late Carboniferous at Bachu-Maikit area in the Tarim Basin, western China, have been examined. Dolostone bodies have stratabound geometry. These dolostones contain a small amount of terrigenous sand grains and clay minerals deposited on carbonate shoals, alternating with sparry granular limestone and partially dolomitzed limestones. Previous studies based on isotope and petrography show that meteoric water has influenced the granular limestones in the Lower Member [40,41], and dolomitization was likely penecontemporaneous [40,42]. However, the mechanism of dolomitization and the pathway of dolomitized fluid flow during the early diagenesis are poorly understood. In this study, the timing of dolomitization, the geochemical feature and hydrology model of dolomitized fluid, and the effect of clay minerals on early dolomitization are discussed, using petrographic observations and geochemical analyses.

2. Geological Setting

Tarim Basin, western China, is located between the Kunlun and Tienshan orogenic belts, with a boundary of the Altyn Tagh Deep Fault (Figure 1). The southern basin was a passive continental margin of the Paleo-Tethys Ocean in the Carboniferous-early Permian [43]. During the Late Carboniferous, Tarim Block drifted to about 20° N [43], which was located in the tropical to northern temperate zone of the Paleo-Tethys Ocean [44,45]. During the Late Devonian-Mississippian period, continents uplifted at the northern, eastern, and southern areas of the Tarim Block (Figure 1). A southwesterly opening lagoonal environment with restricted circulation developed in the central and western areas [46]. In the Early Pennsylvanian period, the depositional environment was restricted lagoonal and tidal flat facies [47,48]. With the rapid sea level rise, a carbonate platform formed in the central and western areas of the basin [40,48]. At the end of the Moscovian, a dramatic sea level fall terminated the growth of the carbonate platform, with obvious sedimentary discontinuities at the top of the Xiaohaizi Formation [48]. Accordingly, the sedimentation from the Early Pennsylvanian period to the Late Muscovite period in the Tarim Basin was mainly controlled by sea level change and experienced little effect from local tectonism [46].
Dolostones from the Xiaohaizi Formation of the Late Carboniferous at the Bachu-Maikit (BM) area in the northwest of Tarim Basin were studied. The BM area is composed by Bachu uplift and Maikit slope. Maikit slope dips southwest, with the burial depth of the top of the Xiaohaizi Formation from 4266 m to 4543 m. Bachu uplift is located in the western part of the central paleo-uplift, formed during the Late Permian [49]. The depth of the Xiaohaizi Formation ranges from 1868 m to 1935 m. There are two gas fields developed in the BM area, named Yasongdi and Bashituo (Figure 1). The Xiaohaizi Formation is the most important gas zone in these two fields. The Xiaohaizi Formation overlies the Kalashayi Formation of the Middle Carboniferous, and is underlain by the Nanzha Formation of the Early Permian. Mudstones and calcareous mudstones of the Kalashayi Formation deposited in the BM area [50]. The geological age of the Xiaohaizi Formation was from the end of the Bashkirian to the Early Moscovian [50]. The carbonate open platform facies with various carbonate shoals developed in the study area in the period of the Xiaohaizi Formation (Figure 1). The BM area is one of the centers of Early Permian volcanic activity in Tarim Basin. There are two deep fractures cutting from Cambrian to Permian (Figure 1). The burial history of the Xiaohaizi Formation in well BT4 was reported by Wang et al. (2013) [40]. In the BM area, the Xiaohaizi Formation is experiencing the maximal burial depth at present, and the other subsidence period happened in the Late Permian (Figure 2).
Figure 1. Location of Bachu-Mikit area and sedimentary facies distribution of Upper Carboniferous Xiaohaizi Formation (modified from [43,50]). (A)The study area is located in the lithofacies paleogeography of the whole basin of Tarim Basin. The red box shows the region of Bachu-Macheti. (B) The distribution of sedimentary facies and data wells in the study area. Note that facies were mainly shoals on shallow open marine area in Bachu-Mikit area. The western part of the study area has uncertain sedimentary facies boundary. Well BT5 is located at the eastern part of the main study area.
Figure 1. Location of Bachu-Mikit area and sedimentary facies distribution of Upper Carboniferous Xiaohaizi Formation (modified from [43,50]). (A)The study area is located in the lithofacies paleogeography of the whole basin of Tarim Basin. The red box shows the region of Bachu-Macheti. (B) The distribution of sedimentary facies and data wells in the study area. Note that facies were mainly shoals on shallow open marine area in Bachu-Mikit area. The western part of the study area has uncertain sedimentary facies boundary. Well BT5 is located at the eastern part of the main study area.
Minerals 15 00419 g001
The upper and lower members developed in the Xiaohaizi Formation [41]. The main lithology of the Upper Member is marly limestones or argillaceous micrites, deposited in lagoon [51,52]. However, the Lower Member mainly contains granular limestones and dolostones, which formed on shoals of the open platform [53] (Figure 2). The lithology of the Lower Member is characterized by porous granular dolostones alternating with tight cemented granular limestone. The grains include ooids, oncoids, bioclastic debris (e.g., algae, bivalve, brachiopod, echinoderm), intraclastic material and full organism (e.g., Fusulinella sp., Taizchoella sp., Climacammina sp., Cribrogenerina sp., and Globivalvulina sp.) [40,50].

3. Methods

The petrographic studies and geochemical measurements are derived from ten wells in the Xiaohaizi Formation at the BM area (Table 1). The core thickness of Xiaohaizi Formation in well M10 is the largest, so more samples are selected from there. A total of 132 thin sections of carbonate rocks were prepared for petrological observations. All thin sections were stained with Alizarin-Red S and potassium ferricyanide, allowing the distinction between calcite and dolomite and the iron content of each mineral. The petrographic features, mineral compositions, elements, stable C and O isotopes, and 87Sr/86Sr values were measured to determine the geochemical characteristic of dolomitized environment, and the homogenization temperature of fluid inclusion was detected to obtain the burial diagenetic condition.
Thin sections were observed with a cathode-luminescent microscope (CL). The polished thin section surface was observed using CL MK5-2 series (CITL, Hertfordshire, UK). The voltage of this equipment was 11 kV, with the current of 380 μA. CL images were captured with the exposure times of 5 s.
A total of 8 samples from well BT2, BT4, and M4 were examined using scanning electron microscope (SEM). Quanta 250 FEG (FEI Company, Hillsboro, OR, USA) and an INCA x-max20 (Energy Disperse Spectroscopy, Oxford Instruments, Abingdon, UK) were used to observe the samples. Under high vacuum, in the model of secondary electron, the resolution for imaging is 2 nm. Under the model of backscattered electron, the resolution for imaging is 2.5 nm. The samples were carbon-coated and the beam diameter was 4 μm.
The content of minerals was determined by X-ray diffraction (XRD) (Rigaku DMAX-3C with Cu Kα radiation)(Rigaku Corporation, Tokyo, Japan), as well as degrees of the order of dolomite. The experimental conditions are 40 kV, 20 mA. The angle accuracy is bigger than 0.02° (2θ). The scanning speed is 0.05 s per step, and the scanning range is 3–50° (2θ). Using the PDF2 (2004) computer using Jada 5.0 software, semi-quantitative phase analysis was completed. When the mineral content is greater than 40%, the relative deviation is less than 10%. The order degree of dolomite is calculated by the ratio d(015)/d (110) of the two diffraction peaks. The molar percentage of CaO and MgO in dolomite from 3 samples (4 collection points) in wells BT3 and BT4 was determined by electron probe microanalysis (EPMA) (EPMA-1720 H Series, Shimadzu Instruments, Kyoto, Japan). The pure metals were used as standards. The resolution is less than 129 eV, and the precision for the major elements is less than 1%.
The weight percentages of insoluble residues within dolostones in well M10 were measured by dissolving the rock with 5% HCl and weighing the residues. These insoluble residues are composed of terrigenous materials (e.g., clay mineral, quartz, and feldspar).
In order to compare the element composition of different lithology on different shoals in the study area, granular dolostones, partially dolomitized limestone, and limestones from the Xiaohaizi Formation were selected in BT2, BT3, BT4, M4, and M10 wells (Figure 1). A total of 13 samples for major elements were measured using Optima 5300 V Inductively Coupled Plasma atomic emission spectroscopy (ICP-AES) (Perkin-Elmer Company, Norwalk, USA). REEs of nine samples were determined by Inductively Coupled Plasma Mass Spectrometry (ICP-MS) (7700 Series, Aglient Technologies, CA, USA). A quantity of 100 mg of 13 bulk sample powder was evenly divided and dried at 100 °C for 2 h and dissolved in a mixed solution consisting of 4 mL hydrofluoric acid, 2 mL hydrochloric acid, 3 mL nitric acid, 1 mL perchloric acid, and triplet of sulfuric acid. The sample dissolved in the solution is then heated to about 200 °C for 4 h until white smoke appears. A quantity of 5 mL of chlorazic acid was added to the solution to extract the element. Transfer the solution to a 50 mL volumetric bottle and dilute it with deionized water. Then, the major element was determined by ICP-AES and the trace element was determined by ICP-MS.
The stable isotopes of dolostones and limestones from ten wells were all measured to make a comparison laterally in the BM area. A total of 37 samples were analyzed for δ18O and δ13C of whole rock, using Thermo Scientific MAT253. The isotopes of relatively pure dolostones without calcite cements may represent that of dolomites. All isotope data are calculated as per mil deviation from the Pee Dee Belemnite (PDB) standard. The samples were treated with phosphoric acid. The digestion device is a rotary disk-type constant temperature reaction device. Phosphoric acid with 99.0% purity is added to the digestion device at 70 °C for 24 h. The precise carbon and oxygen isotope values were 0.0037‰ for 13C, and 0.013‰ for 18O.
A total of 12 samples from wells BT2, BT3, BT4, and M4 in the Yasongdi and Bashituo fields were analyzed for 87Sr/86Sr ratios using Thermo Scientific TRITON Plus Thermal ionization MS. The 50 g samples were put into the flasks to react with 2.5 mL 2.5 N HCl for 12 h, and then they were dried under 120 °C. A quanitty of 1 mL 6 N HCl was added into the flask at the temperature of 100 °C for 20 min, then dried under 120 °C. After the dissolution of sample, a quantity of 1.5 mL 2.5 N HCl was added into the flask, then the solution was transferred into the tube to centrifuge, with 3000 rpm for 10 min. After that, the pre-equilibration column was equilibrated by 5 mL 2.5 N HCl. A quantity of 1 mL liquid supernatant of the centrifuge tube was added on the column. The column was flushed with 0.5 mL 2.5 N HCl four times. Then, a quantity of 12 mL 4 N HCl was added into the column, and the eluent was removed. A new flask was used to receive the solution containing Sr, and the column was flushed again by 15 mL 4 N HCl. Finally, the solution containing Sr was put into the MS, and the 87Sr/86Sr ratios weredetected. The strontium isotope measurement error is expressed as 2σ±.
In order to exclude the influence of late diagenesis and hydrothermal fluid charging, the diagenetic temperature of dolostones adjacent to the deep fracture is estimated. A THMS600 Cooling-Heating Stage (Linkam Scientific, Surrey, UK) was used in the study. The homogenization temperature (TH) of inclusions in calcite veins and calcite cement was measured by using samples from 5 Xiahaizi formations near the deep fault in wells BT4 and M10. It failed to find gas–liquid inclusion within fine crystalline dolomites. The temperature range of the instrument is −196 to 600 °C with an accuracy of <0.1 °C. When approaching the critical point of rapid homogenization of gas–liquid two-phase inclusions, the heating rate can be controlled within 1 °C/min.

