You are on page 1of 21

Marine and Petroleum Geology 163 (2024) 106740

Contents lists available at ScienceDirect

Marine and Petroleum Geology


journal homepage: www.elsevier.com/locate/marpetgeo

Magma-driven thermal convection dolomitization of the Lower-Middle


Ordovician strata caused by Mg sourced from underlying Cambrian
dolomites in the Tarim Basin, China
Haofu Zheng a, c, Feifan Lu b, c, *, Yixin Dong d, Bo Liu e, Xuefeng Zhang e, Kaibo Shi c, Jiajun He a,
Hairuo Qing f
a
College of River and Ocean Engineering, Chongqing Jiaotong University, Chongqing, 400074, China
b
Department of Geological Engineering, Faculty of Geosciences and Environmental Engineering, Southwest Jiaotong University, Chengdu, 611756, China
c
School of Earth and Space Sciences, Peking University, Beijing, 100871, China
d
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation, Chengdu University of Technology, Chengdu, 610059, China
e
Institute of Oil and Gas, Peking University, Beijing, 100871, China
f
Department of Geology, University of Regina, Regina, SK, S4S0A2, Canada

A R T I C L E I N F O A B S T R A C T

Keywords: Thermal convection arises due to a density gradient between geothermally heated formation water within the
Thermal convection underlying strata and the water in the overlying strata. The thermal convection model has long been accepted as
Dolomitization a stable and long-lived flow mechanism for dolomitization, which can affect large volumes of rock. This study
Mg2+ flux
investigates the dolomitization of the Lower-Middle Ordovician strata in the Tarim Basin using petrological
Ordovician
Tarim basin
evidence, geochemical data, and the Calculation of Mg2+ flux. The Ordovician dolomites occurred in several
types, including 1) individual dolomite crystals; 2) patchy dolomites; 3) layered-massive dolomites; 4) cement
dolomites in pores and fractures. The dolomite content decreases vertically with decreasing burial depth.
Geochemically, the dolomitizing fluids of the Ordovician carbonates are suggested as deep-formation water
slightly influenced by hydrothermal fluids, with an obvious genetic connection with the underlying Cambrian
dolomites. In addition, mass balance calculations were conducted to help identify the Mg2+ sources, supporting
that the Mg-rich dolomitizing fluids were derived largely from the underlying strata. Large volumes of Mg-rich
fluids were released when the underlying contemporaneous Cambrian dolomites were physically and chemically
compacted. The released fluids were driven by heat from a high-temperature Early Ordovician magma chamber
and migrated upward through fractures and/or permeable strata due to hydrothermal convection. These pro­
cesses caused thermal convection dolomitization of the overlying Ordovician limestones from bottom to top. This
study explains how these widespread dolomites developed in the Ordovician strata of the Tarim Basin, which can
help identify the locations of the oil-gas reservoir rocks. In addition, this revised thermal convection model
highlights the contribution of Mg-rich fluids derived from the underlying dolomites, and the circulation of fluids
may have been driven by the heat from a magmatic source. This model is believed to be of significance to areas
affected by abnormal geothermal activities.

1. Introduction in an atoll or platform (Fanning et al., 1981; Saller, 1984; Aharon et al.,
1987; Hein et al., 1992; Kaufman, 1994). This model has been proven
Thermal convection occurs when a density gradient generates be­ effective by numerical methods and is believed to be a stable and
tween the geothermally heated formation water within the underlying long-lived flow mechanism for dolomitization, which can affect large
strata and that within the overlying strata. The thermal convection volumes of rock (Sanford et al., 1999; Jones et al., 2003; Whitaker et al.,
model has long been accepted as a typical dolomitization model for use 2004).

* Corresponding author. Department of Geological Engineering, Faculty of Geosciences and Environmental Engineering, Southwest Jiaotong University, Chengdu,
611756, China.
E-mail address: lufeifanbh@163.com (F. Lu).

https://doi.org/10.1016/j.marpetgeo.2024.106740
Received 6 December 2023; Received in revised form 18 January 2024; Accepted 31 January 2024
Available online 9 February 2024
0264-8172/© 2024 Elsevier Ltd. All rights reserved.
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

The commonly used thermal convection dolomitization model, or a addition, Ordovician magma activity is believed to have provided the
variation model called Kohout Convection (Kohout, 1967; Simms, heat source, as evidenced by petrological data, geochemical data, and
1984), demonstrates the convection of original or pressurized seawater the calculation of Mg2+ flux. The effects of special hydrologic models on
after being heated by a normal geothermal gradient at the margin of a thermal convection dolomitization are elucidated in this study,
platform or within a platform (Fig. 1; Kaufman, 1994; Whitaker and providing important insights into the impact of abnormal geothermal
Xiao, 2010; Saller and Dickson, 2011). In addition, the stratum brine can features on dolomitization.
also enter the permeable layer under the control of gravity, and then rise
after being heated by the ground temperature, forming dolomite in the 2. Geological setting
permeable layer of limestone (Wendte et al., 1998; Saller and Dickson,
2011). Considering the large amount of Mg2+ needed for extensive The Tarim Basin, situated in China, is a vast superimposed basin that
dolomitization, it is understandable that a large volume of seawater has formed on a continental crust basement, primarily comprising Pro­
from the sea is required. However, there may be other fluids providing terozoic metamorphic rocks. (Fig. 2A; Tuo and Philp, 2003). The
the large volume of Mg2+ needed for dolomitization. Whether they have structural framework of the basin is divided into three major episodes of
the potential to form large-scale thermal convection dolomites under uplift and four episodes of subsidence based on the complex deformation
specific hydrologic conditions is unknown. and displacement that occurred during its multi-stage tectonic evolution
The Tarim Basin, situated in northwest China, represents a vast and (Fig. 2B and C; Tang et al., 2012).
intricate sedimentary basin that has undergone numerous phases of During the Caledonian stage (Sinian–Ordovician), the Tarim Basin
tectonic activity, including volcanism, and deformation (Chen et al., was an intra-cratonic extensional basin, with a marine carbonate plat­
1997; Tang et al., 2012; Deng et al., 2019). Dolomite predominantly form (study area) in the west and an abyssal basin in the east (Li, 1995;
manifests in the Lower Paleozoic, notably within deeply buried (>4000 Kang and Kang, 1996; Lin et al., 2012). The western platform of the
m) Cambrian and Lower Ordovician strata, subjected to a multifaceted, Tarim Basin is composed of a succession of several kilometers of marine
multi-stage diagenetic evolution. Recent years have witnessed a bur­ carbonates, including the Upper Sinian Qigbulak Formation (Fm.); the
geoning interest in exploring deeply buried extensive dolomites as Lower Cambrian Yurtusi, Xiorbrag and Wusongger Fms.; the Middle
prospective reservoirs, intensifying curiosity surrounding their origins. Cambrian Awatag and Shayilike Fms.; the Upper Cambrian Qiulitage
Extensive research has been conducted on Early Paleozoic dolomite Fm.; the Lower Ordovician Penglaiba Fm.; the Lower-Middle Ordovician
origin in the Tarim Basin, and its origin was once explained as wide­ Yingshan Fm.; the Middle Ordovician Yijianfang Fm.; and the Upper
spread penecontemporaneous dolomitization superimposed on mixed Ordovician strata (Fig. 2D; Zhang et al., 2009; Jiang et al., 2016). A
water dolomitization (Shen et al., 1995; Yang and Wang, 2000). How­ regional unconformity formed between the Yingshan Fm. and the
ever, a growing body of scholarly opinion posits that only limited-scale overlying strata because of the tectonic uplift during the second stage of
penecontemporaneous dolomite may exist within the Ordovician Sys­ the Caledonian movement, which may include the Middle-Ordovician
tem of the Tarim Basin due to the absence of evaporites in this strata Yijianfang Fm., the Upper Ordovician Lianglitage Fm., or the
(Zhang et al., 2008; Dong et al., 2013; Qiao et al., 2021). Through the Lower-Carboniferous Bachu Fm. (Lin et al., 2012).
augmentation of oilfield drilling data, the identification of various hy­ The Cambrian strata are mainly dolomites with Middle Cambrian
drothermal minerals, marbleization, and silicification (Jin et al., 2006; evaporitic anhydrites and mudstones. Most of the Cambrian dolomites
Dong et al., 2013; Zhu et al., 2010; Du et al., 2018) suggests the for­ are believed to have formed contemporaneously (Fig. 2D; Zhang et al.,
mation of dolomite from hydrothermal metasomatic limestone. Addi­ 2009). The 2–3 km thick Ordovician carbonate platform is composed of
tionally, some conjecture exists regarding the burial recrystallization of a completely dolomitized lower unit, a dolomite-limestone interbedded
penecontemporaneous dolomite in a thermal background (Zhu et al., middle unit, and a limestone-dominated upper unit (Fig. 2D; Zhang
2010; Guo et al., 2020; Qiao et al., 2021; Qing et al., 2023). Nonetheless, et al., 2008), many of which preserve hydrocarbons (Wang et al., 2014).
there remains considerable contention surrounding the specific dolo­ Therefore, understanding how these dolomites developed is key to un­
mitization hydrological model (Hu et al., 2019; Guo et al., 2020; Qiao derstanding how the petroleum reservoirs in this area formed and where
et al., 2021), and complex geochemical evidence fails to unequivocally they are located (Zhang and Luo, 2010; Zhu et al., 2010). The Wells used
support the assertion that the extensive Lower Ordovician dolomites are in this study are mainly distributed in the central Tarim Basin (Fig. 2B
exclusively linked to mantle-derived hydrothermal fluids. Furthermore, and C).
multiple magmatic events contribute to dolomitization, the source, na­ Multiple stages of faultings occurred during the Caledonian, Hercy­
ture, and migration pathways of Mg2+, remain unresolved. nian, and Himalayan orogenies in the study area (Fig. 2C; Tang et al.,
This study introduces a model for the hydrothermal convection of 2012; Zhu et al., 2015; Qiu et al., 2019). Moreover, the basin experi­
formation water derived from compaction of the underlying Cambrian enced four periods of magmatic and volcanic activity during the period
dolomite strata, which was responsible for the dolomitization of the of Ediacaran-Cambrian, the Early Ordovician, the Late
overlying Ordovician limestones in the Tarim Basin, NW China. In Carboniferous-Early Permian, and the Cretaceous (Chen et al., 1997;

Fig. 1. Hydrologic model for thermal convection in a carbonate platform with a steep–sided margin (Kaufman, 1994). At the platform’s margin, cold marine waters
are drawn into the platform, heated, and rise to the surface, which is called Kohout convection (Kohout, 1967; Simms, 1984). In the interior of the platform, the heat
flow across the basal boundary in the center of the model, elevated to a value of 170 mW/m2, produces a convection cell in which surficial seawater is drawn into the
platform. A constant basal heat flow of 63 mW/m2 is used.

2
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

Fig. 2. (A) The location of the Tarim Basin in China. (B) A simplified (or schematic) map of the Tarim Basin, showing the major tectonic features and the locations of
wells sampled for this study (Wei et al., 2021). (C) Sketch map of the fault structure distribution of the study area and Cross-section of the fault movement during the
Middle Caledonian, The strike-slip faults in the Ordovician show positive flower structure (Tang et al., 2012; Qiu et al., 2019). (D) The Cambrian–Ordovician strata of
the Tarim Basin, showing the overall vertical distribution of dolomite.

