Sie sind auf Seite 1von 15

bs_bs_banner

Lacustrine sedimentological and geochemical records for the last


170 years of climate and environmental changes in southeastern China
KANDASAMY SELVARAJ, BAO ZHI LIN, JIANN-YUH LOU, WEI LAN XIA, XIANG TONG HUANG AND CHEN-TUNG A.
CHEN

Selvaraj, K., Lin, B. Z., Lou, J.-Y., Xia, W. L., Huang, X. T. & Chen C.-T. A.: Lacustrine sedimentological and
geochemical records for the last 170 years of climate and environmental changes in southeastern China. Boreas .
10.1111/bor.12143. ISSN 0300-9483.
Reconstruction of modern climate and environmental changes in east Asia using inland natural climate archives
can provide valuable insights on decadalmultidecadal climate and environmental patterns that are probably
related to both natural and anthropogenic forcing. Here we investigated an 89-cm-long sediment core (TH1) from
Tian Lake, southeastern China, for sedimentological, physical and geochemical parameters in order to understand climate and environmental changes for the latest two centuries. 137Cs- and 210Pb-based age models show
that the fine sandcoarse silt-dominated core contains ~170 years (c. AD 18422011) of continuous sedimentation. Sediments with fine sands, low MS values, high water content, high TOC content and a high C:N ratio from
c. AD 1842 to 1897 suggest intense hydrological conditions and strong runoff in the catchment, probably because
of a humid climate. From AD 1897 to 1990, sediments with very fine sand and coarse silt, high MS values, low
water content and unchanged TOC and C:N ratios indicate normal hydrological conditions and in-lake algaederived organic matter. During this interval, the chemical weathering indicators show stronger weathering conditions compared with sediments deposited during AD 18421897, supporting the dominance of weathered surface
soil input in the earlier interval and physical erosion dominance in the later period, respectively. Since AD 1990,
the continuous decrease of geochemical proxies suggests human-interacted Earth surface processes in the catchment of Tian Lake. A PCA revealed four dominant geochemical controlling factors detrital input, trophic status,
grain size and early diagenesis , accounting for 26, 20, 18 and 16% of total variance, respectively. This study for
the first time provides lacustrine geochemical evidence for the most recent two centuries of climate and environmental changes in coastal southeastern China, a region that is currently undergoing an inversion of critical zone,
i.e. an overturning of its soil profile, owing to swift modernization.
Kandasamy Selvaraj (selvaraj@xmu.edu.cn) and Bao Zhi Lin, State Key Laboratory of Marine Environmental
Science, Xiamen University, Xiamen 361102, China; Jiann-Yuh Lou, Department of Marine Sciences, Chinese Naval
Academy, Kaohsiung 81345, Taiwan; Wei Lan Xia, Key Laboratory of Lake Sedimentation and Environment, Nanjing Institute of Geography and Limnology, Chinese Academy of Sciences, Nanjing 210008, China; Xiang Tong
Huang, State Key Laboratory of Marine Geology, Tongji University, Shanghai 200092, China; Chen-Tung A. Chen,
Department of Oceanography, National Sun Yat-sen University, Kaohsiung 80424, Taiwan; received 17th December
2014, accepted 16th August 2015.

Lake sediments have been widely regarded as an


important natural archive for the reconstruction of
past climatic and environmental changes over centennialmillennial time scales in both humid and arid
parts of China (Yancheva et al. 2007; Chen et al. 2010,
2013; An et al. 2012; Selvaraj et al. 2012). However,
similar studies on shorter, decadalmultidecadal time
scales focusing on the last millennium and part thereof
using lake sediments from this region are rare. This
may hamper our understanding of potential future
environmental changes, given the drastically changing
land-use pattern of China during the last two decades
that might have a great impact on the sedimentation
pattern of inland aquatic systems in the near future.
Investigation of the hydrodynamic characteristics and
erosion-weathering processes of a lake basin using multi-proxy analyses is common in palaeoenvironmental
research (Selvaraj et al. 2007; Nahm et al. 2010; Chen
et al. 2013). For example, grain-size variations of lake
sediments are generally used as a reliable proxy for
past changes in precipitation and eolian activity (e.g.
Yan et al. 2011; Qiang et al. 2014). Likewise, sediment
magnetic susceptibility (e.g. Warrier & Shankar 2009)

DOI 10.1111/bor.12143

and colour intensities (e.g. Moy et al. 2002) are used to


infer past climate changes. In addition, sedimentary
inorganic (Ti, Al, Si, K, Fe, Mg, Ca and other trace
elements) and organic geochemical proxies (TC, TOC
and C:N ratio) reflect information about physical erosion and chemical weathering processes as well as vegetation types in the catchment (Meyers 2004; Selvaraj
et al. 2007, 2011, 2012; Giguet-Covex et al. 2011; Liu
et al. 2014). Hence, a combination of diverse geochemical parameters is useful to extract information about
past climate and environmental changes depending
upon the sediment accumulation rate as well as the
length of sediment core under investigation.
A substantial increase in climate and environmental
reconstructions using lake sediment records has
recently been achieved in China (e.g. Yancheva et al.
2007; Chen et al. 2010; An et al. 2012). However, most
researchers have mainly focused upon longer-term
climate changes, with only a few studies addressing
environmental changes at shorter, centennialdecadal,
time scales, especially in southeastern China, including
Taiwan Island (Lou & Chen 1997; Selvaraj et al. 2007,
2011, 2012; An et al. 2011), with a complete lack of

2015 Collegium Boreas. Published by John Wiley & Sons Ltd

Kandasamy Selvaraj et al.

geochemical investigation on island lakes along the SE


coast of China. There are around 6500 islands along
the 18 000 km coastline of China, of which 93% are
considered as not suitable for human settlement
because of an acute shortage of fresh water. Nonetheless, Yushan Island, located in Fujian Province
(Fig. 1A), is one island that has a lot of fresh water to
support human habitation, with a current population
of ~6500. Tian Lake, a relatively pristine lake and an
important source of fresh water in this island, has
never been investigated either in terms of water chemistry or sediment geochemistry (Fig. 1B). Here, we
used an integrated sedimentological, physical and geochemical approach to infer the modern climate and
environmental changes of Tian Lake using a short sediment core collected from the lake in May 2012. We
aimed to investigate (i) the hydrological changes in the
catchment surrounding the lake; (ii) the trophic status
within the lake; (iii) the factors that drive the sediment
input; and (iv) the processes that control the geochemical composition of sediments.

Study area
In China, the subtropical region falls between ~22 and
34N latitude and is bordered by the Qingling Mountain and Huaihe River in the north, the Leizhou Peninsula in the south and extends to the Hengduan
Mountain in the west (Wu 1980). Fujian Province,
located at the southeastern boundary of the subtropical area (Fig. 1A), consists mostly of mountains with
undulated topography and faces the East China Sea to
the southeast. Tian Lake (latitude 26560 52N, longitude 120200 56E) is located on Yushan Island of Fuding City, northeast of Fujian Province and Yushan
Islands southeast border is the East China Sea
(Fig. 1A). Covering an area of 21 km2, Yushan Island
A

BOREAS

is the seventh largest island in the province and has a


coastline of ~30 km surrounded by the sea. The local
bedrock of Fuding City consists of various rock types,
including Carboniferous, Jurassic, Cretaceous and
Quaternary ages, of which Jurassic and Cretaceous
rocks are most developed, covering an area of about
1200 km2, and accounting for ~78% of the total land
area. Yushan Island is underlain mainly by Jurassic
granites, dark grey intermediate acidic dacite and tuff
lava (source: Compiling Team for Annals of Fuding
County). The morphology of Yushan Island is largely
defined by an abrasion landmass covered mostly with
both coarse- and fine-grained granites (Fig. 2AC).
Yushan Island is strongly influenced by subtropical
oceanic monsoons and thus has a mild and moist climate, with mildly cold winters and humid summers.
The rainfall is controlled dominantly by the East Asian
monsoon, and has an annual mean of ~1300 mm. The
mean annual evaporation is 1270 mm. The annual
mean temperature is 15.1 C and the relative humidity
is ~100%. The average number of foggy days per month
is 6.2 days and the soil is mostly fertile red clay type,
which is suitable for grass growth. Therefore, Yushan
Island has widespread grassland (Fig. 1B), covering an
area of around 667 km2. The dominant species of
grass is Miscanthus floridulus. Affected by the topography, the land hydrology in this island is mostly
recharged from both groundwater and rainfall.
Oceanographically, Yushan Island is a region of regular semidiurnal tide with mean annual macro-tidal
variation of 5.015.83 m and mean neap tide varying
between 3.37 and 4.10 m. Yushan Island contains
three natural lakes, namely Large Tian Lake, Small
Tian Lake and Jiuzhuxiachao Lake, with a combined
gross water storage capacity of up to 1.7 million m3.
Tian Lake has an average altitude of 306 m a.s.l., with
minimum and maximum altitudes of 303 and 310 m
B

Fig. 1. A. Map showing the location of Tian Lake on Yushan Island (black dot), southeastern China. B. A panoramic view of Tian Lake. Red
arrow shows the location of the sediment core TH1 investigated in the present study.

