Distribution and ecology of planktonic foraminifera from ... · PDF fileDistribution and...

21
Distribution and ecology of planktonic foraminifera from the seas around the Indonesian Archipelago X. Ding a, * , F. Bassinot b , F. Guichard b , Q.Y. Li c , N.Q. Fang a , L. Labeyrie b , R.C. Xin a , M.K. Adisaputra d , K. Hardjawidjaksana d a China University of Geosciences (Beijing), Beijing 100083, PR China b Laboratoire des Sciences du Climat et de l’Environnement, Domaine du CNRS, Gif-sur-Yvette 91198, France c School of Ocean and Earth Sciences, Tongji University, Shanghai 200092, PR China d Marine Geological Institute, Bandung 40174, Indonesia Received 14 August 2005; received in revised form 26 September 2005; accepted 11 October 2005 Abstract Planktonic foraminiferal assemblages in 50 core-top samples from the western and southern areas of the Indonesian Archipelago and 29 core tops retrieved northwest of Australia were grouped using cluster analysis. These assemblages make it possible to sub-divide the studied area in five provinces: 1/ the Banda/Java region (I); 2/ the Timor region (II); 3/ the Java upwelling region (III); 4/ the Indian monsoon Sumatra region (IV), and 5/ the NW Australia margin region (V). The foraminiferal assemblage groups reflect differences in sea-surface temperature, salinity, thermocline depth, and nutrient supply between these five provinces. These differences are related to surface circulation patterns. The carbonate dissolution is rather intense compared to that in other areas of the eastern Indian Ocean. Within the studied area, the strongest dissolution occurs in samples from the Java upwelling region, with the lysocline level rising above ~2800 m. The increase in abundance of Globigerina bulloides at 10–8 ka BP in core SHI-9034 (the Java upwelling region) corresponds to the decrease in core SHI-9006 (the Banda/Java region) which indicates an intensification of upwelling in relation to a strengthened southeastern monsoon over the studied area. D 2005 Elsevier B.V. All rights reserved. Keywords: Indonesian Archipelago; planktonic foraminifera; carbonate dissolution; cluster analysis; upwelling; southeastern monsoon 1. Introduction The Indonesian Archipelago is a pathway for oce- anic heat transfer between the Pacific and Indian oceans (Godfrey and Golding, 1981; Godfrey and Ridgway, 1985). Within this archipelago there are at least eight deep-sea basins connected through relatively shallow sills (Tomczak and Godfrey, 1994; Martinez et al., 1997). The surface circulation shows a strong seasonal variability related to monsoonal forcing. The Indone- sian Archipelago, the only low-latitude connection be- tween two major ocean basins, is a key area along the return branch of the Great Conveyor Belt which ulti- mately brings surface waters from the Pacific to the north Atlantic (Gordon, 1986; Hirst and Godfrey, 1993; Bray et al., 1996; Gordon and Fine, 1996; Mu ¨ller and Opdyke, 2000). The archipelago is situated within the Western Pa- cific Warm Pool (WPWP), and annual mean sea surface temperatures exceed 28 8C(Tomczak and Godfrey, 1994; Martinez et al., 1997). The WPWP not only supplies large amounts of water vapor and latent heat 0377-8398/$ - see front matter D 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.marmicro.2005.10.003 * Corresponding author. Tel.: +86 10 82323501; fax: +86 10 82321540. E-mail address: [email protected] (X. Ding). Marine Micropaleontology 58 (2006) 114 – 134 www.elsevier.com/locate/marmicro

Transcript of Distribution and ecology of planktonic foraminifera from ... · PDF fileDistribution and...

www.elsevier.com/locate/marmicro

Marine Micropaleontolog

Distribution and ecology of planktonic foraminifera from the seas

around the Indonesian Archipelago

X. Ding a,*, F. Bassinot b, F. Guichard b, Q.Y. Li c, N.Q. Fang a, L. Labeyrie b, R.C. Xin a,

M.K. Adisaputra d, K. Hardjawidjaksana d

a China University of Geosciences (Beijing), Beijing 100083, PR Chinab Laboratoire des Sciences du Climat et de l’Environnement, Domaine du CNRS, Gif-sur-Yvette 91198, France

c School of Ocean and Earth Sciences, Tongji University, Shanghai 200092, PR Chinad Marine Geological Institute, Bandung 40174, Indonesia

Received 14 August 2005; received in revised form 26 September 2005; accepted 11 October 2005

Abstract

Planktonic foraminiferal assemblages in 50 core-top samples from the western and southern areas of the Indonesian

Archipelago and 29 core tops retrieved northwest of Australia were grouped using cluster analysis. These assemblages make it

possible to sub-divide the studied area in five provinces: 1/ the Banda/Java region (I); 2/ the Timor region (II); 3/ the Java

upwelling region (III); 4/ the Indian monsoon Sumatra region (IV), and 5/ the NWAustralia margin region (V). The foraminiferal

assemblage groups reflect differences in sea-surface temperature, salinity, thermocline depth, and nutrient supply between these

five provinces. These differences are related to surface circulation patterns. The carbonate dissolution is rather intense compared to

that in other areas of the eastern Indian Ocean. Within the studied area, the strongest dissolution occurs in samples from the Java

upwelling region, with the lysocline level rising above ~2800 m. The increase in abundance of Globigerina bulloides at 10–8 ka

BP in core SHI-9034 (the Java upwelling region) corresponds to the decrease in core SHI-9006 (the Banda/Java region) which

indicates an intensification of upwelling in relation to a strengthened southeastern monsoon over the studied area.

D 2005 Elsevier B.V. All rights reserved.

Keywords: Indonesian Archipelago; planktonic foraminifera; carbonate dissolution; cluster analysis; upwelling; southeastern monsoon

1. Introduction

The Indonesian Archipelago is a pathway for oce-

anic heat transfer between the Pacific and Indian oceans

(Godfrey and Golding, 1981; Godfrey and Ridgway,

1985). Within this archipelago there are at least eight

deep-sea basins connected through relatively shallow

sills (Tomczak and Godfrey, 1994; Martinez et al.,

1997). The surface circulation shows a strong seasonal

0377-8398/$ - see front matter D 2005 Elsevier B.V. All rights reserved.

doi:10.1016/j.marmicro.2005.10.003

* Corresponding author. Tel.: +86 10 82323501; fax: +86 10

82321540.

E-mail address: [email protected] (X. Ding).

variability related to monsoonal forcing. The Indone-

sian Archipelago, the only low-latitude connection be-

tween two major ocean basins, is a key area along the

return branch of the Great Conveyor Belt which ulti-

mately brings surface waters from the Pacific to the

north Atlantic (Gordon, 1986; Hirst and Godfrey, 1993;

Bray et al., 1996; Gordon and Fine, 1996; Muller and

Opdyke, 2000).

The archipelago is situated within the Western Pa-

cific Warm Pool (WPWP), and annual mean sea surface

temperatures exceed 28 8C (Tomczak and Godfrey,

1994; Martinez et al., 1997). The WPWP not only

supplies large amounts of water vapor and latent heat

y 58 (2006) 114–134

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134 115

to the western Pacific atmosphere, but it represents a

major contributor to global climate changes through the

El Nino/Southern Oscillation system. The WPWP dy-

namics are intimately linked to current transport pro-

cesses in the Indonesian Archipelago (Thunell et al.,

1994; Ahmad et al., 1995; Linsley, 1996; Wang, 1998;

Martinez et al., 1999; Wang et al., 1999b).

Thus, researches focused on surface circulation var-

iability in the low-latitude Indian–western Pacific re-

gion are critical for a better understanding of the Asian

monsoons, the dynamics of the WPWP, El Nino events

and thermohaline circulation in the global ocean.

Several paleoceanographic studies have been con-

ducted in the study area. Martinez et al. (1998, 1999)

studied the late Pleistocene paleoceanography of the

northern Australian margin and the eastern Indian

Ocean including the Indonesian Archipelago. Gingele

et al. (2002) analyzed the history of the South Java

Current during the last 80 ka using the distribution of

clay minerals. Hanebuth et al. (2000) studied late-glacial

sea level changes across the Sunda shelf. Visser et al.