Zhang et al., 2010; Yu et al., 2011; Dong et al., 2013). During the Early 3. Samples and methods
Paleozoic, a series of EN-trending thrust-faulted-and fault-bend folds
(Fig. 2C) formed in the middle of the Tarim Basin (Tang et al., 2012). 3.1. Samples and microscopic observation
Simultaneously, the Penglaiba Formation and Yingshan Formation have
been truncated by erosion because of the large-scale uplift movement in A total of 271.04 m of cores from 20 wells were analyzed (Fig. 2B–C).
this area (Lu et al., 2017; Hu et al., 2019). These EN-trending faults The 321 core samples were collected perpendicular to the bedding
formed in the Tangguzibasi Depression during the late Early Paleozoic surfaces at 0.2–2.0 m intervals based on changes in lithology, texture,
were reactivated during the Middle Paleozoic, but the intensity and mineralogy, and porosity. This study primarily focuses on the lime­
degree of faulting activity were significantly less than those in the latest stones, dolomites, and siliceous rocks of the Lower to Middle Ordovician
Ordovician (Tang et al., 2012; Hu et al., 2019). However, before the late in the study area, specifically targeting the Penglai Formation and the
Permian, abnormal thermal event also exerts a great influence on the Yingshan Formation. The limestones encompass argillaceous limestone,
Ordovician strata. During the Early-Middle Permian, the lithosphere of wackestone, packstone, and grain limestone. According to the size of the
basin was broken up and large scale magmatism occurred while the crystal, dolomite is divided into cryptocrystalline (<1 μm), microcrys­
uplift force exceeded the tortuosity of lithosphere (Tang et al., 2012; talline (1–10 μm), very fine crystalline (10–100 μm), finely crystalline
Deng et al., 2019). The basic basalt lavas or intrusive diabases, and in­ (100–250 μm), medium crystalline (250–500 μm), coarsely crystalline
termediate acidic magmas distributed extensively in the (500–4000 μm) and very coarse crystalline (>4 mm) (Bissell and Chi­
central-northern part of basin (Fig. 2C–He et al., 2019), related to the lingar, 1967). The distribution patterns of very fine to finely crystalline
evolution of Paleo-Tethys Ocean (Chen et al., 1997). Therefore, the and medium to coarsely crystalline dolomites may exhibit characteris­
hydrothermal fluids related to the faultings and magmatic episodes may tics of dispersion, patchiness, or layering. The selected dolomite samples
have caused hydrothermal reworking or alteration of the Cambrian and for analytical testing predominantly manifest as patchy or layered do­
Lower Ordovician strata (Zhu et al., 2010, 2015; Dong et al., 2013; Du lomites. Dolomite cements commonly occur within cavities or fractures.
et al., 2018; Hu et al., 2019). Siliceous rocks typically consist of chert or quartz. The chert is often
associated with matrix limestone or dolomite, while the quartz fills
fractures. Two hundred and forty polished and Alizarin Red-S-stained
thin sections were prepared. Some of the thin sections were

3
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

impregnated with blue epoxy for porosity identification. The optical and Ni, 2007; Steele-MacInnis et al., 2011; Lu et al., 2017).
investigation was conducted conventionally using a microscope. Cath­
odoluminescence microscopy (CL) was carried out on the unstained thin 4. Results
sections using a RELION cold cathodoluminescence stage with a 5.8 kV
beam and a current intensity of 0.7–0.8 mA. 4.1. Petrology and Mineralogy

3.2. Geochemical analysis We have categorized dolomite based on its occurrence into individ­
ual dolomite crystals, patchy dolomites, layered-massive dolomites, and
Whole rock samples were powdered to less than 200 mesh. The cement dolomite. These dolomites can be further classified based on the
prepared powder samples were used for major and trace elements crystal size into very fine to finely crystalline dolomites and coarsely to
(containing rare earth elements) and isotope analysis. The petrographic medium crystalline dolomites. Additionally, in the Middle-Lower
characteristics of the Ordovician carbonate rocks are shown in “4.1. Ordovician, there are isolated occurrences of siderite dolomite and
Petrology and Mineralogy”. All Cambrian dolomites are very fine to siliceous dolomite. However, their volumes are relatively small (less
finely crystalline dolomites (50–200 μm), and the specific petrographic than 1% of the total carbonate volume) and are not the focus of our
characteristics of the Cambrian dolomites are shown in Du et al. (2018), study.
Jiang (2022) and Liu et al. (2022). All of the geochemical analysis of
elements were performed at Key Laboratory of Orogenic Belts and 4.1.1. Individual dolomite crystals – Id
Crustal Evolution, Peking University, Beijing. Ids are common but make up a small volume of the Lower-Middle
A total of 94 samples were selected and analyzed for oxygen and Ordovician strata, especially in the upper part of this section. They are
carbon isotopes. Oxygen and carbon isotopes were analyzed using featured by turbid finely crystalline planar dolomites (Fig. 3A), a few
standard analytical techniques and a Finnigan MAT 253 mass spec­ thin sections show siliceous rock, and dolomite floating among crystal­
trometer equipped with a Thermo Finnigan Gasbench II. The Oxygen line quartz (Fig. 3B). They are generally formed by replacing calcite
and carbon isotopic data are reported in per mil (‰) relative to the cement and/or grains. Selective dolomitization of grains can be seen in
Vienna PeeDee Belemnite (VPDB) standard. The precision was better the thin sections (Fig. 3C). Ids are pink-dull red under cath­
than 0.05‰ for both δ18O and δ13C. odoluminescence (CL) and calcites (grains and cement) are dull red
A total of 36 samples were selected and analyzed for strontium iso­ (Fig. 3C and D).
topes. Strontium isotopic ratios were measured using a Finnigan MAT
261 mass spectrometer. For strontium isotope analyses, ~50–100 mg 4.1.2. Patchy dolomites – Pd
powders were dissolved in 1 mL HNO3 in a Teflon container at 190 ◦ C for Pds are very common in the Lower-Middle Ordovician formations.
48 h. Strontium was extracted using conventional ion exchange pro­ Dolomites are usually patchy and coexist with limestone (Fig. 4A). They
cedures utilizing ion exchange resin and then subjected to 87Sr/86Sr are featured by planar finely crystalline dolomites (Fig. 4B–D). Most of
measurement on a Triton Plus thermal ionization mass spectrometer, the crystals are featured by cloudy cores but with no clear rim texture
which was corrected using the SRM-987 standard, yielding a mean error (Fig. 4F). Typically, Pds are observed along the stylolites (Fig. 4A–E).
of ±0.5 × 10− 5 (2σ) for 87Sr/86Sr ratios. Pds are pink-dull red in color under CL (Fig. 4F and G).
A total of 34 samples were selected and analyzed for major elements.
Major elements were analyzed using X-ray fluorescence (XRF). After 4.1.3. Layered-massive dolomites - Ld
acid digestion of the samples in Teflon bombs, the trace and rare earth The dolomites in the lower unit of the Yingshan and Penglaiba Fms.
element (REE) compositions were determined using an HR-ICP-MS are generally layered or massive in texture. Similar to Pds, Lds can be
(Element I; Finnigan MAT) at 22 ◦ C in 30% humidity. further subdivided into planar very fine to finely crystalline dolomites
A total of 23 samples were selected and analyzed for trace and rare (Ld1, Fig. 5A–C) and non-planar coarsely to medium crystalline dolo­
earth elements. The analysis of trace and rare earth elements required mites (Ld2, Fig. 5E–G). Some of the dolomites exhibit clear relict tex­
0.5 mL concentrated HNO3 to dissolve and then dry the 50 mg sample tures (Fig. 5A and B), while others preserve little textural information of
powders. The dried sample was dissolved in 5 mL of HNO3 with a con­ their precursor limestones (Fig. 5C–E, G). We can only infer the original
centration of 1.42 g/mL and was heated for 3 h (at 130 ◦ C). Then, the fabrics from the turbid or cloudy dolomite crystals of these dolomites.
heated solution was diluted to 60 mL by adding ultra-pure H2O for trace The Ld1s and Ld2s are generally dull red in color under CL (Fig. 5D–F).
element testing. The contents of trace and rare earth elements were
determined by using an inductively coupled plasma mass spectrometer 4.1.4. Cement dolomites
(ICP-MS, Agilent 7500, USA). For the standardization of REE data, we Cement dolomites occur as fillings in veins or pores, exhibiting an
used the data of Post-Archean Average Shale (PAAS). overall scarce distribution (Fig. 6A and B). The dolomite veins are
mostly located southeast of the Shunnan (SN) area and are nearly ver­
3.3. Fluid inclusion study tical in occurrence (Fig. 7). The pore-filling dolomite cements are mainly
found along the edges of large pores (Fig. 6C–E). The crystal sizes of the
A total of 86 samples were selected and analyzed for fluid inclusion vein- or pore-filling dolomites are generally larger than those of the
microthermometry. The fluid inclusion microthermometry was per­ matrix dolomites, and the crystal faces are much clearer. Besides, the
formed at the Geofluids Lab, University of Regina. Fluid inclusion ana­ cement dolomites in the pores are saddle-shaped in units with abnor­
lyses were conducted on double-polished thin sections using a Linkam mally coarsely matrix dolomite crystals (Fig. 6D). Cement dolomites
THM600 heating-freezing stage, which was calibrated using synthetic exhibit dull red color under CL (Fig. 6F and G).
fluid inclusions with known compositions. Prior to the start of the in­
clusions thermometry experiment, we conducted a careful petrological 4.2. Dolomite distribution
study of the inclusions and identified the primary and secondary in­
clusion. Inclusion homogenization temperature and freezing tempera­ The dolomite content generally decreases vertically with decreasing
ture were measured by cyclic thermometry. The heating and/or cycling burial depth (Fig. 7), along with systematic textural changes. The Middle
intervals are 0.1 ◦ C for ice-melting temperatures (Tm-ice) and hydro­ Ordovician Yijianfang Fm. contains few Id. The Lower-Middle Ordovi­
halite melting temperatures (Tm-HH) and 1 ◦ C for homogenization cian Yingshan Fm. is the transitional zone between pure limestone and
temperatures (Th). Therefore, the precision of temperature measure­ Lds, with numerous Ids and/or Pd present in those partially-dolomitized
ment is 1 ◦ C for Th and 0.1 ◦ C for Tm-ice and Tm-HH (Bodnar, 1993; Chi units. The Lower Ordovician Penglaiba Fm. is characterized by Lds and

4
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

Fig. 3. Thin section photomicrographs of individual dolomite crystals. (A) finely crystalline planar–e dolomites in grainy limestone, Well SN1, Shunnan area, O1–2y,
6968.3 m, PPL (PPL = plane-polarized light); (B) finely crystalline planar–e dolomites in silicified grainy limestone, Well YB3–1, Yubei area, O1–2y, 5034.8 m, CPL
(CPL = cross-polarized light); C(PPL) and D(CL): individual rhombic dolomite crystals in limy–packstone, with pink–dull red color under CL; calcite within both
grains and non–luminescent cements, Well Z101, Tazhong area, O1–2y, 5535.74 m.

minor Pds. A general trend is that the grain size of the dolomites in­ similarly positive δ13C values (ranging from 0.5 to − 1.7‰VPDB) and
creases with increasing burial depth. Coarser-grained dolomites are negative δ18O values (− 14 to − 5.7‰VPDB).
more common in the Penglaiba Fm. than in the Yingshan Fm.
Horizontally, the dolomite content is significantly heterogeneous 4.3.2. Radiogenic Sr isotopes
throughout the Central Uplift area of the Tarim Basin. In the Shunnan The Ordovician limestones have the lowest 87Sr/86Sr ratios
area, the dolomites are weakly developed except for in the eastern re­ (0.7086–0.7089). These values are similar to those of strontium isotopes
gion (i.e., Gulong (GL) Area). While dolomites are more common in the in Ordovician seawater (Fig. 9). In addition, 87Sr/86Sr ratio increases
Tazhong and Yubei (YB) areas (Fig. 7). The scale of dolomite develop­ with increasing dolomite content, the 87Sr/86Sr ratios of the Ordovician
ment is larger in the position closer to faults of different ages (Fig. 2C). dolomites range from 0.7090 to 0.7096, except for 2 samples exhibiting
Notwithstanding the horizontal heterogeneity of the dolomites, the extraordinarily high 87Sr/86Sr ratios (0.7106–0.7110; Fig. 9). The
87
overall vertical decrease in the dolomite content with decreasing depth Sr/86Sr ratios of the Ordovician siliceous rocks range from 0.7090 to
in the Ordovician strata is likely similar throughout the western Tarim 0.7095, except for 1 sample exhibiting extraordinarily high 87Sr/86Sr
platform (Fig. 6), including outcrops in the northwest area (Dong et al., ratios (0.7105; Fig. 9) The 87Sr/86Sr ratios of several the Middle-Upper
2013), the western Tarim Basin (Zhang et al., 2009), Tazhong (central Cambrian dolomite samples are also plotted in Fig. 9. These values are
Tarim; Liu et al., 2009), Tabei (northern Tarim; Mao et al., 2014), and generally similar to those of the Ordovician dolomites (Fig. 9).
the southeastern platform margin (Wang et al., 2014).
4.3.3. Major and trace elements
4.3. Geochemistry The CaO contents of the total Ordovician dolomites range from
22.72% to 33.20%, with an average of 29.01% (Fig. 10A), which is close
4.3.1. Stable carbon and oxygen isotopes to the theoretical CaO content of stoichiometric dolomite (30.44%). The
The carbon and oxygen isotopic values are plotted in Fig. 8. Most of MgO contents range from 15.05% to 21.61% (average 18.26%)
the host limestones have a relatively small range of δ18O and δ13C (Fig. 10A), which is lower than the theoretical MgO content of stoi­
values, with δ13C ranging from − 2.4 to − 0.9‰VPDB and δ18O ranging chiometric dolomite (21.74%). The Ca and Mg contents of the Ordovi­
from − 9‰ to − 6.5‰VPDB. The very fine to finely crystalline dolomites cian limestones and dolomite exhibit a distinctly negative relationship.
have δ18O values of − 7.8 to − 5.5‰VPDB, δ13C values of − 2.5‰ to The content of Ca and Mg is positively correlated between Cambrian
− 0.5‰VPDB. The coarsely to medium crystalline dolomites have a dolomite and Ordovician dolomite (Fig. 10A).
larger range of δ18O values (− 9.7 to − 4.5‰VPDB) and a more narrow Generally, the MnO contents of the very fine to finely crystalline
δ13C range (− 2.1 to − 0.3‰VPDB) close to that of the Cambrian dolomites are relatively low (0.003–0.012%), while the Fe2O3 contents
replacement dolomites. The δ18O of Cement dolomite ranges from − 9.3 are relatively high (0.04–0.28%). The MnO contents of the coarsely to
to 5.9‰VPDB, and the δ13C ranges from − 1.8 to 0.6‰VPDB. In contrast medium crystalline dolomite vary between 0.004 and 0.012%, and
to the Ordovician replacement dolomites, the siliceous rocks have Fe2O3 contents vary between 0.054 and 0.279% (Fig. 10B).

