Journal Pre-proof Origin of modern dolomite in surface lake sediments on the central and western Tibetan Plateau Jiao Li, Liping Zhu, Minghui Li, Junbo Wang, Qingfeng Ma PII:
S1040-6182(20)30060-4
DOI:
https://doi.org/10.1016/j.quaint.2020.02.018
Reference:
JQI 8150
To appear in:
Quaternary International
Received Date: 20 October 2019 Revised Date:
27 December 2019
Accepted Date: 12 February 2020
Please cite this article as: Li, J., Zhu, L., Li, M., Wang, J., Ma, Q., Origin of modern dolomite in surface lake sediments on the central and western Tibetan Plateau, Quaternary International, https:// doi.org/10.1016/j.quaint.2020.02.018. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2020 Published by Elsevier Ltd.
1
Origin of modern dolomite in surface lake sediments on the central and
2
western Tibetan Plateau
3
Jiao Lia, Liping Zhua,b,c, Minghui Lia, Junbo Wanga,b, Qingfeng Maa
4
a
5
Research, Chinese Academy of Sciences (CAS), Beijing 100101, China
6
b
CAS Center for Excellence in Tibetan Plateau Earth Sciences, Beijing 100101, China
7
c
University of Chinese Academy of Sciences, Beijing, China
8
Abstract
Key Laboratory of Tibetan Environment Changes and Land Surface Processes, Institute of Tibetan Plateau
9
Modern dolomite was found in the sediments of 20 lakes on the central and western Tibetan
10
Plateau during lake surveys, in August to November of 2012-2015. We analyzed the mineralogy of
11
authigenic carbonates and the chemical composition of lake water to investigate the environmental
12
factors affecting dolomite formation and to determine the mechanisms of dolomite formation.
13
X-ray diffraction analysis suggested that the dolomite content ranged from 3% to 10%, with a
14
mean of 5%, and that the majority of the dolomite in the surface lake sediments was Ca-dolomite.
15
Dolomite formed in a wide range of salinity environments. There was a positive linear relationship
16
between salinity and Mg/Ca ratio, whereas the pH and sulfate (SO42-) concentration are not
17
significantly correlated with Mg/Ca ratio. High Mg/Ca ratio, high salinity, high sulfate
18
concentration, and high pH are important environmental factors in favor of dolomite precipitation.
19
Evaporation and microbial activity are potential mechanisms for dolomite formation in the central
20
and western Tibetan Plateau. The positive relationship between δ18O and δ13C indicates that the
21
dolomite formed in lakes has a long residence time and undergoes continuous evaporation. The
22
crystal morphologies of dolomite were scattered euhedral grains with diameters of 2‒10 µm, as
23
well as sub-spherical and spherical grains (< 2 µm) which were present either as dispersed crystals
24
or were aggregated into dolomite clusters. Dolomite crystals were identical both in shape and size
25
to those of microbial dolomite previously discovered in modern lake environments and culture
26
experiments. The dolomite precipitation temperature range calculated using the low-temperature
27
microbial equation was close to the measured temperature, further indicating that microbial
28
activity may be involved in dolomite formation. This study provides natural locations with
29
different water salinities and other environmental factors for dolomite precipitation, which has
30
important significance for studying the mechanism of dolomite formation.
31
Keywords: dolomite, mechanisms of dolomite formation, surface lake sediments, Tibetan Plateau
32
1 Introduction
33
Dolomite (CaMg[CO3]2 ) is a common carbonate mineral in the geological record, but very rare
34
in modern carbonate environments (Warren, 2000). The inability to precipitate dolomite
35
inorganically in laboratory condition under earth surface temperatures and pressures (Land, 1998)
36
makes it difficult to identify the mechanisms of dolomite formation. Controversy over the origin
37
of dolomite has generally been referred to as the “dolomite problem”, which is a complex issue
38
involving a large number of interacting factors such as thermodynamics, chemical kinetics,
39
hydrology, host-rock mineralogy and texture, and microbial activity (Fairbridge, 1957; Hardie,
40
1987; Vasconcelos and McKenzie, 1997). Over the last few decades, dolomite formation has been
41
identified in different environments and settings, such as marine sediments (Saller, 1984; Jose
42
Carballo et al., 1987; Compton, 1988; Chellie S. Teal et al., 2000), marginal marine lagoon and
43
lakes (Aharon et al., 1977; Warren, 1990; Vasconcelos and McKenzie, 1997; Van Lith et al., 2002;
44
Sánchez-Román et al., 2009b), sabkha (Bontognali et al., 2010). In recent years there has been an
45
increasing number of reports of dolomite in lakes, and a number of landmark studies of the origins
46
of dolomite have been proposed based on lacustrine sediments (Talbot and Kelts, 1986; Deckker
47
and Last, 1988; Rosen and Coshell, 1992; Colson and Cojan, 1996; Jiang and Liu, 2007;
48
BrÉHÉRet et al., 2008; Deng et al., 2010; Meister et al., 2011). The formation mechanisms of
49
lacustrine dolomite are diverse. Therefore, lakes are ideal large-scale laboratories that offer a wide
50
range of depositional environment to study the kinetics and define the conditions that promote
51
dolomite formation (Last, 1990).
52
Studies of sedimentary dolomite formation in modern environments have greatly increased our
53
understanding of the physicochemical conditions required for dolomite formation. Kinetic barriers
54
are believed to be the cause of the failure to form dolomite inorganically in laboratory at low
55
temperatures, which may be a sufficient explanation for the paucity of dolomite in modern marine
56
environments (Land, 1998). High salinity, pH, Mg/Ca ratio, and temperature are conducive to
57
dolomite formation due to the long-term replacement of pre-existing calcite (Krauskopf and Bird,
58
1995). The primary precipitation of dolomite can occur in aqueous solutions which are associated
59
with saline evaporate deposits and organic-rich sediments (Compton, 1988; Deckker and Last,
60
1988; Vasconcelos and McKenzie, 1997; Wright, 1999; Wright and Oren, 2005). Although many
61
dolomite formation models have been established, e. g., the seepage refluxion model (Adams and
62
Rhodes, 1960), the mixing-zone model (Badiozamani, 1973), sulfate reduction model (Baker and
63
Kastner, 1981), organogenic model (Slaughter and Hill, 1991), and microbial dolomite model
64
(Vasconcelos and McKenzie, 1997), a large number of culture experiments have proved that
65
microbial activity is the key to dolomite formation at low temperatures (Sánchez-Román et al.,
66
2008).
67
The Tibetan Plateau has a large group of lakes with rich mineral resources. Modern dolomite
68
has also been found in lakes on the Tibetan Plateau. Xia and Li (1986) suggested that dolomite
69
that formed in the Xiaochaidan Salt Lake originated from organic processes during cyanobacterial
70
activity rather than from replacement or pure chemical precipitation. Sub-spherical and spherical
71
dolomite in Bayinchagan Lake and Qinghai Lake are identical both in size and in shape to
72
microbially precipitated dolomite, suggesting a biogenic origin (Jiang and Liu, 2007; Yu et al.,
73
2007; Deng et al., 2010). Low-content dolomite was universal in all selected typical lakes in the
74
Qaidam basin and Central Tibetan Plateau; however, the origin of dolomite has not been discussed
75
(Wang et al., 2008).
76
In this study, we obtained numerous lacustrine surface sediment samples based on field surveys
77
of lakes on the Tibetan Plateau. Among them, dolomite occurred in 20 lakes with different
78
salinities on the central and western Tibetan Plateau (Table 1). The objectives of this study were,
79
therefore, to: (a) investigate the environmental factors affecting dolomite formation, such as the
80
Mg/Ca ratio, salinity, pH, and the SO42- concentration, and; (b) determine the mechanisms of
81
dolomite formation. This study provides new locations of modern dolomite in lacustrine
82
environments which has important significance for understanding dolomite formation.
83 84
2 Geological setting
85
The study area (30.8‒35.0 °N, 80.1‒89.0 °E) is located in the central and western Tibetan
86
Plateau (Fig. 1). It is surrounded by the Kunlun Mountains to the north, and the Gangdese
87
Mountains and Nyainqentanglha Mountains to the south. All the investigated lakes are located in
88
the Tibetan Plateau endorheic drainage basin. The lake area of these endorheic drainage basins
89
accounts for 70.7% of the total lake area of the Tibetan Plateau (Wan et al., 2016). The elevations
90
of these lakes are ~4500‒5000 m above-sea-level (a.s.l.), while the surrounding mountains reach a
91
mean altitude of ~5500‒6000 m (Guan et al., 1980).
