Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/356575663

Recrystallization of dolostones in the Cambrian Xiaoerbrak Formation, Tarim


Basin and possible link to reservoir development

Article  in  Marine and Petroleum Geology · November 2021


DOI: 10.1016/j.marpetgeo.2021.105452

CITATIONS READS

0 51

7 authors, including:

Yinggang Zhang Tao Yang


Nanjing University Nanjing University
4 PUBLICATIONS   11 CITATIONS    53 PUBLICATIONS   981 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

The development of in-situ sulfur isotope analysis technique applied to sulfides View project

All content following this page was uploaded by Yinggang Zhang on 27 November 2021.

The user has requested enhancement of the downloaded file.


1 Recrystallization of dolostones in the Cambrian Xiaoerbrak Formation, Tarim
2 Basin and possible link to reservoir development
3
4 Yinggang Zhang1,4, Wenqing Pan2, Bi Zhu 3, Weiqiang Li1, Liu Willow Yang1,
5 Yongquan Chen2, Tao Yang1*
6
1
7 State Key Laboratory for Mineral Deposits Research, School of Earth Sciences

8 and Engineering, Nanjing University, Nanjing, PR China;

2
9 Research Institute of Exploration and Development, Tarim Oil Field Company,

10 PetroChina, Korla, PR China;

3
11 School of Earth Sciences and Engineering, Hohai University, Nanjing, PR

12 China;

4
13 School of Earth and Environment, University of Leeds, Leeds, UK.

14

15 *Corresponding author: Tao Yang

16 Email: yangtao@nju.edu.cn

17 Postal address: Nanjing University Xianlin Campus, No.163 Xianlin Avenue,

18 Qixia District, Nanjing 210046, PR China.

19

20

21 Revised manuscript submission to Marine and Petroleum Geology

22 November 2021

23
24 Abstract
25 The Cambrian Xiaoerbrak Formation in the Tarim Basin, NW China, is a promising oil
26 and gas reservoir target. Many factors that affect reservoir development, such as
27 sedimentation, paleomorphology, dolomitization, and meteoric water corrosion, have
28 been investigated in previous studies. However, no consensus about the decisive factor
29 limits further hydrocarbon exploration in the Cambrian dolostones. In this study, we
30 present the stratigraphic variation in dolomite crystallization parameters (e.g.,
31 I(015/110), d(104)) and element contents (carbonate-associated-sulfate (CAS),
32 aluminum, manganese, strontium (Sr), and rare earth elements), across the entire
33 Xiaoerbrak Formation (Xiaoerbrak section) in the Keping region, NW Tarim Basin.
34 Along with petrographic evidence, the consistent variations in dolomite crystallization
35 parameters, CAS and Sr concentrations suggest that the dolostones in the Xiaoerbrak
36 Formation experienced progressive recrystallizations. Typical seawater REE patterns
37 and relative heavier oxygen isotopic compositions than contemporaneous carbonates,
38 likely suggest that the primary altering fluid related to this recrystallization event
39 originated from lagoon seawater. In addition, dolomite crystallographic parameters and
40 element contents generally follow the variation trends of porosity and permeability.
41 According to petrographic observations and variations in element concentrations and
42 crystallographic parameters, we think the reservoir quality of the Xiaoerbrak Formation
43 was strongly affected by the variation in the original rock petrophysics that were mainly
44 controlled by microbial structures and depositional facies. Furthermore, we argue that
45 the pre-existing pore network in carbonates could provide pathways for altering fluid
46 and then increase the possibility of fluid alteration, which brings a speculation that
47 primitive carbonates with higher porosity are more likely to be altered.
48 Keywords: Xiaoerbrak Formation; Fluid alteration; Recrystallization; Reservoir
49 development; Tarim Basin.
50
51 1. Introduction
52
53 The Tarim Basin is the largest oil-and-gas-bearing basin in China. The Cambrian
54 Xiaoerbrak Formation, widespread in the Tarim Basin, is a promising oil and gas
55 reservoir (Hu et al., 2014; Huang et al., 2016; Liu et al., 2020; Shen et al., 2016; Song
56 et al., 2014; Z. Wang et al., 2014; Yan et al., 2017; Zheng et al., 2020b, 2020a). Since
57 the discovery of oil and gas in the Cambrian Xiaoerbrak dolostones (well Zhongshen-
58 1) in the Tarim Basin (Z. Wang et al., 2014), the factors that controlled reservoir
59 development in the Xiaoerbrak Formation have been debated vigorously for further
60 exploration (Hu et al., 2014; Huang et al., 2016; Li et al., 2017, 2020; Liu et al., 2020;
61 Shen et al., 2016; Song et al., 2014; Z. Wang et al., 2014; Yan et al., 2017; Zheng et al.,
62 2020a, 2020b). Some researchers believe that the Xiaoerbrak Formation mainly
63 originated from microbialites, and reservoir development was mainly controlled by
64 sedimentary paleomorphology and microbial structures (Huang et al., 2016; Li et al.,
65 2020; Song et al., 2014). Others suggest that the main controls of reservoir development
66 were depositional facies and meteoric water corrosion (Li et al., 2011; Liu et al., 2020;
67 Shen et al., 2016; Yan et al., 2017). In addition, multi-stage dolomitization under
68 hydrothermal influence was proposed as an important factor that influenced reservoir
69 characteristics (Li et al., 2017). The lack of consensus on the factors that controlled
70 reservoir development limits the prospects of oil and gas exploration in the Xiaoerbrak
71 Formation, which outlines the significance of understanding diagenetic processes that
72 affect reservoir quality.
73 Diagenetic imprints, along with their type and intensity, play a significant role in
74 shaping the reservoir quality of carbonates (Abdolmaleki et al., 2016; Amel et al., 2015;
75 Gregg et al., 1992; Tavakoli et al., 2011). Different diagenetic processes have different
76 degrees of impact on the reservoir petrophysical properties, such as porosity and
77 permeability (Abdolmaleki et al., 2016; Amel et al., 2015; Tavakoli et al., 2011).
78 Besides basic petrography studies, researches of element concentrations (e.g. calcium
79 (Ca), magnesium (Mg), manganese (Mn), iron (Fe), and rare earth elements (REEs)),
80 isotopes (carbon (C), oxygen (O), and strontium (Sr)) and crystallographic parameters
81 of dolomite were carried out to recognize the diagenesis history of the Xiaoerbrak
82 Formation (e.g., Liu et al., 2020; Shen et al., 2016). On the one hand, these studies
83 raised many possible factors controlling the reservoir quality of the Xiaoerbrak
84 Formation. On the other hand, limited approaches adopted on distinct samples in the
85 Xiaoerbrak Formation would instead exacerbate the controlling-factor controversy to
86 some extent.
87 Dolomitization of limestone and recrystallization of dolomite are common diagenetic
88 processes in carbonates (e.g., Gregg et al., 1992), playing a significant role for
89 carbonate to develop into a good reservoir (Wierzbicki et al., 2006; Zhu et al., 2010).
90 Although dolomitization and recrystallization have been widely reported in the
91 carbonates of the Xiaoerbrak Formation (e.g., Hu et al., 2014; Jiang et al., 2019; Li et
92 al., 2020; Zheng et al., 2020a), the specific process behind the dolomitization and
93 dolomite recrystallization in the Xiaoerbrak Formation has not been resolved.
94 The crystallographic parameters of dolomite (e.g., cation ordering, mineral crystallinity,
95 and d(104) values) are useful indicators of dolomite recrystallization (Bialik et al., 2018;
96 Demény et al., 2016; Li et al., 2019, 2015). For example, in dolomite X-ray diffraction
97 (XRD) data, a high dolomite peak (015)-to-peak (110) intensity ratio indicates a higher
98 degree of Ca–Mg cation ordering (Goldsmith and Graf, 1958), while the full-width-at-
99 half-maximum (FWHM) of dolomite peak (104) decreases as mineral crystallinity
100 increases (Demény et al., 2016). Recently, the dolomite cation ordering has been
101 proposed to be markedly influenced by the carbonate associated sulphate (CAS)
102 concentration (Baldermann et al., 2015) which is vulnerable to diagenetic influence
103 (Fichtner et al., 2017; Gill et al., 2008). Besides the CAS concentration, other
104 conventional geochemical proxies, including Mn and strontium (Sr) concentration, the
105 REE pattern, and O isotopes of carbonates, are also indicative for diagenetic processes
106 and can indicate possible sources of altering fluid (e.g., Elderfield et al., 1990; Hohl et
107 al., 2015; Ling et al., 2013; Tostevin et al., 2016).
108 In this study, we combine previously published O isotope values from the Xiaoerbrak
109 Formation carbonates (Zhang et al., 2020) with new crystallographic parameter (XRD
110 measurements) and element content (including Mn, Sr, SCAS, Al, Fe, and REEs) data
111 across the Xiaoerbrak Formation (Xiaoerbrak section, Keping region), in the Tarim
112 Basin, NW China (Fig. 1). The main objectives of this study are to 1) document the
113 stratigraphic variation of crystallographic parameters and element content across the
114 entire formation; 2) identify the type(s) of diagenesis in the Xiaoerbrak Formation; and,
115 3) discuss the relationship between the diagenesis and reservoir development.
116
117 2. Geological Setting
118

