ArticlePDF Available

Evolution of Chaka Salt Lake in NW China in response to climatic change during the Latest Pleistocene–Holocene

Authors:

Abstract and Figures

The Late Pleistocene and Holocene hydrologic balance of Chaka Salt Lake in the eastern Qaidam Basin of NW China was studied based on the analysis of lithostratigraphy, mineralogy, total organic carbon, and total nitrogen from a 9.0-m long sediment core. An age–depth model for the lake sediments is based on eight accelerator mass spectrometry (AMS) 14C measurements of organic matter and a 1700-year radiocarbon reservoir correction. The Pitzer model was used to calculate the sequence of minerals precipitated as a function of salinity assuming that the initial lake water was recharged from adjacent rivers and springs. Lake salinity values were derived from a good match between the calculated and observed mineral sequences. Our multi-proxy based hydrologic reconstruction of Chaka Lake indicates that it was a clastic-dominated, freshwater lake between 17.2 and 11.4 cal ka BP, which may have resulted from the input of glacial water into the lake at that time. During the Lateglacial and Holocene, a warm climatic regime predominated between 13.9 and 12.7 cal ka BP and then a cold climatic regime prevailed between 12.7 and 11.4 cal ka BP. These warm and cold periods correlate with the Bølling–Allerød and Younger Dryas events in the region. Beginning at 11.4 cal ka BP, a saline or hypersaline lake developed, which may have resulted from increased summer insolation and temperatures. These conditions persisted throughout the Holocene. Modeling indicates that lake salinity fluctuated between 66 and 223 ppt from 11.4 to 7.2 cal ka BP and then increased to 223–322 ppt between 7.2 and 6.0 cal ka BP, when most regions of China recorded high moisture availability (i.e. the so-called “mid-Holocene Climatic Optimum”). Lake salinity decreased to 66–223 ppt during a short time period between 6.0 and 5.3 cal ka BP, possibly caused by reduced evaporation. Subsequently, salinity values rapidly increased to 223–322 ppt between 5.3 and 5.2 cal ka BP and a hypersaline lake with a salinity higher than 324 ppt, similar to modern Chaka Salt Lake, became established around 5.2 cal ka BP. This establishment may have been related to the weakening of Asian monsoon. Regional comparisons of paleoclimatic records indicate that large temporal and spatial variations exist in the occurrence and intensity of drought over Western China since the Lateglacial.
Content may be subject to copyright.
Quaternary Science Reviews 27 (2008) 867–879
Evolution of Chaka Salt Lake in NW China in response to climatic
change during the Latest Pleistocene–Holocene
Liu Xingqi
a,
, Hailiang Dong
b
, Jason A. Rech
b
, Ryo Matsumoto
c
, Yang Bo
d
, Wang Yongbo
a
a
State Key Laboratory of Lake Science and Environment, Nanjing Institute of Geography and Limnology,
Chinese Academy of Sciences, Nanjing 210008, China
b
Department of Geology, Miami University, Oxford, OH 45056, USA
c
Department of Earth and Planetary Science, University of Tokyo, Tokyo 113-0033, Japan
d
Institute of Salt Lakes, Chinese Academy of Sciences, Xining 810008, China
Received 23 May 2007; received in revised form 21 November 2007; accepted 4 December 2007
Abstract
The Late Pleistocene and Holocene hydrologic balance of Chaka Salt Lake in the eastern Qaidam Basin of NW China was studied
based on the analysis of lithostratigraphy, mineralogy, total organic carbon, and total nitrogen from a 9.0-m long sediment core. An
age–depth model for the lake sediments is based on eight accelerator mass spectrometry (AMS)
14
C measurements of organic matter and
a 1700-year radiocarbon reservoir correction. The Pitzer model was used to calculate the sequence of minerals precipitated as a function
of salinity assuming that the initial lake water was recharged from adjacent rivers and springs. Lake salinity values were derived from a
good match between the calculated and observed mineral sequences. Our multi-proxy based hydrologic reconstruction of Chaka Lake
indicates that it was a clastic-dominated, freshwater lake between 17.2 and 11.4 cal ka BP, which may have resulted from the input of
glacial water into the lake at that time. During the Lateglacial and Holocene, a warm climatic regime predominated between 13.9 and
12.7 cal ka BP and then a cold climatic regime prevailed between 12.7 and 11.4 cal ka BP. These warm and cold periods correlate with the
Bølling–Allerød and Younger Dryas events in the region. Beginning at 11.4 cal ka BP, a saline or hypersaline lake developed, which may
have resulted from increased summer insolation and temperatures. These conditions persisted throughout the Holocene. Modeling
indicates that lake salinity fluctuated between 66 and 223 ppt from 11.4 to 7.2 cal ka BP and then increased to 223–322 ppt between 7.2
and 6.0 cal ka BP, when most regions of China recorded high moisture availability (i.e. the so-called ‘‘mid-Holocene Climatic
Optimum’’). Lake salinity decreased to 66–223 ppt during a short time period between 6.0 and 5.3 cal ka BP, possibly caused by reduced
evaporation. Subsequently, salinity values rapidly increased to 223–322 ppt between 5.3 and 5.2 cal kaBP and a hypersaline lake with a
salinity higher than 324 ppt, similar to modern Chaka Salt Lake, became established around 5.2 calka BP. This establishment may have
been related to the weakening of Asian monsoon. Regional comparisons of paleoclimatic records indicate that large temporal and spatial
variations exist in the occurrence and intensity of drought over Western China since the Lateglacial.
r2008 Elsevier Ltd. All rights reserved.
1. Introduction
The Asian summer monsoon is thought to play a key
role in atmospheric circulation over Asia and globally.
Paleoclimatic proxies from northeastern Tibet and the
Qaidam Basin, a region influenced by precipitation from
both the Asian summer monsoon and moist air masses
entrained in the Westerlies, are essential for determining
the interplay between these components of the climate
system over East Asia (Gao, 1962;Lehmkuhl and Haselein,
2000). Here, we present a record of the Latest Pleistocene
and Holocene hydrologic changes of Chaka Salt Lake in
the Qaidam Basin of northeast Tibet to determine the
potential role of insolation forcing on climate in a region
that is located at the boundary between Westerly and Asian
summer monsoon summer precipitation.
The Qaidam Basin, bounded by the Kunlun, Aljun, and
Qilian Mountains on the north and the Kunlun Mountains
on the south, is a large intermontane depression on the
northern margin of the Qinghai–Xizang Plateau (Fig. 1).
The basin covers an area of 120,000 km
2
with a catchment
ARTICLE IN PRESS
0277-3791/$ - see front matter r2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.quascirev.2007.12.006
Corresponding author. Tel.: +86 25 8688 2142; fax: +86 25 5771 3063.
E-mail addresses: xingqiliu@yahoo.com,
xqliu@niglas.ac.cn (L. Xingqi).
area of about 250,000 km
2
(Chen and Bowler, 1986;Zhang,
1987). Within this vast basin, there are many salt lakes and
playas containing alternating clastic–evaporite sedimentary
sequences. These sequences are potentially important
archives of hydrologic and paleoclimatic changes in
Western China. Wang et al. (1986) established the
preliminary magnetostratigraphy of Dabusan Lake in
the Qaidam Basin. A research project sponsored jointly
by the Chinese Academy of Sciences and Australian
National University in the 1980s focused on the climatic
evolution history of salt lakes in the Qaidam Basin, from
Late Pleistocene to today (Bowler et al., 1986;Chen and
Bowler, 1986). These authors determined the timing of
major climatic events for Kunteyi, Xiao Chaidan, and
Qarhan salt lakes in the Qaidam Basin (Fig. 1). Their
results indicate that expansion of these lakes occurred from
at least 40 to 15
14
C ka BP, large evaporate deposits
formed after 15
14
C ka BP, and dry conditions prevailed
throughout the Holocene (Bowler et al., 1986). Dry
conditions since the Lateglacial were also indicated by a
high content of gypsum in Barkol Lake of Xinjiang (Gu
et al., 1998), 750 km northwest of the Qaidam Basin. No
further paleoclimatic studies have been carried out in the
Qaidam Basin using clastic–evaporite sequences. A recent
study, based on total organic and inorganic carbon, total
sulfur, and carbon and oxygen stable isotopes of Zabuye
Lake sediments (Wang et al., 2002), reconstructed the Late
Pleistocene/Holocene climate conditions of Qinghai–Xi-
zang Plateau. The authors showed that the Zabuye Lake
overflowed during ca 16.2–10.6
14
C ka BP, closed its surface
outflow during ca 10.6–5.0
14
C ka BP, and a hypersaline
environment formed after 5
14
C ka BP.
Previous research suggests that significant temporal and
spatial variations of wet or dry phases reconstructed from
salt lakes or playas most likely existed in Western China
since the Lateglacial. For example, Lehmkuhl and Haselein
(2000) indicated that higher lake levels in the deserts of
Central Asia and on the Qinghai–Xizang Plateau are dated
to 40–25 ka (OIS 3) and to the Lateglacial/Early to mid-
Holocene periods. This is probably due to various local
hydrological or topographic factors that may have
different responses to two different climate systems
(Westerlies and Asian monsoons). However, well-dated
paleoclimatic records from salt lakes or playas in Western
China are rather limited. Thus, in order to fully understand
the mechanism of the observed temporal and spatial
variations, more paleoclimatic proxy records are highly
desirable. Our objective was to study the hydrologic
evolution of Chaka Salt Lake and to infer its response to
climatic changes that have occurred since the Lateglacial.
We employed mineralogical, lithostratigraphic, total or-
ganic carbon (TOC), and total nitrogen (TN) record from a
ARTICLE IN PRESS
Fig. 1. Map of the Qaidam Basin in China showing the location of Chaka Salt Lake. Also shown are other salt lakes, playas, and fresh or saline lakes in
the Basin. The inset map in the upper right corner shows the location of the coring site (CKL-2004). Numbers are locations of other lake records cited in
Fig. 6 (1, Zabuye Salt Lake; 2, Qinghai Lake; 3, Qarhan Salt Lake).
L. Xingqi et al. / Quaternary Science Reviews 27 (2008) 867–879868
9.0-m long continuous sediment core to achieve our
objective.
2. Study area
Chaka Salt Lake (361380–361450N, 991020–991120E;
3200 m above sea level, m a.s.l.) is located on the eastern
edge of the Qaidam Basin about 150 km west of Qinghai
Lake (Fig. 1). The Chaka Salt Lake region, isolated from
oceanic air masses, is characterized by a highly continental
climate. The mean annual temperature is 3.5 1C, with a low
mean monthly temperature of 12.4 1C in January and a
high mean monthly temperature of 14.4 1C in July. The
mean annual precipitation (197.6 mm) is far less than the
mean annual evaporation (2074.1 mm).