(2003) found that SST of waters within the Indonesian

Archipelago increased by 3.5–4.0 8C across the last two

glacial–interglacial transitions, leading the Northern

Hemisphere ice sheets melting by about 2000–3000

years. The tropical Pacific region possibly regulates the

poleward flux of heat and water vapor, thus affecting

global glacial–interglacial climate changes (Cane, 1998),

similar to the mechanisms involved in El Nino/Southern

Oscillation.

Much work remains to be done to unravel the history

of this region, having an intricate geographic setting with

numerous straits and basins, and affected by complex

climatic mechanisms on various timescales (monsoon,

El Nino/Southern Oscillation). Because the Indonesian

Archipelago is located near the equator and has vast,

shallow shelves and coasts, the glacial–interglacial and

seasonal temperature differences are small, but changes

in the amount of fresh water transported are large, mak-

ing paleoclimatologic and paleoceanographic research in

the region difficult. Many attempts have been made

using single cores or single proxies to reconstruct spe-

cific aspects of local environmental changes. However,

these attempts often lack a solid background based on

modern sedimentary and faunal patterns. A comprehen-

sive survey based on seafloor surface samples can pro-

vide us with a key to unravel the interrelation between

modern sedimentation and ocean environment. Once this

interrelation is revealed, we can reconstruct past current

distribution and sediment transport patterns over longer

time periods in the past. Therefore, this paper focuses on

providing a more complete planktonic foraminifera da-

tabase obtained from core-top material recently collected

in the seas around the Indonesian Archipelago, correlat-

ing foraminiferal distribution with environmental vari-

ables and an attempt to apply the results to interpretation

of down-core observations.

2. Modern hydrography and environments in the

Indonesian region

2.1. Surface currents, throughflow and monsoon-

related, seasonal changes

The wind stress between the Pacific and the Indian

oceans maintains a sea level height difference between

these two ocean basins (Bray et al., 1996), leading to a

net inflow of water into the Indian Ocean (Hirst and

Godfrey, 1993). The major components of the Indone-

sian throughflow (ITF) are the Mandanao Current

waters that originate from the upper thermocline of the

north Pacific and are transported into the Indonesia seas

through the Makassar Strait (Gordon, 1986; Gordon and

Fine, 1996) (Fig. 1). Halmahera Eddy water originates

from the upper thermocline of the South Pacific and

seeps into the lower thermocline of the Banda Sea,

making up an important part of the Throughflow (Fig.

1). Only a small portion of the waters flowing through

the Makassar Strait into the Indonesian seas directly

enters the Indian Ocean through the Lombok Strait,

between the islands of Bali and Lombok. The largest

part of these waters turns eastward into the Banda Sea

and Flores Sea before spreading into the Indian Ocean

through the Timor Sea as parts of the west-flowing

South Java Current, the South Equatorial Current, and

the south-flowing Leeuwin Current that runs along the

western Australian margin (Gordon and Fine, 1996;

Siedler et al., 2001) (Fig. 1).

Temperature gradients between the ocean and adja-

cent continents (eastern Asia and Australia) result in

monsoon winds blowing from the southeast during

winter (August), and turning to the opposite direction

during summer (February). Ocean currents in the area

move according to the wind regime. During the summer

monsoon (NW monsoon), the surface currents flow

from the Java Sea into the Banda Sea. The high atmo-

spheric temperatures in the region induce evaporation

of oceanic water at the surface, low-pressure cells and

rain. The overall balance is a gain of freshwater from

precipitation at the sea surface. Consequently, SSTs are

high and sea-surface salinities (SSSs) low (Martinez et

al., 1998). Surface water masses of low salinity reach

the Banda Sea. During the winter monsoon (SE mon-

soon), surface current flows from Arafura and Banda

Fig. 1. Today August currents in the seas around Indonesian Archipelago (Gingele et al., 2002; Gordon and Fine, 1996).

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134116

Sea towards the Java Sea. Surface water masses whose

salinity has been increased by evaporation reach the

Banda Sea (Ahmad et al., 1995).

During the NW monsoon season, the South Java

Current (derived from the Equatorial Counter Current)

moves towards the southeast to meet the Leeuwin Cur-

rent. The mixing of the South Java Current with the

Leeuwin Current gives origin to the South Equatorial

Current that moves towards the west (Martinez et al.,

1998). Near the maximum in eastward flow, salinities in

the SJC can be as low as 32x and extend down to 138S.Runoff from Sumatra and Java, and advection of fresher

Java Sea water through the Sunda Strait may be respon-

sible for this low-salinity btongueQ. In August/Septem-

ber, during the peak of the SE monsoon, the throughflow

from the Pacific into the Indian Ocean through the Timor

Passage and the Lombok Straits is at its maximum. The

flow of the SJC, which incorporates some of the through-

flow water, is weak and its direction possibly westward

from September to October (Gingele et al., 2002).

2.2. Inter-connections with El Nino/Southern Oscillation

Due to the trade winds which force warm surface

waters to pile up in the WPWP (Enfield, 1989), the

steric height in the west Pacific is higher than in the

east, and the annual average sea surface temperature

(SST) is about 29 8C in the west but only 24 8C in the

eastern side. During an El Nino episode, trade winds

weaken and warm waters of the west Pacific flow to the

east causing SST to rise abnormally in the eastern and

central equatorial Pacific. The ITF and El Nino/South-

ern Oscillation are intimately inter-connected: ITF

weakens during El Nino events with a shallower ther-

mocline, and strengthens during La Nina events with a

deeper thermocline. In addition, the ITF transports

temperature anomalies related to El Nino/La Nina

events from the equatorial Pacific into the Indian

Ocean, as evidenced by the strong correlation existing

between upper thermocline temperatures recorded in

the Makassar Strait and the occurrence of El Nino

events (Gordon and Fine, 1996; Siedler et al., 2001).

2.3. Environmental variables in the western and south-

ern Indonesian Archipelago area

2.3.1. Surface temperatures and salinities

Sea-surface temperature (SST) in February varies

from ~30.0 8C in the northwest to ~26.0 8C in the

south of the study area (Fig. 2a). The Timor Sea that

Fig. 2. Sea-surface temperature (8C) in February (a) and August (b); sea-surface salinity (x) in February (c) and August (d); SST seasonality (e) and

SSS seasonality (f) (from Levitus, 1998).

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134 117

separates the east end of the Archipelago from Australia

shows the relative high temperature (29.4 8C) (Fig. 2a).In August, the near-latitudinal SST gradient across the

studied area increases markedly, with SST ranging from

~30.2 8C in the northwest to ~23.0 8C in the south,

whereas the SST in the Timor Sea drops to about 26.5

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134118

8C (Fig. 2b), making it the Indonesian area with the

largest seasonal contrast (Fig. 2e). Sea surface salinity

(SSS) varies from 31.5x to 35.2x in February with a

steep gradient between the north and the south (Fig.

Fig. 3. Annual average sea-surface nitrate (a), phosphate (b) and oxygen (c),

sea-surface nitrate, phosphate are in Amol/l, sea-surface oxygen in ml/l, the

2c). The lowest salinities are recorded in the Java Sea

due to the influence of fresher waters brought from the

South China Sea under the control of the northwest

monsoon. August shows a slightly reduced latitudinal

and 21 8C thermocline depth in February (d) and August (e). Units for

rmocline depths in m (from Levitus, 1998).

Fig. 4. 12 months temperature and salinity vertical profiles for 5 selected localities in the seas around the Indonesian Archipelago. (a)(f) Banda/Java region (7.38S, 1188E); (b)(g) Timor region

(8.38S, 128.38E); (c)(h) Java upwelling region (10.38S, 112.38E); (d)(i) Indian monsoon Sumatra region (2.38S, 98.38E); (e)(j) NWAustralian margin region (14.38S, 121.38E) (from Levitus, 1998).