119
120 Figure 1. Proterozoic and Phanerozoic geology of the Keping region, NW China. The
121 red star shows the location of the Xiaoerbrak section (modified from Zhou et al., 2015).
122
123 The Xiaoerbrak section of the Cambrian strata (GPS location: N 40° 55′ 23″, E 79° 53′
124 53″), identified by the Tarim Oil Company in 2008, is located in the Keping region, in
125 the northwestern part of the Tarim Basin, NW China (Fig. 1). The thickness of the
126 Xiaoerbrak Formation (~172 m in this section) is constant over the northwester of the
127 Tarim Basin. The Xiaoerbrak Formation conformably overlies the Cambrian Yurtus
128 Formation (21.6 m) and is conformably overlain by the Wusonger Formation. On the
129 basis of lithological characteristics, porosity, and thickness, the Xiaoerbrak Formation
130 is divided into Members I, II, and III, in an ascending stratigraphic order (Zheng et al.,
131 2020a, 2020b) (Fig. 2).
132
133
134 Figure 2. Stratigraphic log of the Xiaoerbrak section in the Keping region (modified
135 after Zheng et al., 2020b). Natural gamma ray values and porosity data are from Zheng
136 et al., 2020b. The light blue stars represent the sample position within the stratigraphic
137 log; Member I: from XBK-21 to XBK-27, Member II: from XBK-28 to XBK-50,
138 Member III: XBK-51 to XBK-66. The stratigraphic thickness was measured from the
139 base of the underlying Cambrian Yurtus Formation, which is consistent with an earlier
140 study (Zhang et al., 2020).
141
142 The main lithology of the Xiaoerbrak Formation in the Xiaoerbrak section, alongside
143 natural gamma and porosity data (Zheng et al., 2020b), is shown in Figure 2. Member
144 I mainly consists of black laminated microbialitic dolostones, locally intercalated with
145 dolarenite, suggesting a low-energy, relatively deep-water depositional environment.
146 The lower parts of Member II (from ~43 to ~103 m in the stratigraphic profile; Figure
147 2) comprise laminated microbialite to medium–thick-layered thrombolitic dolostone,
148 indicating a moderate-energy depositional environment (Hu et al., 2014). Overlying
149 dolarenite, spongiostromata biostromes, oncolitic dolostone, and siliciclastic dolostone
150 at the upper part of this member from ~103 to ~156 m in the stratigraphic profile (Yan
151 et al., 2017; Zheng et al., 2020b), suggesting a shallower depositional environment. As
152 reported previously (Hu et al., 2014; Li et al., 2011, 2020), the Xiaoerbrak Formation
153 dolostones consist of predominantly micritic dolomite; however, the top part of
154 Member II from ~138 to ~150 m in the stratigraphic profile (Figure 2), with sparry
155 dolomite in spongiostromate biostromes cavities (Jiang et al., 2019; Yan et al., 2017),
156 is an exception. This lithology and the presence of cross-bedding suggest that the top
157 of Member II was deposited in a high-energy environment (Hu et al., 2014; Jiang et al.,
158 2019; Zheng et al., 2020b). Notably, in Member II, the layer thickness, porosity, and
159 permeability of the dolostone generally increase up-section (Fig. 2; Zheng et al., 2020b).
160 The natural gamma value is fairly stable in Members I and II; however, Member III
161 (from ~156 to ~193 m in Figure 2), which comprises argillaceous dolostones
162 interlayered with stromatolitic dolostones, has much higher natural gamma values than
163 the other two members (Zheng et al., 2020a, 2020b). Subaerial exposure features, such
164 as mud cracks and tepee-like structures, are found in the Member III (Zheng et al.,
165 2020b).
166
167 3. Materials and analytical methods
168
169 In total, 46 dolostone samples in the Xiaoerbrak Formation (samples XBK-21 to XBK-
170 66) were collected from the Xiaoerbrak section. During field sampling, visible
171 alteration such as weathering crusts and visible calcite veins were avoided. The hand
172 specimen was then subsequently crushed into small pieces of ~0.5 cm diameter. The
173 best-preserved pieces (i.e., showing no visible veins to naked eyes) were hand-picked
174 and ground into powder (200-400 mesh) for further geochemical analyses.
175 Sample powders (200-400 mesh) were analyzed for dolomite crystallographic
176 parameter (such as dolomite stoichiometry and cation ordering) with a Bruker D8
177 Advance X-ray diffractometer with a Cu anode at the State Key Laboratory for Mineral
178 Deposits Research, Nanjing University. Mineralogical data were collected using a 2θ
179 range from 20° to 40°, a step size of 0.01°, and a count time of 1 s/step. The obtained
180 raw data were processed with Jade 6.5 software for the calculation of mineralogical
181 parameters and intuitive data visualization (Fig. 3). The abundance of dolomite relative
182 to calcite (dolomite percentage) was calculated using the ratio of the dolomite (104)
183 peak area to the dolomite (104) peak area plus the calcite (104) peak area (Lumsden,
184 1979). The cation ordering was determined using the intensity ratio of the dolomite
185 (015) peak to the (110) peak (I(015/110)) (Goldsmith and Graf, 1958).
186 For REEs and barium (Ba) concentrations analysis, we followed the leaching method
187 of Zhang et al., 2020. In brief, to obtain the carbonate fractions of the collected samples
188 without the influence of silicate dissolution, ~50 mg of carbonate powder was leached
189 with 3 mL of 1 mol/L HAc for 12 h, and then leached twice with pure water at room
190 temperature (~20 °C). The dissolved carbonate fractions were transferred into 3 % v/v
191 nitric acid with 10 ppb internal-standard rhodium. Ba and REE concentrations were
192 determined using a Thermo Fisher Scientific Element XR sector-field inductively
193 coupled plasma mass spectrometer (ICPMS) at the same laboratory. A sample of
194 dolomite reference material (JDo-1) was analyzed once per 10 samples analyses.
195 Repeated analyses of JDo-1 yielded >90 % accuracy and >95 % precision for the Ba
196 and REEs concentrations reported in this study. Anomalies including Ce, Pr anomaly
197 are using conventional calculation equations: Ce/Ce* = 2Cen/(Lan + Prn) and Pr/Pr* =
198 2Prn/(Cen + Ndn)), respectively, where lower subscript n denotes normalization of
199 concentration against the surface Pacific water (SN, Kawabe et al., 1998).
200 A miniaturized CAS extraction protocol (He et al., 2020, 2019) was applied to extract
201 the CAS contents and other elements (Al, Mn, Sr and Fe) in the dissolved carbonate
202 part. ~10 g powder of the bulk samples was firstly bleached with excess 6 % sodium
203 hypochlorite (NaOCl) for 48 hours to oxidize organic matter and metastable sulfides
204 minerals. The solid bleached residue was obtained after filtration through 0.2 µm
205 polypropylene membrane syringe filters. The obtained residue was then washed in 10 %
206 NaCl solution for 24 hours to remove easily soluble sulfate and other non-CAS sulfur-
207 bearing compounds. The NaCl washing process was repeated for five times to guarantee
208 the removal of water-soluble contaminates (Edwards et al., 2019), and samples were
209 constantly agitated using a roller shaker at room temperatures. The five-times NaCl-
210 leached carbonate samples were digested with 6 mol/L HCl, the quantity of which was
211 calculated based on the total carbonate content of these samples (calculated using the
212 Ca, Mg concentration reported in Zhang et al., 2020, Table A.1). This HCl digestion
213 step was completed within 30 mins to minimize the potential for oxidation of any
214 remaining pyrite and other possible contaminates during dissolution. The extracted
215 CAS solution was retained by filtration through 0.2 µm polypropylene membrane
216 syringe filters and then measured for the concentration of SCAS and other elements using
217 the ICP Optical Emission Spectrometer (OES) at School of Earth and Environment,
218 University of Leeds, UK.
219 In addition, we made 4 thin sections for typical samples of each member in the
220 Xiaoerbrak Formation, which are XBK-24 in the Member I, XBK-44 and XBK-49 in
221 the Member II, and XBK-57 in the Member III. Standard thin sections were prepared
222 at Nanjing Hongchuang Geological Exploration Technology Service Co. Ltd., Nanjing,
223 China. The thin sections of these samples were examined with a Zeiss optical
224 microscope at the State Key Laboratory for Mineral Deposits Research, Nanjing
225 University. Microphotographs were taken at a magnification of 100x and 200x under
226 plane polarized light (PPL).
227
228 4. Results
229 4.1 Petrography and mineralogy
230