Climate and environmental changes, SE China

BOREAS

Fig. 2. A. Grey granite rocks on the mountain surrounding Tian Lake. B. Pink granite boulders, with a broken surface, on the mountain
slope. C. Hammered rock pieces of coarse-grained intermediate dacite along the mountain slope.

a.s.l., respectively. As the lake direction is the same


as the rock crevice direction, Tian Lake has been
categorized as a tectonic fissure lake. Small and Large
Tian lakes have areas of 0.1 and 0.7 km2, respectively,
with maximum depths of around 5 and 18.5 m. As
Large Tian Lake has recently been dammed artificially
for drinking and agricultural purposes, we selected the
pristine Small Tian Lake, which is surrounded by
>100200 m high hills (Fig. 1B), for sediment coring,
aiming to investigate climate and environmental
changes for the last ~170 years.

Material and methods


Four short (<1 m long), push sediment cores were
manually collected from chest-deep waters in Tian
Lake during May 2012, using 4-cm diameter transparent plastic tubes. Amongst them, the longest sediment
core (89 cm), named as TH1, was chosen for this study
(Fig. 4). The core was subsampled at 1-cm intervals,
except for the top depth of 12 cm and bottom depth
of 8687.5 cm, for sedimentological (grain size), physical (magnetic susceptibility and red, green and blue
(RGB) colour intensities) and organic (total carbon
(TC), total organic carbon (TOC), total nitrogen
(TN)), as well as inorganic (major and trace elements)
geochemical analyses.
Prior to the detailed laboratory work, all subsamples
were freeze-dried for 24 h to achieve a constant weight
and were homogenized. A low-frequency (0.47 kHz)
magnetic susceptibility analysis was performed using a
Bartington MS-2B sensor attached to a MS-2 susceptibility meter, and each sample was measured in triplicate. Grain-size analysis was carried out with a Coulter

LS-100 laser particle size analyser, as detailed in Janitzky (1987). Briefly, samples were sieved to remove
grains larger than 1000 lm and then treated with HCl
and H2O2 to remove carbonate and organic matter,
respectively. In the final stage, sodium hexametaphosphate was added to prevent the flocculation of clay
minerals and the samples received a short period of
ultrasonic agitation. RGB colour intensities of samples
were measured by digital colour scale. Grain size, magnetic susceptibility and colour intensities were measured using instruments located at the Department of
Marine Sciences, Chinese Naval Academy, Taiwan.
Forty-nine bulk sediment subsamples from every 2cm interval were ground to  200 micrometer in an
agate mortar for geochemical analysis. Sediments were
acidified with 1 N HCl to remove carbonates and then
rinsed three times with distilled water, dried at 50 C
and homogenized using a mortar and pestle. Precisely
20 mg of sediment was weighed into a tin boat, which
was then folded into a pellet. These pellets were measured for TOC and TN using a Vario EL III elemental
analyzer (Selvaraj et al. 2007). The standard reference
materials used were BCSS (2.24% C, 0.24% N) and
NIST2704 (3.34% C, 0.22% N). C:N ratios were calculated as molar proportions. Selected major (Al, Ca, Fe,
Mg, Mn, Na, K, P, S, Si and Ti) and trace (Ba, Rb, Sr
and Zr) elements were determined using an X-ray fluorescence (XRF) spectrometer equipped with an Rh
tube located at the State Key Laboratory of Marine
Geology, Tongji University, China. Details of the XRF
method have been described in Selvaraj & Chen (2006)
and Selvaraj et al. (2010). Activities of 210Pb and 137Cs
radionuclides in subsamples of core TH1 were measured with EG and G Ortec gamma spectrometry using

Kandasamy Selvaraj et al.

BOREAS

a germanium detector located at the Institute of Limnology and Geography, Chinese Academy of Sciences,
Nanjing, China (Yao et al. 2008).

Results and discussion


Chronology: 137Cs dating
The radioisotope caesium-137 (137Cst1/2 = 30.1 years)
is a fission product and was initially introduced into
the environment in significant quantities as a result of
atmospheric nuclear weapons tests conducted in the
early 1950s (Nahm et al. 2010). Widespread global
dispersal of 137Cs to the environment began with
high-yield thermonuclear tests in November 1952  2
(Perkins & Thomas 1980). Also, the maximum fall-out
of 137Cs has been observed in the Northern Hemisphere during 1963, and in the Southern Hemisphere
during 1964 (Longmore et al. 1983). The fall-out of
137
Cs from the 1986 Chernobyl accident was mainly
observed at sites in Europe and has been used as a time
marker for the sediment deposited during that year
(Wieland et al. 1993; Fan et al. 2010).
The profiles of 137Cs in coastal sediments of China
generally have two to three distinct chronological
markers (Chen et al. 2009; Sun et al. 2011). As for
many places in the Northern Hemisphere, the time
markers include the onset of measurable amounts in
soils in 1954 and the peak fall-out in 1963. In some
areas of China, there may also be another smaller peak
of 137Cs fall-out that probably derives from the Chernobyl fall-out in 1986 (Xiang 1998; Zhang 2005). However, in some situations, the 1954 time marker may give
a false result because of the detection limit of 137Cs ,
which has a value of ~0.5 Bq kg1 (Campbell 1983).
Figure 3 shows the variations in 137Cs activity with
depth for our sediment core TH1. The 137Cs pattern is
137

Cs,

226

Ra,

210

very similar to the theoretical fall-out pattern, suggesting that such processes might not have occurred by
chance (Grousset et al. 1999). The first peak, which
had the highest 137Cs activity (8.33 Bq kg1), may
record the nuclear weapons fall-out peak that occurred
in 1963, which appeared at 2526 cm in core TH1. Secondary small peaks at around 1314 cm (~6.05 Bq
kg1) are likely to have been caused by the Chernobyl
fall-out. As sampling was carried out in May 2012, we
assumed that the surface layer of core TH1 corresponds to the sediment accumulated in AD 2011. The
average sedimentation rate (SR) deduced from the
1963 time marker is 0.53 cm a1. The SR derived from
the 1986 time marker is 0.54 cm a1. It is apparent
from Table 1 that the SR given by the 1986 marker is
the same as the SR determined from the 1963 peak
137
Cs activity.
210

Pb dating

Lead-210 (210Pbt1/2 = 22.3 years) is a natural


radioactive by-product of the 238U decay series that is
supplied to the environment via atmospheric precipitation (e.g. Selvaraj et al. 2010). The 210Pb activity of
lake sediments has two components, a supported component (210Pbs) derived from 222Rn decay within the
Table 1. Comparative results of sedimentation rates calculated using
137
Cs and 210Pb dating methods in sediment core TH1 from Tian
Lake, southeast China.
Time of
peak activity

Depth
(cm)

Time-span
(year)

Sedimentation
rate (cm a1)

13.5
25.5
1341

19862011
19632011
19321985

0.54
0.53
0.51

137

Cs
1986
1963
210
Pbex

210

Pb (Bq kg )

10 100 200 300 400 500

100

Pbe x (Bq kg1)


200

300

400

500

10

10

15
20

30

Depth (cm)

Depth (cm)

20

40
50
60
70
80
90

137
CS
210
Pbex
226
Ra
210
Pbtotal

25
30
35
40

y = 112.33 - 16.51Inx
R2 = 0.84
SR = 0.51 cm a 1

45

Fig. 3. Down-core variations in 137Cs, 226Ra, excess 210Pb (210Pbex) and total 210Pb activities in sediment core TH1 from Tian Lake. The profile of 210Pbex is plotted on a logarithmic scale for the depth of 1341 cm, giving a sedimentation rate of 0.51 cm a1.