92
The area has been subjected to arid to semi-arid climatic conditions in modern times, with mean
93
annual precipitation (MAP) of < 50 to 300–400 mm, and mean annual temperatures (MAT)
94
varying from < −8 to 2 °C (Institute of Geography, 1990). Lakes in the endorheic drainage basins
95
are mainly inland saltwater lakes or saline lakes (Guan et al., 1980). Bedrock of the lake basins is
96
primary composed of clastic, metamorphic, and igneous rocks (Fig. 1a-f) consisting mainly of
97
conglomerate, sandstone, siltstone, mudstone, limestone, shale, pyroclastic rock, granite, gneiss,
98
diabase, and ophiolite (Bureau of Geology and Mineral resources of Xizang Autonomous Regions,
99
1993). Sixteen types of carbonate minerals are found in the Tibetan Plateau, of which calcite,
100
dolomite, aragonite, and magnesite are the most abundant (Zhang and Zheng, 2017).
101
3 Methods
102
Sediment and water sampling was mainly carried out in late summer and autumn (August to
103
November) of 2012‒2015. Surface sediment samples were sampled from 55 lakes using an Ekman
104
Bottom Sampler. To reduce the possible influence of lakeshore detrital carbonate, surface lake
105
sediments were sampled from the centers of the lakes. The top 2 cm of sediments were collected
106
from each site. Lake water samples were collected from lake water surfaces. A multi-parameter
107
water detector (HYDROLAB DS5, Hach and EXO2, YSI) was used to measure water quality
108
parameters (i.e., temperature, pH, conductivity, salinity and total dissolved solids [TDS])
109
simultaneously at the water sampling positions. Climate data of the 55 lake sites were obtained
110
from the China Meteorological Forcing Dataset (Yang et al., 2010; Chen et al., 2011). The
111
horizontal resolution of climate data is 0.1°. The MAP and MAT for each site were the 30-year
112
average from 1981 to 2010.
113
Prior to mineralogical analysis and microscopic observation, all sediment samples were sieved
114
and the < 40 µm fractions were collected for further analysis, as the carbonate in the < 40 µm
115
fraction of lake sediment was considered to be authigenic origin (Fontes et al., 1996; Jiang and Liu,
116
2007). Mineral compositions were analyzed using X-ray diffraction (XRD) in the Beijing Micro
117
Structure Analytical Laboratory. Carbonate contents were calculated using the results of TIC,
118
which were measured using a Shimadzu TOC-VCPH analyzer with a solid sample module
119
SSM-5000 at the Institute of Tibetan Plateau Research, Chinese Academy of Sciences. A total of
120
20 samples containing dolomite from 20 lakes (Table 1) on the central and western TP were taken
121
for analyses of the crystal morphology. Scanning electron microscope (SEM) observations with an
122
accompanying energy dispersive spectroscopy (EDS) was performed in the SINOPEC Research
123
Institute of Petroleum Processing. The MgCO3 mole% of dolomite was calculated using the
124
following formula:
125
NMgCO3 mol% =-333.33d 104 +1011.99,
126
where d 104 is the interplanar spacing which can be calculated based on the 2θ value of the (104)
127
face’s diffraction peak from the XRD profiles (Solotchina and Solotchin, 2014).
128
A total of 20 carbon and oxygen isotope sediment samples were performed on an IsoPrime
129
100 mass spectrometer equipped with a MultiPrep system at the Institute of Earth Environment,
130
Chinese Academy of Sciences. The international standard NBS19 and the laboratory standard TB1
131
were included in the analysis to check for homogeneity and reproducibility of results. The
132
analytical results for the isotope values are presented using the δ-notion relative to Vienna Pee Dee
133
Belemnite (VPDB). The standard deviations of both δ13C and δ18O were < 1‰ (2σ).
134
Water samples were filtered and diluted based on the salinity prior to water chemistry analysis.
135
Concentrations of major cations (Na+, K+, Ca+, Mg2+) and anions (Cl- and SO42-) were determined
136
using an ICS-2000 ion chromatography system (DIONEX Company, USA). HCO3- concentrations
137
were estimated by the balance between cations and anions (Wang et al., 2010b). Oxygen isotope
138
analyses of lake water from six lakes were prepared and analyzed using a Wavelength Scanning
139
Cavity Ring-down Spectrometer (WS-CRDS, Picarro-L2120i). Each sample was measured six
140
times and the first three analyses were not used to avoid memory effects from previous samples.
141
All oxygen isotope compositions are reported in standard δ-notion relative to Standard Mean
142
Ocean Water (SMOW). The standard results showed that the external precision of δ18O analysis
143
was better than 0.1‰. All analyses were performed at the Institute of Tibetan Plateau Research,
144
Chinese Academy of Sciences.
145
4 Results
146
4.1 Mineral composition of the surface lake sediments
147
The XRD results showed that the mineralogy of the surface lake sediments was characterized
148
mainly by carbonate, clay, and detrital minerals. Carbonate contents ranged from 18.89‰ to
149
60.71‰, with a mean value of 43.24‰. Carbonates consisted of calcite (CaCO3), aragonite
150
(CaCO3), dolomite (CaMg(CO3)2), and trace hydromagnesite [Mg5(CO3)4(OH)2·H2O]. Ankerite
151
[Ca(Fe, Mg)(CO3)2] only appeared in Serling Co. Clay and detrital minerals were composed of
152
illite and chlorite, quartz, and feldspar, respectively. The XRD patterns of carbonate minerals are
153
shown in Fig. 2.
154
Dolomite contents varied from 3‒10%, with a mean of 5.1% (Table 1). The MgCO3 content in
155
dolomite was approximately 45.10‒50.60% as shown by the XRD analysis (d 104 = 2.8842‒
156
2.9007 Å). Generally, the mol% MgCO3 in Ca-dolomite was > 40%, whereas the MgCO3 content
157
in stoichiometric dolomite is > 50% (Van Lith et al., 2002). Therefore, most dolomite in the
158
surface lake sediments was Ca-dolomite, except for that from Sumxi Co and Dong Co (MgCO3
159
contents of 50.13% and 50.60%, respectively).
160
4.2 Morphology of carbonate minerals
161
Calcite crystals appeared in the form of rhombohedral or blocky crystals with the grains size
162
< 15 µm as shown by SEM observations (Fig. 3a, b). The low magnesium peak in the EDS chart
163
of calcite crystals indicated that these crystals were magnesium carbonate (Mg-calcite). Aragonite
164
crystals showed prismatic morphology (~ 5 µm) (Fig. 3c). Dolomite exhibited two principal forms
165
in the surface lake sediments. Rhombic dolomite appeared as well-formed euhedral crystals (2‒10
166
µm) and was mostly scattered in sediments (Fig. 3d, e). Sub-spherical and spherical dolomite
167
occurred either as dispersed crystals or were attached to the surface of other minerals with grains <
168
2 µm (Fig. 3f, g) and aggregated into clusters (Fig. 3h, i). In some lakes, dolomite was not
169
identified under SEM due to the low content or the inhomogeneity of the mineral distribution.
170
4.3 Chemistry of lake water
171
The water chemistry of dolomitic lakes is quite different from those of lakes without
172
dolomite, especially in terms of Mg/Ca ratio, salinity, and SO42- concentration (Table 2). The
173
salinity of the dolomitic lake water ranged from 0.11 to 135 g/L, with a mean of 25.87 g/L (Fig.
174
4a); and the pH ranged from 7.61 to 10.56, with a mean of 9.02 (Fig. 4b). The major ionic
175
composition showed distinct variability in lake water. For example, the concentrations of K+, Na+,
176
Mg2+, and Ca2+ cations were 0.88‒4853.58 mg/L, 5.85‒74165.60 mg/L, 0‒10690.26 mg/L, and 1‒
177
5172.51 mg/L, respectively. The concentrations of Cl-, SO42-, and HCO3- anions also had large
178
ranges (0‒128777.8 mg/L, 10.36‒51160.79 mg/L, and 0‒13429.05 mg/L, respectively). The
179
resulting Mg/Ca (in mol) ratios ranged between 0.04‒4519.54, with an average of 535.94 (Table 2,
180
3).
181
4.4 Isotopic composition of lake water and sediments
182
The oxygen isotope results of water taken from six lakes showed that all the δ18OSMOW values
183
were negative, spanning a wide range of -4.12‰ to -0.22‰. The δ13CPDB values for carbonates are
184
shown in Fig. 5 (plotted against δ18O) and ranged from 1.7‒8.92‰. The δ18OPDB values for
185
carbonates ranged from -7.82‰ to 0.54‰, and on the SMOW scale were 22.85‰ to 31.47‰.
186
These δ18O values are very similar to those reported previously for Qinghai Lake (from -9.00‰ to
187
-1.59‰, Deng et al., 2010), and fall into the field of normal sedimentary rocks (Zheng and Chen,
188
2000). There was a significantly positive linear relationship between δ13CPDB and δ18OPDB (r =
189
0.70, P < 0.01).