231
232 Figure 3. Representative XRD spectra and microphotographs of the Xiaoerbrak
233 Formation carbonates, showing pure dolomite composition (with a few quartz grains in
234 Member III). Note the decreasing FWHM of the dolomite peak (104) accompanies the
235 increasing recrystallization up-profile from Member I (XBK-24) to the upper part of
236 Member II (XBK-44). XBK-57 is a typical sample of Member III in the Xiaoerbrak
237 Formation. Samples and lithology: A, XBK-24 – fine-crystalline dolomite (~10 μm); B,
238 XBK-44 – coarse-crystalline dolomite (~200 μm); C, XBK-57 – micritic dolomite (<5
239 μm).
240
241 According to our petrographic work, the carbonates of the Xiaoerbrak Formation are
242 predominantly composed of dolomite (e.g., Fig. 3A and Fig. 3B), except a few quartzes
243 in the Member III (Fig. 3C), which is in line with previous study (Hu et al., 2014; Li et
244 al., 2020; Yan et al., 2017). Member I mainly consists of fine-crystalline dolomite (~10
245 μm, Fig. 3A). Member II contains coarser dolomite (10-200 μm, Fig. 3B) but with
246 varied crystal size especially in the spongiostromate biostromes at the top of Member
247 II (Fig. 4). The spongiostromate biostromes could be divided into two groups: (a)
248 euhedral coarse-grained dolomite filling cavities in (b) anhedral fine-crystalline
249 dolomite matrix (Fig. 4). The microscopic photos and XRD spectra of Member III (Fig.
250 3C) show that Member III is composed of micritic dolomites and a few quartzes.
251