X.Dinget

al./Marin

eMicro

paleo

ntology58(2006)114–134

119

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134120

salinity gradient, with SSS varying from 32.5x in the

North to 34.9x in the South (Fig. 2d). The Sunda Strait

and Great Channel have the lowest SSS as a result of

fresh water discharge from coastal regions. The Java

Sea shows the biggest seasonal salinity difference be-

tween February and August (Fig. 2f) (Levitus, 1998).

2.3.2. Nutrient and dissolved oxygen spatial distribution

Mean annual surface phosphate contents are the

highest in the Java Sea and along the Java upwelling

regions close to the south coast of the Java (Fig. 3b),

while mean annual surface nitrate contents are rela-

tively high in the Java Sea, Banda Sea and the Sunda

strait (Fig. 3a), indicating a higher surface water

fertility due either to throughflow waters carried

across the Archipelago and/or to upwelling cells

(Levitus, 1998). Surface dissolved oxygen values

show little variability, with the lowest values mea-

sured in the Java Sea and Sunda straits (Fig. 3c)

(Levitus, 1998).

2.3.3. Thermocline depth

The thermocline depth closely follows the 21 8Cisotherm (Bray et al., 1996; Martinez et al., 1998) and

is on average at about ~120 m depth in February in the

studied area, being slightly shallower to the west of

Sumatra and slightly deeper to the east of the Archi-

pelago including the Timor Sea, NW Australia margin

region (Fig. 3d) (Levitus, 1998). Important changes in

Fig. 5. Phosphate, nitrate, silica and oxygen dissolved vertical profiles for s

Levitus, 1998).

the thermocline depth occur in August: the thermocline

rises to about 90 m along the south coast of Java and

west of Sumatra but deepens rapidly southward (Fig.

3e), indicating the increased upwelling due to southeast

monsoon and Indian monsoon (Levitus, 1998). Mean-

while, the sea level height difference between the south-

ern and northern Archipelago increases, resulting in an

increase volume of upper water being carried across the

Throughflow and resulting in a deepened thermocline

in the Timor Sea, along the NWAustralia coast and the

south Java coast around 158S where the South Equato-

rial Current runs.

2.3.4. Vertical profiles

The vertical profiles of temperature and salinity,

phosphate, nitrate, silica, and dissolved oxygen in 5

localities are presented in Figs. 4 and 5 (Levitus,

1998). The surface temperatures show a slight differ-

ence between February and August to the west of the

Sumatra. The water temperatures below the thermocline

drop rapidly from ~90 to ~150 m, below which the

vertical temperature gradient is small. The thermoclines

are shallow in this region, but markedly deeper along

the NW Australia margin region. A high nutrient con-

tent and low oxygen levels occur at intermediate depths

(between ~100 and ~400 m) to the west of the Sumatra,

Timor and in the Java upwelling regions.

In February, the Java Sea shows the lowest SSS

and the highest salinity difference between surface and

elected localities in the seas around the Indonesian Archipelago (from

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134 121

subsurface waters, followed by the area to the west of

Sumatra. In contrast, the Timor Sea shows the highest

SSS and the smallest salinity difference between sur-

face and subsurface waters, which is the effect of the

northeast winter monsoon and a weakened Through-

flow. In August, the area to the west of the Sumatra

shows the lowest SSS and the biggest salinity differ-

ence between surface and subsurface waters, probably

because of the large amount of fresh water imported

as a result of the southwest monsoon. Rising SSS in

the Java Sea and falling SSS in the Timor Sea and

NW Australia margin occur at times of a stronger

Throughflow.

3. Materials and methods

Cores were collected in the Indonesian Archipelago

during the joint French–Indonesian SHIVA marine geo-

logical cruise in February 1990, and west of Sumatra

and in the northeast Indian Ocean during the French–

Indonesian BARAT cruise onboard the R/V Baruna

Jaya I in 1994. Core-top samples representing the

upper 0–1 cm or 0–2 cm of 35 gravity cores and 15

piston cores were used for this study. BARP- and SHIP-

represent gravity core samples; SHI- and MD-represent

piston core samples. Relevant data from 29 gravity

core-top samples collected west of Australia during

Fig. 6. Map of the top core locations and 5 current regions from the foraminif

III Java upwelling region. IV Indian monsoon Sumatra region. V Australia

Fr10/95 and Fr10/96 cruises onboard the Australian R/

V Franklin was added in our database (Martinez et al.,

1998) (Appendix 1, Fig. 6). The implication of forami-

niferal assemblage changes for reconstructing paleoen-

vironmental is analyzed further using two cores SHI-

9034 (9809.764VS, 111800.721VE, water depth 3330 m)

and SHI-9006 (4833.223VS, 117859.584VE, water depth1999 m).

After being dried and weighted, samples from core

top and core SHI-9034 and SHI-9006 sampled at 10 cm

intervals were washed through a 150 Am-sieve. The

coarse fraction was dried at 40 8C, then split with a

micro-splitter to provide a sub-sample with at least 300

whole specimens of foraminifers, which were identified

and counted (Appendix 2). 10–15 specimens of Globi-

gerinoides ruber in the size range of 250–315 Am of 22

core-top samples and core SHI-9034 and SHI-9006

samples were hand picked, and then washed by an

ultrasonic cleaner in methanol for less than 10 s. Car-

bon and oxygen isotopes were measured using a Finni-

gan Mat-251 mass spectrometer in Laboratoire des

Sciences du Climat et de l’Environnement, Gif-sur-

Yvette, France. Eight samples were selected from core

SHI-9034 for picking ~10 mg of Globorotalia menardii

(in the size range of N250 Am). The specimens were

cleaned in an ultrasonic bath in distilled water, and then

analyzed for AMS 14C ages using the Tandertron Ac-

era assemblage cluster analysis. I Banda–Java region. II Timor region.

margin region.

able 1

lanktonic foraminifera d18O values of core tops

ore d18O (PDB)

ARP-9426 �2.67

ARP-9422 �2.23

ARP-9409 �2.72

ARP-9415 �3.03

ARP-9407 �2.90

ARP-9412 �2.60

ARP-9413 �2.92

ARP-9437 �2.88

ARP-9435 �2.22

ARP-9406 �2.65

HIP-9047 �2.22

ARP-9441 �2.87

HIP-9040 �2.75

HIP-9041 �3.07

HIP-9018 �2.39

D-982172 �2.79

HIP-9037 �2.62

HIP-9024 �2.72

HIP-9034 �2.78

HIP-9029 �2.37

HIP-9038 �2.13

D-982165 �2.10

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134122

celerator in Gif-sur-Yvette. Finally, foraminiferal 14C

ages were adjusted for the apparent reservoir effect on

the ages of surface seawater (400 years). Planktonic

foraminiferal assemblages from the 79 top samples of

the Indonesian Archipelago and the NWAustralia were

grouped using cluster analysis. The cluster analysis was

performed using JMP statistic package (version 3) for

Macintosh, available from SAS Institute, Cary, NC

27513, USA. Similarity in cluster analysis was calcu-

lated with the Complete and/or Ward methods without

weighting. A planktonic foraminifera dissolution index

(FDX) was estimated using the weighted average meth-

od based on different dissolution grades for foraminif-

eral species (Berger, 1979).

4. Results

4.1. The foraminiferal distribution in core-top samples

and oceanographic conditions

The plankton-tow samples could be correlated well

with recent oceanographic condition, but plankton-tow

material may differ substantially from core-top mate-

rial, because these samples represent the average (viz.

seasonal and inter-annual variations) of possibly sev-

eral centuries. Plankton-tow samples are thus of lim-

ited value when reconstruction past oceanographic

conditions rely on core material (Martinez et al.,

1998).