4. Results

4.1. Lithology and Mineral Composition

The main lithologies in the Lower Member of the Xiaohaizi Formation include granular dolostones (Figure 3A,B), granular limestones (Figure 3C,D), partially dolomitized granular limestones (Figure 3E), and dolomicrites (Figure 3F). The granular dolostones are characterized as stratabound, fabric-retentive, and dark brown in color (Figure 3A). In terms of crystal size, the granular dolostones contain fine crystalline dolomites (4–63 μm). In the bottom of a set of granular dolostones, dolostones are partially cemented by poikilotopic calcite in the M10, BT3, and BT8 wells (Figure 3B). Vadose silt could be observed in the moldic pores of dolostones in granular dolostones in well BT8 (Figure 3G,H). The thickness of shoals can be calculated by combining core description with well logging. As shown in Figure 4 and Table 2, dolomitized carbonate rocks are prone to occur in relatively thick shoals, while the proportion of dolomitized shoal to the total thickness of shoal has a negative relationship with the thickness of shoals accounting for the total Xiaohaizi Formation.
Granular limestones have grain-supported texture and contain various grains, including oosparite limestone (Figure 3C), intrasparite limestone, biosparite limestone (Figure 3D), and oncoidsparite limestone. Granular limestones are cemented by two generations of calcite cementation (Figure 3C,D). The first generation of calcite cementation occurs mainly as bladed calcite forming isopachous fringes on grains. The second generation is of blocky spar, filling in the void space. The observation of potassium ferricyanide stained thin section shows the ferroan-calcite is very rare, present in a very small amount of late diagenetic mineral.
Partially dolomitized granular limestones occurred in wells BT3, BT4, BT5, BT9, BC1, and M4 (Figure 3E). Partially dolomitized limestones have similar texture to granular limestones, with the microcrystalline calcite within grains dolomitized.
There are variable amounts of terrigenous material in carbonate rocks of the Xiaohaizi Formation. The detrital grains in granular dolostones are dominated by well-sorted and round quartz (Figure 3A), with minor kaolinite, dickite, illite, and uncertain terrigenous materials. The weight percentages of terrigenous material in granular dolostones range from 1.47 to 4.17 wt.% (Table 3). The size of terrigenous quartz ranges from 0.15 to 0.2 mm. Through the counting of terrigenous grains point by point under microscope, the contents of terrigenous detrital grains in granular dolostones in well M10 were determined to range from 1% to 8%. There are more terrigenous detrital grains in dolostones than in limestones (Figure 5). The contents of terrigenous grains increase upwards vertically in dolostones in the Lower Member of Xiaohaizi Formation on stratigraphy column of well M10 (Figure 5). The diffraction peak of illite (d = 4.28) can be seen on the XRD pattern (Figure 6).

4.2. Types and Characteristics of Dolomites

In terms of dolomite textural classification [7], dolomites from granular dolostones or partially dolomitized limestones all appear to be planar-e or planar-s (Figure 7A–C). According to crystal size classification of dolomites [8], the size fractions of dolomite are determined. Based on the observation of thin sections, and analysis of XRD (Table 4), there are two types of dolomites in the Xiaohaizi Formation, which are described below.

4.2.1. Type A Dolomite

Dolomites within granular dolostones from the Xiaohaizi Formation are characterized by small crystal sizes; cloudy, earthy yellow color; and fabric-retentive texture (Figure 8A). Determined by SEM, the fine crystalline dolomites (about 14–32 μm) occur in granular dolostones. The CL colors of type A dolomites are non-luminescent or dull red (Figure 8B). XRD analyses indicate that dolomites in granular dolostones are relatively well ordered, with ordering ratios from 0.41–0.63 (Table 4). Determined by EMPA, the mole percent of CaO (55.16–58.58%) and MgO (39.21–43.09%) show that type A dolomites are stoichiometric (Table 5). There are some kaolinite and illite-like clay minerals detected in granular dolostones (Figure 9A,B). In addition, rare anhydrite-replaced calcite or anhydrite cement occur in granular dolostones with type A dolomite.

4.2.2. Type B Dolomite

Type B dolomites partially replaced microcrystalline calcite within grains in dolostones, with cloudy center and clear rim (Figure 8C). In the SEM image, type B dolomites have coarse crystal (about 80 μm), larger than type A dolomites (Figure 7A–C). The coarse crystalline type B dolomites under CL have relatively brighter mottled red color than type A dolomites in granular dolostones (Figure 8B,D).

4.3. C, O Isotope

Bulk-rock stable δ18O of granular dolostone ranges from −4.36~2.8‰ (Table 6). The δ13C of granular dolostone is from −2.1‰ to 3.5‰ PDB, ranging from isotopically light to heavy (Table 6).
It’s reported that the tropical seawater temperature in the Middle Pennsylvanian ranges from 19.0 to 30.2 °C. Given average tropical seawater temperature of 25 °C, the δ18O of seawater in fractionation equilibrium with granular dolostones is from −2.8‰ to 1.7‰ PDB, and that with dolomicrites ranges from 1.0‰ to 5.2‰ PDB (Table 6) (Equation (1) from [54].
1000 × ln α d o l o m i t e w a t e r = 2.73 × 10 6 T 2 + 0.26
This equation is suitable for low temperature environments and can be used to calculate the oxygen isotopes of seawater in equilibrium with dolomite.
The δ18O values of granular limestones range from −6.5‰ to −7.9‰ PDB (Table 7), which is obviously more depleted than granular dolostones. The difference in oxygen isotope fractionation between dolomite and limestone can be as high as 2.5‰, even if the main oxygen isotope of dolomite is still heavy [6]. The δ13C of granular dolostones ranges from −2.6‰ to 1.7‰ PDB (Table 7), lighter than that of granular dolostones. There is a positive correlation between carbon and oxygen isotope, implying that meteoric water influence existed at the top of each upward shallower cycle.

4.4. 87Sr/86Sr Isotope

The 87Sr/86Sr values of limestones range from 0.7079 to 0.7083 (Table 8). 87Sr/86Sr values of granular dolostones and calcareous dolostones (0.7081–0.7085) are higher than those of limestones (Table 8). There is a positive correlation between 87Sr/86Sr and δ18O (Figure 10A).

4.5. Rare Earth Elements (REEs)

The REE concentrations of carbonate rocks from the Xiaohaizi Formation are shown in Table 9. Rare earth element data of five carbonate rocks from the Xiaohaizi Formation are normalized to Post-Archean Australian Shale (PAAS) [56]. PAAS can reflect the original REE composition of marine sediments, so PAAS is chosen as a standard. The REE profiles have a flattened heavy REE (HREE) profile (Figure 10B). The abundance of the light rare earth elements (LREEs) is higher than that of the heavy rare earth elements (HREEs). As shown in Figure 10C, the ∑REE of granular dolostones is obvious higher than that of limestones. The composition of REEs is characterized by HREE enrichment (average NdSN/YbSN = 0.83). The (Ce/Ce*)SN [2CeSN/(LaSN + PrSN)] ranges from 0.83 to 1.15, yet without negative anomaly, identified by Figure 10D. There is a positive (La/La*)SN anomaly (1.04–1.28), slightly positive (Gd/Gd*)SN (1.03–1.17) and (Y/Y*)SN anomaly (0.9–1.5) (Table 10). The positive (Eu/Eu*)SN anomaly ((Eu/Eu*) > 1) of four samples in Table 10 is caused by the relatively high concentration of Ba.

4.6. Major and Trace Elements

The major element Al in dolostones is mostly associated with fine-particle clay minerals [57]. Meanwhile, Rb is prone to be absorbed by clay minerals to enrich the terrigenous materials. The concentrations of Al and Rb in the impure dolostones from the Xiaohaizi Formation could be the proxies for the terrigenous fraction. The concentrations of the main major and trace elements are shown in Table 11. The correlation between 1/Sr and Mg is positive (Figure 10E). The cross-plot between Al, Rb, and Mg/Ca concentration ratio shows a relative weak positive correlation (Figure 10F). The relatively high concentration of Fe in these dolostones is caused by the diagenetic pyrite filled in the dissolved pores [52].

4.7. Fluid Inclusion

The result of homogenization temperature of fluid inclusions within calcite veins or cements in granular dolostones was shown in Table 12. In terms of the homogenization temperature of saline fluid inclusions, the lowest formation temperature of calcite cements most frequently occurred at 80–90 °C, with minor at 100–110 °C and 120–130 °C (Figure 11). However, the calcite vein filled with fractures has the lowest formation temperature of 80–90 °C (Figure 11).

5. Discussion

5.1. The Implication of the Textural Characteristic of Dolomites

The formation of dolomite or dolomitization of precursor limestone could be identified based on the texture of dolomites. In the study area, there are at three crystal sizes of dolomites (Figure 3), consistent with three lithologies, e.g., granular dolostones, partially dolomitized granular limestones, and dolomicrites. The small crystal sizes, cloudy earthy yellow color, and planar crystal boundaries of type A dolomites in granular dolostones indicate the dolomitization begins or proceeds at low temperature [7,58] and with multiple-site nucleation and relatively rapid crystal growth on mud-rich limestone precursors [59]. It has been suggested that two dolomite populations with different crystal size (type A and type B dolomite) are closely related to the different temperature of dolomitization and the reactive surface area [60,61]. The coarser crystalline dolomites (type B dolomite) occurs in the partially dolomitized granular limestone, with cloudy center and clear rims, implying they undergo different fluid activity at shallow burial diagenesis under higher formation temperature than type A dolomite. There is ubiquitous two-generation calcite cementation, indicating that partially dolomitized granular limestone has experienced the calcite-saturated diagenetic fluid, compared with dolostones almost without cements. Moreover, the precursor limestones for the fine crystalline type A dolomite and the coarse crystalline type B dolomite have different reactive surface area; fine grain size could lead to more finely crystalline textures. The dolomitization of the type B dolomite with coarse crystalline might have replaced coarse precursor limestones, implying the recrystallized calcite matrix in the partially dolomitized limestone in burial diagenesis might predate dolomitization. Therefore, the formation of type A dolomites predated type B dolomites, with low temperature and large reactive surface area.