190
5 Discussion
191
The exposed rocks in the study area are mainly composed of clastic, metamorphic, and igneous
192
rocks, however, a few limestone and marine carbonatites are also distributed around some lakes,
193
such as Serling Co, Gyeze Caka, Lagkor Co and Bura Co (Fig 1a, b, c, and f). To preclude the
194
presence of detrital carbonate, all sediment samples of the < 40 µm fraction were selected for
195
analyses, which are considered authigenic carbonate (Fontes et al., 1996; Jiang and Liu, 2007).
196
Calcite appeared as rhombohedral or blocky crystals, aragonite appeared as prismatic crystals, and
197
dolomite appeared as rhombic or sub-spherical and spherical crystals, which were similar in
198
morphology to those formed from natural environments and precipitation experiments and argue
199
against a detrital origin (Talbot and Kelts, 1986; Vasconcelos and McKenzie, 1997; Jiang and Liu,
200
2007; Gu et al., 2015). The close relationships between the δ18O of bulk carbonate, calcite, and
201
dolomite and environment indictors (the δ18O of lake water, temperature, salinity,
202
precipitation/evaporation, altitude and latitude) in typical lakes on the Tibetan Plateau also
203
indicate that carbonate minerals are formed directly through chemical precipitation in lake water
204
(Wang et al., 2008). In addition, the cold climate on the Tibetan Plateau leads to weak physical
205
weathering and a limited contribution of detrital carbonate in lake sediments (Li and Kang, 2007).
206
This evidence suggests that the contribution of detrital carbonate can be ignored in most lakes.
207
5.1 Environmental factors affecting the dolomite formation
208
Under low temperature and pressure conditions, there are three main kinetic inhibitors of
209
dolomite formation: (1) the high hydration energy of Mg2+ ions (Lippmann, 1973); (2) the
210
extremely low concentration and activity of CO32- (Garrels and Thompson, 1962), and; (3) the
211
presence of SO42-. Previous studies have suggested that dolomite precipitation is governed by
212
geochemical factors, such as the Mg/Ca ratio, salinity (Folk and Land, 1975; Sibley et al., 1987;
213
Deckker and Last, 1988; Van Lith et al., 2002), pH (Slaughter and Hill, 1991; Hammes and
214
Verstraete, 2002), and the SO42- concentration (Baker and Kastner, 1981; Vasconcelos and
215
McKenzie, 1997). Compared with non-dolomitic lakes, dolomitic lakes are characterized by high
216
Mg/Ca ratio, high salinity, high sulfate concentration, and slightly higher pH (Table 2).
217
In general, a high Mg/Ca ratio and high salinity are regarded as important factors affecting
218
dolomite precipitation (Müller et al., 1972; Wright and Oren, 2005). Mg2+ ions are strongly
219
hydrated by polar water shells, which makes it difficult for them to enter the dolomite lattice, and
220
hinders dolomite precipitation from supersaturated solution (Lippmann, 1973). A high Mg/Ca ratio
221
(i.e. high Mg2+ concentrations of hypersaline waters) means that the hydration barrier is more
222
easily overcome (Warren, 2000). High salinity favors hydrated ions losing their water shell and
223
decreasing the hydration energy of Mg2+, due to direct binding as a function of ion distribution,
224
repartition of fixed charges, and the anions present (Wright and Oren, 2005). In addition, above
225
dolomite saturation, increasing salinity clearly favors dolomite formation because supersaturation
226
increases (Machel and Mountjoy, 1986).
227
Previous studies suggested that hypersaline dolomite begins to precipitate only when the
228
Mg/Ca weight ratio exceeds 5:1 to 10:1(Kinsman, 1966; Folk and Land, 1975). In the lakes of the
229
East Asian monsoon region, a Mg/Ca molar ratio of 1.7 is the threshold for dolomite formation
230
(Gu et al., 2015). With methanogenic metabolic activity, dolomite can precipitate in waters with
231
very low Mg/Ca molar ratios (< 1) (Roberts et al., 2004; Kenward et al., 2009). In our study, only
232
the Mg/Ca (in mol) ratios of Hongshan Lake and Matou Lake were < 1.
233
Our investigation revealed a positive linear relationship between salinity and the Mg/Ca ratio
234
(r = 0.56, P < 0.05; Fig. 4a). The high Mg/Ca ratio and high salinity of most lakes would favor
235
dolomite precipitation. However, in some cases, the salinity was as low as 0.11 mg/L in the
236
freshwater lake where dolomite formed. Although high salinity favors the breakdown of hydration
237
shells, Folk and Land (1975) argued that high salinity results in a more rapid crystallization rate
238
and an increase in the number of interfering ions (e.g. Ca2+), which is not conducive to dolomite
239
formation because of the precise Ca-Mg ordering required. Higher salinity also means higher
240
sodium, which means less free CO32- due to the formation of soluble NaCO30 (Garrels and
241
Thompson, 1962). Taking Nganggun Co as an example, although the salinity was only 1.89 g/L,
242
the Mg2+ concentration in the lake water can reach more than ten times the Ca2+ concentration,
243
which is beneficial for dolomite formation. Therefore, a high Mg/Ca ratio may be more important
244
than high salinity for dolomite precipitation in Tibetan lakes.
245
High pH is important because dolomitization is partly controlled by the activity of carbonate
246
anions in solution. This activity increases the CO32- concentration relative to HCO3- concentration
247
with increased pH (Slaughter and Hill, 1991). pH was not significantly correlated with the Mg/Ca
248
ratio in this study (Fig. 4b), which was revealed by a low correlation coefficient of -0.18 (P =
249
0.454, n = 20). Compared with Qinghai Lake (pH = 9.3, Deng et al., 2010), Coorong Region (pH
250
= 6.82‒9.11, Wright and Wacey, 2005; Wacey et al., 2007), and Lagoa Vermelha (pH = 8.0‒8.5,
251
Vasconcelos and McKenzie, 1997), the pH range of lakes in this study is similar to that of
252
previous studies (Table 3). The high pH (=10.56) of Zhari Namco may be caused by the
253
consumption of large amounts of CO2 in the upper lake due to photosynthesis by aquatic
254
organisms (Wang et al., 2010a).
255
SO42- is regarded as a dolomite formation inhibitor (Baker and Kastner, 1981; Kastner, 1984).
256
Even very low SO42- concentrations can cause the formation of strongly bonded neutral MgSO40
257
ion pairs, which significantly increase Mg2+ solubility (Slaughter and Hill, 1991; Wright and Oren,
258
2005). In contrast, Van Lith et al. (2002) considered that during dolomite precipitation, SO42- does
259
not decrease to zero, therefore the accompanying alkalization due to bacterial SO42- reduction is
260
more important for dolomite formation than the low SO42- concentrations. As for pH, there was no
261
simple linear relationship between the SO42- concentration and the Mg/Ca ratio (Fig. 4c), however
262
it is closely related to the water chemistry of lake water on the Tibetan Plateau (Zheng and Liu,
263
2010). Our results suggest that SO42- need not be eliminated in order to precipitate dolomite.
264
Compared with lakes where microbially-mediated dolomite formation, the SO42- concentration in
265
this study is close to the ranges reported in previous studies (Table 3). In the microbial dolomite
266
model, an abundant and continuous SO42- supply is also considered a necessary factor for
267
maintaining microbial activity to produce dolomite (Vasconcelos and McKenzie, 1997).
268
Many experiments and field cases have demonstrated that metabolic activities of
269
microorganisms, such as SO42--reducing bacteria (SRB) (Compton, 1988; Wright, 1999;
270
Warthmann et al., 2000; Wright and Wacey, 2005), methanogenic Archaea (Kenward et al., 2009),
271
and halophilic bacteria (Sánchez-Román et al., 2009a; Deng et al., 2010) played key roles in
272
mediating low-temperature dolomite precipitation. Metabolic activities of microbial communities
273
in lakes can raise the pH and carbonate and magnesium ion concentrations, and reduce the SO42-
274
concentration, thereby forming a microenvironment that help overcome the kinetic inhibitors of
275
dolomite precipitation (Wright, 1999; Wright and Wacey, 2005). Microbial cell surfaces and
276
excreted extracellular polymeric substances (EPS) carry a net negative electric charge, easily
277
adsorbing metal cations (Mg2+ and Ca2+) on its surface and forming a microenvironment on its
278
surface that facilitates dolomite precipitation (Sánchez-Román et al., 2008). It has also been
279
suggested that microbes such as bacteria, nano-bacteria, and EPS can act as nuclei for dolomite
280
precipitation (Warthmann et al., 2000; Bontognali et al., 2008). Therefore, ubiquitous
281
microorganisms may play an important role in the dolomite formation.