The lake has no outlet but is fed by fresh water from the
Mo River at its northwestern margin, the Hei River at its
southeastern margin, and springs on its northeastern and
southwestern bank (Fig. 1;Table 1). Some glaciers are
distributed in the north and northwest of Chaka Salt Lake
(inset of Fig. 1). It covers an area of 105 km
2
with a
catchment area of 11,600 km
2
. The depth of lake-water
varies greatly from 50–60 cm in the rainy season (June/
July) to just 1 cm in the dry season (January–March).
Similar to many other salt lakes in the Qaidam Basin, the
salinity of the Chaka Salt Lake water at present is between
317 and 347 ppt (about 10 times that of sea water). The
lake water chemistry is dominated with ions Na
+
,Mg
2+
,
Cl
, and SO
4
2
, but also contains smaller amounts of Ca
2+
,
K
+
,CO
3
2
, and HCO
3
(Table 1). The present-day lake
water is saturated with respect to halite, and halite
crystallization occurs during the dry season.
3. Materials and methods
In October 2004, a 9.0-m long sediment core (CKL-
2004) was drilled from the southeastern part of the lake at a
water depth of 2 cm using a strengthened corer made in
China (chamber length 1 m, inner diameter 7 cm). The
stratigraphic sequences recovered from core CKL-2004
generally consist of halite from 0 to 565 cm, a mixture of
soluble and sparingly soluble sulfate salts with dark
siliciclastics from 565 to 693 cm, and dark siliciclastic
layers from 693 to 900 cm. The core was subsampled at a
depth interval of 1 or 2 cm with a total of 590 subsamples.
Subsamples from CKL-2004 were analyzed for TOC, TN,
and mineralogy. A small subset was used for
14
C dating.
Samples for analyses of TOC and TN were treated with
1 N HCl to remove carbonates, rinsed repeatedly with
deionized water to remove soluble salt minerals, and dried.
TOC and TN contents were determined using a CE Model
440 Elemental Analyzer at the Nanjing Institute of
Geography and Limnology, Chinese Academy of Sciences
(CAS).
All samples analyzed for mineralogy were air-dried at
room temperature, disaggregated in a mortar and pestle,
and passed through a 62.5-mm sieve. Mineralogy was
determined at the Institute of Salt Lakes, CAS, using a
Phillips X-pert Pro X-ray diffraction with Cu Karadiation
(l¼1.5406 A
˚) at a scanning rate of 21min
1
for 2yranging
from 101to 801. Mineral identification and concentration
were estimated from the bulk mineral diffractograms using
the intensity of the strongest peak for each mineral
(Schultz, 1964;Chung, 1974;Last, 2001) aided by the use
of an automated search-match computer program X’Pert
HighScore Plus. Duplicate analyses of mineralogy for 15
samples in different lithostratigraphic units indicated that
precision of the mineralogical data is approximately 75%.
To understand the paleohydrological conditions of the
lake during different time periods and to explain the origin
of the evaporite mineral sequence observed in the core, a
computer simulation was carried out. The computer
program PHRQPITZ (Plummer et al., 1988), based on
the Pitzer virial-coefficient approach for activity-coefficient
corrections (Pitzer, 1973), was used to calculate the mineral
sequences through evaporation of initial lake water. This
ARTICLE IN PRESS
Table 1
Chemical composition of Chaka Salt Lake water, river water, and spring (concentration in mg/l)
Water body Sampling date Ca
2+
Mg
2+
Na
+
K
+
HCO
3
+CO
3
2
SO
4
2
Cl
TDS (ppt) Specific gravity pH
Lake water 1986
a
124 26,506 80,231 4473 199 23,625 187,700 322 1.22 6.8
7/1994 210 21,100 86,520 3620 600 22,320 182,710 317 1.21 6.9
11/1994 120 46,170 54,670 7560 1310 45,390 192,020 347 1.25 7.1
10/2004 125 45,160 56,120 6500 1200 44,400 191,230 344 1.23 7.0
Mo River 7/1994 71 84 378 4 192 445 517 1.70 1.00 6.9
11/1994 132 116 438 6 310 450 671 2.10 1.00 7.1
Hei River 7/1994 58. 104 296 15 409 384 113 1.38 1.00 6.8
11/1994 35 73 260 9 259 201 392 1.22 1.00 6.7
Spring 7/1994 52 29 153 4 243 109 193 0.78 1.00 7.1
11/1994 52 23 137 3 212 116 183 0.73 0.99 7.2
Initial water
b
1994 80 69 333 8 279 318 376 1.46
a
Data from Zhang (1987).
b
Data from Liu et al. (2004).
L. Xingqi et al. / Quaternary Science Reviews 27 (2008) 867–879 869
type of thermodynamic-based model is used commonly for
this purpose (Harvie and Weare, 1980;Harvie et al., 1980,
1982). Because presence or absence of a given mineral
assemblage is controlled not only by lake water composi-
tion and hydrological conditions but also by water
temperature, we studied homogenization temperatures of
205 fluid inclusions in halite crystals above 546 cm of the
900-cm core formed in the last 5180 years and 11 fluid
inclusions in halite formed in modern Chaka Salt Lake
(Liu et al., 2007a). Our study suggests that homogenization
temperatures range from 18 to 25 1C with an average of
23.7 1C(Liu et al., 2007a). These temperatures are very
similar to the water temperatures that were measured in the
lake in August 2005, which was the main season for halite
precipitation. Based on the Pitzer virial-coefficient ap-
proach for highly concentrated electrolyte solutions, we
thus calculated the mineral sequence produced from
evaporation of the initial water at 23.7 1C. The initial lake
water was assumed to have the same composition as that
formed by mixing river and spring waters according to
their present-day influxes to the lake.
4. Results and interpretation
4.1. Chronology
Ten samples of bulk organic material were processed for
accelerated mass spectrometry (AMS)
14
C dating to
construct an age model for the Chaka Salt Lake core. Six
samples were chemically pretreated, combusted, and
cryogenically purified at Miami University and then
analyzed at the University of Arizona NSF-AMS Labora-
tory, the United States, and four samples from a parallel
core to CKL-2004 were analyzed at the University of
Tokyo Radiocarbon Laboratory, Japan (Table 2). During
base treatment with 2% NaOH, all samples except for CK-
568 remained clear, indicating that no humic acids were
present. Sample CK-568, however, turned opaque during
base treatment due to high concentrations of humic acids.
The base treatment was repeated three times to remove all
humic acids; however, the
14
C age for CK-568 is 2000
14
C
years younger than sample CK-480 which was located from
1 m above (Table 2). We therefore exclude sample CK-568
from our age model because of the likelihood of secondary
contamination by humic acids. Sample CK-519 also shows
an age reversal of 1000
14
C years. We do not include this
sample in our age model, but we also have no reason to
suspect contamination by secondary humic acids. There is
a good linear relationship between depth and
14
C age for
the remainder of the radiocarbon samples (Fig. 2A). AMS
14
C ages could not be obtained above the depth of 450 cm
because the sediments contain very low TOC content.
The radiocarbon dating of lacustrine sediments in the
arid–semiarid regions of Western China is generally
affected by
14
C reservoir effects (Ren, 1998;Shen et al.,
2005;Morrill et al., 2006). This is especially true for saline
lakes on the Tibetan Plateau (Shen et al., 2005;Herzschuh
et al., 2006b). In an analysis of lake sediments from
Qinghai Lake, Shen et al. (2005) fit a linear regression line
to organic
14
C ages with depth to infer a radiocarbon
reservoir effect of 1039 years. These authors then assumed
a constant
14
C reservoir effect over their 18,000-year lake
record and subtracted 1039 years from all
14
C ages prior to
calibration. Herzschuh et al. (2006b) determined a modern
14
C reservoir effect of 2010
14
C years for Lake Zigetang, by
AMS
14
C dating of organic sediments from the surface
sample of the core. Herzschuh et al. (2006b) also assume a
constant
14
C reservoir effect over their lake record and
ARTICLE IN PRESS
Table 2
AMS
14
C ages and calendar age in CKL-2004 and its parallel core
Sample
no.
Lab I.D. Depth
(cm)
Sample description Dated
material
14
C age
(a BP)
14
C age corrected by 1700
years
Calendar age (cal a BP)
CKL326
a
Tka-
14017
447–453 Middle—grained halite with dark
clayey silt
TOC 4395730 2695730 2754–2851 (2802)
CKL340
a
TKa-
14018
508–511 Coarse halite with dark clayey silt TOC 5070735 3370735 3554–3694 (3624)
CK-308
b
AA67512 546–548 Coarse halite with dark clayey silt TOC 5705745 4005745 4401–4586 (4493)
CKL378
a
TKa-
14019
577–579 Dark clayey silt with gypsum. TOC 6815735 5115735 5749–5830 (5789)
CK-395
b
AA67514 642–643 Dark clayey silt with gypsum. TOC 9035750 7335750 8021–8208 (8114)
CK-435
b
AA67515 695–696 Dark clayey silt with gypsum. TOC 11,740765 10,040765 11,270–11,824 (11,547)
CKL577
a
Tka-
14020
705–706 Dark clayey silt TOC 11,840750 10,140750 11,601–12,041 (11,820)
CK-480
b
AA67516 749–750 Dark clayey silt TOC 12,995785 11,295785 12,997–13,321 (13,159)
CK-
519
b,c
AA67517 798–799 Dark clayey silt TOC 12,054765 10,354765 11,979–12,401 (12,190)
CK-
568
b,c
AA67518 867–868 Dark clayey silt TOC 10,9007120 92007120 10,168–10,707 (10,437)
a
Samples from a parallel core to CKL-2004.
b
Samples from CKL-2004.
c
Rejected ages.
L. Xingqi et al. / Quaternary Science Reviews 27 (2008) 867–879870
subtracted 2010 years from all radiocarbon samples prior
to calibration. In a study of Lake Zabuye, which is a
shallow (o1 m) saline (200–460 g/l) lake similar to Chaka,
Wang et al. (2002) did not determine if a
14
C reservoir
effect was present and therefore did not correct radio-
carbon ages.