5
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

Fig. 4. Core photo and thin section photomicrographs of pds. (A) A photo of typical patchy dolomites (Pds) in a core, Well Z101, Tazhong area, O1–2y,
5551.0–5551.2 m; (B) finely crystalline planar–s dolomites in thin section, separated from the limestone by mud–and–pyrobitumen–filled stylolite, Well Z101,
Tazhong area, O1–2y, 5548.93 m, PPL; (C) finely crystalline planar–s dolomites along mud–and–pyrobitumen–filled stylolite, Well Z19, Tazhong area, O1–2y, 5524.51
m, PPL; (D) finely crystalline planar–e dolomites along mud–and–pyrobitumen–filled stylolite, Well SN1, Shunnan area, O1–2y, 6966.3 m, PPL; (E) finely crystalline
planar–s dolomites developed along mud–and–pyrobitumen–filled stylolite, Well YB3, Yubei area, O1–2y, 5362.3 m, PPL; F(PPL) and G(CL): dolomite with relict
texture, dolomite exhibiting dull red color under CL, and non–luminescent relict limestone, Well YB5, O1–2y, 6839.75 m.

6
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

Fig. 5. Thin section photomicrographs of layered–massive dolomites. (A) fabric–retentive very fine to finely crystalline dolomites showing the texture of its grainy
limestone precursor, Well YB5, Yubei area, O1p, 6841.6 m; (B) very fine to finely crystalline dolomites with parallel bedding, indicating a limy–mudstone precursor,
Well YB5, Yubei area, O1p, 6841.2 m; C(PPL) and D(CL): fabric–destructive dolomite, pyrobitumen developed in the inter–crystalline pores, dolomites exhibit dull
red color under CL, dirty cloudy dolomite crystals and particle morphology indicate the precursor was grainy limestone, Well YB5, Yubei area, O1p, 6605.53 m; E
(PPL) and F(CL): fabric–destructive coarsely crystalline dolomites, original fabrics of precursors have been obliterated, dolomites exhibit dull red color under CL, Well
YB2, Yubei area, O1–2y, 5609.45 m. G(PPL) and H(CPL): coarsely crystalline dolomites, cryptocrystalline quartz is associated with dolomite under cross-polarized
light, Well SN1, Shunnan area, O1–2y, 6966.3 m.

7
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

Fig. 6. Core photo and thin section photomicrographs of cement dolomites. (A) handspecimen of fine grained dolomite with coarsely to medium crystalline dolomite
veins, Well GL1, Shunnan area, O1–2y, 6543.8 m; (B) thin section made from the specimen in (A), showing the fine grained dolomite wall rock and a coarsely to
medium crystalline crystalline dolomite–filled vein, CPL; (C) clear, crystalline dolomite partly cemented pore, Well YB6, Yubei area, O1–2y, 6708.00 m, PPL; (D)
saddle shaped dolomite (SD) cementing a large pore, followed by calcite (stained red) filling the rest of the pore space. Well YB3, Yubei area, O1–2y, 5261.37 m, CPL;
(E) coarsely to medium crystalline dolomites and bitumen filling in the moldic pore of a finely crystalline dolomite, Well Z101, Tazhong area, O1–2y, 5535.74 m, PPL.
F(PPL) and G(CL): coarsely to medium crystalline dolomites filling in the moldic pore of finely crystalline dolomites, all dolomites exhibit the same dull red color
under CL, Well GL1, Shunnan area, O1–2y, 6533.1 m.

4.3.4. Rare earth elements dolomite crystals. A total of 28 Fluid inclusions (FIs) in the coarsely to
All of the replacement dolomites have flat REE patterns after being medium crystalline dolomites and the cement dolomites, 23 FIs in the
normalized by the Post-Archean Australian Shale (PAAS) (McLennan, cement dolomites were measured, 14 FIs in the intercrystalline cement
1989) with slight Ce negative anomaly and slight HREE depletion, calcites were measured, and 11 FIs in the intercrystalline Crystalline
which resembles those of the Ordovician limestone and Cambrian very quartzs were measured.
fine to finely crystalline dolomites (Fig. 11A–C, E). The cement dolo­ The results indicate that coarsely to medium crystalline dolomites
mites and quartz primarily exhibit positive Eu anomalies (Fig. 11D–F). have homogenization temperatures (Th) of 85.7–143.9 ◦ C and salinities
of 2.5–23.1 wt% NaCl equivalent, cement dolomites have homogeni­
4.3.5. Microthermometry and salinity zation temperatures (Th) of 97.4–162.3 ◦ C and salinities of 4.9–21.3 wt
The samples for fluid inclusion analysis encompassed all types of % NaCl equivalent, crystalline quartzs have homogenization tempera­
dolomites as well as granular quartz, with a total of 85 valid data from tures (Th) of 156.1–172.3 ◦ C and salinities of 13.3–23.5 wt% NaCl
the examination of 80 thin-section samples. For most of the very fine to equivalent, Late calcites s have homogenization temperatures (Th) of
finely crystalline dolomites, fluid inclusions could not be found in the 84.9–136.5 ◦ C and salinities of 4–20.9 wt% NaCl equivalent (Fig. 12).

8
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

Fig. 7. A cross section based on well logs and core samples, showing dolomite contents of the Lower–Middle Ordovician strata (see Fig. 2B and C for well locations).
Top of the Yingshan Fm. (O1–2y) was flattened. Note the formations overlying O1–2y may be part of the Yijianfang Fm. (O2yj), the Lianglitag Fm. (O3l), or the Lower
Carboniferous Bachu Fm. (C1b), which contains one of the regional Caledonian unconformities. From bottom to top, the normal stratigraphic sequence should be O1p,
O1–2y, O2yj, O3q (Qiaerbake Fm. or O3t Tumuxiuke Fm.), O3l, and O3s (Sangtamu Fm.).

9
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

Fig. 8. A cross plot of δ13C–PDB (‰) and δ18O–PDB (‰) for the limestones and dolomites from the Middle–Lower Ordovician strata and dolomites from the
Middle–Upper Cambrian strata. Note there is a large overlap in the C–, O–isotopic compositions of the Ordovician dolomites and the Cambrian dolomites (Table S1).
Typical δ18O ranges for estimated Ordoviaian evaporative dolomite (Montañez and Read, 1992), and estimated coeval marine dolomite calculated based on the
equilibrium fractionation value (δ18Odolomite – δ18Ocalcite = + 3‰ VPDB; Land, 1980; Budd, 1997) between dolomite and cogenetic calcite, are given for comparison.

Fig. 9. A cross plot of the 87Sr/86Sr ratios of the Ordovician and Cambrian dolomites on the modified curve for seawater 87Sr/86Sr variation through refined geologic
time, seawater 87Sr/86Sr data from Burke et al. (1982), Nicholas (1996), and Ebneth et al. (2001) (Table S1).

5. Discussion 2015). The carbonate rocks in the study area exhibit extremely low total
rare earth element (ΣREE) concentrations (ranging from 2.382 to
5.1. Evaluation of the fidelity of geochemistry data 25.905 ppm, average 6.759 pmm, Table S3), among them, the average
ΣREE value of limestone is 6.579 ppm, and the average ΣREE content of
The trace elements in marine carbonate rocks are often influenced by dolomite is 6.828 ppm. Furthermore, the concentrations of trace ele­
terrestrial siliciclastic minerals (Nothdurft et al., 2004). Hence, prior to ments of the samples such as Sc (0.38 ppm), Zr (4.44 ppm), Hf (0.30
scrutinizing the trace geochemical data in this study, an assessment of ppm), and Th (0.88 ppm) are notably lower than those of the average
the degree of terrigenous siliciclastic minerals contamination was con­ crustal abundance (Sc = 14.90 ppm, Zr = 240 ppm, Hf = 5.8 ppm, Th =
ducted to ensure the reliability of the data. From a paleogeographic 2.3 ppm; Taylor and McLennan, 1981). In conjunction with the low
perspective, the study area was situated in the central Tarim platform ΣREE, we suggest that the Lower to Middle Ordovician limestones and
during the Ordovician, surrounded by extensive shallow sea and conti­ dolomites in the study area have not contaminated by terrestrial silici­
nental shelf, which is far away from the source regions (Gao and Fan, clastic minerals, thereby representing the characteristics of the original

10
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

Fig. 10. (A) A cross plot of CaO vs. MgO of the Ordovician limestones and dolomites; the Cambrian dolomite and cherts are plotted for comparison. (B) A cross plot
of Fe2O3 vs. MnO of different limestones and dolomites (Table S2).

diagenetic fluids. Wallace, 1997; Fabricius and Borre, 2010). Therefore, dolomitization
should begin around the same time (i.e., shallow burial stage at a burial
5.2. Relative timing of dolomitization depth of 500–800 m) with the pressure dissolution.
The crystalline quartz was found to have cut through dolomites,
The relationship between the formation time of the dolomite and the leaving behind residual rhombohedral dolomite crystals. (Fig. 5G and
timing of the other stages of diagenesis can be determined based on H). The genesis of heterogeneous quartz in the central region of the
macroscopic and microscopic observations. It can be seen from the Tarim Basin is believed to be related to magmatic hydrothermal activ­
microscopic observations that macrocrystalline dolomite occurred at the ities (Chen et al., 2015; He et al., 2019; Wei et al., 2021), and the for­
moldic pores, indicating that the dolomite formed after the moldic pores mation time could be from the Late Devonian to the Permian (Lu et al.,
formed (Fig. 6E and F). Selective dissolution is the main cause of moldic 2017; Lu et al., 2017). This may be related to the activation of Devonian
pore formation, which generally occurs during the early diagenesis of strike-slip faults and Permian volcanic activity during the Late Caledo­
carbonate minerals before stabilization (Gregg et al., 2015). In the Tarim nian–Hercynian period (Chen et al., 2015; Lu et al., 2017; Lu et al.,
Basin, this may be related to meteoric waters (Fu et al., 2020; Jiu et al., 2017). In addition, a large amount of bitumen developed on the surface
2022; Wang et al., 2022). Therefore, the dolomite formation occurred of or between the dolomite crystals on the pores (Fig. 4C and 5C). This
later than dissolution by the penecontemporaneous-early invasion of indicates that the charging of crude oil and postdated dolomite forma­
fresh water. The stylolite has a strong correlation with the distribution of tion. The first episode of oil and gas charging occurred in the Late
dolomite, indicating that the dolomite formed simultaneously with the Caledonian–Early Hercynian. The burial depth of the Middle-Lower
stylolite (Fig. 4A–E; Paganoni et al., 2016; Gomez–Rivas et al., 2022). Ordovician strata was about 2300 m at that time (Zhang et al., 2002).
Pressure dissolution in the clay-poor carbonates generally begins to in­ Therefore, the dolomite formed before it was buried at 2300 m. It is
crease at around 550 m (1800 ft). In addition, many stylolites would consistent with the conclusion of fluid inclusion analysis by predecessors
form at 830 m (2723 ft) (Dutton and Diggs, 1992; Nicolaides and (Wang et al., 2016; Zhang et al., 2019).

11
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

Fig. 11. Histogram of the PAAS normalized REE contents (McLennan, 1989) of the Ordovician limestones and dolomites with the Cambrian dolomites plotted for
comparison (Table S3).

Fig. 12. A cross plot of inclusion homogenization temperature and salinity. The center of the cross-shaped symbol is the average value of temperature and salinity
values, and the end of the cross-shaped symbol is the maximum and minimum value of temperature and salinity values, the color of the cross-shaped symbol is
consistent with the lithology symbol (Table S4).