5.2. Timing of Dolomite Formation and Dolomitizing Fluid

This study focuses on the formation of type A dolomites in granular dolostones, with their high visible porosity making them a high-quality reservoir. As described above, the textural characteristic of stoichiometric type A dolomite implies that the dolomites grew under low temperature in the early diagenesis. Through the observation by SEM, the growth of dolomite was restricted by the detrital kaolinite (Figure 7G), implying that dolomite formed after the input of detrital kaolinite. Kaolinite is distributed in almost all carbonate rocks with visible pores in the Xiaohaizi Formation. It is impossible for kaolinite to be converted from K-feldspar in the burial diagenesis, for there is very limited input of K-feldspar in the carbonate shoals. Kaolinite accounts for up to 6% of bulk rock (Table 4). The occurrence of kaolinite is more likely related to the chemical weathering at subaerial emergence. It is suggested that the kaolinite could occur in carbonate rocks below the uncomformity. The earthy yellow color of type A dolomites (Figure 3A) was possibly caused by infected by organic matter during emergence [62]. Together with the vadose silt in the dolostones (Figure 3G,H), the earthy yellow color of type A dolomites indicates the possibility of subaerial exposure before dolomitization.
The δ18O of the seawater in the period of Early to Late Carboniferous ranges from −6‰ to 0‰ PDB [63], based on analyses of pristine brachiopod shells and other calcareous fossils. According to the average value of 2.5‰ as the fractionation between dolomite and calcite [64], The δ18O values of dolomite precipitated from normal Early to Late Carboniferous seawater range from −3.5‰ to 2.5‰ PDB. As shown in Table 6, in terms of the calculated δ18O of seawater in fractionation equilibrium with granular dolostones, the δ18Odolostone resembles that of the dolostone formed from seawater. However, the δ18O values of granular dolostones are slightly heavier than that in seawater. Unless fluid rock ratios are high, dolomite should inherit the REE characteristics of temporal seawater [65,66]. The rare earth elements in PAAS-normalized (SN) model of granular dolomite can be used as an index to study the source of dolomitization fluid. As shown in Table 10, there is positive (La/La*)SN anomaly, positive (Y/Y*)SN anomaly, and slightly positive (Gd/Gd*)SN anomaly in granular dolostones, implying the dolomitization fluids having affinity with seawater [67]. Moreover, the non-luminescent or dull red CL color of type A dolomites shows low Mn concentration [68], resembling seawater. The homogenization CL patterns of type A dolomites also suggest one phase of dolomite precipitation, and absence of recrystallization in the late diagenesis. Some enriched δ13C compositions of dolomites may indicate varying degrees of evaporation, perhaps early stages of methanogenesis [18,69]. As a result, it is credible that the dolomitizing fluid of type A dolomite is mainly contemporary seawater with slight evaporation. In addition, the absence of primary anhydrite indicates the dolomitizing fluid was below gypsum saturation.
The 87Sr/86Sr values of seawater during the Carboniferous Age were 0.7075 to 7080 [63]. The 87Sr/86Sr values of limestones in this study are higher than that in seawater, indicating the burial diagenesis overprinted these rocks. However, under similar diagenetic background, the 87Sr/86Sr values of granular dolostones ranging from 0.7081–0.7085 are higher than that of limestones (Table 8). The 87Sr/86Sr values indicate the fluid responsible for dolomitization in granular dolostones is explicitly influenced by the radioactive strontium cf. [70]. It is impossible that meteoric water contributes to the radioactive strontium for dolomitization, inasmuch as the flushing of the platform top by meteoric water will decrease the Mg/Ca ratio of seawater to levels insufficient for dolomitization [8]. The terrigenous materials in the mixed shoals of siliciclastic-carbonate could provide radioactive strontium to the dolomitizing fluid. The dolomitized fluid could derive from modified seawater mixed with a small amount of terrigenous input, especially fine components, e.g., clay minerals. Of course, there is another hypothesis for the origin of radioactive strontium. Based on the homogenization temperature of fluid inclusion within calcite cements and veins (Figure 11), the most important diagenetic fluid activity occurred at 80–90 °C. In terms of the burial history [40], the timing of diagenetic fluid activity was about 272–254 Ma in the Late Permian in the study. However, the occurrence of high homogenous temperature with more than 120 °C might be caused by the hydrothermal fluid charging, matching with the fracture activation in the Late Permian [49]. The hydrothermal fluid from underlying siliciclastic rocks entered into the dolostones along two deep fractures will carry the radioactive strontium to the porous dolostones. However, the absence of hydrothermal dolomites and very few hydrothermal minerals indicate the origin of high 87Sr/86Sr values of granular dolostones could not be related to the hydrothermal fluid.
There are few geochemical data of type B dolomite obtained in this study for their complicated mineral assemblies. The formation of type B dolomites in the partially dolomitized limestones has a different way distinguishing from type A dolomites. The brighter CL color of type B dolomites than the type A dolomite lattice indicates the relatively higher Mn concentration, assuming Fe concentration is less than 5000 ppm [71]. Mn2+ has been found to favor Mg2+ site in dolomite [72,73], with mobility in dolomitizing fluid. Meanwhile, the cloudy center and clear rim of type B dolomite show the calcite precursor replacement by dolomite at the burial condition during the late diagenesis [59].

5.3. The Depositional Environment of Dolomite Formation

Through the observation of thin sections, the granular dolostones with very few calcite cements have relatively high visible pores. The various residual dolomitized allochems in granular dolostones indicate they deposited on a carbonate shoal with mixed bioclast and intraclast (Figure 3A). In the vertical profile of the Xiaohaizi Formation, there are some rocks bearing clay (e.g., marlstone and marly limestone), which are interbedded with dolostones. In the upper part of this formation, lagoonal limestone containing clay minerals was deposited.
Multiple periods of sea level fluctuation appeared in the Late Carboniferous in response to Gondwanan glaciation [74,75]. The sea level fluctuation is recognizable in the sequence of granular dolostones alternating with tight granular limestone in the study area [52] (Figure 12). In the cross section, granular dolostones composed of type A dolomites are prone to occur on the top of upward shallower sequence (Figure 12). As shown in Figure 4, there is much thick granular dolostone with type A dolomite deposited when the shoal has a large scale, while the partially dolomitized limestones containing type B dolomites occurr on the relatively thinner shoals. As discussed above, the dolomitization of type A dolomite on the thick shoal occurred at the early diagenesis, while the partial dolomitization of type B dolomite on the thin shoal mostly formed at the late diagenesis. This might be caused by the sufficient dolomitized fluid flow in the interior of the thick shoal and the amount of reflux related to sea level fall. However, there is not a difference of fluid flow capacity between thick and thin shoals, because the carbonate rocks on the two types of shoals have similar grain composition resulting from similar precursor limestone. There is another hypothesis that subaerial exposure might be an important factor of the dolomitization difference between the thick and thin shoals. Based on the identification of thin sections, vadose silt related to subaerial exposure only occurred in well M10, BT8 and BT9, located on the thick shoal. As shown in Table 11, the Sr concentration of dolostones in well M10 has almost no change vertically (the abnormally high Sr concentration of sample at 4397.30 m in well 10 is caused by authigenous gypsum), while Mn concentration increases upward, indicating that meteoric water has influenced the dolostones at the top of shoal. The mechanism for the dolomitization influenced by subaerial exposure will be discussed in the next section.
Meanwhile, as shown in Figure 4, there is a relatively high partial dolomitization proportion on the thin shoal. The partial dolomitization postdating the calcite cementation occurred on the thin shoal at the burial diagenesis. From the CL color and crystal features, the partial dolomitization might be precipitated from connate seawater stored in contemporaneous limestone in burial area [76]. If the connate seawater was responsible for the partial dolomitization, it implies there will be a longer induction time for proto-dolomite precipitation in the partial dolomitization at low temperature.

5.4. The Catalysis Role of Clay Minerals on Dolomitization

Many synthesis experiments have shown that abiogenic stoichiometric dolomites are very difficult to precipitate at low temperatures [33]. The dolomitization has to experience three- or four-stage reaction and produce various intermediate products (e.g., low-magnesium calcite, high-magnesium calcite, very high-magnesium calcite, and nonstoichiometric, poorly ordered dolomite) [32,77,78], beginning with the induction stage. The induction stage is crucial to the dolomitization reaction [79]. The induction stage might last for a very long time [77,79] without the catalysts promoting the reaction. It has been proven that microbes could be effective catalysts for the reaction [25,35,36,37]. However, the inorganic chemical mechanisms that catalyze stoichiometric dolomite formation remain an enigma. Recently, new research has revealed that zinc in saline water can facilitate magnesium ion dehydration, resulting in a dramatic decrease in induction time [11]. The catalysis role of zinc on dolomitization may be a good explanation for the formation of abiogenic dolomites, but it is difficult to evaluate the different zinc concentrations in various depositional environments at the present.
In this study, clay minerals have a pivotal role in dolomitization; meanwhile, control the concentrations of elements like Fe, Sr and Mn etc. The clay minerals covered on the dolomites in granular dolostones might promote the precipitation of dolomite at low temperature. The microbe relicts are absent in the shoal containing various coarse allochems; see Figure 4, Figure 7 and Figure 9. As discussed above, the salinity of the dolomitized fluid responsible for type A dolomite was normal to slightly higher than contemporary seawater. The kinetics for the dolomitization in the fluid with normal salinity at low temperature could be overcome by highly negative-charged clay minerals (i.e., montmorillonite and illite) [10] via electrostatic binding of Mg²⁺ and Ca²⁺ ions and simultaneous desolvation of these strongly hydrated cations. As evidence from SEM images shows, illite-like clay mineral on the surface of fine crystalline dolomite was detected (Figure 9B), and several microcrystalline dolomites occurred on the terrigenous material aggregation (Figure 9C). Meanwhile, the 87Sr/86Sr values of granular dolostones bearing type A dolomites are heavier than limestones. There are no permeable clastic aquifers underlying the Xiaohaizi Formation. The radioactive strontium could originate from the mixed terrigenous clay minerals, rather than the dolomitized fluid interaction with clastic strata. As shown in Figure 10B, the ∑REE concentrations of granular dolostones is higher than that of partially dolomitized limestones and granular limestones, also resulting from the input of mixed clay minerals. The influence of clay mineral contaminant on the REE patterns has been excluded. The obvious positive (La/La*)SN anomaly and relatively high (Y/Ho) ratio indicate there is no obvious clay contaminant [55]. Therefore, the petrographic and geochemical analyses both support that the clay minerals have a close relationship with dolomites. Combined with the simulation experiment from Liu et al. (2019) [10], it can be inferred that the highly negative-charged clay minerals can facilitate the precipitation of dolomites at low temperature and seawater with normal or slightly high salinity.
However, a major problem of this hypothesis is whether there is a sufficient amount of clay minerals for the dolomitization catalysis. In this study, the amount of illite in granular dolostones in the Xiaohaizi Formation is less than 5 wt.% for the very weak diffraction peak in the detection of XRD (Figure 6). However, many terrigenous materials are observed on the surface of dolomites under the SEM images, compared with the clean surface of granular limestones (Figure 7). The terrigenous quartz, feldspar, and lithic fragments could be carried by many geological processes, including offshore current and wind. Meanwhile, detrital clay minerals primarily originated from the hinterland. Keping Uplift, located in the north of the Bachu-Makit area, could provide the terrigenous detrital materials (Figure 1). River entrenchment and downcutting under falling sea level will result in more clastic input to the basin [80]. As discussed above, alling sea level also resulted in the subaerial exposure of the precursor of dolostones on relatively thick shoals. The relatively high contents of terrigenous quartz occurring in granular dolostones vertically in well 10 (Figure 5) show the possibility of dolomitization induced by terrigenous materials. Within these terrigenous materials, a small amount of illite could serve as nucleation centers for dolomite. Once these proto-dolomites are formed, they might act as nuclei for later massive dolomite formation [81].
The dolomitization in the partially dolomitized limestones occurring in the late diagenesis may be caused by the absence of sufficient illite to overcome the kinetic barrier. According to relevant simulation experiments, stoichiometric dolomites could be synthesized mostly under temperatures >175 °C [34,38,39]. Therefore, the elevated temperature could provide the dynamics to overcome the kinetics for proto-dolomite precipitation, resulting in partial dolomitization.

