You are on page 1of 12

Available online at www.sciencedirect.

com

Palaeogeography, Palaeoclimatology, Palaeoecology 261 (2008) 234 – 245


www.elsevier.com/locate/palaeo

A major change in monsoon-driven productivity in the tropical Indian Ocean


during ca 1.2–0.9 Myr: Foraminiferal faunal and stable isotope data
Anil K. Gupta ⁎, M. Sundar Raj, K. Mohan, Soma De
Department of Geology & Geophysics, Indian Institute of Technology, Kharagpur – 721 302, India
Received 16 July 2007; received in revised form 4 January 2008; accepted 7 January 2008

Abstract

Tropical climate is variable on astronomical time scale, driving changes in surface and deep-sea fauna during the Pliocene–Pleistocene. To
understand these changes in the tropical Indian Ocean over the past 2.36 Myr, we quantitatively analyzed deep-sea benthic foraminifera and
selected planktic foraminifera from N 125 μm size fraction from Deep Sea Drilling Project Site 219. The data from Site 219 was combined with
published foraminiferal and isotope data from Site 214, eastern Indian Ocean to determine the nature of changes. Factor and cluster analyses of the
28 highest-ranked species distinguished four biofacies, characterizing distinct deep-sea environmental settings. These biofacies have been named
after their most dominant species such as Stilostomella lepidula–Pleurostomella alternans (Sl–Pa), Nuttallides umbonifer–Globocassidulina
subglobosa (Nu–Gs), Oridorsalis umbonatus–Gavelinopsis lobatulus (Ou–Gl) and Epistominella exigua–Uvigerina hispido-costata (Ee–Uh)
biofacies. Biofacies Sl–Pa ranges from ~ 2.36 to 0.55 Myr, biofacies Nu–Gs ranges from ~ 1.9 to 0.65 Myr, biofacies Ou–Gl ranges from ~ 1 to
0.35 Myr and biofacies Ee–Uh ranges from 1.1 to 0.25 Myr. The proxy record indicates fluctuating tropical environmental conditions such as
oxygenation, surface productivity and organic food supply. These changes appear to have been driven by changes in monsoonal wind intensity
related to glacial–interglacial cycles. A shift at ~ 1.2–0.9 Myr is observed in both the faunal and isotope records at Site 219, indicating a major
increase in monsoon-induced productivity. This coincides with increased amplitude of glacial cycles, which appear to have influenced low latitude
monsoonal climate as well as deep-sea conditions in the tropical Indian Ocean.
© 2008 Elsevier B.V. All rights reserved.

Keywords: DSDP 214 and 219; Tropical Indian Ocean; Faunal and isotope data

1. Introduction This climatic transition is often known as the Mid-Pleistocene


Transition (MPT) (e.g., Raymo et al., 1997) or Mid-Pleistocene
Earth's climate has evolved from a state of relative warmth Revolution (MPR) (e.g., Hayward, 2001; Xu et al., 2005), and is
during the mid-Pliocene, with major ice sheets restricted to well-documented in numerous proxy records (Mix et al., 1995;
Antarctica, to a cooler world during the late Pleistocene marked Raymo et al., 1997; Gupta et al., 2001; Hayward, 2001; Gupta
with extensive bipolar ice sheets and increased pole to equator and Dhingra, 2004; Xu et al., 2005). At ~ 900 Kyr BP,
temperature gradients (Raymo, 1997; Ravelo et al., 2004). the mixed-layer thickness reduced and thermocline shoaled in
These changes are believed to have been driven by changes in the eastern Indian Ocean, and a 100-Kyr component began to
Earth's orbital parameters with a switch from 41-Kyr mode to a dominate in the Indian monsoon variability (Gupta et al., 2001;
100-Kyr rhythm after 900 Kyr BP (Raymo, 1997; Lisiecki and Gupta and Dhingra, 2004). Elongated benthic foraminifera
Raymo, 2005). Changes in polar ice volume brought significant (including Stilostomella lepidula) suffered widespread extinc-
changes in the tropics (Sirocko et al., 1999; Gupta et al., 2001). tion across this climatic transition (Weinholz and Lutze, 1989;
Gupta, 1993; Hayward, 2001; Kawagata et al., 2005, 2006).
Numerous studies have recently been undertaken to under-
⁎ Corresponding author. Tel.: +91 3222 283368; fax: +91 3222 282700. stand paleoclimatic and paleoceanographic evolution of the
E-mail address: anilg@gg.iitkgp.ernet.in (A.K. Gupta). Indian Ocean during the Plio-Pleistocene and the Holocene
0031-0182/$ - see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.01.012
A.K. Gupta et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 261 (2008) 234–245 235

Fig. 1. Location map of Deep Sea Drilling Program (DSDP) Sites 214 and 219 in the tropical Indian Ocean. Also shown are deep-ocean currents (You, 2000). AABW =
Antarctic Bottom Water, NADW = North Atlantic Deep Water, CPDW = Circumpolar Deep Water.

using faunal and geochemical data (Gupta and Srinivasan, 2. Site location: materials and methods
1990; Kroon et al., 1991; Gupta et al., 2003). Most of these
studies have focused on high-productivity areas of the Arabian DSDP Site 214 is located in the eastern Indian Ocean on the
Sea such as the northwestern and northeastern Arabian Sea, Ninetyeast Ridge (11o20.21′S, 88o43.08′E; present water depth
where surface productivity changes are driven by seasonal 1665 m). This site is presently located beneath a hydrological
reversals in the monsoon winds (e.g., Reichart et al., 1998; Jung
et al., 2001; Schmiedl and Leuschner, 2005; Schmiedl and
Mackensen, 2006). The southeastern Arabian Sea is also an
important region where low salinity surface currents from the
Bay of Bengal mix with those of the Arabian Sea.
Changes in organic flux to the seafloor due to variations in
surface productivity modulate deep-sea faunal composition
(e.g. Jorissen et al., 1995; Gupta and Thomas, 1999). The
amount of organic flux to the seafloor not only depends on
surface production but also on the nature of deep-sea column.
Well-oxygenated deep-sea circulation may cause reminera-
lization of organic carbon resulting in little organic material
reaching the seafloor (Schmiedl and Mackensen, 2006). To
understand if the changes in the surface and deep-water column
of the tropical Indian Ocean were driven by the Indian Ocean
climate (monsoon) and deep-sea circulation, we analyzed a
2.36 Myr record of deep-sea benthic foraminifera from Deep
Sea Drilling Project (DSDP) Site 219 from the southeastern
Arabian Sea and combined it with published planktic and
benthic records from Site 214, eastern Indian Ocean (Fig. 1).
The sediment accumulation rate is low to moderate at Sites 214
(0.91 cm/Kyr) and 219 (1.03 cm/Kyr) (Fig. 2). Fig. 2. Numerical age versus depth at Site 219 (Gupta and Thomas, 1999).
236 A.K. Gupta et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 261 (2008) 234–245