12
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

In summary, the primary dolomitization process began during the underlying Cambrian dolomites.
pressure dissolution and continued to develop after pressure dissolution
ended. Recrystallization of dolomites ended during the first phase of oil 5.3.2. Comparison of dolomitizing fluids in the Tarim Basin
and gas charging but before the influence of siliceous fluids. The depth Through a comprehensive regional comparison, both consistency
was approximately 500–2300 m when the dolomite was formed. and difference in dolomite composition and origin across diverse regions
of the Tarim Basin have been identified.
5.3. Dolomitizing fluids The geochemical attributes of the majority of Lower and Middle
Ordovician matrix dolomites within the Tarim Basin exhibit a flat rare
5.3.1. Dolomitizing fluids in the study area earth element (REE) pattern. The carbon-oxygen strontium isotope
Several evidences suggest the primary dolomitizing fluids are deep- values observed in matrix dolomites are also generally consistent (Dong
formation water with some input of hydrothermal fluids. The overall et al., 2017; Du et al., 2018; Guo et al., 2020; Wei et al., 2021). The
flattened REE patterns of the dolomites resemble those of the Ordovician prevailing interpretation attributes the observed phenomenon to burial
limestone (Fig. 11A–C), suggesting that they were dolomitized from the dolomitization induced by altered seawater. This investigation eluci­
original limestone. In addition, the Ordovician dolomites have similar dates a robust correlation between the fluids emanating from matrix
REE patterns to the Cambrian dolomites, implying that the diagenetic dolomite and contemporaneous Cambrian dolomite formations in the
fluid is also related (Fig. 11B–C, E). The low values of MnO and Fe2O3 Central Tarim Basin. Anticipating further validation in subsequent in­
(Fig. 10B) also indicate that it was formed in an anoxic burial diagenetic vestigations across different locales within the Tarim Basin constitutes a
system (Jin et al., 2006). promising avenue for future research.
Most of the non-evaporite-related Ordovician dolomites show Widespread involvement of ultramafic-derived or magmatic fluids
slightly negative δ13C values and moderately negative δ18O values associated with Permian volcanism is observed in the central, northwest,
(Fig. 8), which is the typical feature of dolomite of burial origin (Warren, northern, and northeastern peripheries of the basin, caused by Hercy­
2000; Swart, 2015). It is noteworthy that the δ13C and δ18O composi­ nian tectonic activity. This results in the substantial formation of cement
tions of nearly all the Ordovician dolomite samples overlap with those of dolomite associated with mantle-derived hydrothermal fluids (Dong
the Cambrian dolomite samples, especially the Upper Cambrian dolo­ et al., 2013; Du et al., 2018; Guo et al., 2020; Wei et al., 2021). In both
mite samples (Fig. 8) (Gao and Fan, 2015), which were experienced the study area and the aforementioned regions, a portion of the
contemporaneous dolomitization in an evaporative-restricted platform cemented dolomite, along with nearly all saddle dolomite, exhibits
(Zhang et al., 2009; Zhou et al., 2012). The overlap of the δ13C and δ18O MREE enrichment characteristics akin to mid-acid magma. Further­
values of the Ordovician dolomites and the Cambrian dolomites (Fig. 8) more, these dolomite formations display oxygen isotope values lower
suggests that the dolomitizing fluids derived from the Cambrian strata than those observed in the matrix dolomite (Qiao et al., 2021; Guo et al.,
overprinted the geochemical features of the Ordovician dolomites. 2016; Dong et al., 2017; Wei et al., 2021). In addition, we also paid
The Ordovician dolomites have much higher 87Sr/86Sr ratios relative attention to the geochemical differences between different regions. The
to those of coeval seawater (0.7078–0.7096) (Veizer et al., 1999) δ18O values of very fine to finely crystalline dolomites in the northwest
(Fig. 9). However, the 87Sr/86Sr of seawater decreased from 0.7091 to and north of the Tarim Basin are more enriched than other dolomites.
0.7090 from the beginning of the Late Cambrian to the Early Ordovician The origin of this dolomite variant is likely linked to the processes of
and the 87Sr/86Sr ratios of Ordovician seawater were typically less than evaporation concentration and reflux seepage dolomitization during the
0.7091 (Fig. 9; Burke et al., 1982; Allen and Wiggins, 1993; Veizer et al., penecontemporaneous stage (Du et al., 2018; Qiao et al., 2021, Fig. 8).
1999; Ebneth et al., 2001). The 87Sr/86Sr ratios of the Ordovician do­ This shows that there are differences in ancient sedimentary environ­
lomites plot higher than those of the coeval seawater, but close to those ments in different areas within the Tarim Basin.
of the Middle-Upper Cambrian seawater (Fig. 9). It is highly unlikely
that the 87Sr/86Sr of the dolomites is directly derived from mantle fluids 5.4. Calculation of Mg2+ flux
nor the mixture of mantle fluids and Ordovician formation water,
because the average 87Sr/86Sr ratio of mantle-sourced fluids is ~0.7037 Sufficient Mg2+ has always been a constraint factor for the formation
(Banner, 2004) and the 87Sr/86Sr ratios of the igneous rocks (diabase and of massive dolomites (Machel, 2004). Conversely, the amount of Mg2+
gabbro) in the Tarim Basin are generally less than 0.7075 (Dong et al., required for dolomitization can be estimated from the amount of dolo­
2013). Therefore, the higher 87Sr/86Sr values suggest that the dolomi­ mites and dolomitic limestones. Thus, it could serve to determine the
tizing fluids were a mixture of Middle-Upper Cambrian and Ordovician formation mechanism of dolomite formation by comparing the amount
seawater because of the upward migration of Middle-Upper Cambrian of Mg2+ needed with the amount of Mg2+ derived from various sources
formation water (with high 87Sr/86Sr values). The magma-related hy­ (Gresens, 1967; Compton and Siever, 1986; Grant, 1986; MacLean,
drothermal fluids provided the energy for migration to a large extent. 1990; Gomez-Rivas et al., 2014; Centrella et al., 2023).
Because 87Sr in clastic rocks is enriched in fluids (Winter et al., 1997;
Duggan et al., 2001), so there are very high 87Sr/86Sr values in a few 5.4.1. Mg2+ needed for dolomitization
Ordovician dolomites (Fig. 9), which may be caused by the pollution of The Middle-Lower Ordovician dolomite content decreases with
Precambrian silicic fault gouge developed along fractures (Wu et al., decreasing depth, and the overall distribution pattern of dolomite in the
2020; Lu et al., 2023). western platform of the Tarim Basin is consistent. The Mg2+ content
However, part of the cement dolomites are characterized by coarsely required for the formation of the amount of dolomite per unit area is
crystalline, wavy extinction (Fig. 6C and D), more negative δ18O isotope calculated in this study based on the dolomite content in Well Zhong 4, a
values than those of the other dolomites (Fig. 8), high homogenization well with representative discution of dolomite.
temperatures (Fig. 12), and positive Eu anomalies (Fig. 11D), which Without considering compaction and pressure dissolution, the
were probably related to the influence of mantle-derived hydrothermal thickness of the O1-2 dolomite was calculated based on the cores from
fluids (Bau, 1991; Davies and Smith, 2006; Du et al., 2018; Guo et al., Well Zhong 4. The thicknesses of the dolomite (>90% dolomite), calcitic
2020). But these cement dolomites make up a tiny fraction of the for­ dolomite (>60% dolomite), and dolomitic limestone (<45% dolomite)
mation, roughly less than 1%. are 79 m, 112 m, and 181 m, respectively. The volume of pure dolomite
In conclusion, most of the Ordovician dolomites are interpreted to be was also calculated, and the mass was calculated according to the den­
dolomitized by deep-formation water but have been slightly influenced sity of the dolomite. The number of moles was then calculated from the
by hydrothermal fluids. Moreover, the dolomitizing fluids of the Ordo­ mole fraction of dolomite, which is the number of moles of Mg ions
vician carbonates have a certain degree of genetic connection to those of required. The results show that the amount of Mg2+ required per unit

13
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

area (1 m2) of the strata is about 3.46 × 106 mol (see the detailed steps in Table 2
Table 1). If the effects of compaction and pressure-dissolution are The Mg2+ content the Lower–Middle Ordovician formations in the Tarim Basin
considered, the amount of Mg2+ required should be much higher. can provide.
Example Wells Zhong–4 BT5
5.4.2. Mg2+ supply from multiple sources ① Thickness of Lower–Middle Ordovician (O1–2) strata 372 998
There may be three potential sources of Mg2+ for dolomitizing a (m)
certain stratum: 1) Mg2+ from water bodies above the stratum, including ② Volume of rock per unit (m3) 372 998
concentrated seawater and/or seawater mixed with other water reser­ ③ Present average porosity of the O1–2 <5% <5%
Present formation water volume (m3) = ② × ③
voirs (e.g., fluids flowing downward from the overlying strata); 2) pore ④ <18.6 <49.9
⑤ Thickness of the O1–2 before about 40% of the 620 1663
water in the same stratum; and 3) fluids moving upwards from the un­ original strata was pressure–dissolved (m) = ①/(1 −
derlying strata, including the fluids from the underlying strata and hy­ 40%)
drothermal fluids from the mantle (Davies and Smith Jr, 2006; Whitaker ⑥ Maximum historic porosity of the O1–2 40% 40%
and Xiao, 2010; Koeshidayatullah et al., 2020; Lu et al., 2023). ⑦ Formation water volume decreased by compaction 217 582
(m3) = ⑤ × (40%–5%)
The amount of Mg2+ that the Middle-Lower Ordovician strata could ⑧ Total maximum formation water volume (m3) = ④ 235 632
supply is limited. The current fluid volume per unit stratum area (1 m2) +⑦
does not exceed 50 m3 based on the thickness of the strata and its current ⑨ 3
Mg content of the O1–2 formation water (mol/m ): 2.68–452.09, 87.76 on
porosity. During the burial process, the porosity could have been average
Maximum Mg molar volume of the O1–2 formation 1.06 × 2.86 ×
reduced by approximately 1/3, i.e., by 13%, due to compaction (Hey­ ⑩
water (mol) = ⑧ × 452.09 105 105
dari, 2000). The thickness of the stratum lost due to pressure dissolution
is highly debatable, such as (1) 5%–34% thickness lost for limestone or
marble (Stockdale, 1922); (2) 13% for a platform margin reef and 24%
for lagoonal sediments (Mossop, 1972); (3) >20% for organic-rich car­ Table 3
The Mg2+ content the Middle–Upper Cambrian semi–contemporaneous dolo­
bonates (Di Primio and Leythaeuser, 1995), (4) 43% for grainstones
stone strata of the Tarim Basin can provide.
(Finkel and Wilkinson, 1990); (5) 40%–50% in general (Ramos, 2000);
and (6) 50%–60% for chalk (Safaricz and Davison, 2005). In this study, a Example Wells Zhong–4 TC1 BT5

total of 40% rock loss was assumed (Toussaint et al., 2018). Thus, the ① Thickness of the underlying Upper 1200 1800 994
thickness of the original strata was calculated, and the reduced amount Cambrian Qiulitage Fm. (m)
Volume of rock per unit (m3) 1200 1800 994
of strata fluid due to compaction and pressure dissolution was deter­ ②
Present average porosity of the Qiulitage 3.80% 1.65%
mined (Table 2). Although the Mg2+ concentration of the original
③ <5%
Fm.
Middle-Lower Ordovician limestone is unknown, it can be inferred that ④ Present formation water volume = ② × ③ 45.6 29.7 49.7
its Mg2+ concentration was lower than that of the current dolomite (m3)
strata. Therefore, the maximum amount of Mg2+ that could be provided ⑤ Thickness of the Qiulitage Fm. before 2000 3000 1657
about 40% of the original strata was
per unit area of the strata was only 2.86 × 105 mol based on Mg2+
pressure–dissolved = ①/(1 − 40%) (m)
concentration of the current strata as a reference which is an order of ⑥ Maximum historic porosity of the 40% 40% 40%
magnitude lower than the amount of Mg2+ required (Table 2). Qiulitage Fm. (%)
Similarly, the amount of Mg2+ that could be provided by the several ⑦ Formation water volume decreased by 724 1151 580
compaction = ⑤ × (40%–5%) (m3)
kilometers of underlying Upper Cambrian (∈3) dolomite can also be 3
⑧ Mg content of the Upper Cambrian formation water (mol/m ): 8.07–43.06,
calculated. Taking the Upper Cambrian Qiulitage group (∈3q) of Well 27.81 on average
Zhong4, Well TC1, and Well BT5 as examples, the calculated amount of ⑨ Minimum Mg content provided by the 6.21 × 9.53 × 5.08
minimum Mg2+ that could be provided by the Upper Cambrian strata Qiulitage Fm. water (mol) 103 103 × 103
due to physical compaction was only 5.08 × 103 mol per unit stratum ⑩ Rock loss of the Upper Cambrian 800 1200 663
dolostone due to pressure dissolution = ⑤
area (1 m2) based on the aforementioned method. However, a large
× 40%
amount of Mg2+ could have been released during pressure dissolution of ⑪ Mg provided by pressure dissolution of the 12.6 × 18.9 × 10.4
the underlying strata. This pattern was discovered by Hu et al. (2019) in Upper Cambrian dolostone = ⑩ × 106 106 × 106
their study of dolomite related to stylolite in the Cambrian Tarim Basin. 2900000/184
On average, we are estimated to range between 10.4 × 106–18.9 × 106 ⑫ Total Mg provided by the underlying 12.6 × 18.9 × 10.4
Upper Cambrian dolostone due to 106 106 × 106
mol per unit stratum area (1 m2) (Table 3). Although the fluids dis­ compaction and pressure dissolution = ⑨
charged due to pressure dissolution would precipitate as cements in the +⑪
Cambrian strata (Vandeginste and John, 2013), the Mg2+ consumed by ⑬ Mg consumed by cementation of the 3.78 × 5.67 × 3.13
this cementation constitutes only a small portion of the available Mg2+ Upper Cambrian, assuming 20% dolostone 106 106 × 106
cement = ② × 20% × 2900000/184