BOREAS

sediment column, and an unsupported (or excess) component (210Pbex) derived from the atmospheric fall-out
of 210Pb. 210Pbs can, for most purposes, be approximated by the 226Ra concentration. In the absence of
210
Pb fall-out, 210Pb and 226Ra would be in radioactive
equilibrium. 210Pbex is determined by subtracting
226
Ra from the total 210Pb concentration (Appleby &
Oldfield 1983). The unsupported 210Pb concentration
in each sediment layer declines with its age in accordance with the usual radioactive decay law. This law
can be used to calculate the age of the sediment provided that the initial unsupported 210Pb concentration
when deposited on the lake bottom can be estimated
(Appleby & Oldfield 1983).
Figure 3 shows the activity profiles of 210Pbex, 210Pbt
and 226Ra in core TH1. The profile of excess 210Pb
(210Pbex) shows a tendency to decrease with depth to
~41 cm, but generally homogenous activity occurs at
4186 cm intervals, which is supported by the decay of
226
Ra. The profile of 210Pbex activity at the depth of
1341 cm shows an exponential decrease with minor
fluctuations (Fig. 3). The age of sediments of depth m
is thus calculated using the equation: t = k/ln (0/A),
where t is the age of sediments of depth m; 0 represents the total excess 210Pb in the sediment column; A
is the cumulative residual 210Pb in the depth m in the
sediment and k is the 210Pb radioactive decay constant
(0.03114 a1). The SR calculated from 210Pbex is
0.51 cm a1 and this value is very similar to the SR
(0.53 cm a1) derived from the 137Cs dating (Table 1).
Figure 3 also shows that 210Pbex profile is nonlinear
from the surface to the bottom. This nonlinearity suggests that either sediment accumulation rates have
varied slightly over time, or that a number of other factors are involved, such as migration of 210Pb through
interstitial waters near the sedimentwater interface
(Koide et al. 1973), mixing of near-surface sediments
by physical (Petit 1974) or biological (Robbins et al.
1977) processes, and postdepositional redistribution of
sediments either discontinuously through slumping
(Edgington & Robbins 1976) or more or less continuously by sediment erosion. The disequilibrium between
210
Pbt and 226Ra activity at the bottom of the core may
mean that the lowermost appearance of 210Pb has not
been reached in our sampling core.
Considering the uncertainty of 210Pb dating because
of the above-mentioned factors and the close similarity
of the SRs obtained between the 137Cs and 210Pb methods (Table 1), we adopted 137Cs dating to calculate the
age of core TH1, given that the peak 137Cs activity in
our sediment core is consistent with the peak 137Cs
activity of the Northern Hemisphere reported in a
number of previous studies (Zhang et al. 2005; Arnaud
et al. 2006; Liu & Fan 2011). If we assume that the surface layer of core TH1 corresponds to AD 2011, then
the bottom of the core (89 cm) corresponds to about
AD 1842 (AD 1836), according to a linear extrapola-

Climate and environmental changes, SE China

tion using the SR of 0.53 cm a1 (0.51 cm a1) derived


from the 137Cs (210Pb) method.
Lithology and magnetic susceptibility (v)
Variations in the grain size of lake sediments may be
influenced by changes in lake level, palaeoenvironmental zones of deposition, and transport energy related to
variations in rainfall and/or wind (Chen et al. 2004;
Conroy et al. 2008). As the relative humidity in
Yushan Island is ~100%, the grain size of Tian Lake
may be controlled by overland flows from the catchment. Magnetic susceptibility is widely used to provide
information on the concentration of ferromagnetic
minerals present in soils and sediments (Dearing
1999). Magnetic properties of soil and sediments
depend not only on minerals inherited from parent
rocks, but also on pedogenic processes such as weathering and biogenic/authigenic formation of minerals
(Ortega-Guerrero et al. 2004).
Lithologically, core TH1 consists of find sand, silty
sand and sandy silt (Fig. 4). Mean contents of sand,
silt and clay are about 72, 25 and 4%, respectively,
throughout the core. The mean grain size ranges from
34 to 327 lm with a mean value of 129 lm, indicating
the dominance of finevery fine sand and coarse silt.
The median grain size varies between 35 and 618 lm
with a mean value of 170 lm (Fig. 4). The trend of
median grain size is similar to the mean grain size.
There is not much change in the composition of sediment in the bottom 15 cm and top 13 cm. In the 67
13 cm interval, which roughly equals the time interval
from AD 1885 to 1986 (AD 18801985), the grain size
fluctuated significantly, with two peaks at ~29.5
41.5 cm and ~51.555.5 cm. The average v value in
our core TH1 is 3.2 SI. v is close to zero from the base
of the core to ~70 cm. The middle section (7013 cm)
of the record is characterized by fluctuating v values
(0.811.5 SI), whereas the upper portion of the record
shows a decreasing trend towards the core top.
TC, TN, TOC and C:N ratio
The atomic ratio of total organic carbon to total nitrogen (C:N ratio) of organic matter is little influenced by
decomposition and can be a reliable proxy for organic
matter sources (algae-derived or land-derived), albeit
with some small early diagenetic alterations (Meyers
2004). Algae generally have atomic C:N ratios <10,
whereas land plants commonly have ratios >20 (Meyers & Ishiwatari 1993). The TOC concentration can
provide important information about variations in
organic matter derived from land plants, which is
brought into the lake via runoff when precipitation is
high (Meyers et al. 2006; Selvaraj et al. 2007). Therefore, a combination of high TOC content and high
C:N ratios is probably related to higher inputs of soil

Kandasamy Selvaraj et al.

BOREAS

Fig. 4. Litho-log of core TH1 along with down-core distributions of clay (<4 lm), silt (464 lm) and sand (>64 lm) contents (all in %) as
well as mean grain size (Ms, lm), median grain size (Md, lm) and magnetic susceptibility (v in SI) against year (in AD). The two horizontal
shaded bars mark the dominance of medium-size sand in core TH1 during AD 19061913 and AD 19351953, suggesting a stronger erosion
in the catchment, probably as a result of the influence of typhoons in SE China.

material accompanied by organic matter from the


catchment during enhanced rainfall. By contrast, a
small amount of land plant material was delivered
when rainfall was low and during which the static
water column may also favour aquatic productivity
and thus sediments in general have low C:N ratios.
The TOC content in core TH1 ranges from 0.5 to
10.9%, with an average TOC content of 2.3% (Fig. 5).
In the bottom ~10 cm, the TOC content is high and
ranges from 4.3 to 10.9%, with large oscillations compared with the top 78.5 cm of the core, where the TOC
content is comparatively less, ranging from 0.5 to
3.5%. TOC shows an increasing trend from 78.5 cm,
corresponding to AD 1863 (AD 1857). TN content
(0.090.53%) in the sediment core varies in concert
with TOC toward the top of the core. C:N ratios are
high in the bottom ~10 cm of the core, whereas in the
top 78.5 cm of the core they remain mostly at ~810

with fewer fluctuations. The TN shows a strong linear


correlation with TOC (R2 = 0.88, p < 0.0001; Fig. 6)
with a positive intercept on the TN axis for sediments
from the upper 78.5 cm, suggesting the presence of a
small amount of inorganic N (~0.03%) in TN.
TOC-rich sediments from the singular indicate larger
proportions of land-derived organic matter, as also
evidenced by higher C:N ratios in these sediments
(Fig. 5), although they contain proportional content
of inorganic N (~0.17%; Fig. 6), probably because of
the presence of clay-associated N in terrestrial materials in the bottom part.
Proxies of chemical weathering
As soils and rocks in the catchment represent major
sources of terrigenous input into the lake (Bertrand
et al. 2005), elemental concentrations in sediments are

Depth (cm)

TN (%)
5

8 12

0.0

0.2

0.4

0.6

2011

10

1992

20

1972

30

1954

40

1935

50

1916

60

1897

70

1878

80

1859

Age (year AD)

TC (%)
0

1842

90
0

TOC (%)

5 10

3 4 5 6 7 8 9 10

16

C:N ratio

Fig. 5. Total carbon (TC), total organic carbon (TOC), total nitrogen (TN) and the TOC to TN (C:N) ratio vs. depth and age in sediment
core TH1.

Climate and environmental changes, SE China

BOREAS

0.6

0.4
0.3
0.2

R2 = 0.70, p = 0.0197 (red)


R2 = 0.88, p < 0.0001 (blue)

0.1
0.0
0

10

12

TOC (wt. %)
Fig. 6. Bi-plot showing the relationship between TN and TOC. Red
squares are sediments from the base of the core to 78.5 cm, whereas
blue squares represent sediments from 78.5 cm to the core top.

helpful to reconstruct the intensity of chemical weathering, an important process in the global hydro-geochemical cycle of elements (Gaillardet et al. 1999;
Warrier & Shankar 2009; Gupta et al. 2011). As this
process is strongly influenced by rainfall and temperature (Ollier 1984; Gaillardet et al. 1999), humid tropical and subtropical climates with denser vegetation in
general enhance the weathering of source rocks (Selvaraj & Chen 2006). Conversely, in arid or polar climates, weathering intensity is weak to non-existent
owing to either low precipitation or water locked in ice
and thus erosion predominates over weathering with
few to no new secondary mineral formation (Schoenborn & Fedo 2011).
K/Al
0.28