5.5. The Density-Dependent Convection of Dolomitized Fluid and the Mechanism of Dolomitization

Under the aforementioned dolomitized fluid and catalysts, the paleohydrological conditions for the dolomitization of granular dolostones on the mixed siliciclastic-carbonate shoals have been established (Figure 13). In this study, based on the evidence from geochemistry and petrography, the timing for the dolomitization of type A dolomite was proven to be penecontemporaneous at low temperature. The CL color of type A dolomite resembles that of the dolostones deposited in the peritidal area. However, the different hydrology systems between the peritidal area and shoals leads to a simple seepage–reflux model that could not explain the mechanism of dolomitization on the shoals. Reflux dolomitization is caused by brines percolating into previously deposited limestones, from repeated flooding and reflux of marine waters of slightly increased salinity [5,6]. In this study, the δ18Odolostone of −3.2‰ to 2.8‰ PDB and the presence of free protogenetic calcium sulphates indicate that the brines are penesaline water, i.e., seawater concentrated below evaporate precipitation. Penesaline water is considered to be the most likely mechanism to provide the required magnesium [6,82]. The reflux system facilitated by basal and permeable siliciclastic aquifers was proposed in the past few years [9,14,24]. However, there is an absence of permeable sandstones in the underlying Kalashayi and Bachu formations (Figure 1). The penesaline water cannot flow through the underlying clastic aquifers. Although two deep fractures have developed, the fault-controlled dolomitization [9,24] could not be used to explain the origin of stratabound dolostones in the Xiaohaizi Formation.
The penecontemporaneous dolomitization of type A dolomite could be driven by the density-dependent convection of penesaline water. At the depositional period, limy allochems deposited on the carbonate shoals near the wave base. With the sea level fall, the relatively thick shoals were subjected to subaerial exposure, and the input of terrigenous materials was enhanced by river entrenchment and downcutting in the meantime. During the subaerial emergence, the meteoric water resulted in the dissolution of aragonitic allochems and chemical weathering of feldspar. The terrigenous quartz, feldspar, and clay minerals (e.g., kaolinite and illite) deposited together with limy allochems on the relatively thick shoal, and the shoal was adjacent to the hinterland (see Figure 13). Meanwhile, the restricted seawater among shoals would experience slight evaporation to produce penesaline water. During this evaporation, with water turbulence lessening, clay minerals were allowed to settle on the limy allochems, which could serve as nucleation centers for further dolomitization. After the sea level rise, these penesaline waters would have flowed through the permeable limy allochem sediments resulting from the density difference between penesaline and normal seawater. The proto-dolomites precipitated from the penesaline water with the catalysis of illite at low temperature, resulting in pervasive dolomitization of limy allochems on relatively thick shoals.
There is another dolomitization mechanism for the partially dolomitized limestone. Due to the absence of sufficient negative-charged clay minerals, the induction time of type B dolomites would have lasted for a long time, until the burial temperature increased and the connate seawater carried them, driven by compaction or tectonic uplift of the Bachu area in the Late Permian. Therefore, the partially dolomitized limestone bearing type B dolomites occurred in the late diagenesis on the relatively thin shoal without subaerial exposure and sufficient illite. Even in the lower part of the sequence on the relatively thick shoal, these granular limestone with very little terrigenous materials have been partially dolomitized in the process to some extent.

6. Conclusions

(1) The positive (La/La*)SN and (Y/Y*)SN anomaly, slightly positive (Gd/Gd*)SN anomaly, and δ18O of seawater in fractionation equilibrium with granular dolostones of −2.8‰~1.7‰ PDB suggest that slightly modified seawater was the origin of dolomitization fluids responsible for the granular dolostones bearing a small amount of terrigenous materials from the Xiaohaizi Formation.
(2) The occurrence of vadose silt, detrital kaolinite, and fabric-retentive dissolution show that granular dolostones experienced subaerial exposure predating dolomitization. With the falling sea level, the subaerial exposure of shoal on the platform was accompanied by the enhanced input of terrigenous materials by river entrenchment and downcutting from the north of the study area.
(3) The input of negative-charged clay minerals could have facilitated the dolomitization of allochemical lime precursors under low temperature. Evidence comea from the weak positive correlation between Rb, Al, and Mg/Ca ratio of bulk dolostone, relatively higher 87Sr/86Sr values and ∑REE in granular dolostones than that in granular limestones, and the SEM images of illite-like clay minerals on the surface of dolomites.
(4) Early dolomitization was driven by the density-dependent convection of penesaline water, with the catalysis of illite. In the late diagenesis, the compaction fluid caused the partial dolomitization of limestone on the relatively thin shoal without subaerial exposure.

Author Contributions

Formal analysis, writing—original draft preparation, X.L.; resources, writing—review and editing, funding acquisition, M.F.; methodology, supervision, J.G.; project administration, R.S.; visualization data, curation H.L.; investigation, Y.F., software, D.W. All authors have read and agreed to the published version of the manuscript.

Funding

This project was supported by the National Natural Science Foundation of China (Grant No. 41402096). Natural Science Foundation of China (Grant No. U24B6001).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Acknowledgments

We thank SinoPec of China for providing rock samples. We thank the State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation in China for the measurement. We thank Cathy Hollis and two anonymous reviewers for their constructive reviews of an earlier version of this manuscript, which greatly improved the final manuscript.

Conflicts of Interest

Xuan Liu is an employee of Petroleum Exploration and Production Research Institute, SINOPEC. The paper reflects the views of the scientists and not the company.