Table 1 even if the cement content was as high as 40% (40% is the pore volume
The Mg2+ content needed to form the amount of Ordovician dolostone per unit in the early stage of carbonate deposition, refer to Schomker and Halley,
(1 m2) area, Tarim Basin, using Well Zhong–4 as an example.
1982). Therefore, most of the fluids from the underlying Cambrian strata
Calculation Steps Dolostone Limy–dolostone Dolomitic would be discharged into Middle-Lower Ordovician strata, satisfying the
limestone amount of Mg2+ required for dolomitization of the target strata.
① Thickness (m) 79 112 181 In summary, the Mg2+ needed for the dolomitizing fluids of the
② Percentage of 90 60 45 Ordovician strata should be sufficiently supplied by the underlying
dolostone (%)
Cambrian dolomite through physical compaction and pressure
③ Dolostone volume (m3) 71.1 67.2 81.45
④ Density (g/m3) 2900000 2900000 2900000 dissolution.
⑤ Mass of dolostone (g) 206190000 194880000 236205000
⑥ Molar content of 11.2 × 105 10.6 × 105 12.83 × 105
dolostone (mol) 5.5. Thermal convection dolomitization model
⑦ Mg2+ needed (mol) 11.2 × 105 10.6 × 105 12.83 × 105

Total Mg2+ needed to dolomitize 1 m2 limestone: 3.46 × 106 mol. Three essential elements are required when developing a

14
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

dolomitization model (Warren, 2000): (1) a source of Mg and usually geochemical evidence (e.g., δ13C, δ18O, and 87Sr/86Sr; Section 5.2)
CO3; (2) a fluid flow mechanism; and (3) kinetic and thermodynamic suggest that the dolomitizing fluids derived from the underlying Upper
conditions favorable to dolomitization. Cambrian strata rather than from the Ordovician strata. Furthermore,
the Mg mass balance calculations (Section 5.3) suggest that the under­
5.5.1. Mg2+ sources lying Upper Cambrian dolomite strata provided most of the Mg2+
The Mg2+ source for dolomitization can be inferred first from the required for the dolomitization of the overlying Middle-Lower Ordovi­
distribution of the dolomite bodies. The upward migration of Mg- cian strata.
containing fluids is speculated as the dolomitization capacity weakens The fluid inclusion data show that the Middle-Lower Ordovician
upward, evidenced by the decreased amount of Middle-Lower Ordovi­ coarsely to medium-crystalline dolomites and cement dolomites were
cian dolomite from bottom to top (Fig. 7). Furthermore, the above formed by a relatively high formation temperature (100–160 ◦ C), which

Fig. 13. Photographs of the outcrop showing pathways for the migration of Mg–rich fluids. (A) grey limestone with red–grey bedding–parallel dolomite. (B) locally
enlarged photo of image (A), showing red ferrous dolomite and grey dolomite, both developed parallel to the bedding; C and D: a picture of the entire outcrop and its
sketch map, showing how the Mg–rich fluids migrated upward through the faults and fractures and horizontally through the bedding. Upper section of the Lower
Ordovician Penglaiba Fm., Kamatikan outcrop, NW Tarim Basin (location is shown in Fig. 2B).

15
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

is higher than the formation temperature of the Ca-rich fluids (i.e., developed along the stylolites (Fig. 4A–E).
forming late calcite cement), but significantly lower than that of crys­
talline quartz (Fig. 12). However, these Mg2+-rich high-temperature 5.5.3. Heat sources for thermal convection dolomitization
fluids are significantly different from the silicon-containing hydrother­ Thermal convection is strongly temperature dependent (Kaufman,
mal fluids in this area because only hydrothermal lithology and mineral 1994; Whitaker and Xiao, 2010; Saller and Dickson, 2011; Dong et al.,
association (i.e., Fe-dolomite and quartz) but no dolomite is observed in 2020). The large-scale upward migration of Mg2+-rich fluids from the
some areas (Fig. 13). It is speculated that the silicon-containing hydro­ Cambrian dolomite and the widespread dolomitization, i.e., several
thermal fluids provided large amounts of Si and Ca and low amounts of hundred meters to more than one thousand meters vertically, could not
Mg. Additionally, as described in section 5.2, only a few cement dolo­ be achieved by free geothermal convection in a normal geothermal
mites are affected by the geochemical signal of the hydrothermal fluid, gradient (Jones and Xiao, 2006).
and most dolomites lack the geochemical signal of the mantle-derived Magmatic and volcanic activity could significantly alter the
fluid, suggesting that this hydrothermal activity did affect the pre- geothermal flow and paleo-temperature. The Tarim Basin experienced
existing dolomite, but not directly cause the large-scale hydrothermal four magmatic and volcanic episodes during the period of Ediacaran-
dolomitization of the Middle-Lower Ordovician limestone (Jin et al., Cambrian, the Early Ordovician, the Late Carboniferous-Early
2006; Zhu et al., 2010; Dong et al., 2013). Therefore, the hydrothermal Permian, and the Cretaceous (Chen et al., 1997; Zhang et al., 2010; Yu
fluids appear to suggest a minor contribution to the formation of et al., 2011; Dong et al., 2013; Li et al., 2016). The geothermal flow
Middle-Lower Ordovician dolomite. increased significantly due to magmatic activity pre- and during each
volcanic episode, which simultaneously increased the geothermal
5.5.2. Mg2+-rich fluid flow mechanism gradient (Fig. 14). In other words, the deep strata were “heated” by each
For the dolomitization of the Middle-Lower Ordovician strata, large stage of magmatic activity.
volumes of Mg2+-rich fluids were released from the underlying Extensive magmatic and volcanic activity occurred during and
Cambrian strata due to physical compaction and pressure dissolution at immediately after the Early-Middle Ordovician deposition (Yu et al.,
a depth of approximately 500–2000 m. These high-temperature fluids 2011; Tang et al., 2012; Dong et al., 2013; Li et al., 2016; Qiu et al.,
were partially discharged laterally in the Cambrian strata, while the 2019). The heat from a high-temperature magma chamber and the
other portion migrated upward through channels (e.g., deep faults) related volcanic eruption altered the geothermal gradient. More
under external forces (the heat source is discussed below). This process importantly, the heat provided the driving force needed for the upward
led to large-scale dolomitization because of the relatively high temper­ flowing of the Mg-rich Cambrian fluids along the highly permeable
ature of the fluids, which help overcome the kinetic barrier for dolo­ Caledonian fractures. These heated fluids traverse through fluid path­
mitization (Warren, 2000). ways such as faults and bedding planes, infiltrating the porous and
In the interior of the Ordovician strata, fracture and crack systems permeable layers composed of Ordovician granular rocks. Afterwards,
could act as preferential migration pathways for the upward migration the dolomitizing fluids thermal convected between the relatively cold
of fluids (Cai et al., 2001; Corrêa et al., 2022). This is suggested by the surrounding permeable limestones and the heated fractures (Dong et al.,
dolomite fillings in many fractures in the Middle-Lower Ordovician 2020), and yielded a large volume of water-rock reaction. Moreover, this
strata (Fig. 6A and B; Yu et al., 2022). Li et al. (2016) documented that upward migration of Mg-rich fluid with high temperatures helped
the positions of craters in the study area were jointly controlled by the overcome the kinetic barriers of dolomite formation (Warren, 2000) and
intersection of NNE-trending faults and NW- trending faults formed in formed the main body of Ordovician dolomites. Throughout multiple
the Caledonian- Hercynian tectonic cycle. Accordingly, extensional episodes of volcanic activity during the Ordovician to Permian, this
segments of these NNE-trending strike-slip faults and the NW oriented - cyclical process persists, fostering the formation of extensive dolomite
echelon normal faults in the study area likely act as effective conduits for bodies.
the upward transport of deep-seated high-temperature fluids (Fig. 2C). During the Later Carboniferous and Early Permian, a large igneous
In Fig. 9, Well YB7, YB5, YB1-2 and GL1 near the Paleozoic strike-slip province (LIP) formed in the Tarim Basin (Chen et al., 1997; Zhang et al.,
fault (Fig. 2C) develop thick layers of dolomite body, while Z13 and 2010; Yu et al., 2011). Permian basalt covers an area of over 150,000
SN5 far from the fault system lack layered dolomite. In addition, the km2 in the basin, which indicates a large magma chamber at deep. This
dilatant microfractures that formed around fault tips during the initial intense tectonic and associated magmatic activity is believed to have
stages of strike-slip fault propagation (Baqués et al., 2010), as well as caused marbleization, silication, tectonically controlled hydrothermal
those induced by increasing displacements of sheared joints, may have alteration (dolomite recrystallization and dolomite precipitation) (Zhou
been suitable conduits for further transmission of deep-seated hot fluids et al., 2009; Dong et al., 2013; Guo et al., 2016), and abnormal matu­
to the porous and permeable Ordovician layer (Deng et al., 2019). ration of the Ordovician oil (Pu et al., 2013). However, these diagenetic
The Middle-Lower Ordovician strata were mainly deposited as events postdate the dolomitization.
granular limestones with high original porosity and permeability (Zhang
and Luo, 2010). At a burial depth of 500–1000 m, physical compaction, 5.5.4. Thermal convection dolomitization model
and cementation have not completely let the pores lost, so the porous A new thermal convection dolomitization model for the Lower-
and permeable layer can still be an important percolation channel for Middle Ordovician strata of the Tarim Basin is proposed based on the
diagenetic fluid (Croize et al., 2013; Lu et al., 2020). Moreover, the evidences discussed above (Fig. 15). First, the penecontemporaneous
Lower-Middle Ordovician strata are directly adjacent to the Upper Middle-Upper Cambrian dolomite was formed and contained large
Cambrian dolomite, which is favorable for the entrance of fluids from amounts of residual seawater, which the dolomite can be easily com­
the underlying strata to the target strata. pacted. During the sedimentary period of the Lower Ordovician Pen­
The bedding plane can also serve as a marker for lateral fluid glaiba Fm., the underlying Cambrian dolomite was physically and
migration, and this evidence is clearly visible in the outcrop (Dong et al., chemically compacted, resulting in the release of massive Mg2+. The
2017; Lu et al., 2023). A dolomitized ‘tongue’ was observed as red and newly formed Mg2+-rich fluids were driven by the geothermal heat
purple-red ferroan dolomite and grey non-ferroan dolomite along the provided by a deep magma chamber, moving upward through the
bedding planes (Fig. 13). This indicates that the Fe-containing Mg-rich Caledonian- Hercynian active faults, and the permeable strata to the
fluids migrated laterally along the bedding planes and extended into the Lower Ordovician strata. The Mg-rich fluids could easily dolomitize the
limestone interiors (Chen et al., 1997; Koeshidayatullah et al., 2020). In Ordovician limestones due to their high-temperature nature and trans­
addition, the stylolites played a role in the pathways for the lateral port mechanism through thermal convection. However, the limestones
migration of the Mg-rich fluids, which is indicated by dolomites of the Penglaiba Fm. remained un-dolomitized locally, which was

16
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

Fig. 14. The four stages of thermal evolution (Ediacaran–Cambrian: 673.1 ± 55.4–774 ± 0.18 Ma; Early Ordovician: 460.2 ± 2.2–484 ± 2.2 Ma; Late Carbon­
iferous–Early Permian: 264.3 ± 1.3–282.3 ± 1.4 Ma; and Cretaceous: 100.9 ± 3.4 Ma; Chen et al., 1997; Li et al., 2016) and their relationship to the thermal flow
values (Qiu et al., 2002) and the palaeotemperatures (Li et al., 2005; Lu et al., 2017) of the Tarim basin. Note an episode of regional tectonic uplift occurred in the
Devonian to Early Carboniferous and caused a significant increase in the thermal flow values and geothermal gradient. The pink shadows are magmatic events.