Table 1
VARIMAX rotated factor scores of eight significant factors at DSDP Site 219
Species name Factor1 Factor2 Factor3 Factor4 Factor5 Factor6 Factor7 Factor8
Astrononian umbilicatulum − 0.27678 −0.54435 0.21522 0.12411 0.24152 − 0.09983 − 0.21964 −0.03865
Bolivina pseudoplicata − 0.09604 0.63124 0.14031 0.28713 0.33582 0.01515 0.26011 0.1557
Bolivina pusilla − 0.18667 0.1082 −0.06034 − 0.35088d 0.54077 0.29692 − 0.10807 0.06134
Bulimina aculeata 0.62835 0.34862 0.16157 − 0.0122 − 0.0527 − 0.00905 0.07341 −0.38974c
Bulimina alazanensis 0.14442 0.2046 0.16591 − 0.12433 − 0.33426 0.43545 0.50979 −0.03206
Bulimina striata 0.68119 −0.10787 0.29073 0.03259 − 0.21666 − 0.30203 0.00952 0.2498
Cassidulina carinata − 0.40159a 0.67055 0.24401 0.34092 − 0.07719 0.12485 0.03409 0.05668
Cibicides bradyi 0.41532 −0.50645 0.13069 0.18526 0.17863 0.15353 0.27073 0.0007
Cibicides wuellerstorfi 0.63842 0.02374 0.28183 − 0.08202 0.02421 0.22133 − 0.30961 −0.03629
Discopulvinulina bertheloti − 0.19432 0.73936 0.32118 0.35622 0.14037 0.19348 0.01261 0.05632
Eggerella bradyi 0.5745 −0.06103 0.14104 0.29633 − 0.08572 − 0.14829 − 0.04954 0.48948
Ehrenbergina carinata − 0.45695a −0.27574 0.1506 − 0.19893 0.14098 − 0.17699 0.26047 0.27035
Epistominella exigua 0.29132 0.06991 0.08212 − 0.48336d 0.59049 − 0.0765 0.14786 0.0468
Evolvocassidulina bradyii 0.46502 0.31869 −0.27943 − 0.06257 0.00804 − 0.36928 − 0.07362 −0.03243
Gavelinopsis lobatulus − 0.37221a −0.26279 0.29207 0.12792 0.34095 − 0.10859 − 0.15582 −0.40186c
Globocassidulina pacifica 0.56338 −0.2219 −0.07875 0.43029 0.02424 0.10288 − 0.25646 0.37342
Globocassidulina subglobosa 0.08172 0.0121 −0.65278b 0.33966 0.2013 0.31311 − 0.01501 −0.04683
Gyroidinoides cibaoensis 0.47499 0.20755 −0.4199b 0.00346 0.02495 − 0.29373 0.12424 0.15028
Hoglandulina elegans 0.51997 −0.42207 0.01022 0.47281 0.24827 0.07963 0.06066 −0.18306
Melonis barleeanum 0.32735 0.03988 0.66648 − 0.24837 − 0.11669 − 0.13728 − 0.15883 −0.09138
Nuttallides umbonifer 0.02991 0.06983 −0.68766b − 0.14073 0.20614 0.07212 − 0.05179 0.13001
Oridorsalis umbonatus 0.03694 0.08683 −0.41545b 0.28318 − 0.23448 − 0.25462 − 0.05876 −0.51602c
Osangularia culter 0.19279 −0.47841 0.10761 0.14219 − 0.06176 0.07296 0.60514 −0.16739
Pleurostomella alternans − 0.66316a −0.0648 0.07837 0.06686 0.05568 − 0.18797 0.01202 0.16325
Pullenia bulloides − 0.04405 −0.28612 0.01745 − 0.09139 − 0.28655 0.68298 − 0.3178 0.08143
Stilostomella lepidula − 0.76712a 0.03078 0.14501 0.22505 − 0.18706 − 0.14994 − 0.2423 −0.0025
Uvigerina hispido-costata 0.60918 0.27717 −0.07434 − 0.38523d − 0.17286 0.07584 − 0.17436 −0.09885
Uvigerina proboscidea − 0.57463a −0.2155 −0.28727b − 0.2008 − 0.4666 − 0.02655 0.15959 0.15555
% variance 27.73 15.18 11.50 8.44 7.66 6.0 5.28 5.02
a
Biofacies Sl–Pa, bBiofacies Nu–Gs, cBiofacies Ou–Gl, dBiofacies Ee–Uh.

front in the core of the Indonesian Throughflow (Du et al., samples were dried in an electric oven at ~ 50 °C and transferred
2005), and is under the influence of the South Equatorial into glass vials. Hard sediment samples were treated with 2–3
Current (SEC), which intensifies during northern summer when drops of 2% hydrogen peroxide. We applied the age model for
southeast trade winds are stronger and Indian summer monsoon Site 219 based on planktic foraminifer datums suggested by
begins to strengthen (Fig. 1; Tchernia, 1980). Site 214 is bathed Gupta and Thomas (1999). The sediment accumulation rate at
by the North Indian Deep Water (NIDW) with a dissolved Site 219 shows a major increase at ~ 1.6 Myr and thereafter is
oxygen content of ~ 5.8 ml/L (GEOSECS, 1983), which is a constant (Fig. 2). The interpolated numerical ages are updated to
mixture of Antarctic Bottom Water (AABW), North Atlantic the Berggren et al. (1995) time scale, with an average time
Deep Water (NADW) and deep water of the northern Indian resolution of 33 Kyr per sample.
Ocean origin (Tchernia, 1980). DSDP Site 219 was drilled We generated benthic foraminiferal census data from an
during Leg 23 on the crest of the Laccadive-Chagos Ridge, aliquot of ~ 300 specimens in the N 125 μm size fraction (census
southeastern Arabian Sea (9°01.75′N, 72°52.67′E; water depth data are made available online as Table 1s). This size fraction
1764 m). At present, this site is bathed by the NIDW with allows us to compare our results with recent results from the
dissolved oxygen content of ~ 4.8 ml/L (GEOSECS, 1983), and Atlantic and Indian Oceans (Mackensen et al., 1993, 1995;
lies within an area where surface production is higher due to Schmiedl et al., 1997; Gupta and Thomas, 2003; Gupta et al.,
monsoon-induced upwelling (Gupta and Thomas, 1999; 2004). Factor and cluster analyses were performed on relative
Shankar et al., 2002). abundance data of the 28 highest-ranked species (Table 1;
Seventy one core samples of 10 cm3 volume from Site 219 supplementary Table 1s), which are present in at least three
were analyzed from a sediment sequence extending back in time samples with a percentage of 2% or more in at least one sample.
to ~ 2.36 Myr. Samples were soaked in water with baking soda The analysis performs standardization of the data and calcula-
for 8–12 h, and washed over a 63 µm size sieve. The washed tion of covariance matrix that helps remove noise and reduce the