0.24

The degree of weathering can be obtained by calculation of the chemical index of alteration (CIA):
CIA = 100Al2O3/(Al2O3+CaO*+Na2O+K2O) (Nesbitt
& Young 1982), and the plagioclase index of alteration
(PIA): PIA = 100[(Al2O3K2O)/(Al2O3+CaO*+Na2O
K2O)] (Fedo et al. 1995), where CaO* is the amount of
CaO incorporated in the silicate fraction. These indices
indicate the degree of decomposition of total feldspar
(both alkali and plagioclase) and plagioclase feldspar,
respectively, to secondary clay products, where CIA
(PIA) values of about 50 indicate unweathered bedrock
(plagioclase), and values of 75100 indicate complete
conversion of feldspars to aluminous clay minerals (intense chemical weathering; see Fedo et al. 1995). Intensely weathered rock thus yields kaolinite or gibbsite
with CIA/PIA values of 100 (Fedo et al. 1996, 1997).
Similar to CIA and PIA, ratios such as Rb/Sr, K/Al
and Al/Na are also good proxies of chemical weathering intensity (Selvaraj & Chen 2006; Clift et al. 2008;
Selvaraj et al. 2010), as increasing weathering rapidly
leaches Sr (Na) compared with Rb (Al) (Nesbitt &
Young 1982) and, therefore, the Rb/Sr and Al/Na
ratios increase in the weathered product (Ma et al.
2000). K/Al has been used as a proxy for the illite/
kaolinite ratio (Yarincik & Murray 2000), as illite represents the dominance of physical weathering in temperate to arid climates, whereas kaolinite indicates
intense weathering in tropical, humid climates (e.g.
Bonatti & Gartner 1973).
As shown in Fig. 7, the CIA values for core TH1
range from 79 to 86 with an average of 84. Such high
CIA values indicate that most primary feldspar minerals in the catchment have been converted to aluminous
clay minerals (Fedo et al. 1996). Consistent with the
Plagioclase index of alteration-PIA

0.20

0.16

89 90 91 92 93 94 95 96 97
2011

0
10

1986-2011

1992
1972

Depth (cm)

20
30

1900-1986

1954

40

1935

50

1916

60

1897

70

1878

1842-1900

1859

80

1842

90
78 79 80 81 82 83 84 85 86 87
Chemical index of alteration-CIA

Age (year AD)

TN (wt. %)

0.5

30 40 50 60 70 80 90 100
Al/Na

1.2

1.4

1.6 1.8
Rb/Sr

2.0

2.2

Fig. 7. Age vs. chemical index of alteration (CIA), the ratio of K to Al (K/Al), the ratio of Al to Na (Al/Na), plagioclase index of alteration
(PIA) and the ratio of Rb to Sr (Rb/Sr) in sediment core TH1. The horizontal shaded bar represents the interval of AD 19001986.

Kandasamy Selvaraj et al.

CIA values, the PIA values for core TH1 vary from 89
to 96 with an average of 94, indicating that all primary
plagioclase feldspars have been converted to secondary
clay minerals. High Rb/Sr ratios (1.212.05) with a
mean value of 1.84, which is higher than the Rb/Sr
ratios of different shales (0.800.88; Selvaraj & Chen,
2006), supporting the idea that source rocks in the
catchment might have been intensively weathered
because of favourable climatic conditions (high rainfall
and high temperature) in tropical and subtropical
regions (Selvaraj & Chen 2006). Rb/Sr ratios in illite
minerals are usually >1, with a maximum value of 6.46
(Chaudhuri & Brookings 1979), suggesting that the
high illite content in the lake sediments might be
responsible for the high Rb/Sr ratios. Values for K/Al
range from 0.16 to 0.27, averaging 0.21. The variations
in K/Al follow CIA, with the ratio increasing during
warm periods and decreasing during cold periods. The
degree of chemical weathering of lake sediments is
high, as shown by their very high Al/Na ratios (3295
with an average value of 61), resulting from the
extreme dissolution of plagioclase (Fig. 7). This weathering condition is comparable to lake sediments from
the sub-alpine region of nearby Taiwan Island, which
show extreme weathering because of the high precipitation and dense vegetation (Selvaraj & Chen 2006; Selvaraj et al. 2007). This similarity also suggests regional
uniformity of chemical weathering of rocks in subtropical Taiwan and SE China, although the catchment of
Tian Lake is devoid of meta-sedimentary rocks, which
are dominantly present in Taiwan Island.
Principal component analysis (PCA)
Geochemical factor analyses have been used for determining the factors that control the sources of sediment
(Selvaraj et al. 2010) and processes that control the
geochemical composition of coastal and river sediments (Chen & Selvaraj 2008). Therefore, to disentangle significant grouping of factors and relationships to
possible geochemical controlling processes, all of the
sediment geochemical data, including grain size, magnetic susceptibility and colour intensities, were subjected to factor analysis using SPSS software (version
17.0, Chicago, IL, USA). The factors were extracted
using the Varimax rotation scheme with Kaiser normalization.
The results reveal that four factors account for 80%
of the total variance (eigenvalue >2, Table 2). Factor 1
accounts for 26% of the variance and Ti, Al, Mg, K,
Pb, Ba, Rb and Sr are positively loaded in this factor.
The strong positive correlations of Ti (r = 0.85), Ba
(r = 0.79), Pb (r = 0.75), Rb (r = 0.74), Sr (r = 0.62),
K (r = 0.79) and Mg (r = 0.83) with Al indicate that
they are associated with aluminous clay minerals such
as kaolinite, illite and/or smectite Table 3. In general,
all these elements showed a similar pattern over the

BOREAS

Table 2. Varimax rotated factor matrix (n = 49). Significant loadings (>0.6) are in bold.
Variable

Factor 1

Factor 2

Factor 3

Factor 4

Si
Ti
Al
Fe
Mg
Ca
Na
K
P
S
Mn
Pb
Ba
Zr
Rb
Sr
MGS
Md
Clay
Silt
Sand
Red
Green
Blue
MS
Eigenvalue
Variance (%)

0.17
0.78
0.90
0.28
0.69
0.08
0.06
0.71
0.16
0.09
0.43
0.83
0.94
0.40
0.83
0.83
0.04
0.13
0.22
0.24
0.24
0.36
0.39
0.43
0.04
6.49
26

0.12
0.23
0.23
0.14
0.10
0.75
0.24
0.21
0.51
0.69
0.03
0.30
0.08
0.41
0.38
0.03
0.13
0.04
0.49
0.24
0.28
0.85
0.83
0.80
0.83
5.09
20

0.13
0.15
0.05
0.22
0.12
0.23
0.23
0.05
0.11
0.20
0.22
0.14
0.07
0.17
0.17
0.03
0.95
0.91
0.73
0.90
0.89
0.22
0.21
0.20
0.08
4.42
18

0.90
0.31
0.24
0.86
0.52
0.08
0.71
0.46
0.53
0.21
0.60
0.14
0.03
0.69
0.26
0.37
0.11
0.06
0.01
0.02
0.02
0.12
0.05
0.10
0.01
4.12
16

whole core (Fig. 8). Ti, Rb, Ba, Sr, K and Mg are in
general incorporated into clay minerals during chemical weathering, whereas Ca and Na tend to be leached
(Nesbitt et al. 1980; Fedo et al. 1996). These elements
are mainly derived from the catchment and thus
related to detrital input into the lake.
Variables such as intensities of red, green and blue
colours, and magnetic susceptibility are positively
loaded and Ca and S are negatively loaded in factor
2, which accounts for 20% of the total variance. The
relationships between colour and the physical and
chemical characteristics of soil, such as iron, organic
matter, moisture content, mineralogy and structure,
have been studied (Krishnamurti & Satyanarayana
1971; Davey et al. 1975). It is evident from Fig. 9
that Ca and S may be responsible for the low values
for colour intensities as well as for magnetic susceptibility in the bottom 10 cm of the core. The covariance of Ca, S, RGB and magnetic susceptibility
suggests that similar environmental conditions caused
their accumulations/variations. High C:N ratios are
generally associated with land plants and the increase
in sediment TOC in the bottom part of the core may
reflect an increase in land-derived organic matter in
the lake (Fig. 5). When land-derived organics
increase, lakes become anoxic, resulting in high TOC
preservation in sediment and thus reducing sediment
magnetic susceptibility values. Such an interpretation
receives support from the S profile in our study.

1500

1000

2000

5500

2.4

2000

2500

5000

2.2

3000

3000

4000

3500

5000

1.6
180

6000
300

160

250

140

70

120
60

100

4000
19862011

1900-1986

4500
240

1842-1900

200

200

160

150
120

100

80

50

80

40
50

60
750

0
60

45

700

50

40
14
12
10
8
6
4
2
0

650

40

600

35

550

30

500

25

450

20
0.60

400

Red

50

Ba (ppm)

Pb (ppm)

Sr (ppm)

80

1.8

40
30
20
10
0

0.55
0.45

12

0.40
10

0.35
0.30

10

20

30

40

50

60

70

80

90

Depth (cm)

14

0.50

Green

1842-1900

1900-1986

Blue

19862011

3500
90

MS (SI)

2.0

Rb (ppm)

4500

S (ppm)

2.6

K (%)

2.8

6000

Al (%)

Ti (ppm)

6500

4000

Mg (%)

Ca (ppm)

Climate and environmental changes, SE China

BOREAS

Fig. 9. Down-core variations in elements of Factor 2, representing


lake trophic levels and/or pedogenesis. The vertical shaded bar represents the interval of AD 19001986.