References

  1. Chafetz, H.S.; Rush, P.F. Diagenetically altered sabkha-type Pleistocene dolomite from the Arabian Gulf. Sedimentology 1994, 41, 409–421. [Google Scholar] [CrossRef]
  2. Liu, C.; Wang, Z.; Raub, T.D. Geochemical constraints on the origin of Marinoan cap dolostones from Nuccaleena Formation, South Australia. Chem. Geol. 2013, 351, 95–104. [Google Scholar] [CrossRef]
  3. Wang, R.; Jones, B.; Yu, K. Island dolostones: Genesis by time-transgressive or event dolomitization. Sediment. Geol. 2019, 390, 15–30. [Google Scholar] [CrossRef]
  4. Land, L.S. The origin of massive dolomite. J. Geol. Educ. 1985, 33, 112–125. [Google Scholar] [CrossRef]
  5. Purser, B.H.; Tucker, M.E.; Zenger, D.H. Dolomites-A Volume in Honour of Dolomieu; Special Publications; Blackwell Scientific Publications: Oxford, UK, 1994; p. 21. [Google Scholar]
  6. Qing, H.; Bosence, D.W.J.; Rose, P.F. Dolomitization by penesaline seawater in Early Jurassic pertidal platform carbonates, Gibraltar, western Mediterranean. Sedimentology 2001, 48, 153–163. [Google Scholar] [CrossRef]
  7. Machel, H.G. Concepts and models of dolomitization: A critical reappraisal. In The Geometry and Petrogenesis of Dolomite Hydrocarbon Reservoirs; Braithwaite, C.J.R., Rizzi, G., Darke, G., Eds.; Geological Society London, Special Publications: London, UK, 2004; Volume 235, pp. 7–63. [Google Scholar]
  8. Sena, C.M.; Cedric, M.J.; Anne-Lise, J.; Veerle, V.; Christina, M. Dolomitization of lower cretaceous peritidal carbonates by modified seawater: Constraints from clumped isotopic paleothermometry, elemental chemistry, and strontium isotopes. J. Sediment. Res. 2014, 84, 552–566. [Google Scholar] [CrossRef]
  9. Hollis, C.; Bastesen, E.; Boyce, A.; Corlett, H.; Gawthorpe, R.; Hirani, J.; Rotevatn, A.; Whitaker, F. Fault-controlled dolomitization in a rift basin. Geology 2017, 45, 219–222. [Google Scholar] [CrossRef]
  10. Liu, D.; Xu, Y.; Papineau, D.; Yub, N.; Fan, Q.; Qiu, X.; Wang, H. Experimental evidence for abiotic formation of low-temperature proto-dolomite facilitated by clay minerals. Geochim. Cosmochim. Acta 2019, 247, 83–95. [Google Scholar] [CrossRef]
  11. Vandeginste, V.; Snell, O.; Hall, M.R.; Steer, E.; Vandeginste, A. Acceleration of dolomitization by zinc in saline waters. Nat. Commun. 2019, 10, 1851. [Google Scholar] [CrossRef]
  12. Jacquemyn, C.; Desouky, H.E.; Hunt, D.; Casini, G.; Swennen, R. Dolomitization of the Latemar platform: Fluid flow and dolomite evolution. Mar. Pet. Geol. 2014, 55, 43–67. [Google Scholar] [CrossRef]
  13. Mahboubi, A.; Nowrouzi, Z.; Al-Aasm, I.S.; Moussavi-Harami, R.; Mahmudy-Gharaei, M.H. Dolomitization of the Silurian Niur Formation, Tabas block, east central Iran: Fluid flow and dolomite evolution. Mar. Pet. Geol. 2016, 77, 791–805. [Google Scholar] [CrossRef]
  14. Newport, R.; Hollis, C.; Bodin, S.; Redfern, J. Examining the interplay of climate and low amplitude sea-level change on the distribution and volume of massive dolomitization: Zebbag Formation, Cretaceous, Southern Tunisia. Int. Assoc. Sedimentol. 2017, 3, 38–59. [Google Scholar] [CrossRef]
  15. McKenzie, J.A.; Hsü, K.J.; Schneider, J.F. Movement of subsurface waters under the sabkha, Abu Dhabi, UAE, and its relation to evaporative dolomite genesis. In Concepts and Models of Dolomitization; Zenger, D.H., Dunham, J.B., Ethington, R.L., Eds.; Special Publications; Society of Economic Paleontologists and Mineralogists: Tulsa, OK, USA, 1980; Volume 28, pp. 11–30. [Google Scholar]
  16. Bi, D.; Zhai, S.; Zhang, D.; Liu, X.; Liu, X.; Jiang, L.; Zhang, A. Constraints of fluid inclusions and C, O isotopic compositions on the origin of the dolomites in the Xisha Islands, South China Sea. Chem. Geol. 2018, 493, 504–517. [Google Scholar] [CrossRef]
  17. Aharon, P.; Socki, R.A.; Chan, L. Dolomitization of atolls by sea water convection flow: Test of a hypothesis at Niue, South Pacific. J. Geol. 1987, 95, 187–203. [Google Scholar] [CrossRef]
  18. Mazzullo, S.J.; Bischoff, W.D.; Teal, C.S. Holocene shallow-subtidal dolomitization by near-normal seawater, northern Belize. Geology 1995, 23, 341–344. [Google Scholar] [CrossRef]
  19. Whitaker, F.F.; Smart, P.L.; Vahrenkamp, V.C.; Nicholson, H.; Wogelius, R.A. Dolomitization by near-normal seawater? Field evidence from the Bahamas. In Dolomites-A Volume in Honour of Dolomieu; Purser, B.H., Tucker, M.E., Zenger, D.H., Eds.; Special Publications; Blackwell Scientific Publications: Oxford, UK, 1994; Volume 21, pp. 111–132. [Google Scholar]
  20. Hsü, K.J.; Siegenthaler, C. Preliminary experiments on hydrodynamic movements induced by evaporation and their bearing on the dolomite problem. Sedimentology 1969, 12, 11–25. [Google Scholar] [CrossRef]
  21. Sanford, W.E.; Whitaker, F.F.; Smart, P.L.; Jones, G. Numerical analysis of seawater circulation in carbonate platforms: I. Geothermal convection. Am. J. Sci. 1998, 298, 801–828. [Google Scholar] [CrossRef]
  22. Ngia, N.R.; Hu, M.; Gao, D. Tectonic and geothermal controls on dolomitization and dolomitizing fluid flows in the Cambrian-Lower Ordovician carbonate successions in the western and central Tarim Basin, NW China. J. Asian Earth Sci. 2019, 172, 359–382. [Google Scholar] [CrossRef]
  23. Rustichelli, A.; Iannace, A.; Tondi, E.; Celma, C.D.; Cilona, A.; Giorgioni, M.; Parente, M.; Girundo, M.; Invernizzi, C. Fault-controlled dolomite bodies as palaeotectonic indicators and geofluid reservoirs: New insights from Gargano Promontory outcrops. Sedimentology 2017, 64, 1871–1900. [Google Scholar] [CrossRef]
  24. Hirania, J.; Bastesen, E.; Boyce, A.; Corlett, H.; Gawthorpe, R.; Hollis, C.; John, C.M.; Robertson, H.; Rotevatn, A.; Whitaker, F. Controls on the formation of stratabound dolostone bodies, Hammam Faraun Fault block, Gulf of Suez. Sedimentology 2018, 65, 1973–2002. [Google Scholar] [CrossRef]
  25. Baldermann, A.; Deditius, A.P.; Dietzel, M.; Fichtner, V.; Fischer, C.; Hippler, D.; Leis, A.; Baldermann, C.; Mavromatis, V.; Stickler, C.P.; et al. The role of bacterial sulfate reduction during dolomite precipitation: Implications from Upper Jurassic platform carbonates. Chem. Geol. 2015, 412, 1–14. [Google Scholar] [CrossRef]
  26. Mountjoy, E.W.; Machel, H.G.; Green, D.; Duggan, J.; Williams-Jones, A.E. Devonian matrix dolomites and deep burial carbonate cements: A comparison between the Rimbey–Meadowbrook reef trend and the deep basin of west-central Alberta. Bull. Can. Pet. Geol. 1999, 47, 487–509. [Google Scholar]
  27. Qing, H.; Mountjoy, E.W. Large-scale fluid flow in the Middle Devonian Presqu’ile Barrier, Western Canada Sedimentary Basin. Geology 1992, 20, 903–906. [Google Scholar]
  28. Duggan, J.P.; Mountjoy, E.W.; Stasiuk, L.D. Fault-controlled dolomitization at Swan Hills Simonette oil field (Devonian), deep basin west-central Alberta, Canada. Sedimentology 2001, 48, 301–323. [Google Scholar] [CrossRef]
  29. Bahnan, A.E.; Carpentier, C.; Pironon, J.; Ford, M.; Ducoux, M.; Barré, G.; Mangenot, X.; Gaucher, E.C. Impact of geodynamics on fluid circulation and diagenesis of carbonate reservoirs in a foreland basin: Example of the Upper Lacq reservoir (Aquitaine basin, SW France). Mar. Pet. Geol. 2020, 111, 676–694. [Google Scholar] [CrossRef]
  30. Ren, M.; Jones, B. Genesis of island dolostones. Sedimentology 2018, 65, 2003–2033. [Google Scholar] [CrossRef]
  31. Katz, A.; Matthews, A. The dolomitization of CaCO3: An experimental study at 252–295 °C. Geochim. Cosmochim. Acta 1977, 41, 297–308. [Google Scholar] [CrossRef]
  32. Sibley, D.F.; Nordeng, S.H.; Borkowski, M.L. Dolomitization kinetics of hydrothermal bombs and natural settings. J. Sed. Res. 1994, 64, 630–637. [Google Scholar] [CrossRef]
  33. Land, L.S. Failure to precipitate dolomite at 25 °C from dilute solution despite 1000-fold oversaturation after 32 years. Aquat. Geochem. 1998, 4, 361–368. [Google Scholar] [CrossRef]
  34. Kaczmarek, S.E.; Thornton, B. The effect of temperature on stoichiometry, cation ordering, and reaction rate in high-temperature dolomitization experiments. Chem. Geol. 2017, 468, 32–41. [Google Scholar] [CrossRef]
  35. Wright, D.T. The role of sulphate-reducing bacteria and cyanobacteria in dolomite formation in distal ephemeral lakes of the Coorong region, South Australia. Sediment. Geol. 1999, 126, 147–157. [Google Scholar] [CrossRef]
  36. Roberts, J.A.; Bennett, P.C.; Gonza’lez, L.A.; Macpherson, G.; Milliken, K.L. Microbial precipitation of dolomite methanogenic groundwater. Geology 2004, 32, 277–280. [Google Scholar] [CrossRef]
  37. Krause, S.; Liebetrau, V.; Gorb, S.; Sa’nchez-Roma’n, M.; McKenzie, J.A.; Treude, T. Microbial nucleation of Mg-rich dolomite in exopolymeric substances under anoxic modern seawater salinity: New insight into an old enigma. Geology 2012, 40, 587–590. [Google Scholar] [CrossRef]
  38. Sibley, D.F.; Dedoes, R.E.; Bartlett, T.R. Kinetics of dolomitization. Geology 1987, 15, 1112–1114. [Google Scholar] [CrossRef]
  39. Kessels, L.A.; Sibley, D.F.; Nordeng, S.H. Nanotopography of synthetic and natural dolomite crystals. Sedimentology 2000, 47, 173–186. [Google Scholar] [CrossRef]
  40. Wang, X.; Chen, Z.; Fan, T.; Yu, T.; Cao, Z.; He, H. Integrated Reservoir Characterization of Late Carboniferous Carbonate Inner Platform Shoals in Bamai Region, Tarim Basin. J. Jilin Univ. (Earth Sci. Ed.) 2013, 43, 371–381. [Google Scholar]
  41. Ma, Z.; Yang, S.; Xu, Q.; Zhang, L.; Zhang, J.; Huang, W. Main controlling factors of carbonate reservoirs of Xiaohaizi Formation in Xianbazha area, Tarim Basin. Pet. Geol. Exp. 2015, 37, 300–327. [Google Scholar]
  42. Wang, Q.; Shi, J.; Chen, G.; Xue, L. Characteristics of Diagenetic Environments of Carbonate Rocks in Western Tarim Basin and Their Controls on the Reservoir Property. Acta Sedimentol. Sin. 2001, 19, 548–555. [Google Scholar]
  43. Jia, C. Characteristics of Chinese Petroleum Geology: Geological Features and Exploration Cases of Stratigraphic, Foreland and Deep Formation Traps; Springer Science & Business Media: Berlin, Germany, 2013; p. 15. [Google Scholar]
  44. Zhu, R.X.; Yang, Z.Y.; Wu, H.N.; Ma, X.H.; Huang, B.C.; Meng, Z.F.; Fang, D.J. Paleomagnetic constraints on the tectonic history of the major blocks of China during the Phanerozoic. Sci. China 1998, 41, 1–19. [Google Scholar]
  45. Chen, Z. Devonian–Carboniferous brachiopod zonation of the Tarim Basin, Northwest China: Implications for biostratigraphy and biogeography. Geol. J. 2004, 39, 431–458. [Google Scholar] [CrossRef]
  46. Chen, Z.; Shi, G.R. Late Paleozoic depositional history of the Tarim Basin, NW China: An integration of lithostratigraphic and biostratigraphic constraints. Am. Assoc. Pet. Geol. Bull. 2003, 87, 1323–1354. [Google Scholar]
  47. Jiang, L.; Gu, J. Analysis on the sedimentary environment of the carboniferous carbonate and detrital rock alternating sedimentation in the Maigaiti slope of the Tarim basin. Pet. Geol. Exp. 2002, 24, 41–47, (Chinese, Abstract in English). [Google Scholar]
  48. Chen, Z. A Late Carboniferous algal mound from the Tarim Basin, NW China: Internal structure and palaeoecology. Geol. J. 2012, 47, 477–494. [Google Scholar] [CrossRef]
  49. Zhang, Y.; He, D.; Liu, C. Three-dimensional geological structure and genetic mechanism of the Bachu uplift in the Tarim Basin. Earth Sci. Front. 2019, 26, 134–148. [Google Scholar]
  50. Zhu, R.; Luo, P.; Luo, Z. Lithofacies palaeogeography of the late Devonian and carboniferous in Tarim Basin. J. Palaeogeogr. 2002, 4, 13–24. [Google Scholar]
  51. Yan, J.; Zhao, X.; He, Y.; Zhong, X.; Wei, S.; Zhou, G. Sedimentary facies of Carboniferous carbonate rock in Bashitop Oil Field, Tarim Basin. Pet. Geol. Exp. 2011, 33, 353–358. [Google Scholar]
  52. Fu, M.; Song, R.; Gluyas, J.; Zhang, S.; Huang, Q. Diagenesis and reservoir quality of carbonates rocks and mixed siliciclastic as response of the Late Carboniferous glacio-eustatic fluctuation: A case study of Xiaohaizi Formation in western Tarim Basin. J. Pet. Sci. Eng. 2019, 177, 1024–1041. [Google Scholar] [CrossRef]
  53. Zhang, R.; Liu, C.; Guo, F. Sedimentary Features of Carboniferous in Bashituo Block in Tarim Basin. Xinjiang Pet. Geol. 2011, 32, 42–44, (Chinese, Abstract in English). [Google Scholar]
  54. Vasconcelos, C.; McKenzie, J.A.; Warthmann, R.; Bernasconi, S.M. Calibration of the δ18O paleothermometer for dolomite precipitated in microbial cultures and natural environments. Geology 2005, 33, 317–320. [Google Scholar] [CrossRef]
  55. Bau, M.; Dulski, P. Distribution of yttrium and rare earth elements in the Penge and Kuruman ironformations, Transvaal Supergroup, SouthAfrica. Precambrian Res. 1996, 79, 37–55. [Google Scholar] [CrossRef]
  56. Mclennan, S.M. Rare Earth Elements in Sedimentary Rocks: Influence of Provenance and Sedimentary Processes. Mineral. Soc. Am. Rev. Mineral. 1989, 21, 169–200. [Google Scholar]