282
5.2 Potential mechanism of dolomite formation on the Tibetan Plateau
283
5.2.1 The role of evaporation
284
Modern authentic dolomite was observed mostly in high salinity environments such as
285
lagoons and saline lakes, and was usually accompanied by strong evaporation. Jiang and Liu
286
(2007) suggested that in the lake environment, the primary dolomite can be used as a proxy for
287
climate drying. Therefore, these dolomite formations have been interpreted as typically
288
evaporative in origin (Hsü and Siegenthaler, 1969; Deckker and Last, 1988; Warren, 1988). In
289
hypersaline environments, crystallization of minerals is commonly rapid and accompanied by high
290
concentrations of competing ions, and it is difficult for Ca and Mg to segregate into monolayers to
291
form stoichiometric dolomite (Folk and Land, 1975). In such conditions, dolomite is usually
292
highly disordered Ca- dolomite (Warren, 2000), as found in this study.
293
The positive correlation between δ18O and δ13C (r=0.7) in the carbonates suggests that
294
carbonates precipitated in lake water have long-term residence in closed lakes (Talbot, 1990). Arid
295
to semi-arid climates accelerate the continuous evaporation of lake water, along with the
296
precipitation of calcite and aragonite. Quantitative Ca2+ was consumed, increasing the Mg/Ca ratio.
297
With further evaporation, the death and degradation of organic components would provide large
298
amounts of magnesium for dolomite formation. The metabolism of organic matter by SRB or
299
other microbes would also elevate the pH values of lake water (Wright, 1999; Wacey et al., 2007).
300
However, evaporation alone could not explain dolomite formation in lakes with low Mg/Ca
301
ratios. Evaporation experiments of lake waters from Coorong Lake, a classic location for the study
302
of modern dolomite, show that aragonite is precipitated instead of dolomite (Warren, 1988; Wacey
303
et al., 2007). Lippmann (1973) argued that evaporation alone could not favor dolomitization
304
because CO32--free ions were decreasing with the increase of Mg/Ca ratios. In freshwater lakes,
305
the evaporation and concentration of lake water is weak and the Mg/Ca ratio of lake water is low
306
(< 3), which is insufficient to form the large amount of magnesium-rich brine that is required for
307
dolomite precipitation. Generally, freshwater dolomite formation typically appears in the mixture
308
of dissolved inorganic carbon (DIC)-rich meteoric water with magnesium-rich groundwater
309
(Kenward et al., 2009). Therefore, in addition to evaporation, for lakes on the Tibetan Plateau,
310
there should be other mechanisms involved in the formation of dolomite. The most likely
311
mechanism would be that of microbial activity.
312
5.2.2 The role of microbial activity
313
Under the regulation of microbial metabolism, dolomite can be formed not only in high
314
salinity environments (Vasconcelos and McKenzie, 1997; Wright, 1999; Van Lith et al., 2002), but
315
also in low salinity lake waters (Deng et al., 2010) or even fresh water (Roberts et al., 2004;
316
Kenward et al., 2009).
317
In this study, dolomite was present in two distinct forms. One type is characterized by single,
318
well-formed euhedral and rhombic crystals range in length from 2‒10 µm. These dolomite rhombs
319
with micron grain size are found mostly in modern dolomite as a replacement of aragonite or
320
calcite, such as the Qatar Sabkha (Illing and Taylor, 1993), and the permanent hypersaline
321
environment of Lake Hayward (Rosen and Coshell, 1992). In high-temperature (≥ 175 ℃)
322
synthesis experiments, rhombic dolomite formed by dolomitization of CaCO3 can nucleate on the
323
corners of calcite reactants (Sibley et al., 1987). However, replacement dolomite appears to
324
require long reaction times (≥ 104 yr) at low temperatures (Hardie, 1987). Although Tibetan lakes
325
are generally characterized by a low recent sedimentation rate, such as Serling Co (0.25 mm a-1,
326
Gu et al., 1994), Nam Co (0.43‒0.98 mm a-1, Wang et al., 2011), Mapam Yumco (0.31 mm a-1,
327
Wang et al., 2013), Taro Co (0.49‒0.58 mm a-1, Ma et al., 2014) and Tangra Yumco (0.18‒0.64
328
mm a-1, Wang et al., 2017), lake sediments accumulated faster than the time required for formation
329
of replacement dolomite. Furthermore, no evidence of replacement textures from petrographic
330
analysis suggests that dolomite is most likely directly deposited from dolomite-saturated brines.
331
The widespread occurrence of primary-dolomite grains in the Upper Cretaceous sandstones of
332
the Western Interior and Alaska are single rhombic crystals (< 0.3 mm) formed within the
333
depositional basin prior to final settling-down and burial of the sediment (Sabins, 1962).
334
Experiments have shown that rhombic dolomite can be induced by L. sphaericus at 30 °C under
335
20 MPa pressure and micro-aerobic conditions (Song et al., 2014). Dolomite rhombs also occur in
336
pelagic, organic-rich sediments with active degradation of organic matter (Baker and Kastner,
337
1981; Compton, 1988). Huang et al. (1997) suggests that the primary rhombic dolomite in
338
algae-rich dolostones is characterized by crystal forms precipitated and deposited directly from
339
seawater. The euhedral-subhedral micritic dolomite crystals from Junggar Basin are rapidly
340
crystallized in the contemporaneous-penecontemporaneous period and are closely related to the
341
metabolic activities of methanogens (Zhang et al., 2018). Microbial activity may also be involved
342
in the formation of dolomite rhombs in this study.
343
A series of growth experiments have been carried out with selected bacterial cultured from
344
natural samples at low temperatures (Vasconcelos et al., 1995; Sánchez-Román et al., 2008;
345
Sánchez-Román et al., 2009b; Warthmann et al., 2000; Deng et al., 2010). The results showed that
346
under the mediation of microbial activity, dolomite formed in lakes has the same morphology as
347
dolomite synthesized in the laboratory, which manifested as dumbbell, sub-spherical, and
348
spherical crystals.
349
Under SEM, dolomite appeared as sub-spherical and spherical crystals, dispersed in
350
sediments, or adhered to the surface of feldspar to form a knobbly dolomite coating (Fig. 3f, g).
351
Individual dolomite crystals were fairly uniform in size, and < 2 µm in diameter. In some lakes,
352
sub-spherical or spherical crystals were aggregated into ~20 µm assemblages (Fig. 3h, i). The
353
characteristics of these dolomites are very similar to those of microbial origin reported in Coorong
354
Lake (Wright, 1999), Lagoa Vermelha (Vasconcelos et al., 1995; Vasconcelos and McKenzie, 1997)
355
and culture experiments (Vasconcelos et al., 1995; Sánchez-Román et al., 2008). The occurrence
356
of spherical dolomites even in low-salinity lakes (such as Serling Co and Co Ngoin) further proves
357
the possibility of microbial origin. This discovery is consistent with the precipitation of biogenic
358
dolomite in freshwater from both field and laboratory experiments and suggests a pivotal role of
359
microbial processes in dolomite formation across a wide range of environmental conditions
360
(Roberts et al., 2004).
361
For equilibrium mineral precipitation, δ18O of lacustrine carbonate are controlled by the
362
temperature and isotopic composition of the lake water δ18O (Leng and Marshall, 2004). Using
363
δ18O-lakewater and δ18O-dolomite results, the range of precipitation temperature of dolomite can
364
be calculated, assuming equilibrium. The δ18O of carbonate, however, is a result of the mixture of
365
the oxygen isotopic compositions in different carbonate minerals, which is expressed as follows
366
(Wang et al., 2008): =
×
+
×
+
× + ⋯,
367
where δ18OA, δ18OB, and δ18OC are δ18Odolomite, δ18Ocalcite, and δ18Oaragonite, respectively; and a, b,
368
and c stand for the proportion of dolomite, calcite, and aragonite to the bulk carbonate. Previous
369
studies have suggested that under the same conditions, δ18O of aragonite and dolomite are about
370
0.6‰ and 3‰ more positive than calcite, respectively (Land, 1980). Based on this relationship,
371
six lakes which precipitated dolomite, calcite, and aragonite were used to calculate the isotope of
372
dolomite (Fig. 6).