0.25

6
0

10

20

30

40

50

60

70

80

90

Depth (cm)
Fig. 8. Down-core variations in elements of Factor 1, representing
the detrital input. The vertical shaded bar represents the interval of
AD 19001986.

Therefore, factor 2 may be interpreted as a representation of the trophic status.


Factor 3 explains about 18% of the total variance
and has negative loadings for clay and silt contents
and positive loadings for sand content, mean grain size
and median grain size. As variations in the grain-size
data are primarily attributed to catchment runoff variability, factor 3 can be interpreted as representing rainfall variations in the study area. Two peak values of
mean and median grain sizes around AD 19061913
and AD 19351953 (Fig. 4) perhaps seem to be attributed to the influence of typhoons in the study area,
given that heavy downpours associated with tropical
storms can transport coarser sediments, although a
causal link needs to be clarified in future studies.
Factor 4 accounts for 16% of the total variance and
shows positive loadings with Fe, Na, Zr and Mn and
a negative loading with Si, probably suggesting the

influence of Fe-Mn oxyhydroxides in the sediments.


Figure 10 shows enrichment of Fe and Mn in the surface sediment, which may be attributed to the redoxrelated diagenetic recycling of iron and manganese in
sediments (Boyle et al. 1999; Boyle 2001). Mn and Fe
in sediments are dissolved under reducing conditions,
migrate upward and accumulate in the oxidized surface
sediments. Si is not only associated with many aluminosilicate minerals, but is also related to diatom productivity as a component in their frustules (Peinerud
2000). Based on poor correlations with K, Ti and Rb
(Table 3), Si behaviour seems to be controlled by primary productivity. The negative relationship between
Si and Zr found here substantiates a biogenic source of
Si (Table 3).
Consistently, the depth profile of the Si:Al ratio
(Fig. 11), an approximate indicator of biogenic silica,
shows that the Si:Al ratio of the lake sediments is
mostly less than the mean ratio of 4.13 obtained from
eight rock samples investigated in the catchment
(range: 3.074.66). The plot also includes the Si:Al
ratio of two soil samples (2.01 and 2.32) investigated
from the catchment with a mean ratio of 2.17. The
Si:Al ratio in sediments depends upon the Si:Al ratio
of the settling suspended particles. If the suspended
matter is dominated by detrital particles (i.e. physically
eroded soil or bedrock particles), the Si:Al ratio will be

Si
Ti
Al
Fe
Mg
Ca
Na
K
P
S
Mn
Pb
Ba
Zr
Rb
Sr
MGS
Md
Clay
Silt
Sand
Red
Green
Blue

0.11

Ti

0.01
0.85

Al

Mg

0.27
0.89
0.83
0.49

Fe

0.77
0.35
0.40

0.03
0.31
0.33
0.20
0.31

Ca
0.63
0.23
0.20
0.40
0.37
0.32

Na
0.29
0.80
0.79
0.47
0.89
0.41
0.28

K
0.52
0.33
0.37
0.44
0.32
0.38
0.55
0.38

P
0.31
0.23
0.15
0.08
0.27
0.69
0.45
0.33
0.33

S
0.49
0.36
0.52
0.79
0.40
0.07
0.27
0.45
0.40
0.15

Mn
0.18
0.64
0.75
0.05
0.46
0.32
0.04
0.48
0.33
0.09
0.31

Pb
0.16
0.66
0.79
0.26
0.62
0.03
0.10
0.69
0.08
0.08
0.32
0.70

Ba
0.61
0.20
0.28
0.61
0.02
0.39
0.26
0.07
0.02
0.20
0.27
0.63
0.34

Zr
0.31
0.65
0.74
0.11
0.51
0.44
0.06
0.62
0.23
0.26
0.14
0.87
0.77
0.75

Rb
0.39
0.44
0.62
0.02
0.24
0.09
0.22
0.31
0.06
0.14
0.18
0.82
0.81
0.62
0.81

Sr
0.19
0.10
0.01
0.32
0.04
0.31
0.15
0.02
0.13
0.26
0.27
0.15
0.08
0.29
0.22
0.04

MGS
0.13
0.01
0.11
0.27
0.02
0.21
0.12
0.03
0.01
0.17
0.25
0.00
0.15
0.15
0.07
0.08
0.96

Md
0.07
0.37
0.33
0.10
0.22
0.44
0.19
0.25
0.41
0.37
0.00
0.42
0.10
0.32
0.45
0.26
0.70
0.54

Clay

0.14
0.35
0.29
0.15
0.26
0.34
0.22
0.20
0.25
0.26
0.07
0.41
0.12
0.33
0.43
0.27
0.85
0.71
0.87

Silt

0.13
0.36
0.30
0.15
0.26
0.36
0.22
0.21
0.28
0.28
0.06
0.42
0.12
0.33
0.44
0.27
0.84
0.70
0.90
1.00

Sand

0.10
0.46
0.49
0.11
0.25
0.62
0.11
0.30
0.42
0.45
0.16
0.59
0.22
0.57
0.63
0.39
0.33
0.14
0.66
0.50
0.53

Red

0.05
0.51
0.55
0.03
0.32
0.60
0.14
0.37
0.45
0.46
0.21
0.59
0.26
0.51
0.63
0.39
0.30
0.11
0.68
0.51
0.54
0.99

Green

0.07
0.58
0.61
0.13
0.41
0.55
0.20
0.45
0.49
0.45
0.30
0.57
0.30
0.37
0.59
0.37
0.26
0.07
0.70
0.50
0.54
0.95
0.98

Blue

Table 3. Correlation coefficients (r) between different pairs of variable (n = 49). Bold values are significant at the 0.01 level. MGS = mean grain size; Md = median grain size; MS = bulk
magnetic susceptibility.

0.14
0.21
0.19
0.02
0.10
0.47
0.05
0.24
0.45
0.47
0.02
0.21
0.06
0.26
0.32
0.08
0.08
0.07
0.40
0.15
0.19
0.68
0.68
0.67

MS

10
Kandasamy Selvaraj et al.
BOREAS

Climate and environmental changes, SE China

BOREAS

0.26

0.20
300

0.18

0
Range in soils - 2.01-2.32
Mean - 2.17

0.22

Na (ppm)

0.24

400

1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0

10

0.16

200

0.14

24

20

0.12

1842-1900

1900-1986

10

26

30

28

30
4

32
34

36

900

Depth (cm)

19862011

100

Fe (%)

Zr (ppm)

500

Si (%)

Si:Al

0.28

40

Range of rocks - 3.07-4.66


Mean - 4.13

600

11

50
60
70

Mn (ppm)

800
700

80

600
500

90

400

0.2

300

0.3

200

0.4

0.5

0.6

Total CaO%

100
0

10

20

30

40

50

60

70

80

90

Depth (cm)

Fig. 10. Down-core variations in elements of Factor 3, representing


runoff variability. The vertical shaded bar represents the interval of
AD 19001986.

Fig. 11. Down-core variations in Si:Al ratio and CaO content, showing the presence of biogenic materials in the form of silica in core
TH1. The vertical grey bars represent the ranges of the Si:Al ratios
of eight rock samples and two soil samples investigated in the study
area.

1
0
1
2
19862011

Grain size
PCA3

3
3

1900-1986

1842-1900

3
2
1
0
1
2
3
3

PCA4
Early diagenesis

Detrital input
PCA1

PCA2
Trophic status

1842

1859

1878

1897

1916

1935

1954

1972

1992

2011

Age (year AD)

close to the ratio in the soil or bedrock within the


catchment. If we assume that the soil in the catchment
is free of biogenic silica, then the depth profile of the
Si:Al ratio suggests the presence of biogenic silica
(510%), especially in the sediments deposited from
68.586 cm (20 cm) and sediments deposited from
12.53.5 cm in core TH1 (Fig. 11). However, the total
CaO content of the sediments measured using XRF
was always <0.6% (Fig. 11). This amount is very low
compared with the mean content of CaO (1.99%) in
eight rock samples analysed from the catchment of
Small Tian Lake, suggesting that the sediments of this
lake contain insignificant amounts of biogenic materials in the form of carbonate and thus are dominantly
detrital in nature. The negative relationship of Si with
Fe and Mn therefore indicates the dilution effect of
authigenic Fe-Mn hydroxides on detrital silica rather
than biogenic silica (Boyle et al. 1998). Thus, factor 4
may be regarded as representing the dilution of detrital
input by Fe-Mn enrichment resulting from early diagenetic reactions.
Thus, when arranged all four factors depth-wise
(Fig. 12), changing PCA scores reveals the importance
of relative components to the overall geochemical
record. PCA1 shows larger variation at a depth of 86
56 cm, corresponding to AD 18481905 (AD 1842

2
0

10 20 30 40 50 60 70 80 90
Depth (cm)

Fig. 12. Age vs. the four main components (PCA1, PCA2, PCA3
and PCA4) extracted by the principal component analysis (PCA).
The vertical shaded bar represents the interval of AD 19001986.