Fig. 3. Dendogram based on Q-mode cluster analysis of 71 samples from DSDP Site 219 using Ward's Minimum Variance method. Four clusters have been
identified on the basis of the number of clusters versus semi-partial R2. Each cluster corresponds to a biofacies named after the most dominant species within each
cluster.
A.K. Gupta et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 261 (2008) 234–245 237
238 A.K. Gupta et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 261 (2008) 234–245

Table 2
Environmental preferences of characteristic benthic foraminiferal species defining different biofacies at Site 219
Species Interpretation
B. pusilla High organic carbon flux (Thomas and Gooday, 1996; Gupta and Satapathy, 2000)
B. aculeata No relation with food flux (Hermelin and Shimmield, 1990)
Calm depositional regimes with clayey and organic-rich sediments (Mackensen et al., 1993)
Interglacial, high productivity (Almogi-Labin et al., 2000)
C. carinata High food supply (Gupta and Thomas, 1999)
C. wuellerstorfi Raised epibenthic, suspension feeder, high energy (Lutze and Thiel, 1989)
Oligotrophic (Linke and Lutze, 1993)
Seasonal food supply (Loubere and Fariduddin, 1999)
AABW (Corliss, 1979, 1983)
E. carinata High organic food (Gupta and Satapathy, 2000)
Low-oxygen, high organic carbon (Nomura, 1991)
E. exigua Epibenthic, cosmopolitan, abyssal, opportunistic, phytodetritus feeder (Gooday, 1988; Gooday and Turley, 1990)
Seasonal food fluxes (Smart et al., 1994; Loubere and Fariduddin, 1999)
G. lobatulus Pulsed food supply (Gooday,1994)
Cool waters, partially pulsed food supply (Gupta and Thomas, 1999)
G. subglobosa Warmer AABW (Corliss, 1979)
Circumpolar Deep Water (Schnitker, 1980)
Phytodetritus feeder (Gooday, 1994)
Oligotrophic (Mackensen et al., 1995)
Low-productivity associated with NADW (Fariduddin and Loubere, 1997)
Strong bottom currents, elevated, oligotrophic (Mackensen et al., 1995; Schmiedl et al., 1997)
G. cibaoensis Low-oxygen (Gupta and Thomas, 1999)
N. umbonifera AABW (Streeter 1973)
Corrosive bottom water (Bremer & Lohmann, 1982)
Oligotrophic (Gooday, 1994)
O. umbonatus AABW (Corliss, 1979)
High carbonate saturation, motile (Miao and Thunell, 1993)
High oxygen, low organic carbon (Mackensen et al., 1995)
Cosmopolitan, both oligotrophic and eutrophic (Schmiedl 1995; Schmiedl and Mackensen, 1997)
P. alternans Abyssal species, characteristic of variable organic flux (Gupta, 1994)
S. lepidula Low-oxygen, organic rich (Boersma, 1990)
High productivity (Gupta, 1993)
Low productivity (Kaiho, 1999)
U. hispido-costata High organic carbon (Altenbach and Sarnthein, 1989; Rathburn and Corliss, 1994)
U. proboscidea Shallow infaunal, high organic carbon, variable oxygenation (Lutze and Coulbourn, 1984)
High, year-round productivity (Gupta and Srinivasan, 1992; Gupta, 1997; Jannink et al., 1998; Gupta and Thomas, 1999;
Gupta et al., 2001, 2004; Murgese and De Deckker, 2005)
Low seasonality (Ohkushi et al., 2000)

data to more meaningful variables. The R-mode factor analysis We calculated Globigerina bulloides percentages from Site
identifies the species associations and Q-mode cluster analysis 219 and combined them with published percent U. proboscidea,
discerns the sample groups. Globigerinoides sacculifer, Globorotalia menardii (complex)
R-mode Principal Component Analysis (PCA) was performed and Gs. sacculifer stable isotope data from eastern Indian
on the correlation matrix followed by orthogonal VARIMAX Ocean Site 214 (Gupta and Dhingra, 2004; Fig. 7). G. bulloides
rotation using the SAS software package (SAS Inc., 1988). Based is a well tested monsoon proxy used in several paleomonsoonal
on a scree plot (x–y plot) of eigenvalues versus the number of studies (Kroon et al., 1991; Overpeck et al., 1996; Gupta et al.,
factors and screening of the VARIMAX factor scores and species 2003).
associations, 8 factors were considered that account for 86.8% of
the total variance (Table 1). Factors that do not show relevant 3. Results and environmental preferences of benthic biofacies
species associations due to their low factor scores and rare spe-
cies occurrence, were not considered to constitute assemblages. Multivariate analysis was performed on reduced data of
Q-mode cluster analysis using Ward's Minimum Variance meth- benthic foraminifera from Site 219 to reduce noise from the data
od was run on a covariance matrix to identify sample groups set, which is induced by postmortem taphonomic changes.
belonging to each biofacies (Fig. 3). Four clusters representing The analysis involves standardization of the data and calcula-
four biofacies were identified on the basis of a plot of semi-partial tion of covariance matrix that reduces the aliasing. Given below
R2 values versus the number of clusters. We used information on are details of each biofacies and their inferred environments
environments of modern benthic foraminifera (Table 2) to (Table 3). Each biofacies is named after the most dominant
interpret different biofacies at Site 219 (Table 3, Figs. 4 and 5). species present (Table 4).
A.K. Gupta et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 261 (2008) 234–245 239