Fig. 15. Thermal convection dolomitization model of the Lower–Middle Ordovician, Tarim Basin. Stage–I: The Middle–Upper Cambrian semi–contemporaneous
dolomite precipitation; Stage–II: During deposition of the Lower Ordovician Penglaiba Fm., the underlying Cambrian dolomites were compacted. The released
Mg2+–rich fluids moved upward through the faults, fractures, and permeable strata driven by geothermal heat from a deep magma chamber. Controlled by fluid
conduits, the Penglaiba Fm. remained locally un–dolomitized. Stage–III: As dolomitization continued, the Mg2+ content decreased. The late–stage thermal convection
dolomitization weakened until the formation fluid was under–saturated in dolomite.

largely controlled by the location of the fluid conduits and the perme­ metasomatic dolomite formed has differentiated crystal grain structures
ability of the limestones (Koeshidayatullah et al., 2020; Lu et al., 2020). and different degrees of dolomitization (Koeshidayatullah et al., 2021;
Thermal convective dolomitization may have lasted for a long time Lu et al., 2023, Figs. 3–5). Finally, a minor amount of dolomites formed
during the shallow to medium burial stage. Due to the differentiation of because the Mg2+ contents decreased with continuing dolomitization.
diagenetic fluids and precursor structures during this process, the The late-stage thermal convection dolomitization became weakened

17
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

until the formation fluids were under-saturated with dolomite. Howev­ adjustment of the pore structure of the original porous reservoir. The
er, magmatic activity and the associated thermal convection continued. changes of the crystallization of dolomite can result in the alterations in
As a result, late-stage calcite cements were formed by the Ca-rich the crystal size and morphology of carbonate minerals (Cai et al., 2019).
diagenetic fluids (Dong et al., 2013, 2017; Guo et al., 2016; Du et al., Accordingly, this process could also reshape the pore structures between
2018). mineral crystals. The predominant transformation from intergranular to
This model explains how the widespread dolomites developed in the intercrystalline pores leads to a flatter pore space and reduced tortuos­
Lower-Middle Ordovician strata of the Tarim Basin. In addition, our ity, which promotes the permeability (Ahmadi et al., 2011). Simulta­
study revised the thermal convection model that highlights the contri­ neously, the dolomitization process can also generate relatively
bution of Mg-rich fluids formed by the compaction of the underlying homogenous size of dolomite crystals replacing the original heteroge­
strata, which may be driven by a heat source related to magmatic and neity limestone particles, thereby decreasing the pore-throat ratio (Cai
volcanic episodes (Machel, 2004). This model is of general significance et al., 2019). Therefore, the decreases in pore tortuosity and pore-throat
for studying dolomitization in areas where abnormal geothermal ac­ ratio induced by thermal convection dolomitization significantly facili­
tivities exist. tated the reservoir permeability (Fig. 16).
In conclusion, the prediction of the Lower to Middle Ordovician
thermal convection dolomite reservoirs in the Tarim Basin should focus
5.6. Effects of dolomitization on the reservoir on the following aspects: (1) Regions with greater thickness of under­
lying dolomite may provide more magnesium sources, enhancing the
Dolomitization plays a pivotal role in the preservation of early pores. scale of dolomite body; (2) Locations near basement fault systems,
In the study area, the precursors of Ordovician dolomites are grain shoal especially strike-slip and extensional faults (Davies and Smith, 2006; Xu
facies, characterized by abundant primary intergranular pores (Liu et al., 2022; Chen et al., 2023), are more likely to develop dolomite
et al., 2009; Zhang et al., 2010; Chen et al., 2015; Gao and Fan, 2015). reservoirs (Figs. 2C and 7); (3) Areas where primary pores preserved are
Afterwards, these pores are inherited and preserved due to the dolo­ mainly granular shoals developed on paleo-highlands.
mitization of precursor limestones. This is because the high permeability
of precursor limestones can be served as conduits for the flow of dolo­ 6. Conclusions
mitizing fluids and subsequent water-rock reaction (i.e., dolomitiza­
tion). On the other hand, dolomites tend to be more resistant to Based on petrological evidence, geochemical data, and Calculation of
compaction compared to the particle-supported limestones, which en­ Mg2+ flux, we investigated the widespread Middle-Lower Ordovician
sures the preservation of previous storage spaces (Moore and Wade, dolomites of the Tarim Basin. The dolomites are present as 1) individual
2013). dolomite crystals; 2) Pds; 3) layered-massive dolomite, and 4) cement
The intensification of thermal convection dolomitization and the dolomites in pores and fractures. The dolomite content generally
associated increase in dolomite crystallization degrees contribute to the

Fig. 16. The effect of dolomitization on pore throat ratio and pore tortuosity (modified from Chen et al., 2024). The pore-throat ratio is defined as the ratio of pore
diameter to throat diameter, while tortuosity is the ratio of the average effective flow path length to the straight-line length along the fluid flow direction.

18
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

decreases with decreasing burial depth. Most of the dolomites exhibit Ahmadi, M.M., Mohammadi, S., Hayati, A.N., 2011. Analytical derivation of tortuosity
and permeability of monosized spheres: a volume averaging approach. Phys. Rev. 83
evidence of being dolomitized by burial formation waters and being
(2), 026312 https://doi.org/10.1103/PhysRevE.83.026312.
partially influenced by hydrothermal fluids. The Mg-rich dolomitizing Allen, J.R., Wiggins, W.D., 1993. Dolomite reservoirs: geochemical techniques for
fluids are believed to be largely derived from the underlying Middle- Evaluating origin and distribution. J. Petrol. Sci. Eng. 14, 262–263.
Upper Cambrian dolomite. Banner, J.L., 2004. Radiogenic isotopes: systematics and applications to earth surface
processes and chemical stratigraphy. Earth Sci. Rev. 65, 141–194.
A new thermal convection dolomitization model was proposed. The Baqués, V., Travé, A., Benedicto, A., Labaume, P., Cantarero, I., 2010. Relationships
Mg2+-rich formation fluids were first extracted from the underlying between carbonate fault rocks and fluid flow regime during propagation of the
Cambrian dolomite by physical and chemical compaction. Then, the Neogene extensional faults of the Penedès basin (Catalan Coastal Ranges, NE Spain).
J. Geochem. Explor. 106, 24–33.
dolomitizing fluids were driven by heat that originated from a high- Bau, M., 1991. Rare–earth element mobility during hydrothermal and metamorphic
temperature magma chamber during the Early Ordovician and moved fluid–rock interaction and the significance of the oxidation state of europium. Chem.
upward through faults and fractures and/or permeable strata into the Geol. 93, 219–230.
Bissell, H.J., Chilingar, G.V., 1967. Classification of sedimentary carbonate rocks. Dev.
Ordovician strata. Finally, the overlying Ordovician limestones were Sedimentol. 9, 87–168.
easily, gradually dolomitized from bottom to top because of the thermal Bodnar, R.J., 1993. Revised equation and table for determining the freezing-point
convection circulation and high temperature of massive dolomitizing depression of H2O-NaCl solutions. Geochem. Cosmochim. Acta 57 (3), 683–684.
Budd, D.A., 1997. Cenozoic dolomites of carbonate islands: their attributes and origin.
fluids. Earth Sci. Rev. 42, 1–47.
This revised thermal convection dolomitization model explains how Burke, W.H., Denison, R.E., Hetherington, E.A., Koepnick, R.B., Neilson, H.F., Otto, J.B.,
these widespread dolomites developed in the Lower-Middle Ordovician 1982. Variation of seawater 87Sr/86Sr throughout Phanerozoic time. Geology 10,
516–519.
strata of the Tarim Basin. As a potentially untraditional dolomite for­
Cai, C., Franks, S.G., Aagaard, P., 2001. Origin and migration of brines from Paleozoic
mation model, thermal convection dolomitization may be of general strata in Cetral Tarim, China: constraints from 87Sr/86Sr, δD, δ18O and water
significance to areas affected by abnormal geothermal activities. In chemistry. Appl. Geochem. 16, 1269–1284.
addition, this model provides new insight into the formation of dolo­ Cai, J., Zhang, Z., Wei, W., Guo, D., Li, S., Zhao, P., 2019. The critical factors for
permeability-formation factor relation in reservoir rocks: pore-throat ratio,
mites conducive to hydrocarbon accumulation. tortuosity and connectivity. Energy 188, 116051. https://doi.org/10.1016/j.
energy.2019.116051.
Centrella, S., Hoareau, G., Beaudoin, N.E., Motte, G., Lanari, P., Piccoli, F., Callot, J.P.,
CRediT authorship contribution statement
Gomez-Rivas, E., Martín-Martín, J.D., 2023. Estimating the fluid composition after
dolomitization using mass balance equation: comparison of examples from Spain,
Haofu Zheng: Writing – original draft, Methodology, Investigation, Canada and France. Global Planet. Change 220, 104016.
Data curation. Feifan Lu: Writing – review & editing, Writing – original Chen, H.H., Lu, Z.Y., Cao, Z.C., Han, J., Yun, L., 2015. Hydrothermal alteration of
Ordovician reservoir in northeastern slope of Tazhong uplift, Tarim Basin. Acta
draft, Conceptualization. Yixin Dong: Formal analysis. Bo Liu: Project Petrol. Sin. 37, 43–63.
administration, Funding acquisition. Xuefeng Zhang: Writing – original Chen, H.L., Yang, S.F., Dong, C.W., Zhu, G.Q., Jia, C.Z., Wei, G.Q., Wang, Z.G., 1997.
draft, Visualization. Kaibo Shi: Project administration, Funding acqui­ Geological thermal events in Tarim Basin. Chin. Sci. Bull. 42, 580–584.
Chen, X., Xu, Q., Hao, F., Chen, Y., Yi, Y., Hu, F., Wang, X., Tian, J., Wang, G., 2023.
sition. Jiajun He: Visualization. Hairuo Qing: Validation, Supervision. Dolomite reservoir Formation and diagenesis evolution of the upper ediacaran
qigebrak formation in the Tabei area, Tarim Basin. Sci. China Earth Sci. 66 (10),
2311–2331.
Declaration of competing interest Chen, S., Liu, S., Liu, X., Wei, L., Liu, B., Wang, E., Yu, J., Xiong, Y., 2024. Diagenetic
recrystallization of carbonate and its significance for reservoir. Unpublished results.
The authors declare that they have no known competing financial Chi, G.X., Ni, P., 2007. Equations for calculation of NaCl/(NaCl+CaCl2) ratios and
salinities from hydrohalite-melting and ice-melting temperatures in the H2O-NaCl-
interests or personal relationships that could have appeared to influence CaCl2 system. Acta Petrol. Sin. 23 (1), 33–37.
the work reported in this paper. Compton, J.S., Siever, R., 1986. Diffusion and mass balance of Mg during early dolomite
formation, Monterey Formation. Geochem. Cosmochim. Acta 50 (1), 125–135.
Corrêa, R.S., Ukar, E., Laubach, S.E., Aubert, I., Lamarche, J., Wang, Q., Stockli, D.F.,
Data availability
Stockli, L.D., Larson, T.E., 2022. Episodic reactivation of carbonate fault zones with
implications for permeability–An example from Provence, Southeast France. Mar.
Data will be made available on request. Petrol. Geol. 145, 105905 https://doi.org/10.1016/j.marpetgeo.2022.105905.
Croize, D., Renard, F., Gratier, J.P., 2013. Compaction and porosity reduction in
carbonates: a review of observations, theory, and experiments. Adv. Geophys. 54,
Acknowledgments 181–238.
Davies, G.R., Smith Jr., L.B., 2006. Structurally controlled hydrothermal dolomite
reservoir facies: an overview. AAPG (Am. Assoc. Pet. Geol.) Bull. 90, 1641–1690.
Financially supported by National Natural Science Foundation of
Deng, S., Li, H., Zhang, Z., Zhang, J., Yang, X., 2019. Structural characterization of
China, China (No. U19B6003), Natural Science Foundation of Sichuan intracratonic strike- slip faults in the central Tarim Basin. AAPG (Am. Assoc. Pet.
Province, China (2023NSFSC0804), Chongqing Natural Science Foun­ Geol.) Bull. 103, 109–137.
dation, China (CSTB2023NSCQ-MSX1022) and Chongqing Municipal Di Primio, R., Leythaeuser, D., 1995. Quantification of the effect of carbonate
redistribution by pressure solution in organic–rich carbonates. Mar. Petrol. Geol. 12,
Education Commission Science and Technology Research Project, China 135–739.
(KJQN202100740). We are appreciative for the thorough and valuable Dong, S.F., Chen, D.Z., Qing, H.R., Zhou, X., Wang, D., Guo, Z., 2013. Hydrothermal
comments of Editor Dr. Marco Brandano and two anonymous reviewers, alteration of dolostones in the Lower Ordovician, Tarim Basin, NW China: multiple
constraints from petrology, isotope geochemistry and fluid inclusion
which greatly improved our manuscript. We thank Dr. Ruixin Guo of microthermometry. Mar. Petrol. Geol. 46, 270–286.
Sichuan University for his help with data processing. We are grateful to Dong, S., Chen, D., Zhou, X., Qian, Y., Tian, M., Qing, H., 2017. Tectonically driven
Dr. Qiqi Wang of the University of Texas at Austin, Dr. Weimin Jiang of dolomitization of Cambrian to Lower Ordovician carbonates of the Quruqtagh area,
north-eastern flank of Tarim Basin, north-west China. Sedimentology 64 (4),
China National Petroleum Corporation and Dr. Jinxin Yu of CNOOC for 1079–1106.
their review and assistance with the manuscript. Dong, Y., Chen, H., Wang, J., Hou, M., Xu, S., Zhu, P., Zhang, C., Cui, Y., 2020. Thermal
convection dolomitization induced by the emeishan large igneous province. Mar.
Petrol. Geol. 116, 104308 https://doi.org/10.1016/j.marpetgeo.2020.104308.
Appendix A. Supplementary data Du, Y., Fan, T., Machel, H.G., Gao, Z., 2018. Genesis of upper cambrian-lower ordovician
dolomites in the tahe oilfield, Tarim Basin, NW China: several limitations from
Supplementary data to this article can be found online at https://doi. petrology, geochemistry, and fluid inclusions. Mar. Petrol. Geol. 91, 43–70.
Duggan, J.P., Mountjoy, E.W., Stasiuk, L.D., 2001. Fault-controlled dolomitization at
org/10.1016/j.marpetgeo.2024.106740. Swan Hills Simonette oil field (Devonian), deep basin west-central Alberta, Canada.
Sedimentology 48 (2), 301–323.
References Dutton, S.P., Diggs, T.N., 1992. Evolution of porosity and permeability in the lower
cretaceous travis peak formation, east Texas. AAPG (Am. Assoc. Pet. Geol.) Bull. 76
(2), 252–269.
Aharon, P., Socki, R.A., Chan, L., 1987. Dolomitization of atolls by sea water convection
flow: test of a hypothesis at Niue, South Pacific. J. Geol. 95, 187–204.