1901) and then decreases slightly to a depth of


13.5 cm. PCA1 decreases significantly over the top
13 cm. PCA2 increases from the bottom to a depth of

12

Kandasamy Selvaraj et al.

around 33 cm and then decreases slightly to the top.


PCA3 shows a decreasing trend from the bottom to a
depth of 63.5 cm and then varies relatively significantly. PCA4 decreases from the bottom to a depth of
66.5 cm and then increases to the top.
Palaeolimnological interpretation
The bottom layer of core TH1 is dated at AD
1842 (AD 1836), thus enabling us to reconstruct
the last 170 years of climate and environmental
changes in southeastern China. Three periods can
be distinguished: two short, erosion-dominated terrigenous input intervals at 1842 to 1900 (8958 cm)
and 1986 to 2011 (130 cm), and a relatively long
period (19001986) of weathering-dominated terrigenous input (Fig. 8).
During the period from AD 1842 to 1900, the mineral load was generally low and punctuated by extreme
events of relatively short duration, probably related to
extreme cold fluctuations that occurred during this
time (Figs 7, 8). It has been suggested that southeastern China experienced a very severe cold winter in AD
1892, when many lakes (e.g. Tai Lake and Dongting
Lake) and rivers (e.g. low reaches of Huaihe, Huangpu
and Yangtze Rivers) became frozen for a long duration, as recorded in Chinese historical documents
(Zheng et al. 2012). After AD 1891, the mineral loads
show an abruptly decreasing trend, which may correspond to this severe cold event. In this interval, some
peaks of TOC, TN and the C:N ratio are consistent
with mineral loading, indicating that the organic
source of sediment was mainly terrestrial material
(Fig. 5). Down-core variations in CIA, PIA, K/Al, Rb/
Sr and Al/Na also show relatively lower values
(Fig. 7), indicating relatively weak chemical weathering of rocks in the catchment during this interval. During AD 1842 to 1876 (7186 cm), the relatively high
value and stable level of grain size and high water content suggest wetter than normal hydrological conditions during the formation of sediment, probably
related to high precipitation. High values of TOC
(>4%), TN (>0.2%) and C:N ratios (>10) further support high levels of runoff in the catchment as a result
of high rainfall and thus delivery of terrestrial materials into the lake. The v-value is close to zero from the
base of the core to 71 cm, which is consistent with high
water content and TOC values of sediments, suggesting that the organic matter is diamagnetic in nature
(Figs 4, 5). High concentrations of S and Ca in the
bottom 15 cm suggest a sub-oxic/anoxic environment
within the lake probably because of high water levels,
supporting the theory of heavy rainfall during AD
18421876 (Fig. 9). Therefore, the climate was generally cold from AD 1842 to 1900, although the climate
from AD 1842 to 1876 was more humid than during
AD 18761900. Such an interpretation of less humid

BOREAS

conditions from AD 18761900 is roughly consistent


with a higher number of typhoons observed in c. AD
18701930 in Taiwan when low solar activity prevailed
(Hung 2013).
Overall, from AD 1900 to 1986, the levels of elements related to detrital input show a highly weathered
nature of sediments, although they decreased up-core
slightly (Figs 7, 8). All chemical weathering indices
indicate greater chemical weathering of rocks in the
catchment compared to sediments that accumulated
during AD 18421900, probably because of a warm
and humid climate (Fig. 7). However, distinct variations in grain size suggest that the depositional environment was relatively turbulent (Fig. 4). For instance,
two medium-size sand-dominated layers in the middle
of core TH1 correspond to AD 19061913 and AD
19351953 (137Cs dating; Fig. 4) and AD 19021910
and AD 19311951 (210Pb dating), suggesting that the
hydrodynamic conditions in the catchment were probably wetter with high surface runoff during these intervals. This may imply the influence of typhoons around
these time intervals (Fig. 4). Given that there is no
prominent river input into the lake and that the lake is
surrounded by hills (~100200 m a.s.l.), the deposition
of medium-size sand-dominated layers in these intervals might have contributed to greater hill slope erosion in the catchment. Consistent with this inference,
the reconstruction of temperatures for the last
500 years from historical records in China revealed
that the warmest interval was from the 1920s to the
1940s (Wang et al. 1991). Likewise, the historical
inventory contains 41 typhoon strikes during the 26year interval between AD 1884 and 1909 (Liu et al.
2001). The position of the medium-size sand-dominated layers in core TH1 roughly corresponds to the
warmest epoch and the period of intense typhoon
strikes, indicating that our sedimentary proxy records,
in particular the grain-size proxies, may respond to the
influence of stronger typhoons in SE China and thus
may be an ideal natural archive for the reconstruction
of decadalcentennial scale climate changes in SE
China. The relatively low C:N ratios with few fluctuations suggest an algal source of organic matter and
thus a lacustrine environment during this interval. The
amounts of Fe and Mn increase as a result of precipitation of Fe-Mn oxyhydroxides, resulting in increases in
v and RGB intensity (Figs 9, 10). Low values of Ca
and S also substantiate the existence of a warm and
humid climate during this interval (Fig. 9).
In the last ~25 years (AD 19862011), the mineral
load reduced abruptly and then slightly increased
(Fig. 8). Chemical weathering proxies also show trends
similar to the mineral loads (Fig. 7). The mean grain
size shows smaller fluctuations, probably indicating
only one source of sediment. The median grain size is
also relatively constant, indicating a stable depositional
environment (Fig. 4). These changes may be attributed

BOREAS

to changes in land use patterns over the last two


decades, an interpretation that is roughly consistent
with the expansion of the urban construction area in
nearby Fuzhou City to 64.2 km2 during the AD 1988
2000 period (Bai & Chen 2013). High concentrations
of Fe and Mn, however, suggest a well-oxygenated sedimentwater interface probably because of low rainfall,
resulting in a lower lake water level. When the lake
level is low, it is expected that the core location become
closer to the catchment, resulting in either the accumulation of larger grain sizes at the core site, or the precipitation of Fe-Mn oxyhydorxides, both might have
diluted the detrital input (Figs 7, 8, 10).

Conclusions
Sedimentological, physical, and organic and inorganic
geochemical records of a sediment core from Tian
Lake in SE China have provided unprecedented
records of climatic and environmental changes and
related catchment processes for southeastern coastal
China over the last ~170 years. Our study revealed that
the climate was mostly wet during this time period.
Although generally wet, the geochemical parameters
indicate the dominant input of physically eroded materials during c. AD 18421900, whereas chemically
weathered surface soil input seems to have dominated
over a relatively longer period from c. AD 19001986.
The situation was completely different since AD 1986
until the end of the core dating period in 2012, during
which time geochemical parameters suggest the influence of land use changes on the sedimentation pattern
and/or dilution of detrital elements by early diagenetic
enrichments of Fe and Mn. Given that most coastal
estuaries and shallow coastal systems have already
been altered because of intense economic development
along the coastal zone of SE China, the wide variety of
physical and geochemical data provided here will be
useful for the future assessment of climate and environmental, including man-made, changes to the coastal
and inland aquatic systems of SE China.
Acknowledgements. This work was supported by the National
Science Foundation of China (grant no. 41273083), Shanhai Fund of
Xiamen University, China (grant no. 2013SH012) and Open Fund of
Tongji University, China (grant no. MGK1201). We thank S. J. Kao,
James Ho and T. Y. Lee for their help during the fieldwork and assistance in collecting push sediment cores and B. J. Wang for C and N
analyses. We gratefully acknowledge Phil Meyers and an anonymous
reviewer for their helpful comments and critique of the original
manuscript.

References
An, Z. S., Clemens, S. C., Shen, J., Qiang, X. K., Jin, Z. D., Sun, Y.
B., Prell, W. L., Luo, J. J., Wang, S. M., Xu, H., Cai, Y. J., Zhou,
W. J., Liu, X. D., Liu, W. G., Shi, Z. G., Yan, L. B., Xiao, X. Y.,
Chang, H., Wu, F., Ai, L. & Lu, F. Y. 2011: Glacial-interglacial
Indian summer monsoon dynamics. Science 333, 719723.