373 374 375
Based on high-temperature experiments, Northrop and Clayton (1966) proposed the following equilibrium fractionation relationship for dolomite: 1000 ln
! "#$% − ' $%( = 3.20 × 10, - ./ − 1.50. (1)
376
Here, T is the temperature in Kelvin, and α is the fractionation factor for oxygen isotope between
377
dolomite and lake water. Due to the failure of low temperature synthetic dolomite experiments,
378
Vasconcelos et al. (2005) proposed a new oxygen isotope fractionation equation based on
379
microbial experiments at a low temperature:
380
1000 ln
! "#$% − ' $%( = 2.73 × 10, - ./ + 0.26. (2)
381
Since the field sampling was conducted in late summer and autumn, the water temperature of
382
the lake had a large range (3‒17 °C). Using the equation obtained from the high-temperature
383
experiment (Eq. 1), the temperature range of dolomite formation was calculated to be 21‒58 ℃
384
(Fig. 6a). However, according to the microbial-mediated low-temperature equation (Eq. 2), the
385
temperature favoring dolomite precipitation is approximately 10 ℃ in most lakes and 42 ℃ in
386
Hongshan Lake (Fig. 6b). Compared to the temperature range derived by Eq. 1, the range
387
calculated using Eq. 2 is closer to the temperature measured in the field, further suggesting that
388
microbial mediation is likely mechanism for dolomite formation.
389
In the lakes of the Tibetan Plateau, SO42--reducing bacteria (Deng et al., 2010; Yang et al.,
390
2013), halophilic bacteria (Jiang et al., 2006; Zhu et al., 2017), and cyanobacteria (Dong et al.,
391
2006 Liu et al., 2009; Xing et al., 2009; Liu et al., 2010;) are common microbial communities that
392
can induce dolomite formation. The abundance of the dsrB gene shared by sulfate reducing
393
bacteria ranges from 1. 71 × 108 to 1. 55 × 109 copies per gram of sediments from six lakes on the
394
Tibetan Plateau (Yang et al., 2013). Anaerobic SRB and aerobic halophilic bacteria-mediated
395
dolomite precipitation are not only active in hypersaline waters, but have also been confirmed to
396
have implications for dolomite formation in slightly saline lake environments such as Qinghai
397
Lake (Deng et al., 2010; Wacey et al., 2007). In lakes of the eastern and southern Tibetan Plateau,
398
cyanobacteria are one of the main sources of biological productivity (Dong et al., 2006; Liu et al.,
399
2009; Xing et al., 2009), which are characterized by concentrating Mg preferentially in their
400
sheaths compared to the ambient water throughout their growth, and releasing Mg rapidly during
401
late evaporation (Wright, 1999). Although the mechanisms of microbial dolomite formation
402
appear diverse, different microorganism species exhibit the ability to alter the water chemistry and
403
overcome kinetic inhibitors of dolomite precipitation.
404
It is plausible that in Tibetan Plateau lake sediments, dolomite precipitation in different lakes
405
may be induced by different microorganism and microbial dolomite patterns, such as the sulfate
406
reduction mode, bacterial aerobic oxidation mode, and methanogenesis mode (Jiang et al., 2017).
407
Further studies will demonstrate the above conclusions using sulfur isotopes of lake water and
408
sediments. In addition, it is necessary to strengthen the study of microorganisms in the modern
409
lakes of the Tibetan Plateau to better understand microbial roles in dolomite precipitation.
410 411
6 Conclusion
412
Modern dolomite formed in the sediment of 20 lakes on the central and western Tibetan
413
Plateau. The dolomite content ranged from 3% to 10%, with a mean of 5%. Dolomite exhibited
414
two principal forms in the surface lake sediments: (1) euhedral 2‒10 µm rhombic dolomite
415
(mostly scattered in sediments), and; (2) sub-spherical and spherical dolomite occurring either as
416
dispersed crystals or aggregated into dolomite clusters. The chemical compositions of lake water
417
affecting dolomite precipitation showed high variability, with the salinity ranging from 0.11 g/L to
418
135 g/L, the Mg/Ca molar ratio ranging from 0.04 to 4519.54, the pH ranging from 7.61 to 10.56,
419
and SO42- concentrations of 10.36‒51160.79 mg/L. The results showed a good linear relationship
420
between salinity and the Mg/Ca ratio, while the pH and SO42- concentration were not significantly
421
correlated with the Mg/Ca ratio. The measured δ18O and δ13CPDB values of carbonates in sediment
422
ranged from −7.82‰ to 0.54‰, and from 1.7‰ to 8.92‰, respectively.
423
Environmental factors, such as high Mg/Ca ratio, high salinity, high sulfate concentration,
424
and high pH favor dolomite precipitation. Evaporation and microbial activity play important roles
425
in dolomite formation on the central and western Tibetan Plateau. The positive relationship
426
between δ18O and δ13C indicates that dolomite formed in lakes has a long residence time and
427
undergoes continuous evaporation. Dolomite crystals were identical both in shape and size to
428
those of microbial dolomite discovered in modern lake environments and culture experiments.
429
This was also confirmed by the calculations using temperature-dependent fractionation factors for
430
oxygen isotopes between dolomite and lake water, which indicated the contribution of microbial
431
activities to dolomite precipitation in these lakes.
432 433
Acknowledgments
434
This work was supported by the Strategic Priority Research Program of Chinese Academy of
435
Sciences (XDA20020100), the National Natural Science Foundation of China (41807445), and the
436
Key Project of National Natural Science Foundation of China (41831177). We would like to thank
437
Ping Peng, Jianting Ju, Xiao Lin, Xing Hu, Baojin Qiao, Lei Huang, Chong Liu, Teng Xu, Hao
438
Chen and Jinlei Kai for their participation in the field. We also thank Editage (www.editage.cn) for
439
English language editing.
440
References
441
Adams, J.E., Rhodes, M.L.J.A.B., 1960. Dolomitization by seepage refluxion. AAPG Bull. 44,
442 443
1912-1920. Aharon, P., Kolodny, Y., Sass, E., 1977. Recent Hot Brine Dolomitization in the "Solar Lake,"
444 445 446 447 448
Gulf of Elat, Isotopici, Chemical, and Mineralogical Study. J. Geol. 85, 27-48. Badiozamani, K., 1973. The Dorag dolomitization model application to the middle Ordovician of Wisconsin. J. Sediment. Res. 43, 965-984. Baker, P.A., Kastner, M., 1981. Constraints on the Formation of Sedimentary Dolomite. Science 213, 214-216.
449
Bontognali, T.R.R., Vasconcelos, C., Warthmann, R.J., Bernasconi, S.M., Dupraz, C., Strohmenger,
450
C.J., McKenzie, J.A., 2010. Dolomite formation within microbial mats in the coastal sabkha
451
of Abu Dhabi (United Arab Emirates). Sedimentology 57, 824-844.
452
Bontognali, T.R.R., Vasconcelos, C., Warthmann, R.J., Dupraz, C., Bernasconi, S.M., McKenzie,
453
J.A., 2008. Microbes produce nanobacteria-like structures, avoiding cell entombment.
454
Geology 36, 663-666 .
455
BrÉHÉRet, J.-G., Fourmont, A., Macaire, J.-J., NÉGrel, P., 2008. Microbially mediated carbonates
456
in the Holocene deposits from Sarliève, a small ancient lake of the French Massif Central,
457
testify to the evolution of a restricted environment. Sedimentology 55, 557-578.
458
Bureau of Geology and Mineral resources of Xizang Autonomous Regions., 1993. Regional
459
geology of Xingzang(Tibet) Autonomous Region. Geological Publishing House, Beijing (in
460
Chinese).
461
Chellie S. Teal, S.J. Mazzullo, Bischoff, W.D., 2000. Dolomitization of Holocene Shallow-Marine
462
Deposits Mediated by Sulfate Reduction and Methanogenesis in Normal-Salinity Seawater,
463
Northern Belize. J. Sediment. Res. 70, 649–663.
464
Chen, Y., Yang, K., He, J., Qin, J., Shi, J., Du, J., He, Q., 2011. Improving land surface
465
temperature modeling for dry land of China (1984–2012). J. Geophys. Res, 16 (D20). http://
466
dx.doi.org/10.1029/2011JD015921.
467
Colson, J., Cojan, I., 1996. Groundwater dolocretes in a lake℃marginal environment: an
468
alternative model for dolocrete formation in continental settings (Danian of the Provence
469
Basin, France). Sedimentology 43, 175-188.
470 471 472 473
Compton, J.S., 1988. Degree of supersaturation and precipitation of organogenic dolomite. Geology 16, 318-321. Deckker, P.D., Last, W.M., 1988. Modern dolomite deposition in continental, saline lakes, western Victoria, Australia. Geology 16, 29-32.
474
Deng, S., Dong, H., Lv, G., Jiang, H., Yu, B., Bishop, M.E., 2010. Microbial dolomite
475
precipitation using sulfate reducing and halophilic bacteria: Results from Qinghai Lake,
476
Tibetan Plateau, NW China. Chem. Geol. 278, 151-159.
477
Dong, H., Zhang, G., Jiang, H., Yu, B., Chapman, L.R., Lucas, C.R., Fields, M.W., 2006.