Table 3 characteristic species of this biofacies are O. umbonatus,


Benthic foraminiferal biofacies and their interpreted environments at DSDP Site G. lobatulus and Bulimina aculeata, indicating low to interme-
219, southeastern Arabian Sea
diate organic carbon flux and high seasonality (Tables 2 and 3).
Biofacies Environment
Sl–Pa (Factor 1 negative scores) High and sustained organic flux 3.4. Biofacies Epistominella exigua–Uvigerina hispido-costata
Stilostomella lepidula (− 0.76712) (Ee–Uh)
Pleurostomella alternans (− 0.66316)
Uvigerina proboscidea (− 0.57463)
Ehrenbergina carinata (− 0.45695) Epistominella exigua, Uvigerina hispido-costata and Boli-
Cassidulina carinata (− 0.40159) vina pusilla are the characteristic species of biofacies Ee–Uh,
Gavelinopsis lobatulus (−0.37221) with high negative Factor 4 scores (Figs. 4 and 5; Tables 1
Nu–Gs (Factor 3 negative scores) Low and variable flux of and 3). This biofacies is present from ~ 1.1 to 0.25 Myr and is
Nuttallides umbonifer (− 0.68766) organic matter, high seasonality,
suggestive of intermediate to high organic flux, high seasonality
Globocassidulina subglobosa (− 0.65278) well-oxygenated and cold
Gyroidinoides cibaoensis (− 0.4199) deep waters and well-oxygenated conditions (Tables 2 and 3).
Oridorsalis umbonatus (− 0.41545)
Ou–Gl (Factor 8 negative scores) Low to intermediate organic 4. Discussion and conclusions
Oridorsalis umbonatus (− 0.51602) carbon flux, high seasonality
Gavelinopsis lobatulus (−0.40186)
Deep-sea benthic foraminifera have been used to understand
Bulimina aculeata (− 0.38974)
Ee–Uh (Factor 4 negative scores) Intermediate to high organic flux, changes in deep-water conditions driven by climate forcing
Epistominella exigua (−0.48336) high seasonality, well-oxygenated during the Pliocene and Pleistocene (Corliss et al., 1986;
Uvigerina hispido-costata (− 0.38523) conditions Thomas et al., 1995; Schmiedl and Mackensen, 1997; Gupta
Bolivina pusilla (− 0.35088) et al., 2001). Several studies have shown the relationship be-
tween benthic faunal composition, productivity of the overlying
3.1. Biofacies Stilostomella lepidula–Pleurostomella alternans waters and organic flux to the seafloor (Lutze and Coulbourn,
(Sl–Pa) 1984; Miao and Thunell, 1993; Smart et al., 1994; Thomas and
Gooday, 1996; Gupta et al., 2004; Smart et al., 2007). Others
Biofacies Sl–Pa is defined by species with a high negative score suggested oxygen and food supply are the main factors con-
on Factor 1, in the ~2.36 to 0.55 Myr interval (Figs. 4 and 5). The trolling the spatial and in-sediment distribution of benthic fora-
characteristic species of this biofacies are Stilostomella lepidula, minifera (Hermelin and Shimmield, 1990; Bernhard, 1992;
Pleurostomella alternans, Uvigerina proboscidea, Ehrenbergina Jorissen et al., 1995; Schmiedl et al., 1997; Gupta and Thomas,
carinata, Cassidulina carinata and Gavelinopsis lobatulus, 1999; den Dulk et al., 2000). This group explains seasonal
indicating high and sustained organic flux (Tables 2 and 3). fluctuations in primary production (Thomas et al., 1995;
Thomas and Gooday, 1996; Schmiedl et al., 1997, 2000; den
3.2. Biofacies Nuttallides umbonifer–Globocassidulina Dulk et al., 2000; Wollenburg and Kuhnt, 2000; Gupta and
subglobosa (Nu–Gs) Thomas, 2003). Gooday (1994), Thomas et al. (1995) and
Thomas and Gooday (1996) classified phytodetritus species
Characteristic species of this biofacies are Nuttallides (like E. exigua) and mobile, infaunal species (like species of
umbonifer, Globocassidulina subglobosa, Gyroidinoides Melonis and Uvigerina) as indicators of predominantly sea-
cibaoensis and Oridorsalis umbonatus, with Factor 3 negative sonally pulsed versus continuous primary production. Thus
scores. This biofacies is present from ~ 1.9 to 0.65 Myr, benthic foraminifera are considered useful for estimating paleo-
indicating low and variable flux of organic matter (oligotrophic), fluxes in high-productivity areas and they are also more resis-
high seasonality, well-oxygenated and cold deep waters (Figs. 4 tant to diagenetic change compared to planktic foraminifera.
and 5, Tables 2 and 3). It has been suggested by Gupta et al. However, in oligotrophic areas deep-sea oxygenation plays an
(2006) that N. umbonifer is out-competed by species of Buli- important role in controlling benthic foraminifera over variable
mina and Uvigerina when the food supply is high and sustained, time scales. Gupta and Thomas (1999) suggested that changes
unless the waters become corrosive to CaCO3. Under conditions in oxygenation are linked partially to productivity and partially
of relatively low, seasonal flux of food N. umbonifer may be to changes in deep-water ventilation.
swamped by the opportunistic species such as E. exigua. Nut- Wind driven coastal and open-ocean surface productivity
tallides umbonifer thus dominates those regions where other influences organic carbon flux and oxygenation of deep waters
calcareous benthic foraminiferal taxa cannot grow optimally, controlling benthic populations in the Arabian Sea. In the
either as a result of very low food supply (extreme oligotrophy), northern part of the Indian Ocean, the wind regimes follow
or a high carbonate corrosivity (Gupta et al., 2006). seasonal change in circulation producing widespread upwelling
controlling surface productivity near Site 219 (Fig. 1). The in
3.3. Biofacies Oridorsalis umbonatus–Gavelinopsis lobatulus situ (0–50 m) primary production above Sites 214 and 219 is
(Ou–Gl) b20 mg C m− 3 h− 1 during June–September, when the summer
monsoon and trade winds are stronger. This is less than one
This biofacies is defined by species with negative Factor 8 fifth that of the northwestern Arabian Sea primary production
scores, across the ~ 1 to 0.35 Myr interval (Figs. 4 and 5). The (Krey, 1973). The primary production reduces significantly to
240 A.K. Gupta et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 261 (2008) 234–245

Fig. 4. Percent distribution of benthic foraminiferal species which are most dominant in different biofacies (panels a–d) at DSDP Site 219. Each colour shade
represents a particular species as given at the bottom of the figure. Open arrows mark the shift in the foraminiferal faunal and stable isotope record.