Climate and environmental changes, SE China

13

An, Z. S., Colman, S. M., Zhou, W. J., Li, X. Q., Brown, E. T., Timothy Jull, A. J., Cai, Y. J., Huang, Y. S., Lu, X. F., Chang, H.,
Song, Y. G., Sun, Y. B., Xu, H., Liu, W. G., Jin, Z. D., Liu, X. D.,
Cheng, P., Liu, Y., Ai, L., Li, X. Z., Liu, X. J., Yan, L. B., Shi, Z.
G., Wang, X. L., Wu, F., Qiang, X. K., Dong, J. B., Lu, F. Y. &
Xu, X. W. 2012: Interplay between the Westerlies and Asian monsoon recorded in Lake Qinghai sediments since 32 ka. Scientific
Reports 2, 619, doi:10.1038/srep00619.
Appleby, P. & Oldfield, F. 1983: The assessment of 210Pb data from
sites with varying sediment accumulation rates. Paleolimnology 15,
2935.
Arnaud, F., Magand, O., Chapron, E., Bertrand, S., Boes, X., Charlet, F. & Melieres, M. 2006: Radionuclide dating (210Pb, 137Cs,
241
Am) of recent lake sediments in a highly active geodynamic setting (Lakes Puyehue and IcalmaChilean Lake District). Science
of the Total Environment 366, 837850.
Bai, L. Y. & Chen, Z. Q. 2013: A study by urban construction land
expansion mode in Fuzhou City. Applied Mechanics and Materials
316317, 219222.
Bertrand, S., Xavier, B., Julie, C., Charlet, F., Urrutia, R., Espinoza,
C., Lepoint, G., Charlier, B. & Fagel, N. 2005: Temporal evolution
of sediment supply in Lago Puyehue (Southern Chile) during the
last 600 yr and its climatic significance. Quaternary Research 64,
163175.
Bonatti, E. & Gartner, S. 1973: Caribbean climate during Pleistocene
Ice Ages. Nature 244, 563565.
Boyle, J. F. 2001: Inorganic geochemical methods in palaeolimnology. In Last, W. M. & Smol, J. P. (eds.): Tracking Environmental
Changes Using Lake Sediments, Vol. 2, Physical and Geochemical
Methods. Developments in Paleoenvironmental Research, 83141.
Kluwer Academic Publishers, Dordrecht.
Boyle, J. F., Mackay, A. W., Rose, N. L., Flower, R. J. & Appleby, P.
G. 1998: Sediment heavy metal record in Lake Baikal. Journal of
Paleolimnology 20, 135150.
Boyle, J. F., Rose, N. L., Bennion, H., Yang, H. & Appleby, P. G.
1999: Environmental impacts in the Jianghan plain: evidence from
lake sediments. Water Air and Soil Pollution 122, 2140.
Campbell, B. 1983: Application of environmental caesium-137 for
the determination of sedimentation rates in reservoirs and lakes
and related catchment studies in developing countries. Radioisotopes in Sediment Studies, IAEA 298, 730.
Chaudhuri, S. & Brookings, D. G. 1979: The Rb-Sr systematics in
acid-leached clay minerals. Chemical Geology 24, 231242.
Chen, C. T. & Selvaraj, K. 2008: Evaluation of elemental enrichments in surface sediments off southwestern Taiwan. Environmental Geology 54, 13331346.
Chen, F. H., Chen, J. H., Holmes, J., Boomer, I., Austin, P., Gates, J.
B., Wang, N. L., Brooks, S. J. & Zhang, J. W. 2010: Moisture
changes over the last millennium in arid central Asia: a review,
synthesis and comparison with monsoon region. Quaternary
Science Reviews 29, 10551068.
Chen, J. A., Wan, G., Zhang, D. D., Zhang, F. & Huang, R. 2004:
Environmental records of lacustrine sediments in different time
scales: sediment grain size as an example. Science in China Series
D: Earth Sciences 47, 954960.
Chen, S. Y., Wang, S. M., Chen, Y. Y., Zhang, E. L., Chen, Y. J. &
Shu, Y. X. 2009: Vertical distribution and chronological implication of 210Pb and 137Cs in sediments of Dongping Lake, Shandong
Province. Quaternary Sciences 29, 981987 (in Chinese, with English abstract).
Chen, Y. Y., Chen, S. Y., Liu, J. Z., Yao, M., Sun, W. B. &
Zhang, Q. 2013: Environmental evolution and hydrodynamic
process of Dongping Lake in Shandong Province, China, over
the last 150 years. Environmental Earth Sciences 68, 6975.
Clift, P. D., Hodges, K. V., Heslop, D., Hannigan, R., Long, H.
V. & Calves, G. 2008: Correlation of Himalayan exhumation
rates and Asian monsoon intensity. Nature Geoscience 1, 875
880.
Conroy, J. L., Overpeck, J. T., Cole, J. E., Shanahan, T. M. & Steinitz-Kannan, M. 2008: Holocene changes in eastern tropical Pacific climate inferred from a Galapagos lake sediment record.
Quaternary Science Reviews 27, 11661180.

14

Kandasamy Selvaraj et al.

Davey, B. G., Russell, J. D. & Wilson, M. J. 1975: Iron oxide and clay
minerals and their relation to colours of red and yellow podozolic
soils near Sydney, Australia. Geoderma 14, 125138.
Dearing, J. A. 1999: Magnetic susceptibility. In Walden, J., Oldfield,
F. & Smith, J. (eds.): Environmental Magnetism, a Practical Guide,
3562. Quaternary Research Association, London.
Edgington, D. N. & Robbins, J. A. 1976: Records of lead deposition
in Lake Michigan sediments since 1800. Environmental Science and
Technology 10, 266274.
Fan, Y. X., Chen, F. H., Wei, G. X., Madsen, D. B., Oviatt, C. G.,
Zhao, H., Chun, X., Yang, L. P., Fan, T. L. & Li, G. Q. 2010:
Potential water sources for Late Quaternary Megalake JilantaiHetao, China, inferred from mollusk shell 87Sr/86Sr ratios. Journal
of Paleolimnology 43, 577587.
Fedo, C. M., Eriksson, K. A. & Krogstad, E. J. 1996: Geochemistry
of shales from the Archean (~3.0 Ga) Buhwa Greenstone Belt,
Zimbabwe: implications for provenance and source-area weathering. Geochimica et Cosmochimica Acta 60, 17511763.
Fedo, C. M., Nesbitt, H. M. & Young, G. M. 1995: Unravelling the
effects of potassium metasomatism in sedimentary rocks and paleosols, with implications for paleoweathering conditions and provenance. Geology 23, 921924.
Fedo, C. M., Young, G. M., Nesbitt, H. W. & Hanchar, J. M. 1997:
Potassic and sodic metasomatism in the southern province of the
Canadian Shield: evidence from the Paleoproterozoic Serpent Formation, Huronian Supergroup, Canada. Precambrian Research 84,
1736.
Gaillardet, J., Dupre, B., Louvat, P. & Allegre, C. J. 1999: Global silicate weathering and CO2 consumption rates deduced from the
chemistry of large rivers. Chemical Geology 159, 330.
Giguet-Covex, C., Arnaud, F., Poulenard, J., Disnar, J. R., Delhon, C., Francus, P., David, F., Enters, D., Rey, P. J. & Delannoy, J. J. 2011: Changes in erosion patterns during the
Holocene in a currently treeless subalpine catchment inferred
from lake sediment geochemistry (Lake Anterne, 2063 m asl,
NW French Alps): the role of climate and human activities.
The Holocene 21, 651665.
Grousset, F., Jouanneau, J., Castaing, P., Lavaux, G. & Latouche, C.
1999: A 70 year record of contamination from industrial activity
along the Garonne River and its tributaries (SW France). Estuarine, Coastal and Shelf Science 48, 401414.
Gupta, H., Chakrapani, G. J., Selvaraj, K. & Kao, S. J. 2011: The fluvial geochemistry, contributions of silicate, carbonate and salinealkaline components to chemical weathering flux and controlling
parameters: Narmada River (Deccan Traps), India. Geochimica et
Cosmochimica Acta 75, 800824.
Hung, C. W. 2013: A 300-year typhoon record in Taiwan and the
relationship with solar activity. Terrestrial, Atmospheric and Ocean
Sciences 24, 737743.
Janitzky, P. 1987: Particle size analysis. In Singer, M. & Janitzky, P.
(eds.): Field and Laboratory Procedures Used in Soil Chrono-Sequence Study. USGS Bulletin 1648, 1116.
Koide, M., Bruland, K. W. & Goldberg, E. D. 1973: 228Th/232Th and
210
Pb geochronologies in marine and lake sediments. Geochimica
et Cosmochimica Acta 37, 11711187.
Krishnamurti, G. S. R. & Satyanarayana, K. V. S. 1971: Influence of
chemical characteristics in the developing of soil colour. Geoderma
5, 243248.
Liu, J. B., Chen, J. H., Selvaraj, K., Xu, Q. H., Wang, Z. L. & Chen,
F. H. 2014: Chemical weathering over the last 1200 years recorded
in the sediments of Gonghai Lake, Lvliang Mountains, North
China: a high-resolution proxy of past climate. Boreas 43, 914
923.
Liu, K-b., Shen, C. & Louie, K-s. 2001: A 1000-year history of
typhoon landfalls in Guangdong, southern China, reconstructed
from Chinese historical documentary records. Annals of the Association of American Geographers 91, 453464.
Liu, M. & Fan, D. 2011: Geochemical records in the subaqueous
Yangtze River delta and their responses to human activities in the
past 60 years. Chinese Science Bulletin 56, 552561.
Longmore, M., OLeary, B. & Rose, C. 1983: Caesium-137 profiles in the sediments of a partial-meromictic lake on Great

BOREAS

Sandy Island (Fraser Island), Queensland, Australia. Hydrobiologia 103, 2127.