19
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

Ebneth, S., Shields, G.A., Veizer, J., Miller, J.F., Shergold, J.H., 2001. High–resolution Koeshidayatullah, A., Corlett, H., Hollis, C., 2021. An overview of structurally-controlled
strontium isotope stratigraphy across the Cambrian–Ordovician transition. dolostone-limestone transitions in the stratigraphic record. Earth Sci. Rev. 220,
Geochem. Cosmochim. Acta 65, 2273–2292. 103751 https://doi.org/10.1016/j.earscirev.2021.103751.
Fabricius, I.L., Borre, M.K., 2010. Stylolites, porosity, depositional texture, and silicates Kohout, F.A., 1967. Ground water flow and the geothermal regime of the Floridan
in chalk facies sediments. Ontong Java Plateau – gorm and Tyra fields, North Sea. Plateau. Gulf Coast Association of Geological Societies Transactions 17, 339–354.
Sedimentology 54, 183–205. Land, L.S., 1980. The isotopic and trace element geochemistry of dolomite: the state of
Fanning, K.A., Byrne, R.H., Breland, J.A., Betzer, P.R., 1981. Geothermal springs of the the art. In: Zenger, D.H., Dunham, J.B., Ethington, R.L. (Eds.), Concepts and Models
west Florida continental shelf: evidence for dolomitization and radionuclide of Dolomitization. 28. SEPM Special Publication, pp. 87–110.
enrichment. Earth Planet Sci. Lett. 52, 345–354. Li, D.S., 1995. Hydrocarbon occurrences in the petroliferous basins of western China.
Finkel, E.A., Wilkinson, B.H., 1990. Stylolitization as source of cement in Mississippian Mar. Petrol. Geol. 12, 26–34.
Salem limestone, west–central Indiana. AAPG (Am. Assoc. Pet. Geol.) Bull. 74, Li, H., Qiu, N., Jin, Z., He, Z., 2005. Geothermal history of Tarim basin. Oil Gas Geol. 26,
174–186. 613–617 (in Chinese with English abstract).
Fu, Q., Hu, S., Xu, Z., Zhao, W., Shi, S., Zeng, H., 2020. Depositional and diagenetic Li, J., Li, Z., Qiu, N., Zuo, Y., Yu, J., Liu, J., 2016. Carboniferous–Permian abnormal
controls on deeply buried Cambrian carbonate reservoirs: longwangmiao Formation thermal evolution of the Tarim basin and its implication for deep structure and
in the Moxi–Gaoshiti area, Sichuan Basin, southwestern China. Mar. Petrol. Geol. magmatic activity. Chin. J. Geophys. 59, 3318–3329 (in Chinese with English
117, 104318 https://doi.org/10.1016/j.marpetgeo.2020.104318. abstract).
Gao, Z.Q., Fan, T.L., 2015. Carbonate platform-margin architecture and its influence on Lin, C.S., Yang, H.J., Liu, J.Y., Rui, Z.F., Cai, Z.Z., Zhu, Y.F., 2012. Distribution and
Cambrian-Ordovician reef-shoal development, Tarim Basin, NW China. Mar. Petrol. erosion of the Paleozoic tectonic unconformities in the Tarim Basin, Northwest
Geol. 68, 291–306. China: significance for the evolution of paleo–uplifts and tectonic geography during
Gomez-Rivas, E., Corbella, M., Martín-Martín, J.D., Stafford, S.L., Teixell, A., Bons, P.D., deformation. J. Asian Earth Sci. 46, 1–19.
Griera, A., Cardellach, E., 2014. Reactivity of dolomitizing fluids and Mg source Liu, L.F., Wang, P., Li, Y., Chen, Z.J., Zhao, Y.D., Wang, W.B., Wang, W.L., Wu, L., 2009.
evaluation of fault-controlled dolomitization at the Benicàssim outcrop analogue Paleozoic reservoir beds and their favorableness in tazhong areas of Tarim Basin,
(Maestrat Basin, E Spain). Mar. Petrol. Geol. 55, 26–42. northwest China. J. Petrol. Sci. Eng. 68, 1–18.
Gomez–Rivas, E., Martín–Martín, J.D., Bons, P.D., Koehn, D., Griera, A., Travé, A., Liu, Y.N., Ngia, N.R., Hu, M.Y., Cai, Q.S., 2022. Evaluation of the properties of
Llorens, M., Humphrey, E., Neilson, J., 2022. Stylolites and stylolite networks as dolomitization fluids and diagenetic alterations of mg/Ca ratios in carbonate rocks in
primary controls on the geometry and distribution of carbonate diagenetic the cambrian series-2 to miaolingian strata in Central Uplift Belt, Tarim Basin:
alterations. Mar. Petrol. Geol. 136, 105444. constraints from halogens, REEs and isotope geochemistry. Mar. Petrol. Geol. 144,
Grant, J.A., 1986. The isocon diagram; a simple solution to Gresens’ equation for 105838 https://doi.org/10.1016/j.marpetgeo.2022.105838.
metasomatic alteration. Econ. Geol. 81 (8), 1976–1982. Lu, F., Tan, X., Zhong, Y., Luo, B., Zhang, B., Zhang, Y., Li, M., Xiao, D., Wang, X.,
Gregg, J.M., Bish, D.L., Kaczmarek, S.E., Machel, H.G., 2015. Mineralogy, nucleation and Zeng, W., 2020. Origin of the penecontemporaneous sucrosic dolomite in the
growth of dolomite in the laboratory and sedimentary environment: a review. permian qixia formation, northwestern sichuan basin, SW China. Petrol. Explor. Dev.
Sedimentology 62 (6), 1749–1769. 47 (6), 1218–1234.
Gresens, R.L., 1967. Composition-volume relationships of metasomatism. Chem. Geol. 2, Lu, F., Tan, X., Xiao, D., Shi, K., Li, M., Zhang, Y., Zheng, H., Dong, Y., 2023. Sedimentary
47–65. control on diagenetic paths of dolomite reservoirs in volcanic setting: a case study of
Guo, C., Chen, D., Qing, H., Dong, S., Li, G., Wang, D., Qian, Y., Liu, C., 2016. Multiple Permian Chihsia Formation in Sichuan Basin, China. Sediment. Geol., 106451
dolomitization and later hydrothermal alteration on the Upper Cambrian-Lower https://doi.org/10.1016/j.sedgeo.2023.106451.
Ordovician carbonates in the northern Tarim Basin, China. Mar. Petrol. Geol. 72, Lu, Z.Y., Chen, H.H., Chen, H.R., Chi, G.X., Chen, Q.G., You, D.H., Yin, H., Zhang, S.Y.,
295–316. 2017. Petrography, fluid inclusion and isotope studies in Ordovician carbonate
Guo, R., Zhang, S., Bai, X., Wang, K., Sun, X., Liu, X., 2020. Hydrothermal dolomite reservoirs in the Shunnan area, Tarim basin, NW China: implications for the nature
reservoirs in a fault system and the factors controlling reservoir formation-A case and timing of silicification. Sediment. Geol. 359, 29–43.
study of Lower Paleozoic carbonate reservoirs in the Gucheng area, Tarim Basin. Machel, H.G., 2004. Concepts and models of dolomitization: a critical reappraisal.
Mar. Petrol. Geol. 120, 104506 https://doi.org/10.1016/j.marpetgeo.2020.104506. Geological Society, London, Special Publications 235 (1), 7–63.
He, J., Ding, W., Huang, W., Cao, Z., Chen, E., Dai, P., Zhang, Y., 2019. Petrological, MacLean, W.H., 1990. Mass change calculations in altered rock series. Miner. Deposita
geochemical, and hydrothermal characteristics of Ordovician cherts in the 25, 44–49.
southeastern Tarim Basin, NW China, and constraints on the origin of cherts and McLennan, S.M., 1989. Rare earth elements in sedimentary rocks: influence of
Permian tectonic evolution. J. Asian Earth Sci. 170, 294–315. provenance and sedimentary processes. In: Lipin, B.R., McKay, G.A. (Eds.),
Hein, J.R., Gray, S.C., Richmond, B.M., White, L.D., 1992. Dolomitization of quaternary Geochemistry and Mineralogy of Rare Earth Elements. De Gruyter, Berlin,
reef limestone, aitutake, Cook Islands. Sedimentology 39, 645–661. pp. 169–200.
Heydari, E., 2000. Porosity loss, fluid flow, and mass transfer in limestone reservoirs: Mao, C., Zhong, J.H., Li, Y., Wang, Y.Z., Niu, Y.B., Ni, L.T., Shao, Z.F., 2014. Ordovician
application to the Upper Jurassic Smackover Formation, Mississippi. AAPG (Am. carbonate rock matrix fractured–porous reservoirs in Tahe Oilfield, Tarim Basin, NW
Assoc. Pet. Geol.) Bull. 84, 100–118. China. Petrol. Explor. Dev. 41, 745–753.
Hu, M., Ngia, N.R., Gao, D., 2019. Dolomitization and hydrotectonic model of burial Montañez, I.P., Read, J.F., 1992. Fluid-rock interaction history during stabilization of
dolomitization of the Furongian-Lower Ordovician carbonates in the Tazhong Uplift, early dolomites, upper Knox Group (Lower Ordovician), US Appalachians.
central Tarim Basin, NW China: implications from petrography and geochemistry. J. Sediment. Res. 62, 753–778.
Mar. Petrol. Geol. 106, 88–115. Moore, C.H., Wade, W.J., 2013. Carbonate Reservoirs: Porosity and Diagenesis in a
Jiang, L., Cai, C., Worden, R.H., Crowley, S.F., Jia, L., Zhang, K., Duncan, I.J., 2016. Sequence Stratigraphic Framework. Newnes.
Multiphase dolomitization of deeply buried Cambrian petroleum reservoirs, Tarim Mossop, G.D., 1972. Origin of the peripheral rim, redwater reef, alberta. Bull. Can.
Basin, north-west China. Sedimentology 63 (7), 2130–2157. Bulletin of Canadian Petroleum Geology 20, 238–280.
Jiang, W.M., 2022. Origin of carbonate reservoir in the lower cambrian xiaoerbulake Nicolaides, S., Wallace, M.W., 1997. Pressure–dissolution and cementation in an
formation. Bachu-Tazhong area, Tarim Basin, NW China (Ph.D. thesis). Peking oligo–miocene non–tropical limestone (clifton formation), otway basin, Australia. In:
University 168. China. James, N.P., Clarke, J.A.D. (Eds.), Cool–Water Carbonates, vol. 56. SEPM Special
Jin, Z., Zhu, D., Hu, W., Zhang, X., Wang, Y., Yan, X., 2006. Geological and geochemical Publication, pp. 249–261.
signatures of hydrothermal activity and their influence on carbonate reservoir beds Nicholas, C.J., 1996. The Sr isotopic evolution of the oceans during the ’Cambrian
in the Tarim Basin. Acta Geol. Sin. 80, 245–253 (in Chinese with English abstract). Explosion’. J. Geol. Soc. (Lond.) 153, 243–254.
Jiu, B., Huang, W., Mu, N., Hao, R., 2022. Petrology, mineralogy and geochemistry of Nothdurft, L.D., Webb, G.E., Kamber, B.S., 2004. Rare earth element geochemistry of
Ordovician rocks in the southwest of Tarim Basin, implications for genetic Late Devonian reefal carbonates, Canning Basin, Western Australia: confirmation of
mechanism and evolution model of the hydrothermal reformed-paleokarst carbonate a seawater REE proxy in ancient limestones. Geochem. Cosmochim. Acta 68 (2),
reservoir. Mar. Petrol. Geol. 140, 105687 https://doi.org/10.1016/j. 263–283.
marpetgeo.2022.105687. Paganoni, M., Harthi, A.A., Morad, D., Morad, S., Ceriani, A., Mansurbeg, H., Suwaidi, A.
Jones, G.D., Smart, P.L., Whitaker, F.F., Rostron, B.J., Machel, H.G., 2003. Numerical A., Al–Aasm, I.S., Ehrenberg, S.N., Sirat, M., 2016. Impact of stylolitization on
modeling of reflux dolomitization in the Grosmont platform complex (Upper diagenesis of a lower cretaceous carbonate reservoir from a giant oilfield, Abu Dhabi,
Devonian), Western Canada sedimentary basin. AAPG (Am. Assoc. Pet. Geol.) Bull. United Arab Emirates. Sediment. Geol. 335, 70–92.
87, 1273–1298. Pu, R., Zhong, H., Zhang, Y., 2013. Preliminary study on the effects of Permian volcanism
Jones, G.D., Xiao, Y., 2006. Geothermal convection in the Tengiz carbonate platform, on the Tahe Ordovician oil pools in Tarim basin. Mar. Petrol. Geol. 44, 13–20.
Kazakhstan: reactive transport models of diagenesis and reservoir quality. AAPG Qiao, Z.F., Zhang, S.N., Shen, A.J., Shao, G.M., She, M., Cao, P., Sun, X.W., Zhang, J.,
(Am. Assoc. Pet. Geol.) Bull. 90, 1251–1272. Tan, X.C., 2021. Features and origins of massive dolomite of lower ordovician
Kang, Y.Z., Kang, Z.H., 1996. Tectonic evolution and oil and gas of Tarim Basin. J. Asian Penglaiba Formation in the northwest Tarim Basin: evidence from petrography and
Earth Sci. 13, 317–325. geochemistry. Petrol. Sci. 18 (5), 1323–1341.
Kaufman, J., 1994. Numerical models of fluid flow in carbonate platforms: implications Qing, H., Qiao, Z., Zhang, S., Cosford, J., Hu, A., Liang, F., Wang, Y., Zheng, J., 2023.
for dolomitization. J. Sediment. Res. A64, 128–139. δ26Mg-δ13C-δ18O systems as geochemical tracers for dolomite recrystallization: a
Koeshidayatullah, A., Corlett, H., Stacey, J., Swart, P.K., Boyce, A., Hollis, C., 2020. case study of lower Ordovician dolomite from Tarim Basin. Chem. Geol. 619, 121302
Origin and evolution of fault–controlled hydrothermal dolomitization fronts: a new https://doi.org/10.1016/j.chemgeo.2023.121302.
insight. Earth Planet Sci. Lett. 541, 116291 https://doi.org/10.1016/j. Qiu, H., Deng, S., Cao, Z., Yin, T., Zhang, Z., 2019. The evolution of the complex
epsl.2020.116291. anticlinal belt with crosscutting strike-slip faults in the central Tarim basin, NW
China. Tectonics 38, 2087–2113.