b 0.50 mg C m− 3 h− 1 above Site 219 and b0.10 mg C m− 3 h− 1 to 100-Kyr periodicities, and an increased intensity of ice age
above Site 214 during December–March when variable and cycles in the Northern Hemisphere (Raymo et al., 1997). This
weak winter monsoon winds blow across the Indian Ocean also corresponds to the Stilostomella extinction observed in
(Krey, 1973; Tchernia, 1980; Schott and McCreary, 2001). different ocean basins (Gupta, 1993; Gupta et al., 2001;
Because of the location of Sites 214 and 219 in oligotrophic Hayward, 2001; Kawagata et al., 2005, 2006). During this
areas with higher dissolved oxygen content during the studied time, the East Asian monsoon system strengthened and contrast
interval, the changes in benthic foraminiferal population might between the summer and winter monsoon wind regime
have been linked to the supply of organic food. increased (Xiao and An, 1999). Thus variability in the East
At Sites 214 and 219, the faunal and isotope data show a shift Asian monsoon increased (Qiang et al., 2001).
during ~ 1.2–0.9 Myr very close to the Mid-Pleistocene At Site 219, U. proboscidea was a dominant benthic species
Transition (MPT), coinciding with the transition from 41-Kyr from 2.36 to 1.2 Myr and G. bulloides was rare (Fig. 7). The
A.K. Gupta et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 261 (2008) 234–245 241

Fig. 5. Cumulative percentages of benthic foraminiferal species dominant in different biofacies at DSDP Site 219. Open arrow marks a shift benthic foraminiferal
fauna, coinciding with the strengthening of the Indian monsoon.

benthic biofaces Sl–Pa and Nu–Gs alternated during this time crease in % Gr. menardii complex (Fig. 7). This higher
(Fig. 6). This indicates that the surface productivity was low surface productivity could have been linked to strong trade
(rare G. bulloides) above Site 219 during this time but there was winds, which would have intensified during cold intervals,
an increased lateral advection of nutrient-rich deep water thus causing intense open-ocean upwelling and high surface
(NIDW) from the oxygen minimum zone (OMZ) of the productivity. Site 214 is located beneath the core of the
northwestern Arabian Sea to the southeastern Arabian Sea.
Dickens and Owen (1999) suggested lateral expansion of OMZ
waters to the eastern and southeastern parts of the Indian Ocean Table 4
in the late Neogene. In the northwestern Arabian Sea, the Relationship between biofacies factor scores of each species and its average
abundance in respective clusters
summer monsoon winds cause widespread upwelling and high
surface productivity and thus increased organic flux to the Biofacies Factor 1-ve Factor scores Cluster IV
(average %)
seafloor, intensifying oxygen minimum zone between 500 and
1500 m water depths. The faunal and isotope trends at Site 214 Sl–Pa Stilostomella lepidula − 0.76712 15.92655633
Pleurostomella alternans − 0.66316 2.282477444
show similar conditions (Fig. 7). At Site 214, the mixed-layer
Uvigerina proboscidea − 0.57463 19.06312037
was thick (abundant Gs. sacculifer) and thermocline was deep Ehrenbergina carinata − 0.45695 1.806140548
(rare Gr. menardii complex), indicating low surface productiv- Cassidulina carinata − 0.40159 8.077220475
ity in the eastern Indian Ocean. Gavelinopsis lobatulus − 0.37221 3.76770311
After 1.0 Myr, at Site 219 the benthic biofacies are domi-
Nu–Gs Factor 3-ve Factor scores Cluster III
nated by Ou–Gl and Ee–Uh while U. proboscidea decreased
(average %)
in abundance, indicating low to intermediate organic carbon Nuttallides umbonifer − 0.68766 2.40671494
flux to the seafloor and high seasonality — conditions typical Globocassidulina subglobosa − 0.65278 4.369466469
of winter monsoon season. The percent G. bulloides increased Gyroidinoides cibaoensis − 0.4199 2.418400106
after 1.0 Myr, yet its abundance is not significant enough Oridorsalis umbonatus − 0.41545 4.100768771
Uvigerina proboscidea − 0.28727 30.41905914
to attribute it to summer monsoon but to winter monsoon
strengthening. We suggest strengthening of the winter mon- Ou–Gl Factor 8-ve Factor scores Cluster II
soon since 1.0 Myr, driven by the onset of ice age cyclicity. It (average %)
has been observed by Overpeck et al. (1996) that G. bulloides Oridorsalis umbonatus − 0.51602 4.257120018
population remains at b 4–5% during winter monsoon sea- Gavelinopsis lobatulus − 0.40186 1.792069343
Bulimina aculeata − 0.38974 10.77563706
son and increases to N30% during summer monsoon season.
Fontugne and Duplessy (1986) suggested that during cold Ee–Uh Factor 4-ve Factor scores Cluster I
intervals, the winter monsoon strengthens and summer mon- (average %)
soon weakens. In the eastern Indian Ocean, the productivity Epistominella exigua − 0.48336 12.30504722
has increased during the past 1.0 Myr as suggested by step- Uvigerina hispido-costata − 0.38523 7.116840533
Bolivina pusilla − 0.35088 1.355918378
wise decrease in δ13 C values and % Gs. sacculifer and in-
242 A.K. Gupta et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 261 (2008) 234–245

Fig. 6. Distribution of benthic foraminiferal biofacies in time along with accumulative percentages of their key species at DSDP Site 219. Also plotted are percent
planktic foraminifer Globigerina bulloides and benthic foraminifer Uvigerina proboscidea (top panel). The shades mark the intervals during which a particular
biofacies dominates.

Indonesian Throughflow (ITF) which has its maximum flow Pacific and the Indonesian Sea shoal, thus restricting the
in summer and minimum in winter (Gordon, 1986; Gordon movement of water between the two regions.
and Fine, 1996). We do not link the isotopic and faunal At Sites 214 and 219, oxygen does not appear to have been
changes at Site 214 during the last 1.0 Myr to the ITF flow a limiting factor since these sites lie in a well-oxygenated
because during glacial times the sill depths between the (N3.8 ml/L) environment in the present-day abyssal setting and
A.K. Gupta et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 261 (2008) 234–245 243

Fig. 7. Percent Uvigerina proboscidea and Globigerina bulloides from Site 219 (panel e), plotted with percent U. proboscidea (panel d), mixed-layer species
Globigerinoides sacculifer and thermocline species Globorotalia menardii complex (panel c), Gs. sacculifer δ13C (‰) (panel b) and δ18O (‰) (panel a) from Site
214. Data in panels a–d are from Gupta and Dhingra (2004). A major switch in the proxy record is visible during ~1.2–0.9 Myr ago.