  57. Biscaye, P.E. Mineralogy and sedimentation of recent deep-sea clay in the Atlantic Ocean and adjacent seas and oceans. Geol. Soc. Am. Bull. 1965, 76, 803–832. [Google Scholar] [CrossRef]
  58. Gregg, J.M.; Sibley, D.F. Epigenetic dolomitization and the origin of xenotopic dolomite texture. J. Sediment. Petrol. 1984, 54, 908–931. [Google Scholar]
  59. Sibley, D.F. The origin of common dolomite fabrics: Clues from the Pliocene. J. Sediment. Petrol. 1982, 52, 1087–1100. [Google Scholar]
  60. Whitaker, F.F.; Smart, P.L.; Jones, G.D. Dolomitization: From conceptual to numerical models. Geol. Soc. Lond. Spec. Publ. 2004, 235, 99–139. [Google Scholar] [CrossRef]
  61. Jones, G.D.; Xiao, Y. Dolomitization, anhydrite cementation, and porosity evolution in a reflux system: Insights from reactive transport models. AAPG Bull. 2005, 89, 577–601. [Google Scholar] [CrossRef]
  62. Ueno, K.; Hayakawa, N.; Nakazawa, T.; Wang, Y.; Wang, X. Pennsylvanian–Early Permian cyclothemic succession on the Yangtze Carbonate Platform, South China. Geol. Soc. Lond. Spec. Publ. 2013, 376, 235–267. [Google Scholar] [CrossRef]
  63. Veizer, J.; Ala, D.; Azmy, K.; Bruckschen, P.; Buhl, D.; Bruhn, F.; Carden, G.A.F.; Diener, A.; Ebneth, S.; Godderis, Y.; et al. 87Sr/86Sr, δ13C and δ18O evolution of Phanerozoic seawater. Chem. Geol. 1999, 161, 59–88. [Google Scholar] [CrossRef]
  64. Major, R.P.; Lloyd, R.M.; Lucia, F.J. Oxygen isotope composition of Holocene dolomite formed in a humid hypersaline setting. Geology 1992, 20, 586–588. [Google Scholar] [CrossRef]
  65. Qing, H.; Mountjoy, E.W. Rare earth element geochemistry of dolomites in the Middle Devonian Presqu’ile barrier, Western Canada Sedimentary Basin: Implications for fluid-rock ratios during dolomitization. Sedimentology 1994, 41, 787–804. [Google Scholar] [CrossRef]
  66. Nothdurft, L.D.; Webb, G.E.; Kamber, B.S. Rare earth element geochemistry of Late Devonian reefal carbonates, Canning Basin, Western Australia: Confirmation of a seawater REE proxy in ancient limestones. Geochim. Cosmochim. Acta 2004, 68, 263–283. [Google Scholar] [CrossRef]
  67. Zhang, J.; Nozaki, Y. Rare earth elements and yttrium in seawater: ICP-MS determinations in the east Caroline, Coral Sea, and South Fiji basins of the western South Pacific Ocean. Geochim. Cosmochim. Acta 1996, 60, 4631–4644. [Google Scholar] [CrossRef]
  68. Frank, J.R.; Carpener, A.B.; Oglesby, T.W. Cathodo-luminescence and composition of calcite cement in the Taum Sauk limestone (upper Cambrian), southeast Missouri. J. Sediment. Res. 1982, 52, 631–638. [Google Scholar]
  69. Irwin, H.; Curtis, C.; Coleman, M. Isotopic evidence for source of diagenetic carbonates formed during burial of organic-rich sediments. Nature 1977, 269, 209–213. [Google Scholar] [CrossRef]
  70. Wang, L.C.; Hu, W.X.; Wang, X.L.; Cao, J.; Wu, H.G.; Liao, Z.W.; Wan, Y. Changes in Strontium Content and Strontium Isotope Fractionation during Dolomitization: Implications and Characteristics. Oil Gas Geol. 2016, 37, 464–472. [Google Scholar]
  71. Huang, S. Relationship between cathodoluminescence and concentration of iron and manganese in Carbonate Minerals. Mineral. Petrol. 1992, 12, 74–79. [Google Scholar]
  72. Lumsden, D.N.; Lloyd, R.V. Mn (II) partitioning between calcium and magnesium sites in studies of dolomite origin. Geochim. Cosmochim. Acta 1984, 48, 1861–1865. [Google Scholar] [CrossRef]
  73. Lumsden, D.N.; Shipe, L.G.; Lloyd, R.V. Mineralogy and Mn geochemistry of laboratorysynthesized dolomite. Geochim. Cosmochim. Acta 1989, 53, 2325–2329. [Google Scholar] [CrossRef]
  74. Rygel, M.C.; Fielding, C.R.; Frank, T.D.; Birgenheier, L.P. The magnitude of Late Paleozoic glacioeustatic fluctuations: A synthesis. J. Sediment. Res. 2008, 78, 500–511. [Google Scholar] [CrossRef]
  75. Buggisch, W.; Krainer, K.; Schaffhauser, M.; Joachimski, M.M.; Korte, C. Late Carboniferous to Late Permian carbon isotope stratigraphy: A new record from post-Variscan carbonates from the Southern Alps (Austria and Italy). Palaeogeogr. Palaeoclimatol. Palaeoecol. 2015, 433, 174–190. [Google Scholar] [CrossRef]
  76. Guo, C.; Chen, D.; Qing, H.; Zhou, X.; Ding, Y. Early dolomitization and recrystallization of the Lower-Middle Ordovician carbonates in western Tarim Basin (NW China). Mar. Pet. Geol. 2019, 111, 332–349. [Google Scholar] [CrossRef]
  77. Sibley, D.F. Unstable to stable transformations during dolomitization. J. Geol. 1990, 98, 739–748. [Google Scholar] [CrossRef]
  78. Kaczmarek, S.E.; Sibley, D.F. Direct physical evidence of dolomite recrystallization. Sedimentology 2014, 61, 1862–1882. [Google Scholar] [CrossRef]
  79. Kaczmarek, S.E.; Sibley, D.F. On the evolution of dolomite stoichiometry and cation order during high-temperature synthesis experiments: An alternative model for the geochemical evolution of natural dolomites. Sed. Geol. 2011, 240, 30–40. [Google Scholar] [CrossRef]
  80. Tucker, M.E. Mixed clastic–carbonate cycles and sequences: Quaternary of Egypt and Carboniferous of England. Geologia Croatica 2003, 56, 19–37. [Google Scholar] [CrossRef]
  81. Burns, S.J.; McKenzie, J.A.; Vasconcelos, C. Dolomite formation and biogeochemical cycles in the Phanerozoic. Sedimentology 2000, 47, 49–61. [Google Scholar] [CrossRef]
  82. Maliva, R.G.; Missimer, T.M.; Guo, W. Paleohydrological modeling of penesaline reflux dolomitization: Avon Park Formation (Middle Eocene), East Central Florida. Carbonates Evaporites 2019, 34, 941–954. [Google Scholar] [CrossRef]
Figure 2. Stratigraphy column of Carboniferous and Xiaohaizi Formation in Bachu-Mikit area. There are two members of Xiaohaizi Formation. Lithologies of Lower Member include dolostones, limestones, and partly dolomitized limestones. The color of the lithologic section of the column represents the color of the core.
Figure 2. Stratigraphy column of Carboniferous and Xiaohaizi Formation in Bachu-Mikit area. There are two members of Xiaohaizi Formation. Lithologies of Lower Member include dolostones, limestones, and partly dolomitized limestones. The color of the lithologic section of the column represents the color of the core.
Minerals 15 00419 g002
Figure 3. Images of thin section observation: (A) granular dolostone, well BT8, 4512.4 m; (B) Granular dolostones with intense cementation, well M4 4388.55 m; (C) Oospartie limestone, well BT8, 4504.36 m; (D) Biosparite limestone, well BT6, 4447.04 m; (E) Partially dolomitized biosparite limestone, well BT8 4533.5 m, the red mineral is calcite stained with alizarin red; (F) Dolomicrite, well BT8 4532.05 m; (G) Vadose silt filled in the dissolved pores, BT8 4534.3 m; (H) Vadose silt filled in the dissolved pores, BT8 4534.3 m. DP—dissolved pore; IP—intergranular pore; Bi—bioclast; In—intraclast; Q—quartz; Cc—calcite; Oo—ooid; PDG—partially dolomitized grain; DG—dolomitized grain; Pcc—poikilotopic calcite. vs—vadose silt. The blue area in the image is the cast, indicating the pores.
Figure 3. Images of thin section observation: (A) granular dolostone, well BT8, 4512.4 m; (B) Granular dolostones with intense cementation, well M4 4388.55 m; (C) Oospartie limestone, well BT8, 4504.36 m; (D) Biosparite limestone, well BT6, 4447.04 m; (E) Partially dolomitized biosparite limestone, well BT8 4533.5 m, the red mineral is calcite stained with alizarin red; (F) Dolomicrite, well BT8 4532.05 m; (G) Vadose silt filled in the dissolved pores, BT8 4534.3 m; (H) Vadose silt filled in the dissolved pores, BT8 4534.3 m. DP—dissolved pore; IP—intergranular pore; Bi—bioclast; In—intraclast; Q—quartz; Cc—calcite; Oo—ooid; PDG—partially dolomitized grain; DG—dolomitized grain; Pcc—poikilotopic calcite. vs—vadose silt. The blue area in the image is the cast, indicating the pores.
Minerals 15 00419 g003
Figure 4. Correlation analysis of the dolomitized carbonate rocks and thickness of shoals and proportion of shoals to total thickness of Xiaohaizi Formation. (A) Crossplot of the thickness of shoal and the thickness of dolomitized carbonate rock. (B) Crossplot of the proportion of the shoal to the total thickness of the Xiaohaizi Formation and the thickness of dolomitized shoal to the total thickness of shoal.
Figure 4. Correlation analysis of the dolomitized carbonate rocks and thickness of shoals and proportion of shoals to total thickness of Xiaohaizi Formation. (A) Crossplot of the thickness of shoal and the thickness of dolomitized carbonate rock. (B) Crossplot of the proportion of the shoal to the total thickness of the Xiaohaizi Formation and the thickness of dolomitized shoal to the total thickness of shoal.
Minerals 15 00419 g004
Figure 5. Vertical variation of terrigenous grain contents and C, O isotope with lithology in Lower Member of Xiaohaizi Formation in well M10. Gray-black lithology represents argillaceous limestone, gray lithology represents argillaceous dolomite, orange lithology represents granulated limestone, and earthy yellow lithology represents granular dolostone formed after exposed on the top of shoal.
Figure 5. Vertical variation of terrigenous grain contents and C, O isotope with lithology in Lower Member of Xiaohaizi Formation in well M10. Gray-black lithology represents argillaceous limestone, gray lithology represents argillaceous dolomite, orange lithology represents granulated limestone, and earthy yellow lithology represents granular dolostone formed after exposed on the top of shoal.
Minerals 15 00419 g005
Figure 6. X-ray diffraction spectra of granular dolostone in Lower Member of Xiaohaizi Formation in well M10.
Figure 6. X-ray diffraction spectra of granular dolostone in Lower Member of Xiaohaizi Formation in well M10.
Minerals 15 00419 g006
Figure 7. Textural morphology of minerals and grains in granular dolostones and partly dolomitized limestones. Identification of minerals was performed by energy spectrum: (A) Completely dolomitized grains, well BT2, 1923.6 m; (B) Euhedral dolomite with size of 10–30 μm, planar-e, granular dolostones, well BT2, 1923.6 m; (C) Dolomites with size of about 80 μm replaced the lime, partly dolomitized limestones, well BT2, 1928.7 m. (D) SEM image of quartz in dolostone, well M4, 4387.35 m. (E) Microcrystalline dolomites occur in the particle of terrigenous material, well BT2, 1923.6 m; (F) Illite-like clay minerals cover the surface of dolomites, well M4, 4386.85 m. (G) Kaolinite predated crystalline dolomite, well BT2, 1923.6 m; (H) Microcrystalline dolomites form on the surface of kaolinite, well M4, 4387.35 m. do—dolomite; Cc—calcite; TM—terrigenous material; Q—quartz; K—kaolinite.
Figure 7. Textural morphology of minerals and grains in granular dolostones and partly dolomitized limestones. Identification of minerals was performed by energy spectrum: (A) Completely dolomitized grains, well BT2, 1923.6 m; (B) Euhedral dolomite with size of 10–30 μm, planar-e, granular dolostones, well BT2, 1923.6 m; (C) Dolomites with size of about 80 μm replaced the lime, partly dolomitized limestones, well BT2, 1928.7 m. (D) SEM image of quartz in dolostone, well M4, 4387.35 m. (E) Microcrystalline dolomites occur in the particle of terrigenous material, well BT2, 1923.6 m; (F) Illite-like clay minerals cover the surface of dolomites, well M4, 4386.85 m. (G) Kaolinite predated crystalline dolomite, well BT2, 1923.6 m; (H) Microcrystalline dolomites form on the surface of kaolinite, well M4, 4387.35 m. do—dolomite; Cc—calcite; TM—terrigenous material; Q—quartz; K—kaolinite.
Minerals 15 00419 g007
Figure 8. Images of thin sections and samples under cathode luminescence (CL) and SEM: (A) Fine crystalline granular dolostone, containing dolomitized allochems ghosts, with earthy yellow color. The cements include calcite and fluorite. Well M4, 4390.74 m; (B) Image of (A) under CL. The CL color of dolomitized grains is null or very dark red. (C) Partly dolomitized limestone with dolomicrite filling in intergranular space. There is bladed calcite coating on grains. Well BT2, 1928.7 m; (D) Image of (C) under CL; the CL color of partly dolomitized grains is relative brighter red than dolomites from (B). do-g: dolomitized grain; Fl: fluorite; P-do-g: partly dolomitized grain; do-m: dolomicrite; do: dolomite.
Figure 8. Images of thin sections and samples under cathode luminescence (CL) and SEM: (A) Fine crystalline granular dolostone, containing dolomitized allochems ghosts, with earthy yellow color. The cements include calcite and fluorite. Well M4, 4390.74 m; (B) Image of (A) under CL. The CL color of dolomitized grains is null or very dark red. (C) Partly dolomitized limestone with dolomicrite filling in intergranular space. There is bladed calcite coating on grains. Well BT2, 1928.7 m; (D) Image of (C) under CL; the CL color of partly dolomitized grains is relative brighter red than dolomites from (B). do-g: dolomitized grain; Fl: fluorite; P-do-g: partly dolomitized grain; do-m: dolomicrite; do: dolomite.
Minerals 15 00419 g008