Planktonic foraminifera in deep-sea core-top sam-

ples from the eastern Indian Ocean were compared with

assemblages collected in plankton-tows (Martinez et

al., 1998), and these studies indicated that data on

core-top samples, grouped by principal components

and analyzed by canonical correspondence, are related

significantly to environmental variables (temperature,

salinity and nutrients), although some differences exist

in foraminiferal assemblages in plankton-tow and core-

top samples. Variables correlating with latitude could

account for most of the variance in the species data.

Data on some core-top samples from d18O analyses

(Table 1) were used to compare information with that in

the surface sediment from the studied area. The around

�2.5x d18O values are the same as those in core-top

d18O analyses from the WPWP (Martinez et al., 1997)

that were assumed to represent present-day upper water

conditions.

4.2. Dissolution overprint

Most planktonic foraminifera from deep-sea core-

top samples are affected to some extent by post-depo-

T

P

C

B

B

B

B

B

B

B

B

B

B

S

B

S

S

S

M

S

S

S

S

S

M

sition dissolution. Selective dissolution alters the orig-

inal foraminiferal assemblages by removing the more

fragile foraminiferal species preferentially. Pore water

saturation within the sediments is strongly dependent

upon acidification resulting from oxic respiration of

organic matter and the release of CO2. Thus, the seas

around Indonesian Archipelago are particularly com-

plex as far as dissolution is concerned, owing to 1/ its

complex bathymetry characterized by numerous semi-

isolated basins inter-connected through shallow sills,

with water depths changing rapidly and greatly from

place to place, and 2/ the local variability of its surface

water productivity.

In order to make an estimate of the regional lyso-

cline more quantitative, we performed a moving aver-

age of binned FDX data as a function of depth, leaving

out the cores in the Java upwelling region. All FDX

data from between 400 and 800 m were averaged and

plotted as a triangle at 600 m. Then all FDX data

between 600 and 1000 m core depth were binned,

and plotted as a triangle at 800 m, and so forth (Fig.

7). This procedure gives a semi-quantitative indication

of dissolution with depth. FDX values increase mark-

edly below ~2800 m water depth in all regions (Fig. 7),

indicating a rapid increase in carbonate dissolution

likely corresponding to the foraminiferal lysocline.

This depth is shallower than the lysocline depth in

other areas of the eastern Indian Ocean (3800 m; Mar-

tinez et al., 1998).

Fig. 8. Water depth (a) and FDX (b) in the se

Fig. 7. Vertical profile of FDX in the seas around the Indonesian

Archipelago.

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134 123

FDX values are relatively high in the Timor Sea

region, the west Sumatra region, and, particularly,

along the south coast of Java in the Java upwelling region

(Fig. 8b). The waters in the Java Sea region are the

shallowest, followed by the NW Australia region (Fig.

8a). Most sampling locations in the two regions are

above 2000 m, so carbonate dissolution is relatively

weak, FDX less than 5 (Fig. 8). The Timor Sea region,

Java upwelling region and the west Sumatra region are

relatively deep, especially the latter two, and most of the

sampling locations are deeper than 2000 m, so dissolu-

tion is stronger (Fig. 8). If samples from over 2800 m are

excluded, most samples with an FDXN5 are located in

the Java upwelling region (Fig. 7), which shows that this

region has strong carbonate dissolution, with lysocline

levels shallower than 2800 m. Martinez et al. (1998)

located the lysocline in the eastern Indian Ocean (north

of 158S) at ~2400 m based on a foraminiferal fragmen-

tation index. The Corg /CaCO3 ratio has a crucial impact

on respiration-driven dissolution at the seafloor (Emer-

son and Bender, 1980), so the shallow lysoclinal depth is

probably related to the high productivity in this region.

Planktonic foraminiferal assemblages from the seas

around the Indonesian Archipelago show various

extents of carbonate dissolution. The Timor Sea and

Java upwelling regions are characterized by high abun-

dance of the dissolution-resistant species G. menardii,

and relatively abundances of the high productivity spe-

cies Neogloboquadrina dutertrei and Globigerina bul-

loides, G. menardii is rare in the west Sumatra region,

but the warm-water, dissolution-resistant species Pull-

eniatin obliquiloculata is most abundant. So dissolution

plays a role in our database, but overall the assemblages

retain mainly a hydrographic signal.

as around the Indonesian Archipelago.

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134124

4.3. Species and cluster analysis

Planktonic foraminifera in the studied area consist of

typical tropical–subtropical assemblages, with domi-

Fig. 9. Cluster analysis of the planktonic fora

nant species: G. ruber, N. dutertrei, P. obliquiloculata,

each of which represents between 10% and 20% of

core-top assemblages. Next in abundance are G. menar-

dii, G. bulloides, Globigerinita glutinata, representing

minifera assemblages for the top cores.

Table

2

Ecological

characteristicsoftheim

portantplanktonic

foraminiferaspecies

N.dutertrei

G.ruber

G.bulloides

G.sacculifer

G.glutinata

G.manardii

N.pachyderma(R)

P.obliquiloculata

Atropical

totemperatespecies,abundant

inactivecurrentsystem

s,along

continentalmarginsandin

upwelling

regionswhereit

lives

mainly

within

thethermocline,

nearthechlorophyll

maxim

um

(Fairbankset

al.,1982;

Curry

etal.,

1983;Cannariato

and

Ravelo,1997;Kaw

ahataet

al.,

2002).Abundantin

highproductivity

areas

with

warm

water

and

alow

salinity

(Be,

1977;Prell

and

Curry,

1981;Thompson,1981;Thunelland

Reynolds,

1984;Thiede

and

Junge,

1992;Pflaumannet

al.,1996;

Hilbrecht,1997;Martinez

etal.,1998;

PflaumannandJian,1999).

Awarm-w

ater

species,a

shallow

dweller,most

commonin

warm

mixed

layer

abovethethermocline

(Fairbankset

al.,1982),

symbiont-bearingwith

zooxanthellae,

prefers

oligotrophic

regionswith

a

deepmixed

layer,andis

the

leastresistantto

dissolution

(Be,

1977;Raveloet

al.,

1990;Andreasen

and

Ravelo,1997;Patrick

and

Thunell,1997;Martinez

etal.,1998;Pflaumann

andJian,1999).

Occurs

predominantlyin

subpolarregions,also

commonin

upwelling

areasandboundary

currentsin

low-latitude

regionswheresurface

productivityis

high(Be,

1977;Duplessyet

al.,

1981;PrellandCurry,

1981;Brock

etal.,1992;

Martinez

etal.,1998;

PflaumannandJian,

1999).

Atropical,shallow-dwelling

species,toleratesawiderangeof

temperatures,

butprefers

mixed

layer

waterswithtemperatures

N248C

andlow

seasonal

changes

inSSTandvertical

temperature

gradients,notwell

suited

tolargeseasonal

salinity

changes,andsusceptible

to

dissolution(Be,

1977;Fairbanks

etal.,1982;Thunelland

Reynolds,

1984;Cannariato

and

Ravelo,1997;Patrick

and

Thunell,

1997;Martinez

etal.,

1998;PflaumannandJian,1999).

Has

awidelatitudinal

distribution,toleratesarather

extensiverangeof

temperaturesandsalinities

andismoderatelysusceptible

todissolution.Itis

abundant

inmid

tohighlatitudes

and

inupwellingregionsat

low

latitudes

with

fertile

waters

(Fairbankset

al.,1982;

ThunellandReynolds,1984;

Martinez

etal.,1998;

Pflaumann

and

Jian,1999;

Kaw

ahataet

al.,2002).

Atropical

tosubtropical

speciesthat

prefers

high

SSTandnorm

alSSS,

andisnotvery

susceptible

to

dissolution.Itis

abundantin

upwelling

regionswith

ashallow

thermoclineandhigh

primaryproductivity

(Be,

1977;Fairbanks

etal.,1982;Thunelland

Reynolds,1984;

Martinez

etal.,1998).

Theright-coilingpopulation

ofN.pachyderma

ismost

abundantin

subpolarand

transitional

regionsandonly

rarely

encounteredin

tropical

areas.