In Chaka Salt Lake, we suggest that there is a
14
C
reservoir effect of organic lake sediments similar to other
salt lakes on the Qinghai–Xizang Plateau. However, we
cannot determine the modern reservoir effect due to the
extremely low TOC content of the surface sample from the
core. Moreover, we suggest that it is quite likely that
the
14
C reservoir effect has changed over time, decreasing
in magnitude when there is a more positive hydrologic
budget. We account for the
14
C reservoir effect at Chaka
Lake by correlating a major decrease in the TOC and TN
of Chaka Lake sediments, indicative of colder and drier
conditions, with the globally recognized Younger Dryas
event (Stuiver et al., 1995). TOC and TN contents distinctly
decrease at the depth interval between 693 and 732 cm,
and gypsum begins to precipitate largely at the depth
of 693 cm (Fig. 3). Based on the sedimentation rate
calculated from CK-395 (9035750
14
C a BP, 642 cm)
to CK-480 (12,995785
14
C a BP, 749 cm), the time char-
acterized by decreased TOC and TN content spans from
12.6 to 11.7
14
C ka BP. Gypsum begins to precipitate at
11.7
14
C ka BP. The Younger Dryas event (12.8–11.5 cal k-
a BP), reflected as a cold event in Greenland, has been
recorded throughout most of China, such as Guliya core
(Thompson et al., 1997;Yao et al., 1998) and Qinghai Lake
(Shen et al., 2005) in Qinghai–Xizang Plateau, on the Loess
Plateau (An et al., 1993;Chen et al., 1997;Wang et al.,
1999), salt lakes in Inner Mongolia (Chen et al., 1996), and
even in East China (Wang et al., 2001;Yuan et al., 2004;
DyKoski et al., 2005). The periods at which TOC and TN
contents increase again and gypsum precipitation initiates
ARTICLE IN PRESS
Fig. 2. Plot of
14
C and calendar age versus depth in the core CKL-2004. Circles indicate
14
C ages, squares are
14
C ages that have been corrected for
14
C
reservoir effects by subtracting 1700
14
C years, and triangles are calendar ages.
Fig. 3. Variations in content of gypsum, total organic carbon (TOC), and
total nitrogen (TN) in core CKL-2004.
L. Xingqi et al. / Quaternary Science Reviews 27 (2008) 867–879 871
are thought to correspond with the end of the Younger
Dryas cold event (Stuiver et al., 1995). Thus, it appears that
the end of the Younger Dryas cold event recorded in
Chaka Salt Lake is earlier than the global average by
1700
14
C years. We therefore assume a
14
C reservoir
correction of 1700 years for Chaka Lake. This magnitude
of a
14
C reservoir effect is comparable to the 1000–2000
year range measured or inferred from other salt lakes on
the Qinghai–Xizang Plateau. We then subtracted 1700
years from all measured radiocarbon ages prior to
calibration using Calib 4.4 (Stuiver et al., 1998)(Table 2).
The age of sample CK-519, which was not corrected by the
reservoir effect, falls well into the line of reservoir-corrected
14
C age and depth (Fig. 2A), probably implying that this
section of the lake record, which records a positive
hydrologic balance, has a minimal reservoir effect. Our
age model was then calculated by applying a 1700-year
correction to all radiocarbon samples above CK-519
and then inferring constant sedimentation rates between
radiocarbon samples or the surface in the age–depth model
(Fig. 2).
4.2. TOC and TN
Total organic carbon values are generally low, below
0.4%, but increase to values between 0.8% and 1.2%
between 7.5 and 5.5 m depth in the core (Fig. 3). TN
values are also generally low, below 0.1%, but also show
higher concentrations below 5.5 m depth in the core. TOC
and TN values are both especially low in the halite layer
and in the dark clastic layers. The dark color in clastic
layers is probably due to anoxic conditions, not necessarily
due to organic matter.
Total organic carbon and total nitrogen in lacustrine
sediment, which are closely related to organic matter
signals of biological productivity (Wang and Ji, 1995;
Sifeddine et al., 1996;Nara et al., 2005), are a good
indicator of primary productivity, especially in arid and
semi-arid region (Chen et al., 2001;Zhu et al., 2002;Shen
et al., 2005). Many studies have shown that high TOC and
TN contents indicate enhanced primary productivity due
to warm and/or wet climate, whereas their low contents
imply decreased productivity due to cold and/or dry
climate (Chen et al., 2001;Zhu et al., 2002;Shen et al.,
2005). An increase in TOC and TN contents can also
indicate an increase in the intensity of the summer
monsoon circulation in the monsoonal region (Chen et
al., 2001;Xiao et al., 2002;Shen et al., 2005).
4.3. Detrital minerals
Quartz, clay minerals, and feldspars are the dominant
siliciclastic minerals in Core CKL-2004. These detrital
minerals are allogenic and were derived from outcrops
within the drainage basin of the lake through processes of
weathering and erosion and transported into the lake by
river runoff. Because Chaka Salt Lake is located in a closed
basin, river discharge, controlled by climatic changes, plays
an important role in sediment input to the lake. Humid
conditions, closely related to increased precipitation or
melting glacial water caused by warm temperature, can
increase river runoff and the supply of detrital materials to
the lake.
4.4. Carbonates and evaporite minerals
Calcite and dolomite are the predominant carbonate
minerals in Core CKL-2004. Evaporite minerals include
halite (NaCl), the only halide mineral identified, and sulfate
minerals such as gypsum (CaSO
4
2H
2
O), glauberite
(Na
2
SO
4
CaSO
4
), mirabilite (Na
2
SO
4
10H
2
O), bloedite
[Na
2
Mg(SO
4
)
2
4H
2
O)], polyhalite [K
2
MgCa
2
(-
SO
4
)
4
2H
2
O] and small amounts (o5%) of epsomite
(MgSO
4
7H
2
O). No carbonate or evaporite minerals exist
within the drainage basin, indicating that they formed by
direct precipitation from lake water (i.e. endogenic) or by
post-depositional authigenic/diagenetic precipitation pro-
cesses. The precipitation of evaporite minerals is governed
mostly by lake salinity and the chemical composition of the
lake water, i.e. the balance between precipitation and
evaporation (Kinsman, 1976;Shang and Last, 1999). Our
simulation results show that the calculated mineral
sequence is in good agreement with that determined from
Core CKL-2004 (Figs. 4 and 5), which suggests that (1) no
dramatic changes in water sources or significant variations
in hydrology has occurred since the lake formed; (2) that
ARTICLE IN PRESS
Fig. 4. Calculated mineral sequence based on Pitzer model through
progressive evaporation of initial water (starting with 1000 g) at 23.7 1C
and 1 atm pressure. Minerals precipitated at different salinities are shown,
mass H
2
O is the amount of water remaining and salinity is the total
dissolved salts in water.
L. Xingqi et al. / Quaternary Science Reviews 27 (2008) 867–879872
the evaporatic mineral sequence in Chaka Salt Lake can be
explained by evaporation of present-day river and spring
waters as a major recharge.
4.5. Reconstruction of the lake evolution
Based on lithostratigraphy, mineralogical composition,
and TOC and TN contents, the hydrologic evolution of
Chaka Salt Lake during past 17,000 years is classified into
the following five stages (Fig. 5).
4.5.1. Stage I: 900–693 cm (17.2–11.4 cal ka BP)
The lowest lithostratigraphic unit is characterized by a
firm and dark, laminated, clayey silt. This unit is
predominately detrital siliciclastics, consisting of quartz
(35%), clay minerals (25%), and feldspars (25%). Carbo-
nates are composed of calcite (o10%) with minor amounts
of dolomite (o5%). No evaporite minerals are present in
this unit. TOC and TN content averages 0.43% and 0.11%,
respectively.
The mineral composition of this unit implies a deep lake
with a low salinity between 17.2 and 11.4 cal ka BP.
Relatively low TOC and TN contents indicate low primary
productivity and thus likely low temperatures. However, a
short warm phase and a subsequent cold phase are thought
to have occurred during this time period, as characterized
by a distinct increase then decrease in both TOC and TN
content between 776 and 732 cm (13.9–12.7 cal ka BP) and
between 732 and 693 cm (12.7–11.4 cal ka BP), respectively.
The warm and cold periods are thought to correspond to
the Bølling–Allerød warm event and Younger Dryas cold
event, respectively, which are mirrored in many records
from China (Thompson et al., 1997;Yao et al., 1998;Wang
et al., 2001;Yuan et al., 2004;DyKoski et al., 2005;Shen
et al., 2005;Herzschuh, 2006).
4.5.2. Stage II: 693–582 cm (11.4–6.0 cal ka BP)
Stage II consists of sulfate minerals and faintly bedded
dark clayey silt. This stage is characterized by an abrupt
and simultaneous increase in TOC, TN, and sulfate and a
decrease in detrital mineral content. TOC and TN contents
increase dramatically and almost reach their highest values
at this stage, suggesting that lake primary productivity was
high, possibly due to warmer/wetter temperatures. The
rapid appearance of gypsum at the beginning of this stage
indicates a dramatic transition of the lake from a
freshwater to a saline/hypersaline environment. Computer
modeling suggests that lake salinity probably fluctuated
between 66 and 232 ppt from 11.4 to 7.2 cal ka BP because
evaporitic salt minerals were mostly gypsum with a small
amount of halite (o5%) (Figs. 4 and 5). A short interval
marked by a distinct decrease in gypsum and an increase in
detrital minerals between 670 and 616 cm suggests that lake
salinity decreased at the period from10.0 to 7.2 cal ka BP.
Beginning at the depth of 616 cm (7.2 cal ka BP), gypsum
content increases to about 50%, glauberite appears with an
ARTICLE IN PRESS
Fig. 5. Stratigraphic variation in total organic carbon, total nitrogen, and mineralogy in the core CKL-2004.
L. Xingqi et al. / Quaternary Science Reviews 27 (2008) 867–879 873
average abundance of 21%, and the detrital mineral
content decreases to 34%. All these changes indicate a
decrease in the hydrologic budget of the lake. The salinity
of Chaka Lake is modeled to have been between 232 and
324 ppt between 7.2 and 6.0 cal ka BP (Figs. 4 and 5).
4.5.3. Stage III: 582–567 cm (6.0–5.3 cal ka BP)
This lithostratigraphic unit is characterized by a thin
(15 cm) layer of massive to faintly bedded dark clayey silt
with small amounts of evaporite minerals. Evaporite
concentration decreases to 10%, whereas contents of
detrital minerals increase drastically to 80%. Carbonates
dominated by calcite with a small amount of dolomite also
show a slight increase. TOC and TN still maintain
relatively high values. The increase of detrital mineral
content in this period could be caused by two opposing
conditions: a humid condition (and thus enhanced weath-
ering and transport) or a dry condition (and thus enhanced
eolian deposition). However, the high TOC and TN
contents imply wetter conditions. During this stage, the
lake experienced a short and distinct wet period, but its
salinity was not lower than 66 ppt (Fig. 4), as inferred from
the concentration of gypsum (10%) at this time (Fig. 5).