252
253 Figure 4. Microscopic characteristics of spongiostromate biostromes from the
254 Xiaoerbrak Formation (top of Member II): euhedral coarse-grained dolomite filling
255 cavities in anhedral fine-crystalline dolomite matrix (Sample XBK-49).
256
257 XRD analyses further confirm the mineral composition and crystalline degree of the
258 carbonates in each member (Fig. 3). Dolomite stoichiometry (% dolomite), calculated
259 from the area of peak (104) is relatively stable and high (>97 %) for most of the samples
260 (Table A.1), which is consistent with the high Mg/Ca ratios (>0.56) across the
261 Xiaoerbrak Formation in the previous study (Zhang et al., 2020). As Figure 3 and Figure
262 5 show, in these dolomites, the FWHM of the dolomite peak (104) and d(104) decreases
263 from Member I to the top of Member II; then increases up-profile across the rest of the
264 Xiaoerbrak Formation. The cation ordering (I(015/110)), calculated from the dolomite
265 peak (015)-to-(110) ratio, increases from Member I to member II and decreases again
266 in member III, showing a general opposite trend to the FWHM of the dolomite peak
267 (104) and d(104). In addition, it is noteworthy that samples with higher cation ordering
268 generally show coarser dolomite grains in the microphotographs, and vice versa (Fig. 3
269 and Fig. 5).
270
271 4.2 CAS and Al, Fe, Mn and Sr concentrations
272
273 The Al concentrations of the carbonate leachates (Fig. 5) as well as Fe concentrations
274 (Table A.1) are lower than the common concentration of dolomites (usually >500 ppm
275 Al; Ling et al., 2013) and relative stable at ~50 ppm Al in Member I and Member II.
276 However, they demonstrate high values in the quartz-bearing carbonates of Member III
277 (Fig. 3 and Fig. 5).
278
279
280 Figure 5. Variation in lithology, elemental concentration in carbonate leachates (REEs
281 concentrations are from acetic acid leaching, other elements concentrations are from
282 CAS extraction method), crystallographic parameters of dolomites (I(015/110),
283 FWHM(104) and d(104)), and bulk O-isotopic composition in the Xiaoerbrak
284 Formation dolostones (Zhang et al., 2020). Note: vertical dotted black lines refer to
285 mean values of each dataset, and the transparent grey lines are LOWESS-fitting curves
286 helping to identity the variation trend.
287
288 Overall, the variation trends of the SCAS and Sr concentration in the carbonate leachates
289 are similar to those of d(104) and the FWHM of the dolomite peak (104); the SCAS and
290 Sr concentration decreases up-profile from Member I to the upper part of Member II;
291 then, it increases within Member III (Fig. 5). The Mn concentration seemingly shares a
292 similar trend; however, it shows higher value (>150 ppm) in Member I and varies gently
293 in Member II and III (almost stable at ~95 ppm). The Mn/Sr ratio ranges from 1 to 6
294 but is mainly lower than 3 (Fig. 5), generally opposite to the variation curve of Sr
295 concentration. Note that the variation trends in Mn, Sr concentrations and Mn/Sr ratios
296 using CAS extraction method are consistent with the concentration curves using acetic
297 acid leaching method (reported in Zhang et al., 2020) but with slightly lower
298 concentrations.
299 In addition, to identify the type(s) of fluid alteration, we also plotted the oxygen isotope
300 of bulk samples (δ18O) previously reported in Zhang et al., 2020 in Figure 5. The δ18O
301 values generally increase from −7.5 to −6 ‰ between members I and II and then remain
302 stable at −6 ‰ in Member III. Exceptions to this are some lower values (−9 to −7 ‰)
303 in the middle part of Member II and a few higher values (−6 to −4 ‰) in the lower part
304 of Member III.
305
306 4.3 REE patterns of the carbonates
307
308 The total REE concentration of the carbonate leachates (mean value: ~1.08 ppm, Fig.
309 5) is one order of magnitude lower than that in other Cambrian carbonates (usually >10
310 ppm: e.g., Ling et al., 2013). It decreases sharply from Member I to the base of Member
311 II; it remains relatively stable in Member II; finally, it increases in quartz-bearing
312 Member III (Table A.1 and Fig. 5).
313 In many samples, the concentration of some REEs in the carbonate leachates is below
314 the detection limit of ICPMS (e.g., Ce). For this reason, Figure 6A only includes
315 samples, all the REEs of which were detectable. In addition, to exclude fake positive
316 Ce anomaly because of the enrichment of La (Shields and Stille, 2001), we plotted Ce
317 anomaly against Pr anomaly in Figure 6B. All the positive Ce anomalies in our samples
318 are confirmed to be true because they all have a seawater normalized Pr anomaly less
319 than one (Bau and Dulski, 1996). Relatively depletion or enrichment of LREEs (light
320 REEs) is indicated by NdSN/YbSN ratio.
321 The main features of the REE patterns are: strong LREE enrichment (average
322 NdSN/YbSN= 8.73), notable positive CeSN anomalies (average = 3.79), and no positive
323 EuSN anomaly (average (Eu/Eu*)SN = 0.81, not including XBK-63) (Table A.2, Fig. 6).
324 Sample XBK-63 with a distinctly positive Eu anomaly (Fig. 6) has an abnormally high
325 Ba concentration (~370 ppm Ba, Table A.2). Moreover, XBK-27, XBK-33 and XBK-
326 43 exhibit slightly different REE signatures. As shown in Figure 6A, their total REE
327 concentrations are much lower than the other carbonates in the Xiaoerbrak Formation
328 and their CeSN anomalies are absent or little (average = 0.98). However, they basically
329 share a pattern with LREE enrichment (average NdSN/YbSN= 8.16).
330
331
332 Figure 6. A: Seawater-normalized (Kawabe et al., 1998) REE pattern in the Xiaoerbrak
333 Formation carbonates (Xiaoerbrak section); B: (Ce/Ce*)SN versus (Pr/Pr*)SN (Bau and
334 Dulski, 1996).
335
336 5. Discussion
337 5.1 Dolostone recrystallization
338
339 Previous petrographic studies have shown that recrystallization has occurred in parts of
340 the Member I and II dolostones (Hu et al., 2014; Li et al., 2020; Zheng et al., 2020b),
341 especially in the spongiostromate biostromes at the top of Member II (Jiang et al., 2019).
342 Although dolomite recrystallization has been sporadically reported in the Xiaoerbrak
343 Formation, the specific process behind the dolomite recrystallization in the Xiaoerbrak
344 Formation has not been resolved.
345 Dissolution and recrystallization processes occur during the fluid-induced alteration
346 and dolomitization of carbonates (Pina et al., 2020). These processes gradually increase
347 the cation ordering in dolomite, as has been attested in dolomite synthesis experiments
348 (Li et al., 2015) and natural dolostones (Li et al., 2019; Pina et al., 2020). In Member II
349 of the Xiaoerbrak Formation, the cation ordering in dolomite, measured from the
350 relative intensity of super-lattice X-ray reflection I(015/110) (Goldsmith and Graf,
351 1958), gradually increases. This likely suggests that, in Member II dolostones,
352 recrystallization progressed downwards, from the top to the base of this member.
353 To furtherly study the recyclization process, cross-plots of crystallographic parameters,
354 element content concentrations and δ18O data (Zhang et al., 2020) are applied. Member
355 II shows a strong negative correlation (R2 = 0.87, Fig. 7A) between the FWHM of the
356 dolomite peak (104) and I(015/110), and a weak negative correlation (R2 = 0.14, Table
357 A.1) between d(104) and I(015/110). A negative correlation between d(104) and
358 I(015/110) was observed in many dolomite synthesis experiments and natural
359 diagenetic dolomites (e.g., Li et al., 2019, 2015) and interpreted as ripening towards
360 higher crystallinity. The FWHM of the dolomite peak (104) is also an indicator of
361 ripening, which usually occurs during the dissolution and recrystallization process (Pina
362 et al., 2020). Across Member II, with increasing I(015/110) up-profile, the FWHM of
363 the dolomite peak (104) decreases from ~0.18 to ~ 0.10. This trend strongly supports to
364 the inference that Member II underwent recrystallization that progressed downwards.
365 In brief, the upward increase in I(015/110), the negative correlations between I(015/110)
366 and the FWHM of dolomite peak (104), and d(104) demonstrate a progressive
367 recrystallization trend from the top to the base of Member II.
368 Generally, because of the higher concentration of Mn, lower concentration Sr and lower
369 δ18O values in meteoric water than in seawater, alteration related with meteoric water
370 generally results in an increase in the Mn content and decreases in the Sr concentration
371 and δ18O value in the rock (Brand and Veizer, 1980; Derry et al., 1994; Land, 1985).
372 Although a weak positive correlation (R2 = 0.32, Fig. 7B) between I(015/110) and
373 Mn/Sr ratio is found in Member II, the Mn concentration in Member II dolostones is
374 relatively stable (~95 ppm, Fig. 5), not showing a similar increasing trend with
375 I(015/110) as we expected for a meteoric-type alteration. In addition, if the alteration is
376 related with meteoric water, lighter O isotopes are expected to be found in the samples
377 having high cation ordering, resulting in a negative correlation between I(015/110) and
378 bulk δ18O value in the Member II. However, in our dataset, the weak positive correlation
379 between the I(015/110) and bulk δ18O values (Fig. 7H), instead of a negative correlation,
380 probably argues against that meteoric-type diagenetic process majorly controls the
381 dolomite recrystallization (Li et al., 2011; Liu et al., 2020; Shen et al., 2016; Yan et al.,
382 2017) .
383