Itoccurs

inareasof

upwelling

ofcold

water

in

thetropical/subtropical

belt

(Be,

1977;Thunelland

Reynolds,1984;Reynolds

andThunell,1986;

Pflaumannet

al.,1996;

PflaumannandJian,

1999).

Athick-w

alledtropical

shallow

water

and

dissolutionresistant

species,

which

calcifies

mostofitstestbelow

themixed

layer.It

isas

sociated

withabundant

phytoplankton

and

high

productivity

(Be,

1977;

Raveloet

al.,1990;

PflaumannandJian,

1999;Kaw

ahata

etal.,

2002).

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134 125

from less than 1% to more than 20% of the assem-

blages. Low-abundance species are: Orbulina universa,

Globigerinoides conglobatus, G. tenellus, G. sacculi-

fera, Sphaeroidinella dehiscens, Globigerinella aequi-

lateralis, Globigerina calida, G. falconensis, G.

rubescens, Neogloboquadrina pachyderma (R), Globo-

quadrina conglomerata, and Globorotalia scitula.

Cluster analysis of our planktonic foraminiferal

assemblages revealed the existence of five main

groups (Fig. 9). These groups correspond to provinces

characterized by differences in hydrographic settings:

1/ the Banda/Java region (I); 2/ the Timor region (II);

3/ the Java upwelling region (III); 4/ the Indian mon-

soon Sumatra region (IV), and 5/ the NW Australia

margin region (V) (Fig. 6). The settings of the NW

Australia margin region (V) are the same as those of

Group II— the WPWP assemblage from Martinez et

al. (1998).

5. Discussion

5.1. Planktonic foraminiferal cluster groups: distribu-

tion and characteristics

The ecological preferences of all species are listed in

Table 2, the environmental characteristics of the regions

in Table 3, and the correspondence between foraminif-

eral assemblages and environmental parameters in

Table 4.

5.1.1. Cluster I—Banda/Java region

This cluster is dominated by N. dutertrei (13.0% to

28.3%, average 21.9%) (Fig. 10b, Table 2).

The next dominant species is G. ruber (13.1% to

25.6%, average 21.2%) (Fig. 10g), but high percentages

(~20%) of this species are not typical, because as G.

ruber is abundant in most of the other clusters with the

exception of region III (see below).

The third dominant species is G. bulloides (6.2% to

22.6%, average 15.8%) (Fig. 10a). The tropical, shal-

low-dwelling species G. sacculifer is least abundant in

core-top assemblages from cluster I (1% to 7.3%, av-

erage 3.4%) (Fig. 10h). The rarity or absence of G.

sacculifer indicates important variability in seasonal

SST, SSS and vertical temperature gradients in the

region (Table 2).

Other species show either minor or extremely vari-

able abundance in cluster I. G. glutinata ranges from

4% to 34.1% (average 12%) (Fig. 10c), only slightly

less than in cluster II (see below).

Cluster I corresponds to core tops from the Java Sea

and Banda Sea; the environmental characteristics of this

Table 3

The environmental characteristics of the 5 regions

Environment

indexes

Cluster I—Banda/Java

region

Cluster II—Timor

region

Cluster III—Java

upwelling region

Cluster IV—Indian monsoon

Sumatra region

Cluster V—NW

Australia margin region

Currents This region corresponds to the

main body of the Indonesian

Throughflow coming from the

Makassar strait, and seasonal

reversal of surface currents

due to the monsoon.

The region is along the

pathway of the most important

Throughflow. These waters

flow from the Makassar strait

eastward in the Flores Sea

before entering the Timor Sea

and then the Indian Ocean.

The upwelling from the

southeast monsoon, and some

Throughflow from Lombok

strait affect this region.

Influenced by Indian monsoon

Currents, and the South Java

Current. Fresh waters imported

through the Sunda Strait and

Great Channel in the eastern

and northern parts of this

province affect the region also.

Water spreads into the Indian

Ocean as the west-flowing

South Java and the South

Equatorial Current, and as the

south-flowing Leeuwin

Current.

SST Important seasonal SST

contrasts, with high SST in

February and relatively low

SST in August.

High SST in February, low

SST in August, with a

large seasonal temperature

difference.

High SST in February and low

SST in August with a large

seasonal temperature

difference.

Low SST in August, and slight-

ly higher SST in February, with

small seasonal SST difference.

High SST in February, low

SST in August, with large

seasonal temperature

difference.

SSS Important seasonal SSS

contrasts, the lowest February

SSS from our studied area and

high SSS in August.

High SSS, the SSS in February

is slightly higher. Seasonal

salinity difference is small.

The SSS in the northern part

are relatively low in February

but increase southward to reach

the same high values typical of

August.

Salinity is low in February due

to northeast monsoon winds.

The opposite occurs in August

when southwest monsoon

winds prevail.

Salinity is slightly higher in

February than in August, but

the seasonal difference is small.

Thermocline The thermocline depth remains

constantly at about 115 m,

although it may deepen slightly

in February.

The thermocline deepens

southward, but seasonal

difference is small.

The thermocline is shallow,

especially in August being the

shallowest for the whole

studied area due to

monsoon-induced upwelling,

but deepen rapidly southward.

The thermocline remains

shallow throughout the year,

becoming shallower in some

periods of August.

The thermocline is deep,

especially in August being the

deepest for the whole studied

area due to the increased

Throughflow.

Nitrate Nitrate distribution is highly

variable, with high contents in

the extremity of the Makassar

strait and the east.

Nitrate content is relative high

but less than region I, and is

well distributed.

Nitrate content is not high, only

higher than region IV and V.

Nitrate content is low. Nitrate content is low.

Phosphate Phosphate content reaches its

maximum value in the

extremity of the Makassar

strait, and is usually very large.

Phosphate content is low. Phosphate distribution

variable, with relatively high

contents in the North. Values

are similar to those from

region II.

Phosphate content is low. Phosphate content is low.

X.Dinget

al./Marin

eMicro

paleo

ntology58(2006)114–134

126

Table 4

Planktonic foraminiferal cluster groups: distribution and characteristics

Regions Cluster I—Banda/Java

region

Cluster II—Timor

region

Cluster III—Java

upwelling region

Cluster IV—Indian

monsoon Sumatra region

Cluster V—NWAustralia

margin regionImportant foraminifera species

N. dutertrei 21.9% (13–28.3%) 11.8% (6.8–16.1%) 13.6% (7.3–25.3%) 18.8% (10.1–27.4%) 3.9% (0.8–15.8%)

G. ruber 21.2% (13.1–25.6%) 19.6% (12.9–26.4%) 11.9% (2.3–22.5%) 20% (5.5–30.5%) 30.2% (21–52.5%)

G. bulloides 15.8% (6.2–22.6%) 9% (4.4–13.5%) 8.3% (1.3–17.1%) 4.8% (0.2–16.4%) 2.9% (0.3–7.8%)

G. sacculifer 3.4% (1–7.3%) 4.7% (2.1–6.8%) 8.1% (1.8–16.9%) 13.9% (6.9–26.3%) 13.2% (4.6–21.1%)

G. glutinata 12% (4–34.1%) 14.7% (10.8–22.1%) 7.2% (1.3–17.5%) 6.9% (0.3–13.1%) 9.7% (0–34.2%)

G. menardii 6% (1.1–20.7%) 19.8% (7.5–31.5%) 29.3% (9.2–51.8%) 7.9% (3.2–18.9%) 11.8% (0–28.7%)

N. pachyderma (R) 0.7% (0–2.2%) 1.8% (0–5.5%) 3% (0.3–6.9%) 0.7% (0–2.7%) 2.1% (0–5.6%)

P. obliquiloculata 12.11% (5.9–16%) 9.9% (4.1–15.3%) 10% (5.6–18.2%) 13.3% (6.6–19.7%) 4% (2–17.2%)

The relationship

between the planktonic

foraminiferal assemblages

and environment

characteristics in the

5 regions

The planktonic

foraminifera assemblages

characterize an

environment with strong

current activity, constant

thermocline depth,

abundant nutrient supply,

low salinity, high

productivity, and large

seasonal SST and SSS

changes.