4.5.4. Stage IV: 567–270 cm (5.3–1.7 cal ka BP)
Glauberite and gypsum are abundant (30% and 40%,
respectively) at the base of this unit, suggesting that lake
salinity rapidly increased to 324 ppt (from 66 ppt) between
5.3 and 5.2 cal ka BP (Figs. 4 and 5). Evaporite minerals
account for 98% of material in this section of the core
with halite being the dominant mineral with smaller
proportions of sulfate minerals, such as gypsum (5%),
mirabilite (1–7%) between 554 and 465 cm, bloedite
(3–21%) between 565 and 465 cm, and polyhalite (o5%)
between 549 and 513 cm. These mineral assemblages
indicate that lake salinity was 325 ppt (Figs. 4 and 5),
which is similar to the salinity of modern Chaka Salt Lake.
Biological activity and primary productivity are very low
during these hypersaline conditions, which is characterized
by very low TOC and TN content (0.2% and 0.03%,
respectively).
4.5.5. Stage V: 270–0 cm (1.7 cal ka BP–modern)
This uppermost lithostratigraphic unit is characterized
by almost a pure bedded halite. Halite content increases to
498% and no other evaporite minerals occur. Carbonate
and detrital minerals (o2%) are nearly absent. All data
suggest that the lake salinity was 4326 ppt (Figs. 4 and 5).
TOC and TN contents further decrease to 0.06% and
0.02%, respectively.
5. Discussion
5.1. Paleoclimate
In order to better understand Chaka Salt Lake evolution
and its response to regional and global climatic changes, we
made a comparison between the Chaka Salt Lake record
and other lakes in the Qinghai–Xizang Plateau, as well as
regional and global climate changes (Fig. 6).
5.1.1. Lateglacial
Studies on air trapped in polar ice show that the last
glacial period was terminated by an abrupt warming event
15.0 ka ago (Severinghaus and Brook, 1999). Similarly,
tropical climate became warmer or wetter (or both) 20 to
80 years after the onset of the Greenland warming event
(Severinghaus and Brook, 1999). Abrupt warming in the
North Atlantic may also have led to slightly greater
warming over the Qinghai–Xizang Plateau, resulting in
melting of glacial ice and the onset of Asian summer
monsoon circulation after the LGM (Prell and Kutzbach,
1992;Meehl and Washington, 1993;Overpeck et al., 1996).
Evidence of a high lake level between 17.2 and 11.4 cal k-
a BP in Chaka Salt Lake is probably attributed to increased
moisture availability in response to warmer temperatures,
melting glacial ice, and the onset of Asian summer
monsoon during the Lateglacial. A wet climate during this
time is consistent with other records from the Qinghai–
Xizang Plateau, such as the lake overflow in Zabuye Salt
Lake between ca 16.2 and 10.6
14
CkaBP (Wang et al.,
2002), the climate amelioration in Qinghai Lake between
16.9, especially 14.1, and 10.8 cal ka BP (Shen et al., 2005),
and higher effective moisture conditions in Western China
during the Lateglacial (Herzschuh, 2006)(Fig. 6). How-
ever, very arid conditions characterized by evaporite
sequences occurred in other salt lakes of the Qaidam Basin
since the Lateglacial, such as Qarhan, Kunteyi, Xiao
Chaidan (Bowler et al., 1986). A possible interpretation for
this discrepancy is that evaporation caused by warmer
temperature in Chaka Salt Lake may have not been
sufficient to balance the quantity of melting glacial water
and the precipitation, but it may have been sufficient in
other lakes (Qarhan, Kunteyi, and Xiao Chaidan). These
lakes are further west and north of Chaka Salt Lake, and
thus are less influenced by the Asian monsoon.
Frequent climatic fluctuations are evident in Chaka Salt
Lake during the transitional period from the Lateglacial to
Holocene. Warm and cold phases are indicated by
variations of TOC and TN contents between 14 and
12.7 cal ka BP and from 12.7 to 11.4 cal ka BP, respectively.
These periods are likely correlated with the Bølling/Allerød
warm event between 14.8 and 12.8 cal ka BP, and the
Younger Dryas cold event between 12.8 and 11.5 cal ka BP
recorded in the GISP2 ice core (Stuiver et al., 1995). These
short climatic oscillations are also reflected in Qinghai lake
sediments (Liu et al., 2002;Shen et al., 2005), the Chinese
loess sequence (An et al., 1993;Chen et al., 1997;Wang et
al., 1999), and Guliya ice core (Thompson et al., 1997;Yao
et al., 1998).
5.1.2. Early Holocene to Early mid-Holocene
The Asian summer monsoon was intensified between 12
and 6 ka ago due to enhanced summer insolation in the
ARTICLE IN PRESS
L. Xingqi et al. / Quaternary Science Reviews 27 (2008) 867–879874
Northern Hemisphere and a stronger thermal contrast
between land and sea (COHMAP Members, 1988). The
intensified Asian Monsoon caused a much wetter climate
during the early to mid-Holocene than during the
Lateglacial and today in arid regions such as the Sahara
Desert (Ritchie and Haynes, 1987;Kutzbach et al., 1996;
de Menocal et al., 2000), the Arabian Peninsula (Hoelz-
mann et al., 1998), and the Thar Desert, India (Enzel et al.,
1999). Similarly, this wet period was also recorded in the
monsoon areas in Western China such as from Qinghai
Lake (Lister et al., 1991;Liu et al., 2002;Shen et al., 2005;
Liu et al., 2007b), Bangong Lake (Gasse et al., 1991, 1996),
and the Chinese Loess Plateau (Kukla et al., 1988;Feng
et al., 2005). Furthermore, most lakes in northern
Mongolia experienced higher levels during the middle
Holocene (Harrison et al., 1996;Tarasov, 1996;Tarasov
and Harrison, 1998;Grunert et al., 2000), which is
attributed to the expansion of the Asian monsoon to
Mongolia. However, our data from Chaka Salt Lake show
that an abrupt transition of the lake from fresh to saline or
hypersaline water occurred at the beginning of the
Holocene. Likewise, Zabuye Salt Lake closed its surface
outflow at ca 10.6
14
CkaBP (Wang et al., 2002). Pollen
data from two of the closest records to Chaka Salt Lake,
Zigetang Lake on the Central Qinghai–Xizang Plateau
(Herzschuh et al., 2006b) and Luanhaizi Lake on northeast
Qinghai–Xizang Plateau (Herzschuh et al., 2006a), indicate
warm conditions during the first half of the Holocene.
Similar early Holocene transitions from fresh to saline and/
or from hydrologically open to closed, evaporitic condi-
tions have been documented in large arid and semiarid
areas throughout the world including Lake Frome in
Australia (Bowler et al., 1986), Lake Van in Turkey
(Degens et al., 1984), and lakes in Canada such as
Ingebright Lake, Clearwater Lake, Lake Manitoba, Kill-
arney Lake, Moore Lake (Alberta), and lakes in United
States such as Pickerel Lake, Moon Lake, Medicine Lake,
Devils Lake, and Coldwater Lake (Vance and Last, 1994;
Valero-Garce
´s and Kelts, 1995;Laird et al., 1996;Valero-
Garce
´s et al., 1997;Last et al., 1998;Aitken et al., 1999;
Shang and Last, 1999;Dean and Schwalb, 2000;Last and
Vance, 2002). This widespread appearance of evaporite
minerals or the formation of a closed lake under warmer
but drier conditions at the beginning of the Holocene is
ARTICLE IN PRESS
Fig. 6. Comparison between the Chaka Salt Lake record and other regional salt lakes, as well as regional and global climate summaries and forcings: (A)
Chaka Salt Lakes, dashed line is TOC content. (B) Insolation at 301N(Berger and Loutre, 1991). (C) The GRIP d
18
O record from Greenland (Johnsen et
al., 2001). (D) The mean effective moisture inferred from many records of lakes located in the Asian monsoonal region (Herzschuh, 2006). (E) Qinghai
Lake (Shen et al., 2005;Liu et al., 2007b). Solid and broken lines are pollen concentration and d
18
O of ostracode shells, respectively. (F) Zabuye Lake
(Wang et al., 2002). Solid line with circle, broken line with square, and dot line with solid circle are d
18
O and d
13
C of carbonate, and TOC content,
respectively. (G) Qarhan Salt Lake (Chen and Bowler, 1986).
L. Xingqi et al. / Quaternary Science Reviews 27 (2008) 867–879 875
almost synchronous with the increased summer insolation
(Fig. 6). Thereby, increased summer insolation may have
acted as a trigger to increase summer temperatures and
evaporation. Subsequent to this increased summer insola-
tion, the salinity of Chaka Salt Lake decreased slightly
between 10 and 7.2 cal ka BP, centered at 8.3 cal ka BP,
probably responding to the intensified summer monsoon
which lagged behind peak June insolation by about
2–3 cal ka (Fig. 6). However, the salinity of Chaka Salt
Lake increased significantly again at 7.2 cal ka BP, which
coincides with the so-called ‘‘mid-Holocene Climatic
Optimum’’ when both megathermal and megahumid
climate occurred in the period from 7.2 to 6
14
CkaBP
(Shi et al., 1993), further demonstrating that evaporation at
high temperatures exceeds precipitation.
5.1.3. Later mid-Holocene to Late Holocene
The weakening of the Asian summer monsoon appears
to have been linked more directly to decreased insolational
forcing since ca 6 ka BP, which caused a cooler and drier
climate in the Asian Monsoon region (COHMAP Mem-
bers, 1988;Gasse et al., 1991, 1996;Overpeck et al., 1996;
Shen et al., 2005;Herzschuh, 2006;Liu et al., 2007b).
A decreased but still high salinity occurred in Chaka Salt
Lake between 6 and 5.3 cal ka BP, which may have been
related to less evaporation caused by cooler temperatures.
Subsequently, the salinity of the lake increased and a more
hypersaline lake environment dominated by halite pre-
cipitation finally formed in Chaka Salt Lake between
5.3 cal ka BP and present day. Similarly, the sediments in
Zabuye Lake on the Qinghai–Xizang Plateau are pre-
dominantly mirabilite and halite (Wang et al., 2002). Since
5.0
14
C ka BP, a dry playa environment of Frome in
Australia fully formed after 4
14
CkaBP (Bowler et al.,
1986), and Van Lake in Turkey experienced a period with
stable but slightly low lake-levels between 6
14
C ka BP and
present day (Degens et al., 1984). Simultaneously, the
climatic deterioration characterized by dry or cold condi-
tions was widely documented in most lakes from Western
China (Lister et al., 1991;Gasse et al., 1991, 1996;
Kashiwaya et al., 1995;Liu et al., 2002;Shen et al., 2005;
Herzschuh, 2006;Liu et al., 2007b), although they all
recorded a warm and humid phase in the early and middle
Holocene. Therefore, the further desiccation of Chaka Salt
Lake after the middle Holocene may have been related to
the weakening of the Asian summer monsoon.