478
Microbial diversity in sediments of saline Qinghai Lake, China: linking geochemical controls
479
to microbial ecology. Microb. Ecol. 51, 65-82.
480
Fairbridge, R.W., 1957. The dolomite question. In: LeBlanc, R.J., Breeding, J.G. (Eds.), Regional
481
Aspects of Carbonate Deposition: A Symposium. Tulsa, Oklahoma, USA. 5. Society of
482
Economic Paleontologists and Mineralogists, SEPM Special Publication, pp. 125–178.
483 484
Folk, R.L., Land, L.S., 1975. Mg/Ca ratio and salinity: two controls over crystallization of dolomite. AAPG Bull. 59, 60-68.
485
Fontes, J.C., Gasse, F., Gibert, E., 1996. Holocene environmental changes in Lake Bangong basin
486
(Western Tibet). Part 1: Chronology and stable isotopes of carbonates of a Holocene
487
lacustrine core. . Palaeogeogr. Palaeoclimatol. Palaeoecol.120, 25-47.
488 489
Garrels, R.M., Thompson, M.E., 1962. A chemical model for seawater at 25°C and one atmosphere total pressure. Am. J. Science 260, 57-66.
490
Gu, N., Jiang, W., Wang, L., Zhang, E., Yang, S., Xiong, S., 2015. Rainfall thresholds for the
491
precipitation of carbonate and evaporite minerals in modern lakes in northern China.
492
Geophys. Res. Lett. 42, 5895-5901.
493
Gu, Z., Liu, J., Yuan, B., Liu, T., Zhang, G., 1994. Lacustrine authigenic deposition expressiveof
494
environment and the sediment recordfrom Siling co, Xizang (Tibet), China. Quat. Sci. 2,
495
162-174 (in Chinese with English abstract).
496 497 498 499 500 501 502 503
Guan, Z., Chen, C., Qu, Y., Fan, Y., Zhang, Y., Chen, Z., Bao, S., Zu, Y., He, X., Zhang, M., 1980. Rivers and lakes of Xizang. Science Press, Beijing (in Chinese with English abstract). Hammes, F., Verstraete, W., 2002. Key roles of pH and calcium metabolism in microbial carbonate precipitation. Rev. Environ. Sci. Bio. 1, 3-7. Hardie, L.A., 1987. Dolomitization: a critical view of some current views. J. Sediment. Res. 57, 166-183. Hsü, K.J., Siegenthaler, C., 1969. Preliminary experiments on hydrodynamic movement induced by evaporation and their bearing on the dolomite problem. Sedimentology 12, 11-25.
504 505
Huang, Z., Yang, S., Chen, Z., 1997. Mineralogical correlation between primary and replacement dolomites. Sci. China. Ser. D. 40, 91-98.
506
Illing, L.V., Taylor, J.C., 1993. Penecontemporaneous dolomitization in Sabkha Faishakh, Qatar;
507
evidence from changes in the chemistry of the interstitial brines. J. Sediment. Res. 63,
508
1042-1048.
509 510
Institute of Geography, 1990. Map of the Qinghai-Tibetan Plateau. Science Press, Beijing (in Chinese)
511
Jiang, H., Dong, H., Zhang, G., Yu, B., Chapman, L.R., Fields, M.W., 2006. Microbial diversity in
512
water and sediment of Lake Chaka, an athalassohaline lake in northwestern China. Appl.
513
Environ. Microb. 72, 3832-3845.
514 515
Jiang, Q., Liu, B., Guo, R., Gao, X., Li, Y., Zhang, S., 2017. Microbial mechanism of lacustrine primary dolomite. J. Palaeogeog. 19, 257-269 (in Chinese with English abstract).
516
Jiang, W.Y., Liu, T.S., 2007. Timing and spatial distribution of mid-Holocene drying over northern
517
China: Response to a southeastward retreat of the East Asian Monsoon. J. Geophys. Res. 112.
518
Jose Carballo, Lynton S. Land, Miser, D.E., 1987. Holocene dolomitization of supratidal
519
sediments by active tidal pumping, Sugarloaf Key, Florida. J. Sediment. Res. 57, 153–165.
520
Kastner, M., 1984. Sedimentology: Control of dolomite formation. Nature 311, 410-411.
521
Kenward, P.A., Goldstein, R.H., Gonzalez, L.A., Roberts, J.A., 2009. Precipitation of
522
low-temperature dolomite from an anaerobic microbial consortium: the role of methanogenic
523
Archaea. Geobiology 7, 556-565.
524
Kinsman, D.J.J., 1966. Gypsum and anhydrite of recent age, Trucial Coast, Persian Gulf. In Rau,
525
J.L. (ed.), Proceedings of the Second Symposium on Salt. Northern Ohio Geological Society,
526
Cleveland, pp. 302-326.
527
Krauskopf, K.B., Bird, D.K., 1995. Introduction to geochemistry. McGraw-Hill New York.
528
Land, L.S., 1980. The isotopic and trace element geochemistry of dolomite: the state of the art. In:
529
Zenger, D.H., Dunham, J.B., Ethington, R.L. (Eds.), Concepts and Models of Dolomitization.
530
SEPM Special Publication, vol. 28, pp. 87-110.
531 532 533
Land, L.S., 1998. Failure to Precipitate Dolomite at 25 °C fromDilute Solution Despite 1000-Fold Oversaturation after 32 Years. Aquati. Geochem. 4, 361-368. Last, W.M., 1990. Lacustrine dolomite-an overview of modern, Holocene, and Pleistocene
534 535 536 537 538
occurrences. Earth-Sci. Rev. 27, 221-263. Leng, M.J., Marshall, J.D., 2004. Palaeoclimate interpretation of stable isotope data from lake sediment archives. Quat. Sci. Rev. 23, 811-831. Li, M., Kang, S., 2007. Responses of Lake Sediments to Paleoenvironmental and Paleoclimatic Changes in Tibetan Plateau. J. Salt. Lake. Res. 15, 63-72 (in Chinese with English abstract).
539
Lippmann, F., 1973. Sedimentary Carbonate Minerals. Springer-Verlag, New York.
540
Liu, X., Yao, T., Kang, S., Jiao, N., Zeng, Y., Liu, Y., 2010. Bacterial community of the largest
541
oligosaline lake, Namco on the Tibetan Plateau. Geomicrobiology Journal 27, 669-682.
542
Liu, Y., Yao, T., Zhu, L., Jiao, N., Liu, X., Zeng, Y., Jiang, H., 2009. Bacterial diversity of
543
freshwater alpine lake Puma Yumco on the Tibetan Plateau. Geomicrobiol. J. 26, 131-145.
544
Ma, Q., Zhu, L., Lü, X., Guo, Y., Ju, J., Wang, J., Wang, Y., Tang, L., 2014. Pollen-inferred
545
Holocene vegetation and climate histories in Taro Co, southwestern Tibetan Plateau. Chinese.
546
Sci. Bull. 59, 4101-4114.
547 548
Machel, H.G., Mountjoy, E.W., 1986. Chemistry and environments of dolomitization—a reappraisal. Earth-Sci. Rev. 23, 175-222.
549
Meister, P., Reyes, C., Beaumont, W., Rincon, M., Collins, L., Berelson, W., Stott, L., Corsetti, F.,
550
Nealson, K.H., 2011. Calcium and magnesium-limited dolomite precipitation at Deep Springs
551
Lake, California. Sedimentology 58, 1810-1830.
552 553 554 555 556 557 558 559 560 561 562 563
Müller, G., Irion, G., Förstner, U., 1972. Formation and diagenesis of inorganic Ca–Mg carbonates in the lacustrine environment. Naturwissenschaften 59, 158–164. Northrop, D.A., Clayton, R.N., 1966. Oxygen-isotope fractionations in systems containing dolomite. The J. Geol. 74, 174-196. Roberts, J.A., Bennett, P.C., González, L.A., Macpherson, G.L., Milliken, K.L., 2004. Microbial precipitation of dolomite in methanogenic groundwater. Geology 32, 277-280. Rosen, M.R., Coshell, L., 1992. A new location of Holocene dolomite formation, Lake Hayward, Western Australia. Sedimentology 39, 161-166. Rosen, M.R., Miser, D.E., Starcher, M.A., Warren, J.K., 1989. Formation of dolomite in the Coorong region, South Australia. Geochim. Cosmochim. Acta. 53, 661-669. Sabins, F.F., 1962. Grains of Detrital, Secondary, and Primary Dolomite from Cretaceous Strata of the Western Interior. Geol. Soc. Am. Bull. 73, 1183-1196.