384
385 Figure 7. Cross-plots of crystallographic parameters of dolomite (I(015/110), d(104)
386 and FWHM(104)), CAS, Mn, Sr (CAS extraction method), total REE abundance (acetic
387 acid method), and δ18O (Zhang et al., 2020) in the Xiaoerbrak dolostones. Only linear
388 fitting for Member II is plotted since no obvious correlations found among most of these
389 parameters.
390
391 We think that the negative correlation between I(015/110) and Mn/Sr ratio (Fig. 7B) is
392 probably not due to meteoric water corrosion but due to the relative loss of Sr during
393 the altering-fluid-induced recrystallization. Because the radius of Sr2+ is larger than that
394 of Ca2+ compared with Mn2+, the recrystallization of dolomite results in a relative Sr
395 loss and a lower Sr concentration in recrystallized dolomite crystals (Brand and Veizer,
396 1980). The Sr concentration is negatively correlated with I(015/110) (R2 = 0.36, Fig.
397 7C) and positively correlated with d(104) (R2 = 0.48, Fig. 7D) in Member II, suggesting
398 that the Member II dolostones were subjected to fluid alteration and Sr loss (Li et al.,
399 2019).
400 Besides, a weak negative correlation between SCAS concentration and I(015/110) (R2 =
401 0.25, Fig. 7E) and a strong positive correlation between SCAS concentration and Sr
402 concentration (R2 = 0.81, Fig. 7F) furtherly confirm the ejection of other ions during
403 dolomite recrystallization. This is consistent with the knowledge that the degree of
404 crystallographic order is controlled by the presence of substituted ions which can be
405 leached out of the crystal lattice during fluid-rock interaction (Fichtner et al., 2017).
406 This could be explained by more Sr and SO4 ejection (or fewer Sr and SO4 incorporation)
407 during the formation of well-ordered dolomite compared with low-ordered dolomite
408 (Baldermann et al., 2015; Fichtner et al., 2017). Therefore, based on this assumption,
409 the lower total REEs concentration in the Member II dolostones than the other two
410 members is also probably a result of the rejection of REE ions during dolomite
411 recrystallization (Wang et al., 2009).
412 The REEs in marine dolomites are inherited mainly from seawater (e.g., Ling et al.,
413 2013; Tostevin et al., 2016), therefore seawater normalized REE patterns of dolomite
414 can better indicate the source of the post-diagenesis fluids (L. Wang et al., 2014).
415 Seawater-normalized positive Ce anomaly, and LREE enrichments are indicative of
416 seawater origin (L. Wang et al., 2014), whereas the REE pattern can be changed and
417 shows distinguished features when dolomites are altered by meteoric water and
418 hydrothermal fluids (Kawabe et al., 1998; L. Wang et al., 2014). For example, since Eu
419 is enriched in highly reducing hydrothermal fluids (Michard and Albarède, 1986) and
420 thus Eu anomalies are utilized to identify the influence of hydrothermal fluids (L. Wang
421 et al., 2014). As shown in Figure 6, almost all the Xiaoerbrak dolostones show apparent
422 CeSN anomaly (average = 3.79) and strong LREE enrichment (average NdSN/YbSN=
423 8.73), except three samples (XBK-27, XBK-33, and XBK-43) without obvious positive
424 CeSN anomaly (< 2). These “seawater-like” features suggest that fluid alteration did not
425 substantially change the pristine REE pattern and (or) the altering fluid has a “seawater-
426 like” REE pattern.
427 In addition, no obvious positive CeSN anomalies in the XBK-33 and XBK-43, along
428 with low δ18O values (−9 to −7 ‰) in the middle part of Member II, indicate that the
429 middle Member II indeed experienced the meteoric water alteration to some extent (Li
430 et al., 2011; Liu et al., 2020; Shen et al., 2016; Yan et al., 2017). It is because that
431 leaching by Ce-depleted meteoric water (Frimmel, 2009) can significantly decrease the
432 Ce content in carbonates. However, as discussed above, the overall Mn concentration
433 and δ18O values probably suggest that the primary alteration fluids did not originate
434 from meteoric water; in other words, meteoric water alteration is not the first-order
435 control in this progressive recrystallization event. Moreover, no positive EuSN
436 anomalies are found in our dataset (Table A.2, Fig. 6A), with an exception of sample
437 XBK-63. XBK-63’s positive Eu anomaly is probably fake because high Ba content in
438 this sample (~370 ppm Ba, Table A.2) likely drove strong BaO interference on Ba
439 element during the ICPMS analysis. Accordingly, the absence of positive Eu anomalies
440 in the REE pattern (Fig. 6A) contradicts the hypothesis of hydrothermal fluid
441 contribution (Li et al., 2017).
442 Based on this “seawater-like” REE pattern (Fig. 6A) and the evidence from Mn
443 concentration and O isotope discussed above, we infer that the primary fluid that altered
444 the Xiaoerbrak Formation carbonates originated from seawater, whereas the meteoric
445 water only altered the middle part of Member II.
446
447 5.2 Fluid alteration of the Xiaoerbrak dolostones and its effects on reservoir
448 development.
449
450 Due to absence of systematic studies on the diagenetic alteration of the Xiaoerbrak
451 Formation in other sections over the Tarim Basin, the influence scope of this identified
452 progressive diagenetic alteration cannot be determined conclusively. However, because
453 almost entire northwestern part of the Tarim Basin developed the same depositional
454 environment in the Xiaoerbrak Formation (Feng et al., 2006) with constant layer
455 thickness, we assume that the progressive diagenetic alteration likely occurred in the
456 other contemporary sections over the northwestern basin.
457 During the deposition of the Xiaoerbrak carbonates, the Tarim Block was located in a
458 low–middle latitude region (Huang et al., 2000; Merdith et al., 2017); therefore, it was
459 potentially subjected to relatively strong evaporation. Earlier works on the lithology
460 and paleogeography of Members II and III of the Xiaoerbrak Formation suggested that
461 lagoonal and other shallow-water environments were widespread at the Member II–III
462 boundary (e.g., Feng et al., 2007; Zheng et al., 2020b), by the extensive presence of
463 cross-bedding and mud cracks at the top of Member II and at the base of Member III
464 over the northwest of the Tarim Basin (Zheng et al., 2020b, 2020a). Accordingly, we
465 propose a mechanism of fluid alteration of the Xiaoerbrak Member II dolostone, as
466 shown in Figure 8.
467