The characteristics of

the planktonic foraminifera

assemblages are typical

of an environment with

large seasonal different

SST, high salinity,

and a relatively high

productivity, relatively

strong dissolution

and deep thermocline.

The characteristics of

the planktonic foraminifera

assemblage indicate an

environment with

upwelling activity, low

temperature, high salinity,

high productivity, and

high dissolution.

The planktonic foraminiferal

characteristics indicate an

environment with relatively

oligotrophy, low salinity,

small seasonal temperature

differences and warm in

winter, shallow thermocline,

and high dissolution.

The characteristics of the

planktonic foraminifera

assemblage indicate an

environment with oligotrophy,

high salinity, low SST in

winter, deep thermocline,

and weak dissolution.

X.Dinget

al./Marin

eMicro

paleo

ntology58(2006)114–134

127

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134128

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134 129

region are listed in Table 3, and the correspondence

between foraminiferal assemblages and environmental

parameters in Table 4.

5.1.2. Cluster II—Timor region

The planktonic foraminifera assemblages in cluster

II resemble those in cluster I, but the relative abun-

dances of N. dutertrei and G. bulloides are lower (6.8–

16.1%, average 11.8% and 4.4–13.5%, average 9%,

respectively) (Fig. 10b, a).

The abundance of G. glutinata is the highest of all

clusters, varying from 10.8% to 22.1% (average 14.7%)

(Fig. 10c), possibly reflecting relatively high produc-

tivity and higher salinity (Table 2).

The percentage of G. menardii is significantly

higher than in cluster I, from 7.5% to 31.5%, averaging

19.8% (Fig. 10e), but lower than in cluster III. The

abundances of G. ruber and G. sacculifer are similar to

those in region I, averaging 19.6% and 4.7% respec-

tively (Fig. 10g, h).

The environmental characteristics of the Timor re-

gion are listed in Table 3, and the correspondence

between foraminiferal assemblages and environmental

parameters in Table 4.

5.1.3. Cluster III—Java upwelling region

The two dominant species N. dutertrei and G. bul-

loides are less abundant than in cluster I but more

abundant than in the other clusters, and show large

fluctuations: 7.3–25.3% (average 13.6%) and 1.3–

17.1% (average 8.3%), respectively (Fig. 10b, a). The

warm water species G. menardii reaches its highest

abundance in this cluster, with large fluctuations:

9.2% to 51.8%, averaging 29.3% (Fig. 10e). G. ruber

is not very abundant, from 2.3% to 22.5%, average

11.9% (Fig. 10g). N. pachyderma (R) only averages

about 3% in this cluster, but this is a higher abundance

than in other clusters (Fig. 10f). G. sacculifer is slightly

more abundant than in clusters I and II, with percen-

tages varying from 1.8% to 16.9%, and averaging 8.1%

(Fig. 10h).

This region corresponding to cluster III is situated in

the south coast of the Java, affected by the upwelling

from the southeast monsoon wind, and small Through-

flow from Lombok strait (Table 3). Its environmental

characteristics are listed in Table 3, and the correspon-

dence between foraminiferal assemblages and environ-

mental parameters in Table 4.

Fig. 10. Planktonic foraminifera relative abundances (%) from top-core

abliquiloculatar; (e) G. menardir; (f) N. pachyderma (R); (g) G. ruber; (h)

5.1.4. Cluster IV—Indian monsoon Sumatra region

The high productivity species G. bulloides shows

abundances ranging from 0.2% to 16.4%, averaging

4.8%, which is low for the study region, although abun-

dances are lower in cluster V (Fig. 10a). The N. dutertrei

abundance ranges from 10.1% to 27.4%, average

18.8%, slightly less than in region I (Fig. 10b). The

distribution and abundance changes of N. dutertrei in-

dicate that it is influenced not only by productivity but

also by salinity. G. glutinata abundances average 6.9%,

the lowest in the studied area, whereas G. ruber is

relatively abundant, averaging 20%. G. sacculifera has

relative abundances between 6.9% and 26.3% (average

13.9%), the highest for the studied area (Fig. 10c, g, h).

The low G. bulloides and G. glutinata content, and the

presence of abundant G. ruber and G. sacculifera indi-

cate that this region is oligotrophic; the high N. dutertrei

content indicates a low salinity and shallow thermocline.

P. obliquiloculata has an average abundance of

around 10% with minor variability between samples.

In region IV, P. obliquiloculata shows its highest abun-

dances of the whole studied area, from 6.6% to 19.7%,

average 13.3% (Fig. 10d), which probably reflects se-

lective dissolution. The other dissolution-resistant spe-

cies, G. menardii, is much less abundant (average

7.9%), probably (Fig. 10e).

The environmental characteristics of this region are

listed in Table 3, and the correspondence between

foraminiferal assemblages and environmental para-

meters in Table 4.

5.1.5. Cluster V—NW Australia margin region

This cluster corresponds to core tops from the NW

Australian margin, as published in Martinez et al.

(1998). This cluster is almost the same as Group II—

the WPWP assemblage in Martinez et al. (1998). We

include these data for comparison. The high productiv-

ity species N. dutertrei and G. bulloides are least

abundant in this cluster, with an average of 3.9% and

2.9%, respectively (Fig. 10b, a). G. ruber reaches it

maximum abundance (average 30.2%), and G. saccu-

lifera is also abundant (average 13.2%), only slightly

less than in cluster IV (Fig. 10g, h). The low G.

menardii and P. obliquiloculata content, and the rela-

tively high N. pachyderma (R) abundance reflect the

weak dissolution and lower winter SST.

The environmental characteristics of this region are

listed in Table 3, and the correspondence between

samples, (a) G. bulloides; (b) N. duterire; (c) G. glutinata; (d) P.

G. sacculifer.

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134130

foraminiferal assemblages and environmental para-

meters in Table 4.

5.2. Java upwelling

The upwelling indicator species G. bulloides and the

high productivity species N. dutertrei are less abundant

in the Java upwelling region (III) than in the Banda/

Java region (I). Though the distribution of N. dutertrei

in the studied area is also influenced by salinity, the low

abundance of G. bulloides may indicate that upwelling

is not active presently in the Java region. Selective

dissolution may have changed the species composition

and abundance of planktonic foraminifera, causing a

decrease in relative abundance of solution-susceptible

species (G. ruber, G. sacculifer and G. bulloides),

whereas the solution-resistant species N. dutertrei and

G. menardii may have increased in abundance. How-

ever, G. bulloides apparently shows similarly low per-

centages in deep-sea sediments from the Java upwelling

system and plankton-tow material from the eastern

Indian Ocean (Martinez et al., 1998). Carbonate disso-

lution may explain its absence close to Java, but not its

low abundance elsewhere (Martinez et al., 1998). Phy-

toplankton productivity in the Java upwelling system

could have been reduced as compared to other upwell-

ing regions, because the ITF limits sea surface upwell-

ing of cold, nutrient-rich subsurface water.

The implication of G. bulloides abundance changes

between region I and region III for reconstructing up-

welling is analyzed further using samples from cores

SHI-9034 and SHI-9006 retrieved from region I and III,

respectively. The AMS 14C ages, curves of d18O values

(Ding et al., 2002) and important planktonic foraminif-

era species abundance changes from the two cores are

presented in Table 5 and Figs. 11 and 12.