Lou, J. Y. & Chen, C. T. A. 1997: Paleoclimatological and paleoenvironmental records since 4000 a B.P. in the sediments of alpine lakes
in Taiwan. Science in China D Series: Earth Sciences 40, 424431.
Ma, Y., Liu, C. & Huo, R. 2000: Strontium isotope systematic during
chemical weathering of granitoids: importance of relative mineral
weathering rates. Journal of Goldschmidt Conference Abstracts 5,
657.
Meyers, P. A. 2004: Application of elemental and isotopic source
identification of sedimentary organic matter. Chemical Geology
114, 289302.
Meyers, P. A. & Ishiwatari, R. 1993: Lacustrine organic geochemistry
- an overview of indicators of organic matter sources and diagenesis in lake sediments. Organic Geochemistry 20, 867900.
Meyers, P. A., Bernasconi, S. M. & Forster, A. 2006: Origins and
accumulation of organic matter in expanded Albian to Santonian
black shale sequence on the Demerara Rise, South American margin. Organic Geochemistry 37, 18161830.
Moy, M. C., Seltzer, G. O., Rodbell, D. T. & Anderson, D. M. 2002:
Variability of El Nino/Southern Oscillation activity at millennial
timescales during the Holocene epoch. Nature 420, 162165.
Nahm, W. H., Lee, G. H., Yang, D. Y., Kim, J. Y., Kashiwaya, K.,
Yamamoto, M. & Sakaguchi, A. 2010: A 60-year record of rainfall
from the sediments of Jinheung Pond, Jeongeup, Korea. Journal of
Paleolimnology 43, 489498.
Nesbitt, H. W. & Young, G. M. 1982: Early Proterozoic climates and
plate motions inferred from major element chemistry of lutites.
Nature 299, 715717.
Nesbitt, H. W., Markovics, G. & Price, R. C. 1980: Chemical
processes affecting alkalis and alkaline earths during continental weathering. Geochimica et Cosmochimica Acta 44, 1659
1666.
Ollier, C. 1984: Weathering. 270 pp. Longman Press, London.
Ortega-Guerrero, B., Sedow, S., Solleiro-Rebolledo, E. & Soler, A.
2004: Magnetic mineralogy in Barranca Tlalpan exposure paleosols, Tlaxcala, Mexico. Revista Mexicana de Ciencias Geologicas
21, 120132.
Peinerud, E. K. 2000: Interpretation of Si concentrations in lake sediments: three case studies. Environmental Geology 40, 16371653.
Perkins, R. W. & Thomas, C. W. 1980: Worldwide fallout. In Hanson,
W. C. (ed.): Transuranic Elements in the Environment, 5382. Technical Information Center, U. S. Department of Energy, Washington, D. C.
Petit, D. 1974: 210Pb et isotopes stables du plomb dans des sediments
lacustres. Earth and Planetary Science Letters 23, 199205.
Qiang, M. R., Liu, Y. Y., Jin, Y. X., Song, L., Huang, X. T. & Chen,
F. H. 2014: Holocene record of eolian activity from Genggahai
Lake, northeastern Qinghai-Tibetan Plateau, China. Geophysical
Research Letters 41, 589595.
Robbins, J. A., Krezoski, J. R. & Mozley, S. 1977: Radioactivity in
sediments of the Great Lakes: post-depositional redistribution by
deposit-feeding organisms. Earth and Planetary Science Letters 36,
325333.
Schoenborn, W. A. & Fedo, C. M. 2011: Provenance and paleoweathering reconstruction of the Neopreoterozoic Johnnie Formation, southeastern California. Chemical Geology 285, 231255.
Selvaraj, K. & Chen, C. T. A. 2006: Moderate chemical weathering
of subtropical Taiwan: constraints from solidphase geochemistry
of sediments and sedimentary rocks. Journal of Geology 114, 101
116.
Selvaraj, K., Chen, C. T. A. & Lou, J. Y. 2007: Holocene East Asian
monsoon variability: links to solar and tropical Pacific forcing.
Geophysical Research Letters 34, L01703, doi:10.1029/
2006GL028155.
Selvaraj, K., Chen, C. T. A., Lou, Y. J. & Kotlia, B. S. 2011: Holocene weak summer East Asian monsoon intervals in Taiwan and
plausible mechanisms. Quaternary International 229, 5766.
Selvaraj, K., Parthiban, G., Chen, C. T. A. & Lou, J. Y. 2010: Anthropogenic effects on sediment quality offshore southwestern Taiwan: assessing the sediment core geochemical record. Continental
Shelf Research 30, 12001210.

BOREAS

Selvaraj, K., Wei, K. Y., Liu, K. K. & Kao, S. J. 2012: Late Holocene
monsoon climate of northeastern Taiwan inferred from elemental
(C, N) and isotopic (d13C, d15N) data in lake sediments. Quaternary Science Reviews 37, 4860.
Sun, Q., Liu, D., Liu, T., Di, B. & Wu, F. 2011: Temporal and spatial
distribution of trace metals in sediments from the northern Yellow
Sea coast, China: implications for regional anthropogenic processes. Environmental Earth Sciences 66, 697705.
Wang, R., Wang, S. & Fraedrich, K. 1991: An approach of reconstruction of temperature on seasonal basis using historical documents from China. International Journal of Climatology 11, 381
392.
Warrier, A. K. & Shankar, R. 2009: Geochemical evidence for the
use of magnetic susceptibility as a paleorainfall proxy in the tropics. Chemical Geology 265, 553562.
Wieland, E., Santschi, P., H
ohener, P. & Sturm, M. 1993: Scavenging
of Chernobyl 137Cs and natural 210Pb in Lake Sempach, Switzerland. Geochimica et Cosmochimica Acta 57, 29592979.
Wu, Z. Y. 1980: Vegetation of China. 1385 pp. Science Press, Beijing
(in Chinese).
Xiang, L. 1998: Dating sediments on several lakes inferred from
radionuclide profiles. Journal of Environmental Sciences 10,
5663.

Climate and environmental changes, SE China

15

Yan, H., Sun, L. G., Oppo, D. W., Wang, Y. H., Liu, Z. H., Xie, Z.
Q., Liu, X. D. & Cheng, W. H. 2011: South China Sea hydrological
changes and Pacific Walker Circulation variations over the last
millennium. Nature Communications 2, 15.
Yancheva, G., Nowaczyk, N. R., Mingram, J., Dulski, P., Schettler,
G., Negendank, J. W., Liu, J. Q., Sigman, D. M., Peterson, L. C. &
Haug, G. H. 2007: Influence of the intertropical convergence zone
on the East Asian monsoon. Nature 445, 7477.
Yao, S. C., Li, S. J. & Zheng, H. C. 2008: 210Pb and 137Cs dating of
sediments from Zigetang Lake, Tibetan Plateau. Journal of Radioanalytical and Nuclear Chemistry 278, 5558.
Yarincik, K. M. & Murray, R. W. 2000: Climatically sensitive eolian
and hemipelagic deposition in the Cariaco Basin, Venezuela, over
the last 578 000 years: result from Al/Ti and K/Al. Paleoceanography 15, 210228.
Zhang, X. B. 2005: Discussion on interpretations of 137Cs depth distribution profiles of lake deposits. Journal of Mountain Science 23,
294299 (in Chinese with English abstract).
Zhang, Y., Peng, B., Chen, J. & Lu, J. 2005: Evaluation of sediment
accumulation in Dianchi Lake, using 137Cs dating. Acta Geographica Sinica 60, 7178.
Zheng, J. Y., Ding, L. L., Hao, Z. X. & Ge, Q. S. 2012: Extreme cold
winters in southern China during AD 16502000. Boreas 41, 112.

Das könnte Ihnen auch gefallen