5.2. Some implications
Some salt lakes in the Qaidam Basin, such as Qarhan,
Kunteyi, Xiaocaidan, formed at the beginning of the
Lateglacial when temperature increased. Chaka Salt Lake
and many other saline or salt lakes in Australia (Bowler
et al., 1986), Turkey (Degens et al., 1984), Canada, and
United States discussed above (Vance and Last, 1994;
Valero-Garce
´s and Kelts, 1995;Laird et al., 1996;Valero-
Garce
´s et al., 1997;Last et al., 1998;Aitken et al., 1999;
Shang and Last, 1999;Dean and Schwalb, 2000;Last and
Vance, 2002), formed at the beginning of the Holocene
when temperature was higher than that during the
Lateglacial. Halite formed in lake water after the late
mid-Holocene, with a very high salinity in both Chaka Salt
Lake and Zabuye Salt Lake (Wang et al., 2002), when
temperature decreased but was still higher than Lateglacial
temperatures. Low monsoon precipitation due to decreased
insolation forcing may have accelerated the increase of
salinity in the two lakes. Therefore, it appears that warm
temperature plays an important role in the formation of the
salt lake. According to this model, we can assume that the
drought will be severe and lakes are threatened to shrink
further, at least in part of the Western China, if
temperatures continue to increase in the future.
6. Conclusions
Lithostratigraphy, mineralogy, TOC, and TN contents,
along with 10 AMS
14
C ages, of a 9.0-m long continuous
lake sediment sequence from Chaka Salt Lake were used to
reconstruct the history of lake evolution since the
Lateglacial, and thus to discuss its response to regional
and global climatic changes. Our main conclusions are as
follows.
(1) It was a freshwater lake during the Lateglacial. The
climate fluctuated frequently during the transitional
period from the Lateglacial to the Early Holocene.
Short-term climatic oscillations such as the Bølling–Al-
lerød warm event and Younger Dryas cold event
appear to be recorded in the lake sediments. Beginning
at 11.4 ka BP, a saline or hypersaline lake formed. The
salinity of Chaka Salt Lake further increased at
7.2 cal ka BP, which corresponds to the middle Holo-
cene Climatic Optimum when most regions of China
recorded availability of high moisture content. During
the Late–middle to Late Holocene, the lake salinity
progressively increased and a hypersaline environment
similar to modern conditions finally established around
5.3 cal ka BP.
(2) Evidence of a high lake level during the Lateglacial is
probably attributed to increased moisture availability
in response to warmer temperatures, melting glacial ice,
and the onset of Asian summer monsoon during the
Lateglacial. Warmer temperatures caused by increased
summer insolation during the early Holocene may have
triggered the formation of a saline or hypersaline lake
at the beginning of the Holocene. Decreased precipita-
tion due to the weakening of Asian monsoon after the
middle Holocene, probably accelerated the formation
of a more hypersaline lake during the middle-Holocene
to Late Holocene. Generally, the hypersaline environ-
ment at Chaka Lake was formed at the beginning of the
Holocene when temperatures were higher than during
the Lateglacial. This suggests that higher temperature
plays an important factor in the formation of salt lakes.
ARTICLE IN PRESS
L. Xingqi et al. / Quaternary Science Reviews 27 (2008) 867–879876
(3) The evolutionary history of Chaka Salt Lake and its
response to climatic changes are characterized by
multiple stages with similarities and differences to
other salt lakes in the region. Chaka and Zabuye Salt
Lake experienced similar evolution and climatic re-
sponse since the Lateglacial. Chaka Salt Lake and
Qarhan, Kunteyi, and Xiaocaidan experienced a similar
history only during the Holocene. Chaka Salt Lake and
other lakes in Western China evolved and responded to
climatic changes in a different way, only with some
broad similarities during the Lateglacial and during the
late mid-Holocene to Late Holocene. Clearly, there
remain many discrepancies among lake sediment
records from the region. In order to determine if these
discrepancies are the result of local differences in past
climate, or are due to problems with chronology or
interpretation, more well-dated paleoclimatic records
from salt lakes in Western China are needed.
Acknowledgments
We thank Prof. Colin Murray-Wallace and two anon-
ymous reviewers for their helpful comments. This research
was supported by the National Natural Science Founda-
tion of China (Grant nos. 40772108, 40373016, 40472064,
and 40672079) and the National Basic Research Program
of China (Grant nos. 2004CB720200 and 2005CB422002).
References
Aitken, A.E., Last, W.M., Burt, A.B., 1999. The lithostratigraphic record
of late Pleistocene–Holocene environmental change at the Andrews
site near Moose Jaw, Saskatchewan. In: Lemmen, D.S., Vance, R.E.
(Eds.), Holocene Climate and Environmental Change in the Palliser
Triangle: A Geoscientific Context for Evaluating the Impacts of
Climate Change on the Southern Canadian Prairies. Bulletin—
Geological Survey of Canada 534, 173–181.
An, Z., Porter, S.C., Zhou, W., Lu, Y., Donahue, D.J., Head, M.J., Wu,
X., Ren, J., Zheng, H., 1993. Episode of strengthened summer
monsoon climate of Younger Dryas age on the loess plateau of Central
China. Quaternary Research 39, 45–54.
Berger, A., Loutre, M.F., 1991. Insolation values for the climate of the last
10 million years. Quaternary Science Reviews 10, 297–317.
Bowler, J.M., Huang, Q., Chen, K.Z., Head, M.J., Yuan, B.Y., 1986.
Radiocarbon dating of playa-lake hydrologic changes: examples from
Northwestern China and Central Australia. Palaeogeography, Palaeo-
climatology, Palaeoecology 54, 241–260.
Chen, K.Z., Bowler, J.M., 1986. Late pleistocene evolution of salt lakes in
the Qaidam Basin, Qinghai Province, China. Palaeogeography,
Palaeoclimatology, Palaeoecology 54, 87–104.
Chen, Y.C., Wei, D.Y., Wang, J.J., Guan, S.Z., Qian, Z.H., Yang, Q.T.,
Liu, Z.M., Hsu
¨, K.J., Mckenzie, J.A., Bernassconi, S., Dobson, J.P.,
Niessen, F., 1996. Determination and reorganization of Younger
Dryas and its important significance in the salt lakes’ sediments of
Inner Mongolian. Geology of Chemical Minerals 18 (3), 163–169 (in
Chinese with English abstract).
Chen, F., Bloemendal, J., Wang, J., Oldfield, F., 1997. High-resolution
multi-proxy climate records from Chinese loess: evidence for rapid
climatic changes over the last 75 k. Palaeogeography, Palaeoclimatol-
ogy, Palaeoecology 130, 323–335.
Chen, F.H., Zhu, Y., Li, J.J., Shi, Q., Jin, L.Y., Wu
¨nemann, B., 2001.
Abrupt Holocene changes of the Asian monsoon at millennial-and
centennial-scales: Evidence from lake sediment document in Minqin
Basin, NW China. Chinese Science Bulletin 46, 1942–1947.
Chung, F.H., 1974. Quantitative interpretation of X-ray diffraction
patterns of mixtures. II. Adiabatic principle of X-ray diffraction
analysis of mixtures. Journal of Applied Crystallography 7, 526–531.
COHMAP Members, 1988. Climatic changes of the last 18,000 years:
observations and model simulations. Science 241, 1043–1052.
de Menocal, P., Ortiz, J., Guilderson, T., Sarnthein, M., 2000. Coherent
high- and low-latitude climate variability during the Holocene warm
period. Science 288, 2198–2202.
Dean, W.E., Schwalb, A., 2000. Holocene environmental and climatic
changes in the Northern Great Plains as recorded in the geochemistry
of sediments in Pickerel Lake, South Dakota. Quaternary Interna-
tional 67, 5–20.
Degens, E.T., Wong, H.K., Kempe, S., Kurtman, F., 1984. A geological
study of Lake Van, Eastern Turkey. Geol Rundschau 73, 701–734.
Dykoski, C.A., Edwardsa, R.L., Cheng, H., 2005. A high-resolution,
absolute-dated Holocene and deglacial Asian monsoon record from
Dongge Cave, China. Earth and Planetary Science Letters 233, 71–86.
Enzel, Y., Ely, L.L., Mishra, S., Ramesh, R., Amit, R., Lazar, B.,
Rajaguru, S.N., Baker, V.R., Sandler, A., 1999. High resolution
Holocene environmental changes in the Thar Desert, Northwestern
India. Science 284, 125–128.
Feng, Z.D., Wang, W.G., Guo, L.L., Khosbayar, P., Narantsetseg, T.,
Julld, A.J.T., 2005. Lacustrine and eolian records of Holocene climate
changes in the Mongolian Plateau: preliminary results. Quaternary
International 136, 25–32.
Gao, Y., 1962. Some Problems on East-Asia Monsoon. Science Press,
Beijing (in Chinese).
Gasse, F., Arnold, M., Fontes, J.C., Fort, M., Gibert, E., Huc, A.,
Bingyan, L., Yuanfang, L., Qing, L., Melieres, F., van Campo, E.,
Fubao, W., Qingsong, Z., 1991. A 13.000-year climate record from
western Tibet. Nature 353, 742–745.
Gasse, F., Fontes, J.C., van Campo, E., Wei, K., 1996. Holocene
environmental changes in Bangong Co basin (Western Tibet): 4.
Discussions and conclusions. Palaeogeography, Palaeoclimatology,
Palaeoecology 120, 79–92.
Grunert, J., Lehmkuhlb, K., Walther, M., 2000. Paleoclimatic evolution of
the Uvs Nuur basin and adjacent areas (Western Mongolia).
Quaternary International 65/66, 171–192.
Gu, Z.Y., Zhao, H.M., Wang, Z.H., Yuan, B.Y., 1998. Evaporation salt
records of environmental response to climate change in Barkol
Lake Basin, Northwestern China. Quaternary Sciences 4, 328–334
(in Chinese with English abstract).
Harrison, S.P., Yu, G., Tarasov, P.E., 1996. Late Quaternary
lake-level record from northern Eurasia. Quaternary Research 45,
138–159.