20
H. Zheng et al. Marine and Petroleum Geology 163 (2024) 106740

Qiu, N., Jin, Z., Li, J., 2002. Discussion on thermal wave model used in the thermal platform and bank carbonates, Wild River area, west–central Alberta. Bull. Can.
evolution analysis in the Tarim Basin. Chin. J. Geophys. 45 (3), 411–419 (in Chinese Petrol. Geol. 46 (2), 210–265.
with English abstract). Winter, B.L., Clark, D.L., Johnson, C.M., 1997. Late Cenozoic Sr isotope evolution of the
Ramos, J.R.D.A., 2000. Stylolites: measurement of rock loss. Rev. Bras. Geociencias 30, Arctic Ocean: constraints on water mass exchange with the lower latitude oceans.
432–435. Deep Sea Res. Part II Top. Stud. Oceanogr. 44 (8), 1531–1542.
Safaricz, M., Davison, I., 2005. Pressure solution in chalk. AAPG (Am. Assoc. Pet. Geol.) Whitaker, F.F., Smart, P.L., Jones, G.D., 2004. Dolomitization: from conceptual to
Bull. 89, 383–401. numerical models. In: Braithwaite, C.J.R., Rizzi, G., Darke, G. (Eds.), The Geometry
Saller, A.H., 1984. Petrologic and geochemical constraints on the origin of subsurface and Petrogenesis of Dolomite Hydrocarbon Reservoirs, vol. 235. Geological Society,
dolomite, Enewetak Atoll: an example of dolomitization by normal seawater. London, Special Publications, pp. 99–139.
Geology 12, 217–220. Whitaker, F.F., Xiao, Y., 2010. Reactive transport modeling of early burial dolomitization
Saller, A.H., Dickson, J.A.T.D., 2011. Partial dolomitization of a Pennsylvanian limestone of carbonate platforms by geothermal convection. AAPG (Am. Assoc. Pet. Geol.)
buildup by hydrothermal fluids and its effect on reservoir quality and performance. Bull. 94, 889–917.
AAPG (Am. Assoc. Pet. Geol.) Bull. 95, 1745–1762. Wu, G., Zhao, K., Qu, H., Scarselli, N., Zhang, Y., Han, J., Xu, Y., 2020. Permeability
Sanford, W.E., Whitaker, F.F., Smart, P.L., Jones, G., 1999. Numerical analysis of distribution and scaling in multi-stages carbonate damage zones: insight from strike-
seawater circulation in carbonate platforms. I. Geothermal Convection. Am. J. Sci. slip fault zones in the Tarim Basin, NW China. Mar. Petrol. Geol. 114, 104208
298, 801–828. https://doi.org/10.1016/j.marpetgeo.2019.104208.
Shen, Z., Chen, Y., Guo, J., 1995. Discussion on genetic mechanism and model for Xu, Q.L., Cheng, Y.Z., You, D.H., Zhang, Y., Li, X.L., Xu, J., Chen, X.D., 2022.
dolomitization, lower paleozoic, Tarim Basin. Xinjing Pet. Geol. 16 (4), 368 (in Characteristics and Formation Mechanism of Multiscale Storage Spaces in Ancient
Chinese with English abstract). Deep Tight Reservoirs: Examples from the Cambrian Yurtus Formation in the
Simms, M., 1984. Dolomitization by groundwater flow systems in carbonate platforms. Northern Tarim Basin, 8449840. https://doi.org/10.2113/2022/8449840. China.
Gulf Coast Association of Geological Societies Transactions 34, 411–420. Lithosphere.
Steele-MacInnis, M., Bodnar, R.J., Naden, J., 2011. Numerical model to determine the Yang, W., Wang, Q., 2000. Dolomite origin of lower ordovician in hetian river gas field,
composition of H2O–NaCl–CaCl2 fluid inclusions based on microthermometric and Tarim Basin. Acta Sedimentol. Sin. 18 (4), 545–548 (in Chinese with English
microanalytical data. Geochem. Cosmochim. Acta 75 (1), 21–40. abstract).
Stockdale, P.B., 1922. Stylolites: Their Nature and Origin, vol. 9. Indiana University Yu, J.X., Shi, K.B., Wang, Q.Q., Liu, B., Han, J., Song, Y.C., Kong, Y., Jiang, W.M., 2022.
Studies, pp. 1–97. Structural diagenesis of deep carbonate rocks controlled by intra-cratonic strike-slip
Swart, P.K., 2015. The geochemistry of carbonate diagenesis: the past, present and faulting: an example in the Shunbei area of the Tarim Basin, NW China. Basin Res.
future. Sedimentology 62, 1233–1304. 34, 1601–1631.
Tang, L.J., Qi, L.X., Qiu, H.J., Yun, L., Li, M., Xie, D.Q., Yang, Y., Wang, G.M., 2012. Poly- Yu, X., Yang, S., Chen, H., Chen, Z., Li, Z., Batt, G.E., Li, Y., 2011. Permian flood basalts
phase differential fault movement and hydrocarbon accumulation of the Tarim from the Tarim Basin, Northwest China: SHRIMP zircon U–Pb dating and
Basin, NW China. Acta Petrol. Sin. 28 (8), 2569–2583. geochemical characteristics. Gondwana Res. 20, 485–497.
Taylor, S.R., McLennan, S.M., 1981. The composition and evolution of the continental Zhang, C., Xu, Y., Li, Z., Wang, H., Ye, H., 2010. Diverse permian magmatism in the tarim
crust: rare earth element evidence from sedimentary rocks. Phil. Trans. Roy. Soc. block, NW China: genetically linked to the permian tarim mantle plume? Lithos 119,
Lond. Math. Phys. Sci. 301 (1461), 381–399. 537–552.
Toussaint, R., Aharonov, E., Koehn, D., Gratier, J.P., Ebner, M., Baud, P., Rolland, A., Zhang, F.Q., Dilek, Y., Cheng, X.G., Wu, H.X., Liu, X.B., Cheng, H.L., 2019. Late
Renard, F., 2018. Stylolites: a review. J. Struct. Geol. 114, 163–195. Neoproterozoic–early Paleozoic seismic structure–stratigraphy of the SW Tarim
Tuo, J., Philp, R.P., 2003. Occurrence and distribution of high molecular weight Block (China), its passive margin evolution and the Tarim–Rodinia breakup.
hydrocarbons in selected non–marine source rocks from the Liaohe, Qaidam and Precambrian Res. 334, 105456 https://doi.org/10.1016/j.precamres.2019.105456.
Tarim Basins, China. Org. Geochem. 34, 1543–1558. Zhang, J., Hu, W., Qian, Y., Wang, X., Cao, J., Zhu, J., Li, Q., Xie, X., 2009. Formation of
Vandeginste, V., John, C.M., 2013. Diagenetic implications of stylolitization in pelagic saddle dolomites in Upper Cambrian carbonates, western Tarim Basin (northwest
carbonates, canterbury basin, offshore New Zealand. J. Sediment. Res. 83, 226–240. China): implications for fault–related fluid flow. Mar. Petrol. Geol. 26, 1428–1440.
Veizer, J., Ala, D., Azmy, K., Bruckschen, P., Buhl, D., Bruhn, F., Carden, G.A.F., Zhang, J., Luo, P., 2010. Genesis of Ordovician matrix–porosity dolomite reservoir in the
Diener, A., Ebneth, S., Godderis, Y., Jasper, T., Korte, C., Pawellek, F., Podlaha, O.G., Tarim Basin. Petroleum Geology & Experiment 32, 470–474+503 (in Chinese with
Strauss, H., 1999. 87Sr/86Sr, δ13C and δ18O evolution of Phanerozoic seawater. English abstract).
Chem. Geol. 161, 59–88. Zhang, S., Liang, D., Li, M., Xiao, Z., He, Z., 2002. Molecule fossil and oilsource rock
Wang, K., Hu, S.Y., Hu, Z.Y., Liu, W., Huang, Q.Y., Shi, S.Y., Ma, K., Li, M., 2016. correlation of Tarim Basin. Chin. Sci. Bull. 47, 16–23.
Cambrian hydrothermal action in Gucheng area, Tarim Basin and its influences on Zhang, X., Hu, W., Jin, Z., Zhang, J., Qian, Y., Zhu, J., Zhu, D., Wang, X., Xie, X., 2008.
reservoir development. Acta Petrol. Sin. 37, 439–453. REE compositions of lower ordovician dolomites in central and north Tarim Basin,
Wang, Q., Liu, J., Lin, C., Li, H., 2022. Depositional evolution, sequence stratigraphic NW China: a potential REE proxy for ancient seawater. Acta Geol. Sin. 82, 610–621
framework and its response to relative sea-level change, the Middle and Lower (in Chinese with English abstract).
Ordovician carbonate system on outcrops, north-western margin of Tarim Basin. Zhou, B., Li, M., Duan, S., Zhu, C., Wu, G., Zeng, C., Gao, L., 2012. Dolomitization
Mar. Petrol. Geol. 143, 105909 https://doi.org/10.1016/j.marpetgeo.2022.105909. mechanism of cambrian carbonates in the Bachu area, Tarim Basin, NW China.
Wang, Z.M., Yang, H.J., Qi, Y.M., Chen, Y.Q., Xu, Y.L., 2014. Ordovician gas exploration Petrol. Explor. Dev. 39, 212–217.
breakthrough in the Gucheng lower uplift of the Tarim Basin and its enlightenment. Zhou, M.F., Zhao, J.H., Jiang, C.Y., Gao, J.F., Wang, W., Yang, S.H., 2009. OIB-like,
Nat. Gas. Ind. B 1 (1), 32–40. heterogeneous mantle sources of Permian basaltic magmatism in the western Tarim
Warren, J., 2000. Dolomite: occurrence, evolution and economically important Basin, NW China: implications for a possible Permian large igneous province. Lithos
associations. Earth Sci. Rev. 52, 1–81. 113 (3–4), 583–594.
Wei, D., Gao, Z., Fan, T., Niu, Y., Guo, R., 2021. Volcanic events-related hydrothermal Zhu, D., Jin, Z., Hu, W., 2010. Hydrothermal recrystallization of the Lower Ordovician
dolomitisation and silicification controlled by intra-cratonic strike-slip fault systems: dolomite and its significance to reservoir in northern Tarim Basin. Sci. China Earth
insights from the northern slope of the tazhong uplift, tarim basin, China. Basin Res. Sci. 53, 368–381.
33 (4), 2411–2434. Zhu, D., Meng, Q., Jin, Z., Liu, Q., Hu, W., 2015. Formation mechanism of deep Cambrian
Wendte, J., Qing, H., Dravis, J.J., Moore, S.L., Stasiuk, L.D., Ward, G., 1998. dolomite reservoirs in the Tarim basin, northwestern China. Mar. Petrol. Geol. 59,
High–temperature saline (thermoflux) dolomitization of Devonian Swan Hills 232–244.

21

You might also like