468
469 Figure 8. Proposed mechanism of fluid alteration in the Xiaoerbrak Formation.
470
471 In such a lagoon environment, evaporation processes preferentially removed
472 isotopically light O (e.g., Jouzel et al., 1994), leading to an increased δ18O value of the
473 lagoon seawater (Zhu et al., 2021), as registered in the lower part of Member III (Fig.
474 5; Zhang et al., 2020). This lagoon seawater recharged into the pre-deposited Member
475 II carbonates and diffused downwards, gradually altering them. As evaporated seawater
476 has relatively higher δ18O value, the evaporated seawater altered dolomite is expected
477 to have high δ18O values (Zhu et al., 2021), which has been evidence by the extensive
478 occurrence of higher δ18O values in the Xiaoerbrak dolostone than contemporaneous
479 seawater (Shen et al., 2016). This mechanism is further supported by the consistent
480 decreases in I(015/110), Mn/Sr ratios, and δ18O values from the top to the base of
481 Member II (Fig. 5 and Fig. 8). In this case (Figure 8), the downward lagoon seawater
482 would continuously alter the Member II carbonates, may leading to over-dolomitization
483 and consequently destroying the carbonates’ porosities (Consonni et al., 2010).
484 However, in our study, samples with higher porosity as well as permeability generally
485 have a higher dolomite crystalline degree (Fig. 5, Fig. 8 and Fig. 9), probably arguing
486 against the influences of over-dolomitization on the Member II carbonates. In addition,
487 the carbonate δ13C record of the Xiaoerbrak Formation shows similarities with the
488 global δ13C record (Zhang et al., 2020), probably indicating this lagoon occasionally
489 connected with the open ocean. When the lagoon was well linked to the open ocean,
490 the open ocean would have supplied seawater into the lagoon thus likely diluting the
491 lagoon evaporated seawater, and consequently avoiding the occurrence of over-
492 dolomitization (Consonni et al., 2010). In addition, the recharge of open ocean seawater
493 would also contend with the evaporation of lagoon seawater to sustain the fluid
494 alteration process.
495 During the deposition of Member III, while the sea level continued to fall and detrital
496 inputs increased (Zheng et al., 2020b) as shown with increased Al, Fe concentrations
497 (Table A.1, Fig. 5) and the occurrence of quartzes in the Member III (Fig. 3C), the
498 lagoon stopped receiving seawater from the open ocean (Fig. 8). Fluid alteration
499 continued until the lagoon brine was exhausted and the lagoon was filled with detrital
500 inputs, forming a giant terrestrial lens in the sediment succession. The decreased
501 I(015/110) and increased SCAS and Sr concentrations in Member III (Fig. 5 and Fig. 8)
502 indicate that the sediments continued to be altered until the lagoon disappeared. This
503 process may have also resulted in the higher SCAS, Sr and REE concentration in the
504 Member III carbonate leachates (Fig. 5), because more silicates (Fig. 3C), with higher
505 Sr and REE content, entered the ocean at that time.
506
507
508 Figure 9. Stratigraphic variation in, and possible cause for the correlation between,
509 porosity and dolomite cation ordering (I(015/110)) in Member II dolostones, the
510 Xiaoerbrak Formation. Note that porosity data are from Zheng et al., 2020b.
511
512 Figure 8 (bottom-left corner) and Figure 9 outline the relationship between porosity and
513 the fluid alteration event in Member II of the Xiaoerbrak Formation. Previous studies
514 show that the petrophysical properties such as porosity, permeability, and pore-throat
515 size can be substantially affected by diagenetic alterations (Abdolmaleki et al., 2016;
516 Amel et al., 2015; Tavakoli et al., 2011). Understanding diagenetic processes that affect
517 reservoir quality in carbonates is useful for porosity evolution studies and therefore
518 important for hydrocarbon explorations. As shown in study of Tavakoli et al., 2011,
519 carbonate cementation may have the maximum effect on the porosity of carbonates. It
520 is supported by the fact that the spongiostromata biostromes with recrystallized
521 dolomites in the cavity (Fig. 4) have the highest porosity across the Xiaoerbrak
522 Formation.
523 Differences in porosity-permeability distribution of Member II correspond to different
524 cation ordering in dolomite (Fig. 9): samples with higher porosity as well as
525 permeability usually have a higher dolomite crystalline degree. However,
526 recrystallization is suggested to have little effect on the porosity of carbonate reservoirs,
527 as shown in an example of Holocene dolomites (Gregg et al., 1992). Absence of obvious
528 pores found in the thin sections of our samples might also support this suggestion.
529 Therefore, we think the correlation between the porosity and the dolomite
530 crystallization degree is probably because that the altering fluid diffused into the rock
531 through a pre-existing pore network without distinct influence on the original pore
532 system (Fig. 8 and Fig. 9). Variations in porosity and permeability may have controlled
533 the flux of the reactive fluid, resulting in varied water–rock ratio (Fig. 9). Higher water–
534 rock ratio— more brine around the previously deposited carbonate—enabled brine with
535 high Mg concentration to alter the rock further and resulted in a higher Ca–Mg cation
536 ordering in the recrystallized dolomite, and vice versa (Fig. 9). Accordingly, reservoir
537 quality is strongly affected by variations in the original rock fabrics; in other words,
538 reservoir quality is strongly affected by variations in the original rock structure, such as
539 the porosity and pore-throat structure, as shown in the study of Amel et al., 2015.
540 Therefore, we think that the reservoir quality of the Xiaoerbrak Formation was mainly
541 controlled by microbial structures and depositional facies (Huang et al., 2016; Li et al.,
542 2020; Song et al., 2014), which probably controlled the original rock fabrics during its
543 formation and recrystallization may have little effect on the primary fabric of the
544 Xiaoerbrak reservoir. For studying the relationship between diagenetic alteration and
545 reservoir quality, we suggest that the original rock petrophysics takes a very important
546 role in reservoir development. In this study, pre-existing pore network in carbonates
547 can provide pathways for altering fluid and then increase the possibility of fluid
548 alteration. This conclusion further brings a speculation that original carbonates with
549 original high porosity, are more likely to be altered. Original pore network, as an
550 accelerator in its diagenesis history, provides a pathway for altering fluid to more severe
551 chemical and physical reform.
552
553 6. Conclusion
554
555 Dolomite crystallographic parameters (dolomite cation ordering [I(015/110)], dolomite
556 percentage, d(104), FWHM of the dolomite peak (104)) and elements (Al, Fe, SCAS,
557 Mn, and Sr content, REE pattern) were analyzed to study the diagenesis and its relation
558 with reservoir quality in the Cambrian Xiaoerbrak Formation in the Keping area of the
559 Tarim Basin, NW China. In Member II of the Xiaoerbrak Formation, the
560 stratigraphically downward decrease in both the dolomite cation ordering and the Mn/Sr
561 ratio, and the correlations between these crystallographic parameters and elemental
562 contents (Sr and SCAS) suggest that this member was subjected to a recrystallization
563 event due to a downward fluid alteration. In addition, the concentrations of these
564 elements are controlled by the leaching process out of the crystal lattice during fluid-
565 rock interaction, and therefore show tight connections with the dolomite
566 crystallographic parameters.
567 The O isotope values, Mn contents in the Member II dolostones and the “seawater-like”
568 REE pattern of the carbonate leachates imply that the primary altering fluid originated