have remained so during the past 2.4 Myr. Organic carbon flux (2003) suggesting that changes in the tropical monsoons were
linked to changes in the monsoon strength and trade wind closely related to changes in the Northern Hemisphere ice vol-
intensity appears to be the primary factor controlling changes in ume both at orbital and suborbital time scales. The high vari-
benthic foraminiferal assemblages in the southeastern Arabian ability in the tropical deep-sea environments occurred at a time
Sea and eastern Indian Ocean. Our study thus supports the recent when the Earth's climate was experiencing large scale turnovers
observations by Gupta et al. (2003) and Gupta and Thomas due to the increased intensity of glacial–interglacial cycles.
244 A.K. Gupta et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 261 (2008) 234–245

Acknowledgements Gooday, A.J., Turley, C.M., 1990. Response by benthic organisms to inputs of
organic material to the Ocean floor: a review. Philosophical Transactions of
the Royal Society of London A 331, 119–138.
Deep Sea Drilling Project is thanked for providing samples Gordon, A.L., Fine, R.A., 1996. Pathways of water between the Pacific and
for the present study. This study was funded by DST, New Delhi Indian oceans in the Indonesian seas. Nature 379, 146–149.
(No. SR/S4/ES-46/2003). Gordon, A.L., 1986. Interocean exchange of thermocline water. Journal of
Geophysical Research 91, 5037–5046.
Gupta, A.K., 1993. Biostratigraphic vs. paleoceanographic importance of Stilosto-
Appendix A. Supplementary data
mella lepidula (Schwager) in the Indian Ocean. Micropaleontology 39, 47–52.
Gupta, A.K., 1994. Taxonomy and bathymetric distribution of Holocene deep-
Supplementary data associated with this article can be found, sea benthic foraminifera in the Indian Ocean and the Red Sea.
in the online version, at doi:10.1016/j.palaeo.2008.01.012. Micropaleontology 40, 351–367.
Gupta, A.K., 1997. Paleoceanographic and paleoclimatic history of the Somali
Basin during the Pliocene-Pleistocene: multivariate analyses of benthic
References foraminifera from DSDP Site 241 (Leg 25). Journal of Foraminiferal
Research 27, 196–208.
Almogi-Labin, A., Schmiedl, G., Hemleben, Ch., Siman-Tov, R., Segl, M., Gupta, A.K., Satapathy, S.K., 2000. Latest Miocene-Pleistocene abyssal benthic
Meischner, D., 2000. The influence of the NE winter monsoon on foraminifera from west-central Indian Ocean DSDP Site 236: Paleoceano-
productivity changes in the Gulf of Aden, NW Arabian Sea, during the last graphic and paleoclimatic inferences. Jorunal of the Palaeontological
530 ka as recorded by foraminifera. Marine Micropaleontology 40, 295–319. Society of India 45, 33–48.
Altenbach, A., Sarnthein, M., 1989. Productivity record in benthic foraminifera. Gupta, A.K., Anderson, D.M., Overpeck, J.T., 2003. Abrupt changes in the
In: Berger, W.H., Smetacek, V., Wafer, G. (Eds.), Productivity of the Ocean: Asian southwest monsoon during the Holocene and their links to the North
Present and Past. Wiley, New York, NY, pp. 255–270. Atlantic Ocean. Nature 421, 354–356.
Berggren, W.A., Kent, D.V., Swisher, C.C., Aubry, M.P., 1995. A revised Gupta, A.K., Das, M., Bhaskar, K., 2006. South Equatorial Current (SEC)
Cenozoic geochronology and Chronostratigraphy. In: Berggren, W.A., driven changes at DSDP Site 237, Central Indian Ocean, during the Plio-
Kent, D.V., Aubry, M.-P., Hardenbol, J. (Eds.), Geochronology Timescales Pleistocene: evidence from benthic foraminifera and stable isotopes. Journal
and Global Stratigraphic Correlation: Framework for and Historical of Asian Earth Sciences 28, 276–290.
Geology. SEPM (Society for Sedimentary Geology) Special Publication, Gupta, A.K., Dhingra, H., 2004. Middle Pleistocene Transition (MPT) in the
54, pp. 129–212. Eastern Indian Ocean: a 2000 Kyr planktic faunal and isotope record from
Bernhard, J.M., 1992. Benthic foraminiferal distribution and biomass related to DSDP Site 214. Journal Geological Society of India 63, 29–38.
pore-water oxygen content: Central California Continental Slope and Rise. Gupta, A.K., Joseph, S., Thomas, E., 2001. Species diversity of Miocene deep-
Deep Sea Research 39, 585–605. sea benthic foraminifera and watermass stratification in the northeastern
Boersma, A., 1990. Late Oligocene to late Pliocene benthic foraminifers from Indian Ocean. Micropaleontology 47, 111–124.
depth traverses in the central Indian Ocean. In: Duncan, R.A., et al. (Eds.), Gupta, A.K., Singh, R.K., Joseph, S., Thomas, E., 2004. Indian Ocean high-