The abundance curve of G. bulloides from core SHI-

9034 shows that this species reaches its peak abundance

in the interval 9.5–7 ka BP, with values increasing

rapidly from 22.7% to 43.5% (averaging 32.3%),

Table 5

The AMS 14C ages of core SHI-9034

Samples (cm) Age (ka BP)

34 2.71

44 3.53

105 5.56

125 6.21

304.5 11.22

360 11.96

390 12.36

430 13.95

above the average 15% abundance typical for the rest

of the core (Fig. 11). The abundance of the warm water

and solution-resistant N. dutertrei, G. menardii and P.

obliquiloculata decreased markedly in this interval

(Fig. 11), indicating enhanced productivity, higher

SSS, lower SST and weaker dissolution. This stronger

upwelling in the Java upwelling system at ~8 ka BP

likely resulted from the coeval enhanced east Asian

monsoon as described by several authors (Shi et al.,

1993; Sirocko et al., 1993; Blunier et al., 1995; Porter

and An, 1995; Wang et al., 1999a).

The age model of Core SHI-9006 is less accurate

than that of SHI-9034 because it is not based on AMS14C dating, but was developed by tuning the d18Orecord of SHI-9006 according to the oxygen isotope

chronology of Martinson et al. (1987). The maximum

abundance of G. bulloides in core SHI-9034 corre-

sponds to an interval of low G. bulloides abundance

in core SHI-9006 (averaging 11.5%) (Fig. 12). In that

interval, the abundance of the high productivity species

G. glutinata decreased also, the oligotrophic species G.

sacculifer became more abundant, and the solution-

resistant G. menardii and P. obliquiloculata increased

in abundance in core SHI-9006 (Fig. 12). The results

indicate low nutrients and strong dissolution, thus pos-

sibly a weaker ITF in the Banda/Java region between

9.5 and 7 ka BP.

Reacting to the seasonal control of strong southeast

monsoons winds, the south Java current flows from the

southeast to the northwest, so that the sea level in the

south of the Archipelago drops. In general, a lower sea

level may lead to enhanced upwelling but it may also

lead to an enhanced ITF that may limit upwelling

activity. We argue that the ITF was indeed relatively

weak during 10~8 ka BP, thus not limiting upwelling

notably. Core SHI-9034 on the northern margin of the

Java upwelling region, to the western of the south

Lombok strait, was not influenced by the weaker ITF,

and upwelling due to the southeast monsoon was vig-

orous in that region.

6. Conclusions

Planktonic foraminifera from 50 deep-sea core-top

samples from the seas around Indonesian Archipelago

and adjoining areas were studied, and the data com-

pared with data on 29 core tops off NWAustralia, with

the following results.

1) The planktonic foraminifera dissolution index

(FDX) shows strong carbonate dissolution in the stud-

ied area compared to rest of the eastern Indian Ocean.

Within the Archipelago, dissolution is strongest in the

Fig. 11. Curves of d18O values and the abundance changes of the important planktonic foraminifera species from the core SHI-9034.

X.Dinget

al./Marin

eMicro

paleo

ntology58(2006)114–134

131

Fig. 12. Curves of d18O values and the abundance changes of the important planktonic foraminifera species from the core SHI-9006.

X.Dinget

al./Marin

eMicro

paleo

ntology58(2006)114–134

132

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134 133

Java upwelling region, with lysocline levels above

~2800 m.

2) Cluster analysis of the planktonic foraminifera

assemblages separates the area into 5 regions with

different oceanographic condition: I—Banda/Java re-

gion, II—Timor region, III—Java upwelling region,

IV—Indian monsoon Sumatra region, and V—NW

Australia margin region.

3) The planktonic foraminifera assemblages from

region I are dominated by N. dutertrei, G. ruber and

G. bulloides, characterizing an environment with abun-

dant nutrient supply, low salinity, high productivity, and

large seasonal SST and SSS changes. Region II is

dominated by G. glutinata, G. menardii and N. duter-

trei, and is a typical environment with high salinity, a

relatively high productivity, relatively strong dissolu-

tion and a deep thermocline. Region III is dominated by

G. menardii and N. dutertrei, and is an environment

with upwelling activity, high productivity, and strong

dissolution. Region IV is dominated by N. dutertrei, P.

obliquiloculata, G. sacculifer and G. ruber, indicating

a relatively oligotrophic environment with low salinity,

a shallow thermocline, and high dissolution. Region V

is dominated by G. sacculifer and G. ruber and indi-

cates an environment with oligotrophy, deep thermo-

cline, and weak dissolution.

3) The abundance changes of G. bulloides in core

SHI-9006 from the Banda/Java region and core SHI-

9034 from the Java upwelling region indicate an

enhanced southeastern monsoon and stronger upwell-

ing in the Java upwelling system at 10–8 ka BP,

corresponding to a period of strengthened east Asian

monsoon.

Acknowledgements

This work was supported by the K. C. Wang Foun-

dation of Center National de la Recherche Scientifique

of France, and the National Natural Science Foundation

of China (Grant No. 40476022, 40272074).

Appendix A. Supplementary data

Supplementary data associated with this article can be

found, in the online version, at doi:10.1016/j.marmicro.

2005.10.003.

References

Ahmad, S.M., Guichard, F., Hardjawidjaksana, K., Adisaputra, M.K.,

Labeyrie, L.D., 1995. Late Quaternary paleoceanography of the

Banda Sea. Mar. Geol. 122, 385–397.

Andreasen, D.J., Ravelo, A.C., 1997. Tropical Pacific Ocean thermo-

cline depth reconstructions for the last glacial maximum. Paleo-

ceanography 12 (3), 395–413.

Be, A.W.H., 1977. An ecological, zoogeographic and taxonomic

review of recent planktonic foraminifera. In: Ramsay, A.T.S.

(Ed.), Oceanic Micropaleontology, vol. 1. Academic Press, Lon-

don, pp. 1–100.

Berger, W.H., 1979. Preservation of foraminifera. Foraminiferal eco-

logy and paleoecology. SEPM Short Course 6, 105–155.

Blunier, T., Chappellaz, J., Schwander, J., Stauffer, B., Raynaud, D.,

1995. Variations in atmospheric methane concentration during the

Holocene epoch. Nature 374, 46–49.

Bray, N.A., Hautala, S., Chong, J., Pariwono, J., 1996. Large-scale

sea level, thermocline, and wind variations in the Indonesian

Throughflow region. J. Geophys. Res. 101 (C5), 12239–12254.

Brock, J.C., McClain, C.R., Anderson, D.M., Prell, W.L., Hay, W.W.,

1992. Southwest monsoon circulation and environments of recent

planktonic foraminifera in the northwestern Arabian Sea. Paleo-

ceanography 7 (6), 799–813.

Cane, M.A., 1998. A role for the tropical Pacific. Science 282,

59–61.

Cannariato, K.G., Ravelo, A.C., 1997. Pliocene–Pleistocene evolution

of eastern tropical Pacific surface water circulation and thermo-

cline depth. Paleoceanography 12 (6), 805–820.

Curry, W.B., Thunell, R.C., Honjo, S., 1983. Seasonal changes in the

isotopic composition of planktonic foraminifera collected in Pan-

ama Basin sediment traps. Earth Planet. Sci. Lett. 64 (1), 33–43.

Ding, X., Guichard, F., Bassinot, F., Labeyrie, L., Fang, N.Q., 2002.

Evolution of heat transport pathways in the Indonesian Archipel-

ago during last deglaciation. Chin. Sci. Bull. 47 (22), 1912–1917.

Duplessy, J.C., Be, A.W.H., Blanc, P.L., 1981. Oxygen and carbon

isotopic composition and biogeographic distribution of planktonic

foraminifer in the Indian Ocean. Palaeogeogr. Palaeoclimatol.

Palaeoecol. 33, 9–46.

Emerson, S., Bender, M., 1980. Carbon fluxes at the sediment water

interface of the deep-sea: calcium carbonate preservation. J. Mar.

Sci. 39, 139–162.

Enfield, D.B., 1989. El Nino, past and present. Rev. Geophys. 27 (1),

159–187.

Fairbanks, R.G., Sverdlove, M., Free, R., Wiebe, P.H., Be, A.W.H.,

1982. Vertical distribution and isotopic fractionation of living

planktonic foraminifera from the Panama Basin. Nature 298,

841–844.