564 565
Saller, A.H., 1984. Petrologic and geochemical constraints on the origin of subsurface dolomite, Enewetak Atoll An example of dolomitization by normal seawater. Geology 12, 217-220.
566
Sánchez-Román, M., McKenzie, J.A., de Luca Rebello Wagener, A., Rivadeneyra, M.A.,
567
Vasconcelos, C., 2009a. Presence of sulfate does not inhibit low-temperature dolomite
568
precipitation. Earth Planet. Sci. Lett. 285, 131-139.
569
Sánchez-Román, M., Vasconcelos, C., Schmid, T., Dittrich, M., McKenzie, J.A., Zenobi, R.,
570
Rivadeneyra, M.A., 2008. Aerobic microbial dolomite at the nanometer scale: Implications
571
for the geologic record. Geology 36. 879-882.
572
Sánchez-Román, M., Vasconcelos, C., Warthmann, R., Rivadeneyra, M.A., McKenzie, J.A., 2009b.
573
Microbial Dolomite Precipitation under Aerobic Conditions: Results from Brejo do Espinho
574
Lagoon (Brazil) and Culture Experiments. IAS Spec. Publ. 41, 167–178.
575 576 577 578 579 580
Sibley, D.F., Dedoes, R.E., Bartlett, T.R., 1987. Kinetics of dolomitization. Geology 15, 1112-1114. Slaughter, M., Hill, R.J., 1991. The influence of organic matter in organogenic dolomitization. J. Sediment. Res. 61, 296-303. Solotchina, È.P., Solotchin, P.A., 2014. Composition and structure of low-temperature natural carbonates of the calcite-dolomite series. J. Struct. Chem. 55, 779-785.
581
Song, Q., Xu, J., Zhang, Y., 2014. Dolomite precipitation mediated by Lysinibacillus sphaericus
582
and Sporosarcina psychrophila. Microb. China 41, 2155-2165 (in Chinese with English
583
abstract).
584 585 586 587
Talbot, M.R., 1990. A review of the palaeohydrological interpretation of carbon and oxygen isotopic ratios in primary lacustrine carbonates. Chem. Geol 80, 261-279. Talbot, M.R., Kelts, K., 1986. Primary and diagenetic carbonates in the anoxic sediments of Lake Bosumtwi, Ghana. Geology 14, 912-916.
588
Van Lith, Y., Vasconcelos, C., Warthmann, R., Martins, J.C.F., McKenzie, J.A., 2002. Bacterial
589
sulfate reduction and salinity two controls on dolomite precipitation in Lagoa Vermelha and
590
Brejo do Espinho (Brazil). Hydrobiologia 485, 35-49.
591
Vasconcelos, C., McKenzie, J.A., 1997. Microbial mediation of modern dolomite precipitation and
592
diagenesis under anoxic conditions (Lagoa Vermelha, Rio de Janeiro, Brazil). J. Sediment.
593
Res. 67, 378-390.
594
Vasconcelos, C., McKenzie, J.A., Bernasconi, S., Grujic, D., Tiens, A.J., 1995. Microbial
595
mediation as a possible mechanism for natural dolomite formation at low temperatures.
596
Nature 377, 220-222.
597
Vasconcelos, C., McKenzie, J.A., Warthmann, R., Bernasconi, S.M., 2005. Calibration of the δ18O
598
paleothermometer for dolomite precipitated in microbial cultures and natural environments.
599
Geology 33, 317-320.
600 601 602 603 604 605
Wacey, D., Wright, D.T., Boyce, A.J., 2007. A stable isotope study of microbial dolomite formation in the Coorong Region, South Australia. Chem. Geol. 244, 155-174. Wan, W., Long, D., Hong, Y., Ma, Y., Yuan, Y., Xiao, P., Duan, H., Han, Z., Gu, X., 2016. A lake data set for the Tibetan Plateau from the 1960s, 2005, and 2014. Sci. Data. 3, 160039. Wang, J., Peng, P., Ma, Q., Zhu, L., 2010a. Modern limnological features of Tangra Yumco and Zhari Namco, Tibetan Plateau. J. Lake. Sci. 22, 629-632 (in Chinese with English abstract).
606
Wang, J., Peng, P., Ma, Q., Zhu, L., 2013. Investigation of water depth, water quality and modern
607
sedimentation rate in Mapam Yumco and La'ang Co, Tibet. J. Lake. Sci. 25, 609-616 (in
608
Chinese with English abstract).
609
Wang, J., Zhu, L., Wang, Y., Gao, S., Daut, G., 2011. Spatial variability of recent sedimentation
610
rate and variations in the past 60 years in Nam Co, Tibetan Plateau, China. Quat. Sci. 31,
611
535-543 (in Chinese with English abstract).
612
Wang, J., Zhu, L., Wang, Y., Ju, J., Xie, M., Daut, G., 2010b. Comparisons between the chemical
613
compositions of lake water, inflowing river water, and lake sediment in Nam Co, central
614
Tibetan Plateau, China and their controlling mechanisms. J. Great. Lakes. Res. 36, 587-595.
615
Wang, J., Zhu, L., Wang, Y., Peng, P., Ma, Q., Haberzettl, T., Kasper, T., Matsunaka, T., Nakamura,
616
T., 2017. Variability of the 14C reservoir effects in Lake Tangra Yumco, Central Tibet (China),
617
determined from recent sedimentation rates and dating of plant fossils. Quat. Int. 430, 3-11.
618
Wang, N., Liu, W.G., Xu, L.M., An, Z.S., 2008. Oxygen isotopic compositions of carbonates of
619
modern surface lacustrine sediments and their affecting factors in Tibet Plateau. Quat. Sci. 28,
620
591-600 (in Chinese with English abstract).
621 622 623
Warren, J., 2000. Dolomite: occurrence, evolution and economically important associations. Earth-Sci. Rev. 52, 1-81. Warren, J.K., 1988. Sedimentology of coorong dolomite in the Salt Creek region, South Australia.
624 625 626 627 628
Carbonate. Evaporite. 3, 175-199. Warren, J.K., 1990. Sedimentology and Mineralogy of Dolomitic Coorong Lakes, South Australia. J. Sediment. Res. 60, 843-858. Warthmann, R., Van Lith, Y., Vasconcelos, C., McKenzie, J.A., Karpoff, A.M., 2000. Bacterially induced dolomite precipitation in anoxic culture experiments. Geology 28, 1091-1094.
629
Wright, D.T., 1999. The role of sulphate-reducing bacteria and cyanobacteria in dolomite
630
formation in distal ephemeral lakes of the Coorong region, South Australia. Sediment. Geol.
631
126, 147-157.
632 633
Wright, D.T., Oren, A., 2005. Nonphotosynthetic Bacteria and the Formation of Carbonates and Evaporites Through Time. Geomicrobiol. J. 22, 27-53.
634
Wright, D.T., Wacey, D., 2005. Precipitation of dolomite using sulphate-reducing bacteria from the
635
Coorong Region, South Australia: significance and implications. Sedimentology 52,
636
987-1008.
637
Xia, W., Li, X., 1986. The discovery of primary dolomite from beach rock in the Xiaochaidan salt
638
lake of Qinghai and its significance. Acta. Sediment. Sin. 4, 19-26 (in Chinese with English
639
abstract).
640
Xing, P., Hahn, M.W., Wu, Q.L., 2009. Low taxon richness of bacterioplankton in high-altitude
641
lakes of the eastern tibetan plateau, with a predominance of Bacteroidetes and
642
Synechococcus spp. Appl. Environ. Microb. 75, 7017-7025.
643
Yang, J., Jiang, H., Sun, Y., Wu, G., Hou, W., Dong, H., Lai, Z., 2013. Abundance and diversity of
644
sulfate-reducing bacteria in Qinghai-Tibetan lakes. J. Salt. Lake. Res. 21, 7-13 (in Chinese
645
with English abstract).
646
Yang, K., He, J., Tang, W., Qin, J., Cheng, C.C.K., 2010. On downward shortwave and longwave
647
radiations over high altitude regions: observation and modeling in the Tibetan Plateau. Agric.
648
For. Meteorol. 150, 38–46.
649
Yu, B., Dong, H., Jiang, H., Li, S., Liu, Y., 2007. Discovery of spheric dolomite aggregations in
650
sediments from the bottom of Qinghai Lake and its significance for dolomite problem.
651
Geoscience 21, 66-70
(in Chinese with English abstract).
652
Zhang, S., Liu, Y., Jiao, X., Zhou, D., Zhang, X., Lu, S., Zhou, N., 2018. Sedimentary
653
environment and formation mechanisim of dolomitic rocks in the Middle Permian Lucaogou
654 655 656
Formation, Jimusar Depression,Junggar Basin. J. Palaeogeog. 20, 33-48. Zhang, X., Zheng, M., 2017. Research progress of salt minerals in Qinghai-Tibetan Plateau. Sci. Technol. Rev. 35, 72-76 (in Chinese with English abstract).