Proceedings of the ODP, Scientific Results 115, pp. 315–380. productivity event (10–8 Ma): linked to global cooling or to the initiation of
Bremer, M.L., Lohmann, G.P., 1982. Evidence for primary control of the the Indian monsoons? Geology 32, 753–756.
distribution of certain Atlantic Ocean benthonic foraminifera by degree of Gupta, A.K., Srinivasan, M.S., 1990. Response of northern Indian Ocean deep-
carbonate saturation. Deep-Sea Research 29, 987–998. sea benthic foraminifera to global climates during Pliocene–Pleistocene.
Corliss, B.H., 1979. Recent deep-sea benthonic foraminiferal distributions in the Marine Micropaleontology 16, 77–91.
southeastern Indian Ocean: inferred bottom water routes and ecological Gupta, A.K., Srinivasan, M.S., 1992. Uvigerina proboscidea abundances and
implications. Marine Geology 31, 115–138. paleoceanography of the Northern Indian Ocean DSDP Site 214 during the
Corliss, B.H., 1983. Distribution of Holocene deep-sea benthonic foraminifera late Neogene. Marine Micropaleontology 19, 355–367.
in the southwest Indian Ocean. Deep-Sea Research 30, 95–117. Gupta, A.K., Thomas, E., 1999. Latest Miocene through Pleistocene
Corliss, B.H., Martinson, D.G., Keffer, T., 1986. Late Quaternary deep-ocean paleoceanographic evolution of the northwestern Indian Ocean (DSDP
circulation. Bulletin of the Geological Society of America 97, 1106–1121. Site 219): global and regional factors. Paleoceanography 14, 62–73.
den Dulk, M., Reichart, G.J., van Heyst, S., Zachariasse, W.J., Van der Zwaan, Gupta, A.K., Thomas, E., 2003. Initiation of Northern Hemisphere glaciation
G.J., 2000. Benthic foraminifera as proxies of organic matter flux and and strengthening of the northeast Indian monsoon: Ocean Drilling Program
bottom water oxygenation? A case history from the northern Arabian Sea. Site 758, eastern equatorial Indian Ocean. Geology 31, 47–50.
Palaeogeography, Palaeoclimatology, Palaeoecolology 161, 337–359. Hayward, B.W., 2001. Global deep-sea extinctions during the Pleistocene ice
Dickens, G.R., Owen, R.M., 1999. The latest Miocene–early Pliocene biogenic ages. Geology 29, 599–602.
bloom: a revised Indian Ocean perspective. Marine Geology 161, 75–91. Hermelin, J.O.R., Shimmield, G.B., 1990. The importance of the oxygen
Du, Y., Qu, T., Meyers, G., Masumoto, Y., Sasaki, H., 2005. Seasonal heat minimum zone and sediment geochemistry on the distribution of recent
budget in the mixed layer of the southeastern tropical Indian Ocean in a benthic foraminifera from the Northwest Indian Ocean. Marine Geology 91,
high-resolution ocean general circulation model. Journal of Geophysical 1–29.
Research 110. doi:10.1029/2004JC002845 C04012. Jannink, N.T., Zachariasse, W.J., van der Zwaan, G.J., 1998. Living (Rose
Fariduddin, M., Loubere, P., 1997. The surface ocean productivity response of Bengal stained) benthic foraminifera from the pakistan continentla margin
deeper water benthic foraminifera in the Atlantic Ocean. Marine Micro- (northern Arabian sea). Deep-Sea Research I 45, 1483–1513.
paleontology 32, 289–310. Jorissen, F.J., De Stigter, H.C., Widmark, J., 1995. A conceptual model explaining
Fontugne, M.R., Duplessy, J.C., 1986. Variations of the monsoon regime during benthic foraminiferal microhabitats. Marine Micropaleontology 26, 3–15.
the Upper Quaternary: evidence from carbon isotopic record of organic Jung, S.J.A., Ganssen, G.M., Davies, G.R., 2001. Multidecadal variations in the
matter in North Indian Ocean sediment cores. Palaeogeography, Palaeocli- early Holocene outflow of Red Sea Water into the Arabian Sea.
matology, Palaeoecology 56, 69–88. Paleoceanography 16, 658–668.
GEOSECS, 1983. Indian Ocean Expedition, Hydrographic data, 1977–1978, 5. Kaiho, K., 1999. Effect of organic flux and dissolved oxygen on the benthic
U.S. Govt. Printing Office, Washington, D.C., pp. 1–48. foraminiferal oxygen index (BFOI). Marine Micropaleontology 37, 67–76.
Gooday, A.J., 1988. A response by benthic foraminifera to the deposition of Kawagata, S., Hayward, B.W., Grenfell, H.R., Sabaa, A., 2005. Mid-Pleistocene
phytodetritus in the deep sea. Nature 332, 70–73. extinction of deep-sea foraminifera in the North Atlantic Gateway (ODP
Gooday, A.J., 1994. The biology of deep-sea foraminifera: A review of some sites 980 and 982). Palaeogeography, Palaeoclimatology, Palaeoecology
advances and their applications in paleoceanography. Palaios 9, 14–31. 221, 267–291.
A.K. Gupta et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 261 (2008) 234–245 245