Gingele, F.X., De Deckker, P., Girault, A., Guichard, F., 2002. History

of the South Java Current over the past 80 ka. Palaeogeogr.

Palaeoclimatol. Palaeoecol. 183, 247–260.

Godfrey, J.S., Golding, T.J., 1981. The Sverdrup relation in the

Indian Ocean, and the effect of Pacific–Indian Ocean through-

flow on Indian Ocean circulation and on the East Australian

Current. J. Phys. Oceanogr. 11, 771–779.

Godfrey, J.S., Ridgway, K.R., 1985. The large-scale environment of

the poleward-flowing Leeuwin Current, western Australia: long-

shore steric height gradients, wind stresses and geostrophic flow.

J. Phys. Oceanogr. 15, 481–495.

Gordon, A.L., 1986. Interocean exchange of thermocline water.

J. Geophys. Res. 91, 5037–5046.

Gordon, A.L., Fine, R.A., 1996. Pathways of water between the

Pacific and Indian Oceans in the Indonesian seas. Nature 379,

146–149.

Hanebuth, T., Stattegger, K., Grootes, P.M., 2000. Rapid flooding of

the Sunda Shelf: a late-glacial sea-level record. Science 288,

1033–1035.

X. Ding et al. / Marine Micropaleontology 58 (2006) 114–134134

Hilbrecht, H., 1997. Morphologic gradation and ecology in Neoglo-

boquadrina pachyderma and N. dutertrei (planktic foraminifera)

from core top sediments. Mar. Micropaleontol. 31, 31–44.

Hirst, A.C., Godfrey, J.S., 1993. The role of Indonesian throughflow

in a global ocean GCM. J. Phys. Oceanogr. 23, 1057–1086.

Kawahata, H., Nishimura, A., Gagan, M.K., 2002. Seasonal change in

foraminiferal production in the western equatorial Pacific warm

pool: evidence from sediment trap experiments. Deep-Sea Res. II

49, 2783–2800.

Levitus, 1998. World Ocean Atlas. http://www.ifremer.fr/sismer/

references/climatologies/.

Linsley, B.K., 1996. Oxygen-isotope record of sea level and climate

variations in the Sulu Sea over the past 150,000 years. Nature 380,

234–237.

Martinez, J.I., De Dekker, P., Chivas, A.R., 1997. New estimate for

salinity changes in the Western Pacific Warm Pool during the Last

Glacial Maximum: oxygen-isotope evidence. Mar. Micropaleon-

tol. 32, 311–340.

Martinez, J.I., Taylor, L., De Deckker, P., Barrows, T., 1998. Plank-

tonic foraminifera from the eastern Indian Ocean: distribution and

ecology in relation to the Western Pacific Warm Pool (WPWP).

Mar. Micropaleontol. 34, 121–151.

Martinez, J.I., De Deckker, P., Barrows, T.T., 1999. Palaeoceanogra-

phy of the last glacial maximum in the eastern Indian Ocean:

planktonic foraminiferal evidence. Palaeogeogr. Palaeoclimatol.

Palaeoecol. 147, 73–99.

Martinson, D.G., Pisias, N.G., Hays, J.D., Imbrie, J., Moore, T.C.,

Shackleton, N.J., 1987. Age dating and the orbital theory of the

ice age: development of a high resolution 0 to 300,000 year

chronostratigraphy. Quat. Res. 27, 1–29.

Muller, A., Opdyke, B.N., 2000. Glacial–interglacial changes in

nutrient utilization and paleoproductivity in the Indonesian

Throughflow sensitive Timor Trough, easternmost Indian Ocean.

Paleoceanography 15, 85–94.

Patrick, A., Thunell, R.C., 1997. Tropical Pacific sea surface tem-

peratures and upper water column thermal structure during the last

glacial maximum. Paleoceanography 12, 649–657.

Pflaumann, U., Jian, Z., 1999. Modern distribution patterns of plank-

tonic foraminifera in the South China Sea and western Pacific: a

new transfer technique to estimate regional sea-surface tempera-

tures. Mar. Geol. 156, 41–83.

Pflaumann, U., Duprat, J., Pujol, C., Labeyrie, L.D., 1996. SIMMAX:

a modern analog technique to deduce Atlantic sea surface tem-

peratures from planktonic foraminifera in deep-sea sediments.

Paleoceanography 11 (1), 15–35.

Porter, S.C., An, Z., 1995. Correlation between climate events in the

North Atlantic and China during the last glaciation. Nature 375,

305–308.

Prell, W.L., Curry, W.B., 1981. Faunal and isotope indices of mon-

soonal upwelling: western Arabian Sea. Oceanol. Acta 4, 91–98.

Ravelo, A.C., Fairbanks, R.G., Phillander, S.G.H., 1990. Reconstruct-

ing tropical Atlantic hydrography using planktonic foraminifera

and an ocean model. Paleoceanography 5, 409–431.

Reynolds, L.A., Thunell, R.C., 1986. Seasonal production and mor-

phologic variation of Neoglaoboquadrina pachyderma (Ehren-

berg) in the northeast Pacific. Micropaleontology 32, 1–18.

Shi, Y., Kong, Z., Wang, S., Tang, L., Yao, T., Zhao, X., Zhang, P.,

Shi, S., 1993. Mid-Holocene climates and environments in China.

Glob. Planet. Change 7, 219–233.

Siedler, G., Church, J., Gould, J., 2001. Ocean Circulation and

Climate: Observing and Modeling the Global Ocean, International

Geophysics Series, vol. 77. Academic Press, A Harcourt Science

and Technology Company, pp. 303–314.

Sirocko, F., Sarnthein, M., Erlenkeuser, H., Lange, H., Arnold, M.,

Duplessy, J.C., 1993. Century-scale events in monsoonal climate

over the past 24,000 years. Nature 364, 322–324.

Thiede, J., Junge, B., 1992. Faunal and floral indicators of ocean

coastal upwelling (NW African and Peruvian continental mar-

gins). In: Summerhayes, C.P., Prell, W.L., Emeis, K.C. (Eds.),

Upwelling Systems: Evolution Since the Miocene, Geological

Society London Special Publication, vol. 64, pp. 47–76.

Thompson, P.R., 1981. Planktonic foraminifera in the western north

Pacific during the past 150,000 years: comparison of modern and

fossil assemblages. Palaeogeogr. Palaeoclimatol. Palaeoecol. 35,

241–279.

Thunell, R.C., Reynolds, L.A., 1984. Sedimentation of planktonic

foraminifera: seasonal changes in species flux in the Panama

Basin. Micropaleontology 30, 243–262.

Thunell, R., Anderson, D., Gellar, D., Miao, Q., 1994. Sea-surface

temperature estimates for the tropical western Pacific during the

Last Glaciation and their implications for the Pacific Warm Pool.

Quat. Res. 41, 225–264.

Tomczak, M., Godfrey, J.S., 1994. Regional Oceanography: An In-

troduction. Pergamon, Oxford. 422 pp.

Visser, K., Thunell, R., Stott, L., 2003. Magnitude and timing of

temperature change in the Indo-Pacific warm pool during degla-

ciation. Nature 421, 152–155.

Wang, P.X., 1998. Western Pacific in glacial cycles: seasonality in

marginal seas and variability of Warm Pool. Sci. China (Ser. D) 4

(1), 35–41.

Wang, L.J., Sarnthein, M., Grootes, P.M., Erlenkeuser, H., 1999a.

Millennial reoccurrence of century-scale abrupt events of East

Asian monsoon: a possible heat conveyor for the global deglaci-

ation. Paleoceanography 14 (6), 725–731.

Wang, L., Sarnthern, M., Erlenkeuser, H., Grimalt, J., Grootes, P.,

Heilig, S., Ivanova, E., Kienast, M., Pelejero, C., Pflaumann, U.,

1999b. East Asian monsoon climate during the late Pleistocene:

high-resolution sediment records from the South China Sea. Ma-

rine Geology 156, 245–284.