Harvie, C.E., Weare, J.H., 1980. The prediction of mineral solubilities in
natural waters: the Na–K–Mg–Ca–SO
4
–Cl–H
2
O system from zero to
high concentration at 25 1C. Geochimica et Cosmochimica Acta 44,
981–997.
Harvie, C.E., Weare, J.H., Hardie, L.W., Eugster, H.P., 1980. Evaporate
of seawater: calculated mineral sequences. Science 208, 498–500.
Harvie, C.E., Eugster, H.P., Weare, J.H., 1982. Mineral equilibria in the
six-component seawater system. Na–K–Mg–Ca–SO
4
–Cl–H
2
Oat251C
II: composition of the saturated solutions. Geochimica et Cosmochi-
mica Acta 46, 1603–1618.
Herzschuh, U., 2006. Palaeo-moisture evolution in monsoonal Central
Asia during the last 50,000 years. Quaternary Science Reviews 25,
163–178.
Herzschuh, U., Ku
¨rschner, H., Mischke, S., 2006a. Temperature
variability and vertical vegetation belt shifts during the last 50,000
in the Qilian Mountains (NE margin of the Tibetan Plateau, China).
Quaternary Research 66, 133–146.
Herzschuh, U., Winter, K., Wu
¨nnemann, B., Li, S.J., 2006b. A general
cooling trend on the central Tibetan Plateau throughout the Holocene
ARTICLE IN PRESS
L. Xingqi et al. / Quaternary Science Reviews 27 (2008) 867–879 877
recorded by the Lake Zigetang pollen spectra. Quaternary Interna-
tional 154/155, 113–121.
Hoelzmann, P., Jolly, D., Harrison, S.P., Laarif, F., Bonnefille, R.,
Pachur, H.J., 1998. Mid-Holocene land-surface conditions in northern
Africa and the Arabian peninsula: a data set for AGCM simulations.
Global Biogeochemical Cycles 12, 35–52.
Johnsen, S.J., Dahl-Jensen, D., Gundestrup, N., Steffensen, J.P., Clausen,
H.B., Miller, H., Masson-Delmotte, V., Sveinbjs
ˇrndottir, A.E., White,
J., 2001. Oxygen isotope and palaeotemperature records from six
Greenland ice-core stations: Camp Century, Dye-3, GRIP, GISP2,
Renland and NorthGRIP. Journal of Quaternary Science 16, 299–307.
Kashiwaya, K., Masuzawa, T., Morinaga, H., Yaskawa, K., Yuan, B.,
Liu, J., Gu, Z., 1995. Changes in hydrological conditions in the central
Qing–Zang (Tibetan) Plateau inferred from lake bottom sediments.
Earth and Planetary Science Letters 135, 31–39.
Kinsman, D.J.J., 1976. Evaporites: relative humidity control of primary
facies. Journal of Sedimentary Petrology 46, 273–279.
Kukla, G., Heller, F., Liu, X.M., Xu, T.C., Liu, T.S., An, Z.S., 1988.
Pleistocene climates in China dated by magnetic susceptibility.
Geology 16, 811–814.
Kutzbach, J., Bonan, G., Foley, J., Harrison, S.P., 1996. Vegetation and
soil feedbacks on the response of the African monsoon to orbital
forcing in the early to middle Holocene. Nature 384, 623–626.
Laird, K.R., Fritz, S.C., Grimm, E.C., Mueller, P.G., 1996. Centuryscale
paleoclimatic reconstruction from Moon Lake, a closedbasin lake in
the northern Great Plains. Limnology and Oceanography 41, 890–902.
Last, W.M., 2001. Mineralogical analysis of lake sediments. In: Last,
W.M., Smol, J.P. (Eds.), Tracking Environmental Change Using Lake
Sediments. Volume 2: Physical and Geochemical Techniques. Kluwer
Academic Publishers, Dordrecht, The Netherlands, pp. 143–187.
Last, W.M., Vance, R.E., 2002. The Holocene history of Oro Lake, one of
the western Canada’s longest continuous lacustrine records. Sedimen-
tary Geology 148, 161–184.
Last, W.M., Vance, R.E., Wilson, S., Smol, J.P., 1998. A multiproxy
limnologic record of rapid early-Holocene hydrologic change on the
northern Great Plains, southwestern Saskatchewan, Canada. Holocene
8, 503–520.
Lehmkuhl, F., Haselein, F., 2000. Quaternary paleoenvironmental change
on the Tibetan Plateau and adjacent areas (Western China and
Western Mongolia). Quaternary International 65/66, 121–145.
Lister, G.S., Kelts, K., Chen, K.Z., Yu, J.Q., Niessen, F., 1991. Lake
Qinghai, China: closed-basin lake levels and the oxygen isotope record
for ostracoda since the last Pleistocene. Palaeogeography, Palaeocli-
matology, Palaeoecology 84, 141–162.
Liu, X.Q., Shen, J., Wang, S.M., Yang, X.D., Tong, G.B., Zhang, E.L.,
2002. A 16000-year pollen record of Qinghai Lake and its Paleoclimate
and Paleoenvironment. Chinese Science Bulletin 47, 1931–1937.
Liu, X.Q., Cai, K.Q., Yu, S.S., 2004. Geochemical simulation of the
formation of brine and salt minerals based on Pitzer model in Chaka
Salt Lake. Science in China Series D 8, 720–726.
Liu, X.Q., Ni, P., Dong, H.L., Wang, T.G., 2007a. Homogenization
temperature and its significance for primary fluid inclusion in halite
formed in Chaka Salt Lake, Qaidam Basin. Acta Petrologica Sinica 23,
113–116.
Liu, X.Q., Shen, J., Wang, S.M., Wang, Y.B., Liu, W.G., 2007b.
Southwest monsoon changes indicated by oxygen isotope of ostracode
shells from sediments in Qinghai Lake since the Lateglacial. Chinese
Science Bulletin 52, 539–544.
Meehl, G.A., Washington, W.M., 1993. South Asian summer monsoon
variability in a model with doubled atmospheric carbon dioxide
concentrations. Science 260, 1101–1103.
Morrill, C., Overpeck, J.T., Cole, J.E., Liu, K.-B., Shen, C., Tang, L.,
2006. Holocene variations in the Asian monsoon inferred from the
geochemistry of lake sediments in central Tibet. Quaternary Research
65, 232–243.
Nara, F., Tani, Y., Soma, Y., Soma, M., Naraoka, H., Watanabe, T.,
Horiuchi, K., Kawai, T., Oda, T., Nakamura, T., 2005. Response of
phytoplankton productivity to climate change recorded by sedimen-
tary photosynthetic pigments in Lake Hovsgol (Mongolia) for the last
23,000 years. Quaternary International 136, 71–81.
Overpeck, J., Anderson, D., Trumbore, S., Prell, W., 1996. The southwest
Indian monsoon over the last 18000 years. Climate Dynamics 12,
213–225.
Pitzer, K.S., 1973. Thermodynamics of electrolytes.1.Theroretical basis
and general equations. Journal of Physical Chemistry 77, 268–277.
Plummer, L.N., Parkhurst, D.L., Fleming, G.W., Dunkle, S.A., 1988.
A computer program incorporation Pitzer’s equations for
calculation of geochemical reactions in brine. US Geological
Survey, Water Resources Investigations Report 88-4153, Virginia,
pp. 1–191.
Prell, W.L., Kutzbach, J.E., 1992. Sensitivity of the Indian monsoon to
forcing parameters and implications for its evolution. Nature 360,
647–652.
Ren, G., 1998. A finding of the influence of ‘‘hard water’’ on radiocarbon
dating for lake sediments in Inner Mongolia, China. Journal of Lake
Sciences 10, 80–82 (in Chinese with English abstract).
Ritchie, J.C., Haynes, C.V., 1987. Holocene vegetation zonation in the
eastern Sahara. Nature 330, 645–647.
Schultz, L.G., 1964. Quantitative interpretation of mineralogical composi-
tion from X-ray and chemical data for the Pierre Shale. US Geological
Survey, Professional Paper 391-C, p.31.
Severinghaus, J.P., Brook, E.J., 1999. Abrupt climate change at the end of
the last glacial period inferred from trapped air in polar ice. Science
286, 930–934.
Shang, Y., Last, W.M., 1999. Mineralogy, lithostratigraphy, and inferred
geochemical history of North Ingebrigt Lake, Saskatchewan. In:
Lemmen, D.S., Vance, R.E. (Eds.), Holocene Climate and Environ-
mental Change in the Palliser Triangle: A Geoscientific Context for
Evaluating the Impacts of Climate Change on the Southern Canadian
Prairies. Bulletin—Geological Survey of Canada 534, 95–110.
Shen, J., Liu, X.Q., Wang, S.M., Matsumoto, R., 2005. Palaeoclimatic
changes in the Qinghai Lake area during the last 18,000 years.
Quaternary International 136, 131–140.
Shi, Y.F., Kong, Z.Z., Wang, S.M., Tang, L.Y., Wang, F.B., Yao, T.D.,
Zhao, X.T., Zhang, P.Y., Shi, S.H., 1993. Mid-Holocene climates
and environments in China. Global and Planetary Change 7,
219–233.
Sifeddine, P., Bertrand, E., Lallier-Verges, E., Patience, A.J., 1996.
Lacustrine organic fluxes and palaeoclimatic variations during the last
15 ka: Lac du Bouchet (Massif central, France). Quaternary Science
Reviews 15, 203–211.
Stuiver, M., Grootes, P.M., Braziunas, T.F., 1995. The GISP2 climate
record of the past 16,500 years and the role of the sun, ocean, and
volcanoes. Quaternary Research 44, 341–354.
Stuiver, M., Reimer, P.J., Bard, E., Beck, J.W., Burr, G.S., Hughen, K.A.,
Kromer, B., McCormac, F.G., van der Plicht, J., Spurk, M., 1998.
INTCAL98 radiocarbon age calibration, 24,000–0 cal BP. Radio-
carbon 40, 1041–1083.
Tarasov, P.E., 1996. Lake Status Records from the Former Soviet Union
and Mongolia: Documentation of the Second Version of the Database.
World Data Center-A for Paleoclimatology, Boulder, CO.
Tarasov, P.E., Harrison, S.P., 1998. Lake status records from the former
Soviet Union and Mongolia: a continental-scale synthesis. Palaeokli-
maforschung 25, 115–130.