Figure 9. Minerals determined by Energy Disperse Spectroscopy. (A) Microcrystalline round dolomite on the surface of kaolinite; (B) Illite-like clay mineral on the surface of fine crystalline dolomite; (C) Several microcrystalline rhombic dolomites occur on a terrigenous grain. These images imply the precipitation of dolomite might be related to clay minerals.
Figure 9. Minerals determined by Energy Disperse Spectroscopy. (A) Microcrystalline round dolomite on the surface of kaolinite; (B) Illite-like clay mineral on the surface of fine crystalline dolomite; (C) Several microcrystalline rhombic dolomites occur on a terrigenous grain. These images imply the precipitation of dolomite might be related to clay minerals.
Minerals 15 00419 g009
Figure 10. Cross-plot of geochemical parameters of Upper Carboniferous Xiaohaizi Formation in Bachu-Makit area. (A) Cross-plot of 87Sr/86Sr and δ18O (PDB) of dolostones and limestones. (B) PAAS-normalized REE patterns of samples from Xiaohaizi Formation. PAAS data is from McLennan, 1989. Dark full lines represent REE patterns of granular dolostone. Short dash lines represent REE patterns of partially dolomitized limestones, while dotted lines represent REE patterns of limestone. Positive La anomalies, slightly positive Gd and Y anomalies are displayed in these samples. (C) Comparison among dolostones and limestons shows granular dolostones of granular dolostones have the highest amount of ∑REE. I-1: Granular dolostone; I-2: Partially dolomitized limestone; I-3: Granular limestone; II: samples with Ba contamination. (D) Dots located in a,b,c areas indicate there is no Ce negative anomaly. The template is from Bau and Dulski (1996) [55]. (E) Strong positive correlation between 1/Sr and Mg, showing the Sr concentration is related to mineral facies. (F) Weak positive correlation between Mg/Ca concentration ratio between Al and Rb. Input of terrigenous sediments seems to increase Mg/Ca ratio.
Figure 10. Cross-plot of geochemical parameters of Upper Carboniferous Xiaohaizi Formation in Bachu-Makit area. (A) Cross-plot of 87Sr/86Sr and δ18O (PDB) of dolostones and limestones. (B) PAAS-normalized REE patterns of samples from Xiaohaizi Formation. PAAS data is from McLennan, 1989. Dark full lines represent REE patterns of granular dolostone. Short dash lines represent REE patterns of partially dolomitized limestones, while dotted lines represent REE patterns of limestone. Positive La anomalies, slightly positive Gd and Y anomalies are displayed in these samples. (C) Comparison among dolostones and limestons shows granular dolostones of granular dolostones have the highest amount of ∑REE. I-1: Granular dolostone; I-2: Partially dolomitized limestone; I-3: Granular limestone; II: samples with Ba contamination. (D) Dots located in a,b,c areas indicate there is no Ce negative anomaly. The template is from Bau and Dulski (1996) [55]. (E) Strong positive correlation between 1/Sr and Mg, showing the Sr concentration is related to mineral facies. (F) Weak positive correlation between Mg/Ca concentration ratio between Al and Rb. Input of terrigenous sediments seems to increase Mg/Ca ratio.
Minerals 15 00419 g010
Figure 11. Homogenization temperature of fluid inclusion within calcite veins or cements in granular dolostones.
Figure 11. Homogenization temperature of fluid inclusion within calcite veins or cements in granular dolostones.
Minerals 15 00419 g011
Figure 12. Distribution of different dolostones on shoal facies in Xiaohaizi Formation.
Figure 12. Distribution of different dolostones on shoal facies in Xiaohaizi Formation.
Minerals 15 00419 g012
Figure 13. Model of dolomitization of granular dolostones in Xiaohaizi Formation.
Figure 13. Model of dolomitization of granular dolostones in Xiaohaizi Formation.
Minerals 15 00419 g013
Table 1. Sampling information of the cores from Xiaohaizi Formation.
Table 1. Sampling information of the cores from Xiaohaizi Formation.
WellThickness of the Cored Intervals (m)Total Thickness of Tested (m)The Number of Thin SectionsTHElements CompositionSEM
Observation
δ13Cδ18O87Sr/86SrXRD
M410.88.257
M1051.6236.2236
BC111.658.768
BT29.548.45
BT315.8914.799
BT418.9216.2318
BT521.51215
BT616.215.4229
BT776.263
BT837.2328.7531
Table 2. Relevant information of stratigraphy and lithology in Xiaohaizi Formation.
Table 2. Relevant information of stratigraphy and lithology in Xiaohaizi Formation.
ItemsBachu UpliftSouth of the Maigaiti SlopeNorth of the Maigaiti Slope
WellBC1BT2BT3BT5BT6BT8BT9M4M10BT4
The interval of burial depth (m)1935–20331876–19711868–19441887.5–19704405–44864458–45444543–46554343–44174361.5–44574266–4333.5
The total thickness of the Xiaohaizi Formation (m)98957682.581861127495.567.5
The thickness of shoal (m)30.3536.538.53336.3754.444945.557.531.81
The thickness of dolomitized carbonate rock (m)19.7535.517339.428.122411.938.721.58
The proportion of the shoal to the total thickness of the Xiaohaizi Formation (%)30.9738.4250.664044.963.343.7561.4960.2147.13
The proportion of dolomitized shoal to the total thickness of shoal (%)65.0797.2644.1610025.8551.6548.9826.1523.8267.84
Table 3. Contents of acid-insoluble residues within dolostones from Xiaohaizi Formation.
Table 3. Contents of acid-insoluble residues within dolostones from Xiaohaizi Formation.
WellDepth (m)Insoluble Residues Content # (wt.%)
M104397.303.63
M104398.103.97
M104400.004.17
M104404.941.47
M104406.541.69
BT31918.023.15
# samples were measured by dealing with 5% HCl.
Table 4. Determination of mineral composition by XRD.
Table 4. Determination of mineral composition by XRD.
WellDepth (m)LithologyContent of Minerals (wt.%)Degree of Order of Dolomite
KaoliniteQuartzCalciteDolomiteIllitePyrite
M104397.30granular dolostoneT1396T00.53
M104398.10calcareous granular dolostone223363T00.63
M104400.00calcareous granular dolostone122770T00.52
M104406.54calcareous granular dolostoneTT2575T00.5
M104404.94calcareous granular dolostoneTT2179T00.59
BT44312.11microcrystalline dolostone6T984T10.41
BT31918.02granular dolostoneT1297T00.6
Note: T—trace.
Table 5. Mole percent of oxides in type A dolomite, determined by EMPA.
Table 5. Mole percent of oxides in type A dolomite, determined by EMPA.
WellBT3BT3BT3BT4
Depth (m)1913.621913.621911.724312.11
LithologyGranular DolostoneGranular DolostoneGranular DolostoneDolomicrite
Crystal size15 μm30 μm12 μm10 μm
Type of dolomiteAAAA
F0.250.090.000.00
SrO0.040.000.000.01
K2O0.000.000.010.09
SO30.280.030.070.16
Na2O0.020.000.020.01
BaO0.000.000.000.00
MnO0.000.010.010.02
CaO58.5858.0357.7255.16
MgO39.2140.6240.2443.09
TiO20.000.000.140.04
FeO0.850.630.841.39
SiO20.320.290.450.03
Al2O30.460.290.490.00
Table 6. The composition of C and O isotope and the δ18Oseawater in fractionation equilibrium with dolomites.
Table 6. The composition of C and O isotope and the δ18Oseawater in fractionation equilibrium with dolomites.
WellDepth (m)Lithologyδ18Odol-PDBδ13Cdol-PDBδ18Oseawater-PDB
BT84532.05Granular dolostone2.8−1.43.41
BT51933.10Granular dolostone−0.43.50.12
BT51937.01Granular dolostone0.10.10.63
BT51936.56Granular dolostone1.10.11.66
BT84512.40Granular dolostone−1.61.8−1.12
BT84525.20Granular dolostone−2.0−2.1−1.53
BT84525.66Granular dolostone−2.2−2.0−1.74
BT84533.82Granular dolostone−2.6−2.0−2.15
BT84534.30Granular dolostone−1.4−1.8−0.92
BT84534.65Granular dolostone−3.2−2.0−2.77
BT44317.55dolomicrite1.31.91.87
BT51938.30dolomicrite1.7−0.12.28
BT51944.46dolomicrite1.0−2.21.56
BT51933.25dolomicrite0.50.41.04
Table 7. Composition of C and O isotope of granular limestones from Xiaohaizi Formation.
Table 7. Composition of C and O isotope of granular limestones from Xiaohaizi Formation.
WellDepth (m)Lithologyδ18Ocal-PDBδ13Ccal-PDB
M104410.75granular limestone−7.902.30
BC11976.76granular limestone−6.90−1.20
BT64450.60granular limestone−7.601.70
BT84515.35granular limestone−6.80−1.30
BT74434.51granular limestone−6.501.60
BT74432.07granular limestone−7.401.50
M104430.42granular limestone−6.50−2.60
Table 8. 87Sr/86Sr values for dolostones and limestones from Xiaohaizi Formation.
Table 8. 87Sr/86Sr values for dolostones and limestones from Xiaohaizi Formation.
WellDepth (m)Lithologyδ18O V-PDB87Sr/86SrError (2σ)
BT21926.30Granular dolostone−4.130.70854.63 × 10−5
BT21923.60Granular dolostone−4.360.70833.43 × 10−5
BT21928.70Partially dolomitized limestone−6.730.70827.05 × 10−5
BT31920.27Granular limestonenone0.70814.28 × 10−5
BT44303.37Granular limestone−8.620.70799.49 × 10−6
BT44311.52Granular limestonenone0.70802.34 × 10−5
BT44316.40Granular limestone−7.670.70837.23 × 10−5
BT44319.60Granular limestone−7.830.70822.41 × 10−5
BT44309.34Granular limestonenone0.70793.23 × 10−5
M44388.50Granular dolostone−3.480.70824.05 × 10−5
M44386.85Calcareous dolostone−6.560.70812.03 × 10−5
M44393.29Granular limestonenone0.70803.81 × 10−5
Table 9. REE concentrations of 9 samples from Xiaohaizi Formation. All concentrations in parts per million (ppm).
Table 9. REE concentrations of 9 samples from Xiaohaizi Formation. All concentrations in parts per million (ppm).
WellDepth (m)LithologyΣLREEΣHREEΣLREE/ΣHREEY/HoLaCePrNdSmEuGdTbDyHoErTmYbLu
BT21926.3Granular dolostone8.284.161.9937.462.133.460.441.730.330.190.380.060.360.070.220.030.210.03
BT21923.6Granular dolostone7.703.981.9439.881.953.190.401.610.340.210.350.060.340.070.200.030.170.03
BT44316.4Granular limestone5.893.121.8940.901.482.490.311.250.250.110.270.040.250.050.160.020.150.02
M44386.85Granular dolostone32.2012.452.5931.598.0514.601.606.381.320.251.420.231.280.250.720.100.590.09
M44388.5Granular dolostone20.9910.352.0332.435.029.671.054.180.910.171.030.171.020.210.590.080.500.07
BT21928.7Partially dolomitizedlimestone17.526.172.8434.504.168.540.873.230.600.110.610.100.570.120.350.050.330.05
M44387.35Granular dolostone4.011.712.3533.290.961.760.220.860.180.030.190.030.170.030.100.010.090.01
BT44314.46Granular limestone11.864.102.8934.912.835.560.592.330.440.100.450.070.390.080.220.030.190.03
BT44303.37Granular limestone9.322.793.3526.182.014.700.441.750.350.070.370.060.330.060.170.020.140.02
Table 10. The indices for the REE patterns of 9 samples from the Xiaohaizi Formation.
Table 10. The indices for the REE patterns of 9 samples from the Xiaohaizi Formation.
WellDepth
(m)
LitholoyNb/YbSNLa/La*SN(Ce/Ce*)SN(Eu/Eu*)SN(Pr/Pr*)SN(Gd/Gd*)SN(Y/Y*)SNBa
(ppm)
M44386.85Granular
dolostone
0.911.270.940.850.971.151.115.44
M44388.5Granular
dolostone
0.701.190.970.810.971.171.1614.23
BT21928.70Partially
dolomitized limestone
0.821.041.040.870.971.121.244.65
M44387.35Partially
dolomitized limestone
0.821.110.890.901.031.111.174.76
BT44314.46Granular
Limestone
1.021.150.991.100.971.031.22175.74
BT44303.37Granular
Limestone
1.031.101.150.940.911.120.903.49
BT21926.30Granular
dolostone
0.691.220.832.551.040.621.36374.08
BT21923.60Granular
dolostone
0.771.280.832.821.030.571.44542.79
BT44316.40Partially
dolomitized limestone
0.691.170.842.011.040.721.50191.12
Table 11. Concentration of elements in granular dolostones from Xiaohaizi Formation.
Table 11. Concentration of elements in granular dolostones from Xiaohaizi Formation.
WellDepth (m)LithologyMg/CaSr
(μg/g)
Ba
(μg/g)
Ti
(μg/g)
Mn
(μg/g)
Fe
(μg/g)
Al
(μg/g)
Rb
(μg/g)
BT31918.02Granular dolostone0.534119114976358149931.624
BT31921.48Calcareous dolostone0.3751685455611259236090/
BT44306.97Granular dolostone0.531207179916814,15367922.363
BT44316.09Micrite0.0322681783836136936980.956
BT44317.55Calcareous dolostone0.0862216473415003880/
M104397.30Granular dolostone0.5556741295119357544803.271
M104398.10Calcareous dolostone0.52616620120160710260171.445
M104400.00Calcareous dolostone0.4591769107152842352602.056
M104404.94Calcareous dolostone0.45516734661413834001.038
M104406.54Calcareous dolostone0.432159294871529053151.664
Table 12. Homogenization temperature of fluid inclusions within calcite in granular dolostones from Xiaohaizi Formation.
Table 12. Homogenization temperature of fluid inclusions within calcite in granular dolostones from Xiaohaizi Formation.
WellDepth (m)Host MineralTypeGas/Liquid RatioTh (°C)
BT44312.11calcite cementsaline8107.3
BT44312.11calcite cementsaline6106.7
M104407.54calcite cementsaline7103.4
M104407.54calcite cementsaline696.3
BT44316.09calcite veinsaline686.1
BT44316.09calcite veinsaline781.8
BT44316.09calcite veinsaline694.2
BT44316.09calcite veinsaline588.1
BT44316.09calcite veinsaline885.7
M104400.00calcite cementsaline881.7
M104400.00calcite cementsaline776.3
M104400.00calcite cementsaline684.3
M104400.00calcite cementsaline695.2
M104400.00calcite cementsaline782.4
BT44306.97calcite cementsaline1073.6
BT44306.97calcite cementsaline8122.1
BT44306.97calcite cementsaline786.3
BT44306.97calcite cementsaline7102.1
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, X.; Fu, M.; Gluyas, J.; Song, R.; Lan, H.; Fan, Y.; Wu, D. Dolomitization Facilitated by Clay Minerals on Mixed Siliciclastic-Carbonate Shoals of Carboniferous Age in the Tarim Basin, China: Constraints on Element Mobility and Isotope Geochemistry. Minerals 2025, 15, 419. https://doi.org/10.3390/min15040419