657
Zheng, M., Liu, X., 2010. Hydrochemistry and Minerals Assemblages of Salt Lakes in the
658
Qinghai-Tibet Plateau, China. Acta. Geol. Sin. 84, 1585-1600 (in Chinese with English
659
abstract).
660
Zheng, Y., Chen, J., 2000. Stable isotope geochemistry. Science Press, Beijing (in Chinese).
661
Zhu, D., Han, R., Shi, Q., Shen, G., Long, Q., Shuang, J., 2017. Correlation analysis of bacterial
662
community and hypersaline environmental factors in extreme salt lakes on the Qinghai-Tibet
663
Plateau. China. Environ. Sci. 37, 4657-4666 (in Chinese with English abstract).
664
Table.1 Locations, salinity, mineralogy, carbonate and dolomite and content from the sampled lakes on the central and western Tibetan Plateau Latitude
Longitude
Altitude
(°N)
(°E)
(m a.s.l.)
Sumxi Co
34.588
80.233
5051
Freshwater lake
0.26
2
Hongshan Lake
34.827
80.062
5065
Freshwater lake
0.59
3
Matou Lake
34.644
80.893
5217
Freshwater lake
4
Co Ngoin
31.627
88.726
4547
Freshwater lake
5
Nganggun Co
31.200
85.448
4663
Brackish lake
6
Serling Co
31.801
88.993
4547
7
Bura Co
34.417
85.779
8
Zhangne Co
31.547
9
Gomang Co
10
Larung Co
No.
Lake Name
1
Carbonate
Dolomite
Portion of
content
content
dolomiteb
dolomite, calcite
18.89%
10%
53%
dolomite, calcite
37.77%
4%
11%
0.11
dolomite, calcite
42.92%
10%
23%
0.21
dolomite, aragonite, calcite
41.30%
3%
7%
1.89
dolomite, aragonite, calcite
55.48%
3%
5%
Brackish lake
7.85
dolomite, aragonite, calcite, ankerite
48.50%
7%
14%
5172
Brackish lake
5.62
dolomite, aragonite, calcite
26.07%
7%
27%
87.385
4602
Brackish lake
4.06
dolomite, aragonite, calcite, hydromagnesite
59.45%
3%
5%
31.583
87.282
4658
Brackish lake
6.35
dolomite, aragonite, calcite, hydromagnesite
56.99%
5%
9%
34.337
85.228
4890
Brackish lake
12.01
dolomite, aragonite, calcite
28.91%
4%
14%
c
c
dolomite, aragonite, calcite, hydromagnesite
40.36%
5%
12%
(g/L)
Mineralogy
11
Zhari Namco
30.883
85.539
4638
Brackish lake
13.90
12
Yunbo Co
30.804
84.827
4614
Saltwater lake
16.09
dolomite, aragonite, calcite, hydromagnesite
42.94%
3%
7%
13
Dogze Co
31.866
87.555
4528
Saltwater lake
13.87
dolomite, aragonite, calcite
30.61%
4%
13%
14
Dong Co
32.146
84.750
4396
Saltwater lake
46.25
dolomite, aragonite, calcite
47.19%
7%
15%
15
Bero Zeco
32.423
82.955
4395
Saline lake
27.38
dolomite, aragonite, calcite, hydromagnesite
49.34%
4%
8%
16
Bamgdog Co
34.953
81.536
4909
Saline lake
30.74
dolomite, aragonite, calcite
52.68%
3%
6%
17
Lagkor Co
32.051
84.170
4470
Saline lake
40.27
dolomite, aragonite, calcite
60.71%
5%
8%
18
Mang Co
34.509
80.447
5027
Saline lake
84.11
dolomite, aragonite, calcite, hydromagnesite
42.43%
6%
14%
19
Longmu Co
34.615
80.480
5010
Saline lake
70.74
dolomite, aragonite, calcite
41.73%
4%
10%
Gyeze Caka
33.942
80.881
4524
Saline lake
135.00
dolomite, aragonite, calcite, hydromagnesite
40.61%
5%
12%
20
665 666 667
Salinity
Lake typea
a
Here the lake is classified according to salinity. The salinity is <1g/L, 1-35g/L, 35-50g/L and > 50g/L for freshwater lake, brackish lake, saltwater lake and saline lake, respectively.
b
The portion of dolomite refers to the proportion of dolomite to bulk carbonate.
c
The lake type and salinity are based on Wang and Dou (1998).
668 669
670 671 672 673 674
675 676 677
Table.2 Comparison of water chemistry and climatic conditions between dolomitic lakes and non-dolomitic lakes on the central and western Tibetan Plateau Mg/Ca Salinity pH (in mol) (g/L) Dolomitic lakes (n=20) 535.94 9.02 25.75 Non-dolomitic lakes (n=35) 23.19 8.95 12.56 Here water chemistry and climate parameters are mean values.
SO42(mg/L) 6959 2992
MAP (mm)
MAT (℃)
189.19 264.6
-4.3 -2.76
Table 3 Comparison with precipitation environments of typical lakes which biogenic dolomite formed Salinity Mg/Ca Location pH SO42-(mg/L) Reference (g/L) (in mol) Qinghai Lake (Saline lake) 12.5 61.0 9.3 1718.4 (Deng et al., 2010) Coorong Region (hypersaline dolomitic lakes)
15~141
6.82~9.11
9667~56591
(Wacey et al., 2007; Wright and Wacey, 2005)
Lagoa Vermelha (hypersaline costal lagoon)
6
8.0~8.5
3936~5760
(Vasconcelos and McKenzie, 1997; Warthmann et al., 2000)
0.04~45 19
7.61~10.56
10.36~5116 0.79
This study
Central and western Tibetan Plateau
0.11~135
Fig.1 Map of the sampling locations on the central and western Tibetan Plateau. Lakes are distinguished by salinity and marked with different colors. The sequence number of lakes in the figure is consistent with that in the table 1. To make the bedrock investigation convenient, six regions are divided based on the distance of lakes. Clastic rocks, metamorphic and igneous rocks are the most widely bedrock in the study area(a-f). Marine carbonatite is exposed around Sumxi Co, Longmu Co, Gyeze Caka and Bura Co(a-b). Fig.2. X-ray diffractograms showing the occurrence of detrital minerals (quartz, feldspar, clay), carbonates (calcite, aragonite, dolomite, hydromagnesite), respectively, in lake surface sediments. The peaks of each mineral are indicated by the following letters: Qtz=Quartz, Il=Illite, Ab=Albite, Cal=Calcite, Dol=Dolomite, Arg=Aragonite, Hm=Hydromagnesite, Hl=Halite. Fig.3. Scanning electronic microscope(SEM) images of <40 mm fraction from lake surface sediments: (a) Rhombohedral calcite from Hongshan Lake; (b) Blocky calcite stacked together from Nganggun Co; (c) Prismatic aragonite from Zhangne Co;(d)(e) Rhombohedral dolomite from Bamgdog Co and Bura Co; (f) dispersed spherical dolomite from Mang Co as marked by circles; (g) spherical dolomite precipitated on the surface of feldspar as marked by circles from Serling Co; (h) (i) aggregates of sub-spherical or spherical dolomite from Lagkor Co and Zhari Namco. Yellow circles represent the location of Energy dispersive spectrums (EDS) results acquisition (insert of each figure). Fig.4. Mg/Ca molar ratio plotted against (a) salinity; (b) pH; and (c) sulfate concentration. Fig.5 Relationship between δ18O and δ13C for carbonates from surface lake sediments. Fig.6 Plot of temperature versus δ18O of lake water. The curved lines represent isotopic compositions of dolomite in equilibrium with lake water at the given temperature. The range of δ18O of dolomite is 27.54‰ to 33.80‰ (in SMOW) from this study. (a) Profile of δ18O of dolomite using Eq.1; (b) Profile of δ18O of dolomite using and Eq.2. Take Bamdog Co as an example to show how the precipitation temperature of dolomite is calculated: the dolomite from Bamdog Co has a measured δ18O value of 33.80‰ and is marked as red pentagram on the diagram. The measured δ18O of lake water is -1.76‰ and so the precipitation temperature to which these two isotopic values correspond is read off the x-axis of the diagram.
Declaration of interest
We declare that we have no financial and personal relationships with other people or organizations that could inappropriately influence (bias) our work. There are no potential conflicts of interest include employment, consultancies, stock ownership, honoraria, paid expert testimony, patent applications/registrations, and grants or other funding. Signed by all authors as follows: Jiao Li Liping Zhu Minghui Li Junbo Wang Qingfeng Ma