Thompson, L.G., Yao, T., Davis, M.E., Henderson, K.A., Mosley-
Thompson, E., Lin, P.-N., Beer, J., Synal, H.-A., Cole-Dai, J., Bolzan,
J.F., 1997. Tropical climate instability: the last Glacial cycle from a
Qinghai-Tibetan ice core. Science 276, 1821–1825.
Valero-Garce
´s, B., Kelts, K.R., 1995. A sedimentary facies model for
perennial and meromictic saline lakes: Holocene medicine Lake basin
South Dakota, USA. Journal of Paleolimnology 15, 123–149.
Valero-Garce
´s, B., Laird, K.R., Fritz, S.C., Kelts, K., Ito, E., Grimm,
E.C., 1997. Holocene climate in the Northern Great Plains inferred
from sediment stratigraphy, stable isotopes, carbonate geochemistry,
diatoms, and pollen at Moon Lake, North Dakota. Quaternary
Research 48, 359–370.
ARTICLE IN PRESS
L. Xingqi et al. / Quaternary Science Reviews 27 (2008) 867–879878
Vance, R.E., Last, W.M., 1994. Paleolimnology and Global Change on
the Canadian prairies, Current Research 1994-B. Geological Survey of
Canada, Ottawa, Ont., Canada, pp. 49–58.
Wang, S.M., Ji, L., 1995. Hulun Lake. Hefei, Press of Chinese University
of Science and Technique, 85–93 (in Chinese).
Wang, J.T., Derbyshire, E., Shaw, J., 1986. preliminary magnetostrati-
graphy of Dabusan Lake, Qaidam Basin, central Asian. Physics of the
Earth and Planetary Interiors 44, 41–46.
Wang, J.M., Shi, Q., Chen, F.H., Duan, S.X., 1999. Rapid oscillations of
Chinese monsoon climate since the last deglaciation and its
comparison with GISP2 record. Chinese Science Bulletin 44,
284–286.
Wang, Y.J., Cheng, H., Edwards, R.L., An, Z.S., Wu, J.Y., Shen, C.-C.,
Dorale, J.A., 2001. A high-resolution absolute-dated late Pleistocene
monsoon record from Hulu Cave, China. Science 294, 2345–2348.
Wang, R.L., Scarpitta, S.C., Zhang, S.C., Zheng, M.P., 2002. Later
Pleistocene/Holocene climate conditions of Qinghai–Xizhang Plateau
(Tibet) based on carbon and oxygen stable isotopes of Zabuye Lake
sediments. Earth and Planetary Science Letters 203, 461–477.
Xiao, J.L., Nakamura, T., Lu, H.Y., Zhang, G.Y., 2002. Holocene climate
changes over the desert/loess transition north-central China. Earth and
Planetary Science Letters 197, 11–18.
Yao, T.D., Thompson, L.G., Shi, Y.F., 1998. Climate variation since the
last interglacial recorded in the Guliya ice core. Science in China Series
D 40, 662–668.
Yuan, D.X., Cheng, H., Edwards, R.L., Dykoski, C.A., Kelly, M.J.,
Zhang, M.L., Qing, J.M., Lin, Y.S., Wang, Y.J., Wu, J.Y., Dorale,
J.A., An, Z.S., Cai, Y.J., 2004. Timing, duration, and transitions of the
last interglacial Asian monsoon. Science 304, 575–578.
Zhang, P.X., 1987. Salt Lakes in Qaidam Basin. Science press, Beijing (in
Chinese).
Zhu, L.P., Chen, L., Li, B.Y., Li, L.F., Xia, W.L., Li, J.G., 2002.
Environmental changes reflected by the lake sediments of the South
Hongshan Lake, Northwest Tibet. Science in China Series D 45, 430–439.
ARTICLE IN PRESS
L. Xingqi et al. / Quaternary Science Reviews 27 (2008) 867–879 879
... The deposition ages of these sampled layers were determined to be ∼1.9 ka for KLK (Fan et al., 2012), ∼5.4 ka for TS (Fan et al., 2012), ∼3.1 ka for XCD (Sun et al., 2010), ∼10.4-11.8 ka for CK (Liu et al., 2008), and <33-57.9 ka for XTJ (Zeng & Xiang, 2017) by optically stimulated luminescence (OSL) and accelerator mass spectrometry 14 C dating. ...
... The stratigraphic ages of these 11 sections in this study are between 1.9 and 113 ka (Fan et al., 2012;Lai et al., 2014;Liu et al., 2008;Madsen et al., 2014;Sun et al., 2010;Zeng & Xiang, 2017). There is an obvious time gap between the deposition time and the present day. ...
Article
Full-text available
Plain Language Summary Knowledge of the present‐day crustal stress field is important for understanding seismic activity and plate tectonics. Earthquake focal mechanisms and borehole breakouts are the traditional methods used to investigate the present‐day stress field. The former depends on the frequency of seismic activities, the magnitude, and the location of each earthquake, none of which can be controlled or adjusted. The latter is time‐consuming and laborious. Magnetic fabric is an economical, rapid, and sensitive indicator of paleostrain. If the magnetic fabric of recently deposited mudstone is measurable, anisotropy of magnetic susceptibility (AMS) has the potential to be used to investigate the present‐day stress. In this study, we compared the magnetic fabric of freshly consolidated lacustrine mudstones with the present‐day stress field derived from traditional stress indicators and found a rough consistency between them. This suggests that magnetic fabric can be used as an indicator of the present‐day stress field, which will add much more new data to the world stress database.
... The Qaidam Basin lies on a major climatic boundary between the westerlies and the ASM Chen et al., 2020a), and Holocene climate evolution has long been considered a "westerlies model" (Chen et al., 2008;He et al., 2013;Song et al., 2020). However, the paleoclimate reconstruction in this study shows a typical "monsoon model", which is more consistent with moisture patterns in Qinghai Lake (Liu et al., 2002;Shen et al., 2005;An et al., 2012;Liu et al., 2017), Chaka Lake (Liu et al., 2008), and Gahai Lake (Chen, 2007;Li, 2008;Chen, 2010;Guo, 2012) in the Qaidam Basin ( Fig. 1A and 6a-c; Table 1). Both the effective moisture curve of monsoonal Central Asia (MCA) based on 75 paleoclimatic records and the speleothem δ 18 O data from Dongge Cave also show a similar climatic evolution as that of the Hurleg Lake area and are consistent with the Northern Hemisphere summer insolation (NHSI) (Fig. 6d, e and f) (Dykoski et al., 2005;Herzschuh, 2006). ...
... The decrease in NHSI during the Holocene may have greatly reduced the thermal contrast between the Asian continent and the surrounding oceans, resulting in Fig. 6. Comparison of the MAT and pH records in the core HL with ASM-dominated regions and westerlies-dominated regions (a, b); variation in TOC (%) in the Chaka Lake sediment core (c, Liu et al., 2008); mean effective moisture in the ASM area (d, Herzschuh, 2006); δ 18 O record of the Dongge Cave stalagmite (e, Dykoski et al., 2005); summer insolation at 35 o N (f, Laskar et al., 2004); peat α-cellulose δ 13 C temperature record from SSP (g, Rao et al., 2019); A/C ratio of pollen in Balikun Lake (h, Tao et al., 2010); north Xinjiang moisture index (i, Wang and Feng, 2013). the displacement of the Intertropical Convergence Zone (ITCZ) and a weakening of the ASM, and ultimately reduced precipitation in the ASM, and its marginal regions (Haug et al., 2001;Yang et al., 2021). ...
... The Center for Applied Isotopic Studies at the University of Georgia analyzed TOC samples. Researchers have satisfactorily dated TOC from similar bedded halite and gypsum sand beds (Cupper, 2006;Xingqi et al., 2008;Zhang and Liu, 2020). We placed saline pan samples in beakers with deionized water to dissolve halite crystals. ...
Article
Full-text available
The depositional history of the Bonneville Salt Flats, a perennial saline pan in Utah's Bonneville basin, has poor temporal constraints, and the climatic and geomorphic conditions that led to saline pan formation there are poorly understood. We explore the late Pleistocene to Holocene depositional record of Bonneville Salt Flats cores. Our data challenge the assumption that the saline pan formed from the desiccation of Lake Bonneville, the largest late Pleistocene lake in the Great Basin, which covered this area from 30 to 13 cal ka BP. We test two hypotheses: whether climatic transitions from (1) wet to arid or (2) arid to wet led to saline pan deposition. We describe the depositional record with radiocarbon dating, sedimentological structures, mineralogy, diatom, ostracode, and portable X-ray fluorescence spectrometer measurements. Gypsum and carbonate strontium isotope ratio measurements reflect changes in water sources. Three shallow saline lake to desiccation cycles occurred from >45 and >28 cal ka BP. Deflation removed Lake Bonneville sediments between 13 and 8.3 cal ka BP. Gypsum deposition spanned 8.3 to 5.4 cal ka BP, while the oldest halite interval formed from 5.4 to 3.5 cal ka BP during a wetter period. These findings offer valuable insights for sedimentologists, archaeologists, geomorphologists, and land managers.
... The material used in this study is a natural brine obtained from Chaka (CK) Lake (Xingqi et al., 2008), situated in the Qaidam Basin on the Qinghai-Tibet Plateau, as depicted in Fig. 1. The Qinghai-Tibet Plateau is renowned for its elevated altitude and is often referred to as the "roof of the world." ...
Article
Full-text available
Natural salt particles show higher SO2 uptake than pure NaCl particles. • The hygroscopicity is influenced by chemical complexity, SO2 exposure, and light conditions. • Light exposure suppresses hygroscopic growth, potentially leading to the formation of less hygroscopic species. • Natural salt particles exhibit rounded shapes with cubic NaCl cores encased in sulfate coatings. • Particles exhibit morphological and spectral changes after exposure to SO2 , light, and high RH. A R T I C L E I N F O Keywords: SO2 Salt particles Chamber Chemical morphology STXM Salt lake Ionic strength A B S T R A C T The interactions between SO 2 and natural salt aerosol particles represent complex and crucial dynamics within atmospheric processes and the broader climate system. This study investigated the SO 2 uptake, hygroscopicity, morphology and mixing states of natural salt particles, which are generated from brines sampled from the Chaka salt lake located in the Qinghai-Tibet plateau. A comparison with atomized pure NaCl particles is included as reference. The results show that NaCl particles exhibit the lowest SO 2 uptake, while Chaka salt particles demonstrate higher uptake due to their complex composition. The hygroscopicity of salt particles is influenced by several factors, including chemical complexity, SO 2 exposure and light conditions. In comparison to pure NaCl, Chaka salt displays higher hygroscopicity, which is further enhanced in the presence of SO 2. However, when exposed to light, mass growth is suppressed, suggesting the formation of species with lower hygroscopicity, such as Na 2 SO 4. Analysis of particle morphology and mixing states reveals notable distinctions between NaCl crystals and Chaka salt particles, where the Chaka salt particles exhibit rounded shapes with a structure composed of cubic NaCl cores surrounded by sulfate materials as a coating. In addition, the chemical morphology analysis also reveals that the particles show morphological and spectral changes before and after the exposure to SO 2 , light and high RH. Therefore, this research highlights the intricate interactions between SO 2 and natural salt aerosol particles in diverse environmental settings, underscoring their multifaceted impacts on atmospheric processes.