569 from seawater, while the meteoric water partly altered the middle part of the Member
570 II. Based on previous lithological and paleogeographic studies, we propose that the
571 primary altering fluid were derived from lagoon seawater. And, the reservoir quality of
572 the Xiaoerbrak Formation is strongly affected by variations in the original rock
573 petrophysics, which supports the assumption, that the reservoir quality of the
574 Xiaoerbrak Formation was mainly controlled by microbial structures and depositional
575 facies. Accordingly, our study shows that the original rock petrophysics can work as an
576 accelerator in its diagenesis history, providing a pathway for altering fluid and
577 increasing the vulnerability of carbonates to chemical and physical reformation.
578
579 Acknowledgement
580 The authors thank reviewers and editors for their helpful suggestions in improving this
581 paper. Y. Zhang appreciates Yuguan Pan’s help at the XRD laboratory at Nanjing
582 University. This work was financially supported by the National Science and
583 Technology Major Project of China (No. 2016ZX05004-004) and Strategic Priority
584 Research Program (B) of the Chinese Academy of Sciences (CAS) (No.
585 XDB26000000). Y. Zhang was funded by a state scholarship grant from China
586 Scholarship Council (CSC).
587
588 References
589 Abdolmaleki, J., Tavakoli, V., Asadi-Eskandar, A., 2016. Sedimentological and
590 diagenetic controls on reservoir properties in the Permian-Triassic successions of
591 Western Persian Gulf, Southern Iran. J. Pet. Sci. Eng. 141, 90–113.
592 https://doi.org/10.1016/j.petrol.2016.01.020
593 Amel, H., Jafarian, A., Husinec, A., Koeshidayatullah, A., Swennen, R., 2015.
594 Microfacies, depositional environment and diagenetic evolution controls on the
595 reservoir quality of the Permian Upper Dalan Formation, Kish Gas Field, Zagros
596 Basin. Mar. Pet. Geol. 67, 57–71.
597 https://doi.org/10.1016/j.marpetgeo.2015.04.012
598 Baldermann, A., Deditius, A.P., Dietzel, M., Fichtner, V., Fischer, C., Hippler, D., Leis,
599 A., Baldermann, C., Mavromatis, V., Stickler, C.P., Strauss, H., 2015. The role of
600 bacterial sulfate reduction during dolomite precipitation: Implications from Upper
601 Jurassic platform carbonates. Chem. Geol. 412, 1–14.
602 https://doi.org/10.1016/j.chemgeo.2015.07.020
603 Bau, M., Dulski, P., 1996. Distribution of yttrium and rare-earth elements in the Penge
604 and Kuruman iron-formations, Transvaal Supergroup, South Africa. Precambrian
605 Res. 79, 37–55. https://doi.org/10.1016/0301-9268(95)00087-9
606 Bialik, O.M., Wang, X., Zhao, S., Waldmann, N.D., Frank, R., Li, W., 2018. Mg isotope
607 response to dolomitization in hinterland-attached carbonate platforms: Outlook of
608 δ26Mg as a tracer of basin restriction and seawater Mg/Ca ratio. Geochim.
609 Cosmochim. Acta 235, 189–207. https://doi.org/10.1016/j.gca.2018.05.024
610 Brand, U., Veizer, J., 1980. Chemical diagenesis of a multicomponent carbonate system
611 - 1. Trace elements. J. Sediment. Petrol. 50, 1219–1236.
612 https://doi.org/10.1306/212F7BB7-2B24-11D7-8648000102C1865D
613 Consonni, A., Ronchi, P., Geloni, C., Battistelli, A., Grigo, D., Biagi, S., Gherardi, F.,
614 Gianelli, G., 2010. Application of numerical modelling to a case of compaction-
615 driven dolomitization: A Jurassic palaeohigh in the Po Plain, Italy. Sedimentology
616 57, 209–231. https://doi.org/10.1111/j.1365-3091.2009.01103.x
617 Demény, A., Czuppon, G., Kern, Z., Leél-Őssy, S., Németh, A., Szabó, M., Tóth, M.,
618 Wu, C.C., Shen, C.C., Molnár, M., Németh, T., Németh, P., Óvári, M., 2016.
619 Recrystallization-induced oxygen isotope changes in inclusion-hosted water of
620 speleothems – Paleoclimatological implications. Quat. Int. 415, 25–32.
621 https://doi.org/10.1016/j.quaint.2015.11.137
622 Derry, L.A., Brasier, M.D., Corfield, R.M., Rozanov, A.Y., Zhuravlev, A.Y., 1994. Sr
623 and C isotopes in Lower Cambrian carbonates from the Siberian craton: A
624 paleoenvironmental record during the “Cambrian explosion.” Earth Planet. Sci.
625 Lett. 128, 671–681. https://doi.org/10.1016/0012-821X(94)90178-3
626 Edwards, C.T., Fike, D.A., Saltzman, M.R., 2019. Testing carbonate-associated sulfate
627 (CAS) extraction methods for sulfur isotope stratigraphy: A case study of a
628 Lower–Middle Ordovician carbonate succession, Shingle Pass, Nevada, USA.
629 Chem. Geol. 529, 119297. https://doi.org/10.1016/j.chemgeo.2019.119297
630 Elderfield, H., Upstill-Goddard, R., Sholkovitz, E.R., 1990. The rare earth elements in
631 rivers, estuaries, and coastal seas and their significance to the composition of ocean
632 waters. Geochim. Cosmochim. Acta 54, 971–991. https://doi.org/10.1016/0016-
633 7037(90)90432-K
634 Feng, Z., Bao, Z., Wu, M., Jin, Z., Shi, X., 2006. Lithofacies palaeogeography of the
635 Cambrian in Tarim area. Front. Earth Sci. China 8, 265–274.
636 https://doi.org/10.1007/s11707-007-0033-2
637 Fichtner, V., Strauss, H., Immenhauser, A., Buhl, D., Neuser, R.D., Niedermayr, A.,
638 2017. Diagenesis of carbonate associated sulfate. Chem. Geol. 463, 61–75.
639 https://doi.org/10.1016/j.chemgeo.2017.05.008
640 Frimmel, H.E., 2009. Trace element distribution in Neoproterozoic carbonates as
641 palaeoenvironmental indicator. Chem. Geol. 258, 338–353.
642 https://doi.org/10.1016/j.chemgeo.2008.10.033
643 Gill, B.C., Lyons, T.W., Frank, T.D., 2008. Behavior of carbonate-associated sulfate
644 during meteoric diagenesis and implications for the sulfur isotope paleoproxy.
645 Geochim. Cosmochim. Acta 72, 4699–4711.
646 https://doi.org/10.1016/j.gca.2008.07.001
647 Goldsmith, J.R., Graf, D.L., 1958. Structural and Compositional Variations in Some
648 Natural Dolomites. J. Geol. 66.
649 Gregg, J.M., Howard, S.A., Mazzullo, S.J., 1992. Early diagenetic recrystallization of
650 Holocene (<3000 years old) peritidal dolomites, Ambergris Cay, Belize.
651 Sedimentology 39, 143–160. https://doi.org/10.1111/j.1365-3091.1992.tb01027.x
652 He, T., Dal Corso, J., Newton, R.J., Wignall, P.B., Mills, B.J.W., Todaro, S., Di Stefano,
653 P., Turner, E.C., Jamieson, R.A., Randazzo, V., Rigo, M., Jones, R.E., Dunhill,
654 A.M., 2020. An enormous sulfur isotope excursion indicates marine anoxia during
655 the end-Triassic mass extinction. Sci. Adv. 6, 1–26.
656 https://doi.org/10.1126/sciadv.abb6704
657 He, T., Zhu, M., Mills, B.J.W., Wynn, P.M., Zhuravlev, A.Y., Tostevin, R., Pogge von
658 Strandmann, P.A.E., Yang, A., Poulton, S.W., Shields, G.A., 2019. Possible links
659 between extreme oxygen perturbations and the Cambrian radiation of animals. Nat.
660 Geosci. 12, 468–474. https://doi.org/10.1038/s41561-019-0357-z
661 Hohl, S. V., Becker, H., Herzlieb, S., Guo, Q., 2015. Multiproxy constraints on
662 alteration and primary compositions of Ediacaran deep-water carbonate rocks,
663 Yangtze Platform, South China. Geochim. Cosmochim. Acta 163, 262–278.
664 https://doi.org/10.1016/j.gca.2015.04.037
665 Hu, W., Zhu, J., Wang, X., You, X., He, K., 2014. Characteristics, origin and geological
666 implications of the Cambrian microbial dolomite in Keping area, Tarim Basin (in
667 Chinese with English abstract). Oil Gas Geol. 35, 860–869.
668 https://doi.org/10.11743/ogg20140613
669 Huang, B., Zhu, R., Otofuji, Y., Yang, Z., 2000. The early paleozoic paleogeography
670 of the North China block and the other major blocks of China. Chinese Sci. Bull.
671 45, 1057–1065. https://doi.org/10.1007/BF02887174
672 Huang, Q., Hu, S., Pan, W., Liu, W., Chi, Y., Wang, K., Shi, S., Liu, Q., 2016.
673 Sedimentary characteristics of intra-platform microbial mounds and their
674 controlling effects on the development of reservoirs: A case study of the Lower
675 Cambrian Xiaoerbulake Fm in the Keping-Bachu area, Tarim Basin (in Chinese
676 with English abstract). Nat. Gas Ind. 36, 21–29.
677 https://doi.org/10.3787/j.issn.1000-0976.2016.06.003
678 Jiang, W., Liu, B., Shi, K., Gao, X., Liu, F., Yu, J., 2019. rofacies and Characteristics