AMA Style

Liu X, Fu M, Gluyas J, Song R, Lan H, Fan Y, Wu D. Dolomitization Facilitated by Clay Minerals on Mixed Siliciclastic-Carbonate Shoals of Carboniferous Age in the Tarim Basin, China: Constraints on Element Mobility and Isotope Geochemistry. Minerals. 2025; 15(4):419. https://doi.org/10.3390/min15040419

Chicago/Turabian Style

Liu, Xuan, Meiyan Fu, Jon Gluyas, Rongcai Song, Haoxiang Lan, Yunjie Fan, and Dong Wu. 2025. "Dolomitization Facilitated by Clay Minerals on Mixed Siliciclastic-Carbonate Shoals of Carboniferous Age in the Tarim Basin, China: Constraints on Element Mobility and Isotope Geochemistry" Minerals 15, no. 4: 419. https://doi.org/10.3390/min15040419

APA Style

Liu, X., Fu, M., Gluyas, J., Song, R., Lan, H., Fan, Y., & Wu, D. (2025). Dolomitization Facilitated by Clay Minerals on Mixed Siliciclastic-Carbonate Shoals of Carboniferous Age in the Tarim Basin, China: Constraints on Element Mobility and Isotope Geochemistry. Minerals, 15(4), 419. https://doi.org/10.3390/min15040419

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