... Controlled by various atmosphere circulation systems, the Qilian Mountains are the sensitive zone for the synergy of the EASM and the WW, which is a key area for understanding the climate interaction (Wang K et al. 2005;Zhang et al. 2007) (Fig. 2). Pollen reconstructions from the Dunde ice-core in the western Qilian Mountains showed that the enhanced summer monsoon may extend beyond the northernmost monsoon margin and even reach the westernmost Qilian Mountains, resulting in high (Liu et al. 2008); (b) Ephedra percentages from a sediment core of Bosten Lake (Huang et al. 2009); (c) XARM/SIRM ratio of Xinjiang loess (Chen et al. 2016); (d) reconstructed precipitation of Achit Nuur ; (e) Lake Qinghai Westerlies climate index (WI, flux of >25 μm fraction) ); (f) δ 18 O in QH-2000 of Lake Qinghai (Liu et al. 2007); (g) δ 18 O of Sanbao Cave SB43 (Dong et al. 2010); (h) δ 18 O of Lianhua Cave (Zhang et al. 2013); (i) δ 18 O of Jiuxian Cave C996-1 and C996-2 (Cai et al. 2010); (j) δ 18 O of Dongge Cave D4 and Hulu Cave H82 (Wang et al. 2001 andYuan et al. 2004). ...
Article
Full-text available
The Qilian Mountains, located in the northeastern Qinghai-Tibet Plateau, is a sensitive zone of both East Asian summer monsoon (EASM) and westerly winds (WW). The evolution history and driving mechanism of the ecosystem and hydrologic cycle in this region on long-term timescales have not yet been clarified. In this study, we comprehensively study the hydrologic and ecological evolution history in the sensitive zone since the Last Glacial Maximum (LGM) by integrating surface sediments, paleoclimate records, TraCE-21ka transient simulations, and PMIP3-CMIP5 multi-model simulation. Results show that hydrologic and ecological proxies from surface sediments are significantly different from west to east and mainly divided into three sections: the monsoon-affected region in the eastern Qilian Mountains, the intersection region in the central Qilian Mountains, and the westerly-affected region in the western Qilian Mountains. Meanwhile, paleo-ecological and paleohydrologic reconstructions from the surroundings uncover a synchronous climate evolution that the EASM mainly controls the eastern Qilian Mountains and penetrates the central Qilian Mountains in monsoon intensity maximum, while the WW dominates the central and western Qilian Mountains on both glacial-interglacial and millennial timescales. The simulation results further bear out the glacial humid climate in the central and western Qilian Mountains caused by the enhanced WW, and the humidity maximum in the eastern Qilian Mountains controlled by the strong mid-Holocene monsoon. In general, east-west differences in climate pattern and response for the EASM and the WW are integrally stable on both short-term and long-term timescales.
... The results of the present study describe a common scenario where saline lakes experienced significant level drops during the transition from the last glacial to the Holocene, associated with deposition of evaporites [South-west United States (Li et al., 1996;Janick & Demicco, 2019;Olson & Lowenstein, 2021); Australia (Herczeg & Chapman, 1991); North-west China (Xingqi et al., 2008;Du et al., 2019)]. This emphasizes the contribution of this work for global palaeo-hydroclimatic interpretation of evaporite sequences, which can be aided with the presented (or even improved) coupled hydroclimatic-sedimentary modelling framework. ...
Article
Establishing accurate palaeo‐hydroclimatic reconstructions from lacustrine and marine archives is a long‐standing challenge in palaeoenvironment studies. Closed‐basin evaporites, and especially halite, record episodes of extremely arid conditions during rapid climate change. However, the complex limnological behaviour of deep hypersaline water bodies and the stochastic nature of the hydroclimatic regime and its variations limit detailed palaeo‐hydroclimatic interpretations from such records. Therefore, a mass‐balance model was developed to explore hydrology–limnology–sedimentology relationships in hypersaline environments under both deterministic and stochastic approaches that generates synthetic halite–mud sequences. Applying the model to the Holocene Dead Sea halites yields novel insights into palaeoenvironmental conditions in the Levant. The deterministic framework indicates that: (i) under a series of similar hydroclimatic cycles, the thickness of each subsequent halite interval decreases, due to the depletion of dissolved‐ions storage in the brine; (ii) halite deposition requires lake levels to drop below the minimal lake level of the preceding cycle; (iii) the time interval between halite deposition and the hydrological minimum is increasingly longer in subsequent cycles. Thus, counter‐intuitively, halite deposition mostly takes place as water discharge increases, providing that the water balance is still negative. The stochastic approach produced random sequences comparable to the observed Dead Sea sedimentary record. It demonstrates that some hydrological minima are not represented by halite deposition at all. Furthermore, the thickness and number of halite beds at each hydrological cycle vary substantially, depending on the specific hydrological conditions realized. Finally, these results imply that the major Dead Sea level drop at the pre‐Holocene deglaciation ( ca 14 ka bp ), previously assumed to be a record minimum, could not have been as pronounced as suggested, and must have been milder than the subsequent drop at the early Holocene ( ca 11–10 ka bp ).
Article
Full-text available
The northern Tibetan Plateau is a climatically sensitive zone influenced by monsoon and westerly winds. In summer, water vapor transport can reach Qinghai Lake and the eastern section of the Qilian Mountains; in winter, westerly winds mainly control the climate. This article compares the wet/dry changes in the region during the mid-Holocene (MH) warm period, the medieval climate anomaly (MCA), the current warm period (CWP), and the future warm period from the perspective of paleoclimate. We found that the MH warm period was mainly affected by the orbit-controlled East Asian summer monsoon, and the region showed warm and humid climate characteristics. The MCA was mainly controlled by solar radiation, and there was a warm and dry phenomenon. The CWP and the future warm period are mainly controlled by the rise in temperature caused by the increase in greenhouse gases, and the climate is becoming more arid. The wet/dry patterns in the CWP and the future warm period in the next century on the northern Tibetan Plateau are similar to those in the MCA. Continued warming will lead to the expansion of the westerly belt and a gradually humid climate. The future wet/dry changes will be more similar to the MH warm period. S. 2024. Transformation and mechanisms of climate wet/dry change on the northern Tibetan Plateau under global warming: A perspective from paleoclimatology. Science China Earth Sciences, 67, https://doi.
Article
The Tibetan Plateau (TP) is a key region for environmental and climatic research due to its significant linkages with large-scale atmospheric circulation. Understanding the long-term moisture evolution pattern and its forcing mechanisms on the TP during the Holocene may provide insights into the interaction between low-latitude climate systems and midlatitude westerlies. Here, we synthesized 27 paleoclimate proxy records covering the past 9500 years. The results of the rotated empirical orthogonal function analysis of the moisture variation revealed spatial-temporal heterogeneity, which was classified into 5 subregions. Proxy records were then compared with the results from the Kiel Climate Model and other paleorecords. The results showed that moisture evolution on the western-southern-central TP was controlled by the Indian summer monsoon (ISM). On the south-eastern TP, moisture change was affected by the interplay between the East Asian summer monsoon (EASM) and the westerlies, as well as the ISM. With diverse patterns of circulation system precipitation, moisture changes recorded in the paleorecords showed spatial-temporal discrepancies, especially during the early to middle Holocene. Moreover, given the anti-phase pattern of summer precipitation in the EASM area under El Niño/Southern Oscillation (ENSO) conditions and the unstable relationship between the ISM and ENSO, it is reasonable to conclude that relatively strong ENSO variability during the late Holocene has contributed to these discrepancies as Asian summer monsoon precipitation has declined.
Article
Full-text available
Based on the data from China's second soil census and the organic carbon data of 51 lakes, the organic carbon sedimentary models of lakes and soils were analyzed by dividing China into inflow areas and outflow areas. The results showed that there were different organic carbon deposition patterns in the inner and outer flow areas of China. The burial rate of organic carbon in lakes and the density of soil organic carbon in the inner flow area changed synchronously, and the organic carbon deposition pattern of the basin aggregation type was presented; the content and density of soil organic carbon in the outflow areas of the northeast and southwest were high while the content and deposition rate of lake organic carbon were low, showing an in-situ deposition pattern of organic carbon; the soil organic carbon data in the outflow areas of the central and eastern region was lower than that in other regions while the lake organic carbon data was much higher than that in other regions. In combination with the intensive human activities in those regions, the organic carbon deposition model thereof was defined as the type of human activities. To sum up, the deposition patterns of organic carbon in lakes and soils in China's inner and outer flow areas are quite different, and the inner part of the outer flow area is affected by different factors, showing two types of organic carbon deposition patterns: in-situ deposition and human activities.
Article
Halite precipitation with water and air temperature was observed in detail, and homogenization temperature of fluid inclusions in halite formed in ancient and modern Chaka Salt Lake was studied. Halite precipitates mainly in August every year and largely precipitates between 13 and 15 pm at one day when water temperatures reach 20°C but can seldom reach 30°C. Homogenization temperatures of fluid inclusions in halite formed in Chaka Salt Lake range from 14°C to 38°C with an average of 23.7°C. The number of inclusions appears an obvious peak value at homogenization temperatures between 18 - 25°C, which probably represent the water temperature in which halite mainly precipitates when water temperatures reach 20°C. Therefore, homogenization temperatures of fluid inclusions in halite formed in Chaka Salt Lake can well reflect the water temperature.
Article
Oxygen isotope records of five stalagmites from Hulu Cave near Nanjing bear a remarkable resemblance to oxygen isotope records from Greenland ice cores, suggesting that East Asian Monsoon intensity changed in concert with Greenland temperature between 11,000 and 75,000 years before the present (yr. B.P.). Between 11,000 and 30,000 yr. B.P., the timing of changes in the monsoon, as established with 230Th dates, generally agrees with the timing of temperature changes from the Greenland Ice Sheet Project Two (GISP2) core, which supports GISP2&apos;s chronology in this interval. Our record links North Atlantic climate with the meridional transport of heat and moisture from the warmest part of the ocean where the summer East Asian Monsoon originates.