679 of Carbonate Reservoir in Xiaoerbulake Formation of Keping Area, Tarim Basin
680 (in Chinese with English abstract). Xinjiang Pet. Geol. 40, 437–447.
681 https://doi.org/10.7657/XJPG20190407
682 Jouzel, J., Koster, R.D., Suozzo, R.J., Russell, G.L., 1994. Stable water isotope
683 behavior during the last glacial maximum: a general circulation model analysis. J.
684 Geophys. Res. 99, 25791–25801. https://doi.org/10.1029/94jd01819
685 Kawabe, I., Toriumi, T., Ohta, A., Miura, N., 1998. Monoisotopic REE abundances in
686 seawater and the origin of seawater tetrad effect. Geochem. J. 32, 213–229.
687 https://doi.org/10.2343/geochemj.32.213
688 Land, L.S., 1985. The origin of massive dolomite. J. Geol. Educ. 33, 112–125.
689 https://doi.org/10.5408/0022-1368-33.2.112
690 Li, B., Peng, J., Yang, S., Xia, Q., Xu, Q., Hao, Y., 2017. Genetic model and
691 characteristics of the Cambrian Shorebulake reservoir in Bachu area, Tarim Basin
692 (in Chinese with English abstract). Pet. Geol. Exp. 39.
693 https://doi.org/10.11781/sysydz201706797
694 Li, Q., Hu, W., Qian, Y., Zhang, J., Wang, X., Zhu, J., 2011. Features of dissoloved
695 dolostone reservoirs in the Xiaoerbulak Formation,Tarim Basin (in Chinese with
696 English abstract). Oil Gas Geol. 32.
697 Li, W., Beard, B.L., Li, C., Xu, H., Johnson, C.M., 2015. Experimental calibration of
698 Mg isotope fractionation between dolomite and aqueous solution and its
699 geological implications. Geochim. Cosmochim. Acta 157, 164–181.
700 https://doi.org/10.1016/j.gca.2015.02.024
701 Li, W., Bialik, O.M., Wang, X., Yang, T., Hu, Z., Huang, Q., Zhao, S., Waldmann,
702 N.D., 2019. Effects of early diagenesis on Mg isotopes in dolomite: The roles of
703 Mn(IV)-reduction and recrystallization. Geochim. Cosmochim. Acta 250, 1–17.
704 https://doi.org/10.1016/j.gca.2019.01.029
705 Li, Y., Pan, W., Wu, Y., Yang, G., Sun, C., Jiang, H., 2020. Fabric types of
706 microbialites from the Stage 3 of Cambrian Series 2 in Sugaitebulake section
707 Tarim Basin (in Chinese with English abstract). J. Palaogeography 22, 663–679.
708 Ling, H., Chen, X., Li, D., Wang, D., Shields-Zhou, G.A., Zhu, M., 2013. Cerium
709 anomaly variations in Ediacaran-earliest Cambrian carbonates from the Yangtze
710 Gorges area, South China: Implications for oxygenation of coeval shallow
711 seawater. Precambrian Res. 225, 110–127.
712 https://doi.org/10.1016/j.precamres.2011.10.011
713 Liu, W., Huang, Q., Bai, Y., Shi, S., 2020. Controls of meteoric water dissolution on
714 microbial carbonate reservoir formation in penecontemporaneous stage:a case
715 study of Lower Cambrian in Tarim Basin (in Chinese with English abstract). Earth
716 Sci. Frotiers. https://doi.org/10.13745/j.esf.sf.2020.5.21
717 Lumsden, D.N., 1979. Discrepancy Between Thin-Section and X-Ray Estimates of
718 Dolomite in Limestone. J. Sediment. Petrol. 49, 429–436.
719 Merdith, A.S., Collins, A.S., Williams, S.E., Pisarevsky, S., Foden, J.D., Archibald,
720 D.B., Blades, M.L., Alessio, B.L., Armistead, S., Plavsa, D., Clark, C., Müller,
721 R.D., 2017. A full-plate global reconstruction of the Neoproterozoic. Gondwana
722 Res. 50, 84–134. https://doi.org/10.1016/j.gr.2017.04.001
723 Michard, A., Albarède, F., 1986. The REE content of some hydrothermal fluids. Chem.
724 Geol. 55, 51–60. https://doi.org/10.1016/0009-2541(86)90127-0
725 Pina, C.M., Pimentel, C., Crespo, Á., 2020. Dolomite cation order in the geological
726 record. Chem. Geol. 547, 119667.
727 https://doi.org/10.1016/j.chemgeo.2020.119667
728 Shen, A., Zheng, J., Chen, Y., Ni, X., Huang, L., 2016. Characteristics, origin and
729 distribution of dolomite reservoirs in Lower-Middle Cambrian, Tarim Basin, NW
730 China (in Chinese with English abstract). Pet. Explor. Dev. 43, 340–349.
731 https://doi.org/10.11698/PED.2016.03.03
732 Shields, G., Stille, P., 2001. Diagenetic constraints on the use of cerium anomalies as
733 palaeoseawater redox proxies: An isotopic and REE study of Cambrian
734 phosphorites. Chem. Geol. 175, 29–48. https://doi.org/10.1016/S0009-
735 2541(00)00362-4
736 Song, J., Luo, P., Yang, S., Yang, D., Zhou, C., Li, P., Zhai, X., 2014. Reservoirs of
737 Lower Cambrian microbial carbonates, Tarim Basin, NW China (in Chinese with
738 English abstract). Pet. Explor. Dev. 41, 404–414.
739 https://doi.org/10.11698/PED.2014.04.03
740 Tavakoli, V., Rahimpour-Bonab, H., Esrafili-Dizaji, B., 2011. Qualité du réservoir
741 contrôlé par la diagenèse du champ de gaz de South Pars, une approche intégrée.
742 Comptes Rendus - Geosci. 343, 55–71. https://doi.org/10.1016/j.crte.2010.10.004
743 Tostevin, R., Shields, G.A., Tarbuck, G.M., He, T., Clarkson, M.O., Wood, R.A., 2016.
744 Effective use of cerium anomalies as a redox proxy in carbonate-dominated marine
745 settings. Chem. Geol. 438, 146–162.
746 https://doi.org/10.1016/j.chemgeo.2016.06.027
747 Wang, L., Hu, W., Wang, X., Cao, J., Chen, Q., 2014. Seawater normalized REE
748 patterns of dolomites in Geshan and Panlongdong sections, China: Implications
749 for tracing dolomitization and diagenetic fluids. Mar. Pet. Geol. 56, 63–73.
750 https://doi.org/10.1016/j.marpetgeo.2014.02.018
751 Wang, X., Jin, Z., Hu, W., Zhang, J., Qian, Y., Zhu, J., Li, Q., 2009. Using in situ REE
752 analysis to study the origin and diagenesis of dolomite of Lower Paleozoic, Tarim
753 Basin. Sci. China, Ser. D Earth Sci. 52, 681–693. https://doi.org/10.1007/s11430-
754 009-0057-4
755 Wang, Z., Xie, H., Chen, Y., 2014. Discovery and exploration of Cambrian subsalt
756 dolomite original hydrocarbon reservoir at Zhongshen-1 Well in Tarim Basin (in
757 Chinese with English abstract). China Pet. Explor. 19, 1–13.
758 Wierzbicki, R., Dravis, J.J., Al-Aasm, I., Harland, N., 2006. Burial dolomitization and
759 dissolution of Upper Jurassic Abenaki platform carbonates, Deep Panuke reservoir,
760 Nova Scotia, Canada. Am. Assoc. Pet. Geol. Bull. 90, 1843–1861.
761 https://doi.org/10.1306/03200605074
762 Yan, W., Zheng, J., Chen, Y., Huang, L., Zhou, P., Zhu, Y., 2017. Characteristics and
763 Genesis of Dolomite Reservoir in the Lower Cambrian Xiaoerblak Formation,
764 Tarim Basin (in Chinese with English abstract). Mar. Orig. Pet. Geol. 22, 35–43.
765 Zhang, Y., Yang, T., Hohl, S. V, Zhu, B., He, T., Pan, W., Chen, Y., Yao, X., Jiang, S.,
766 2020. Seawater carbon and strontium isotope variations through the late Ediacaran
767 to late Cambrian in the Tarim Basin. Precambrian Res. 345.
768 https://doi.org/10.1016/j.precamres.2020.105769
769 Zheng, J., Huang, L., Yuan, W., Zhu, Y., Qiao, Z., 2020a. Geochemical features and its
770 significance of sedimentary and diagenetic environment in the Lower Cambrian
771 Xiaoerblak Formation of Keping area, Tarim Basin (in Chinese with English
772 abstract). Nat. Gas Geosci. 31.
773 Zheng, J., Pan, W., Shen, A., Yuan, W., Huang, L., Ni, X., Zhu, Y., 2020b. Reservoir
774 geological modeling and significance of Cambrian Xiaoerblak Formation in
775 Keping outcrop area, Tarim Basin, NW China (In Chinese with English abstract).
776 Pet. Explor. Dev. 47, 499–511. https://doi.org/https://doi.org/10.1016/S1876-
777 3804(20)60071-4
778 Zhou, X., Chen, D., Dong, S., Zhang, Y., Guo, Z., Wei, H., Yu, H., 2015. Diagenetic
779 barite deposits in the Yurtus Formation in Tarim Basin, NW China: Implications
780 for barium and sulfur cycling in the earliest Cambrian. Precambrian Res. 263, 79–
781 87. https://doi.org/10.1016/j.precamres.2015.03.006
782 Zhu, D., Jin, Z., Hu, W., 2010. Hydrothermal recrystallization of the Lower Ordovician
783 dolomite and its significance to reservoir in northern Tarim Basin. Sci. China Earth
784 Sci. 53, 368–381. https://doi.org/10.1007/s11430-010-0028-9
785 Zhu, D., Liu, Q., Wang, J., Ding, Q., He, Z., 2021. Stable carbon and oxygen isotope
786 data of Late Ediacaran stromatolites from a hypersaline environment in the Tarim
787 Basin (NW China) and their reservoir potential. Facies 67, 1–25.
788 https://doi.org/10.1007/S10347-021-00633-0/FIGURES/18
789

View publication stats

You might also like