Kawagata, S., Hayward, B.H., Gupta, A.K., 2006. Benthic foraminiferal Reichart, G.J., Lourens, L.J., Zachariasse, W.J., 1998. Temporal variability in
extinctions linked to late Pliocene–Pleistocene deep-sea circulation changes the northern Arabian Sea Oxygen Minimum Zone (OMZ) during the last
in the northern Indian Ocean (ODP Site 722 and 758). Marine Micro- 225,000 years. Paleoceanography 13, 607–621.
paleontology 58, 219–242. SAS Institute, Inc., 1988. SAS/STAT Users' Guide: Release 6.03 Edition, Carry,
Krey, J., 1973. Primary production in the Indian Ocean I. In: Zeitzschel, B. (Ed.), N.C., 1-1003.
The Biology of the Indian Ocean. Springer, pp. 115–130. Schmiedl, G., Mackensen, A., 1997. Late Quaternary paleoproductivity and
Kroon, D., Steens, T.N.F., Troelstra, S.R., 1991. Onset of monsoonal related deep water circulation in the eastern South Atlantic Ocean: evidence from
upwelling in the western Arabian Sea as revealed by planktonic fora- benthic foraminifera. Palaeogeography, Palaeoclimatology, Palaeoecology
minifers. In: Prell, W.L., Niitsuma, N., et al. (Eds.), Proceedings of the 130, 43–80.
Ocean Drilling Program, Scientific Results, 117. Ocean Drilling Program, Schmiedl, G., Leuschner, D.C., 2005. Oxygenation changes in the deep western
College Station, TX, pp. 257–264. Arabian Sea during the last 190,000 years: productivity versus deep-water
Linke, P., Lutze, G.F., 1993. Microhabitat preferences of benthic foraminifera – circulation. Paleoceanography 20. doi:10.1029/2004PA001044 PA2008.
a static conceptor a dynamic adaptation to optimize food acquisition? Schmiedl, G., Mackensen, A., 2006. Multi-species stable isotopes of benthic
Marine Micropaleontology 20, 215–234. foraminifers reveal past changes of organic matter decomposition and deep-
Lisiecki, L.E., Raymo, M.E., 2005. A Pliocene–Pleistocene stack of 57 globally water oxygenation in the Arabian Sea. Paleoceanography 21. doi:10.1029/
distributed benthic δ18O records. Paleoceanography 20. doi:10.1029/ 2006PA001284 PA4213.
2004PA001071 PA1003. Schmiedl, G., Mackensen, A., Müller, P.J., 1997. Recent benthic foraminifera
Loubere, P., Fariduddin, M., 1999. Quantitative estimates of global patterns of from the eastern South Atlantic Ocean: dependence on food supply and
surface ocean biological productivity and its seasonal variation on time water masses. Marine Micropaleontology 32, 249–288.
scales from centuries to millennia. Global Biogeochemical Cycles 13, Schmiedl, G., de Bove´e, F., Buscail, R., Charrie`re, B., Hemleben, C.h.,
115–133. Medernach, L., Picon, P., 2000. Trophic control of benthic foraminiferal
Lutze, G.F., Coulbourn, W.T., 1984. Recent benthic foraminifers from abundance and microhabitat in the bathyal Gulf of Lions, western
continental margin of Northwest Africa: community structure and distribu- Mediterranean Sea. Marine Micropaleontology 40, 167–188.
tion. Marine Micropaleontology 8, 361–401. Schnitker, D., 1980. Quaternary deep-sea benthic foraminifers and bottom water
Lutze, G.F., Thiel, H., 1989. Epibenthic foraminifera from elevated micro- masses. Annual Reviews of Earth and Planetary Sciences 8, 343–370.
habitats: Cibicides wuellerstorfi and Planulina ariminensis. Journal of Schott, F.A., McCreary Jr., J.P., 2001. The monsoon circulation of the Indian
Foraminiferal Research 19, 153–158. Ocean. Progress in Oceanography 51, 1–123.
Mackensen, A., Fütterer, D.K., Grobe, H., Schmiedl, G., 1993. Benthic Shankar, D., Vinayachandran, P.N., Unnikrishnan, A.S., 2002. The monsoon
foraminiferal assemblages from the eastern South Atlantic Polar Front currents in the north Indian Ocean. Progress in Oceanography 52, 63–120.
region between 35° and 57°S distribution, ecology and fossilization Sirocko, F., Leuschner, D., Staubwasser, M., Maley, J., Heusser, L., 1999. High-
potential. Marine Micropaleontology 22, 33–69. frequency oscillations of the last 70,000 years in the tropical/subtropical and
Mackensen, A., Schmiedl, G., Harloff, J., Giese, M., 1995. Deep-sea fora- polar climates. Mechanisms of Global Climate Change at Millenial Time
minifera in the south Atlantic Ocean: ecology and assemblage generation. Scales. American Geophysical Union Monograph, 112, pp. 113–126.
Micropaleontology 41, 342–358. Smart, C.W., King, S.C., Gooday, A.J., Murray, J.W., Thomas, E., 1994. A
Miao, Q., Thunell, R.C., 1993. Recent deep-sea benthic foraminiferal benthic foraminiferal proxy of pulsed organic matter paleofluxes. Marine
distributions in the South China and Sulu Seas. Marine Micropaleontology Micropaleontology 23, 89–99.
22, 1–32. Smart, C.W., Thomas, E., Ramsay, A.T.S., 2007. Middle–late Miocene ben-
Mix, A.C., Pisas, N.G., Rugh, W., Wilson, J., Morey, A., Hagelberg, T.K., 1995. thic foraminifera in a western equatorial Indian Ocean depth transect:
Benthic foraminifer stable isotope record from Site 849 (0–5 Ma): local and paleoceanographic implications. Palaeogeography, Palaeoclimatology,
global climate changes. In: Pisas, N.G., Mayer, L.A., Janecek, T.R., Palmer- Palaeoecology 247, 402–420.
Julson, A., van Andel, T.H. (Eds.), Proc. ODP, Sci. Results, 138, pp. 371–412. Streeter, S., 1973. Bottom water and benthonic foraminifera in the North
Murgese, D.S., De Deckker, P., 2005. The distribution of deep-sea benthic Atlantic-glacial-interglacial contrasts. Quaternary Research 3, 131–141.
foraminifera in core tops from the eastern Indian Ocean. Marine Tchernia, P., 1980. Descriptive Regional Oceanography. Pergamon, New York,
Micropaleontology 56, 25–49. p. 253.
Nomura, R., 1991. Oligocene to Pleistocene benthic foraminifer assemblages at Thomas, E., Gooday, A.J., 1996. Cenozoic deep-sea benthic foraminifers:
Site 754 and 756, Eastern Indian Ocean. Proceedings of ODP, Scientific tracers for changes in oceanic productivity? Geology 24, 355–358.
Research 121, 31–75. Thomas, E., Booth, L., Maslin, M., Shackleton, N.J., 1995. Northeastern
Ohkushi, K., Thomas, E., Kawahata, H., 2000. Abyssal benthic foraminifera Atlantic benthic foraminifera during the last 45,000 years: productivity
from the northwestern Pacific (Shatsky Rise) during the last 298 kyr. Marine changes as seen from the bottom up. Paleoceanography 10, 545–562.
Micropalentology 38, 119–147. Weinholz, P., Lutze, G.F., 1989. The Stilostomella extinction. In: Ruddiman, W.,
Overpeck, J., Anderson, D., Trumbore, S., Prell, W., 1996. The southwest Indian Sarnthein, M., et al. (Eds.), Proceedings of the Ocean Drilling Program,
Monsoon over the last 18000 years. Climate Dynamics 12, 213–225. Scientific Results, 108, pp. 113–117.
Qiang, X.K., Li, Z.X., Powell, C.A., Zheng, H.B., 2001. Magnetostratigraphic Wollenburg, J., Kuhnt, W., 2000. The response of benthic foraminifers to carbon
record of the late Miocene onset of the East Asian monsoon, and Pliocene flux and primary production in the Arctic Ocean. Marine Micropaleontology
uplift of northern Tibet. Earth and Planetary Science Letters 187, 83–93. 40, 189–231.
Rathburn, R., Corliss, B.H., 1994. The ecology of living (stained) deep-sea Xiao, J., An, Z., 1999. Three large shifts in East Asian monsoon circulation
benthic foraminifera from the Sulu Sea. Paleoceanography 9, 87–150. indicated by loess-paleosol sequences in China and late Cenozoic deposits in
Ravelo, A.C., Andreasen, D.H., Lyle, M., Lyle, A.O., Wara, M.W., 2004. Japan. Palaeogeography, Palaeoclimatology, Palaeoecology 154, 179–189.
Regional climate shifts caused by gradual global cooling in the Pliocene Xu, J., Wang, P., Huang, B., Li, Q., Jian, Z., 2005. Response of planktonic
epoch. Nature 429, 263–267. foraminifera to glacial cycles: Mid-Pleistocene change in the southern South
Raymo, M.E., Oppo, D.W., Curry, W., 1997. The mid-Pleistocene climate China Sea. Marine Micropaleontology 54, 89–105.
transition: a deep sea carbon isotopic perspective. Paleoceanography 12, You, Y., 2000. Implications of the deep circulation and ventilation of the Indian
546–559. Ocean on the renewal mechanism of North Atlantic Deep Water. Journal of
Raymo, M.E., 1997. The timing of major climate transitions. Paleoceanography Geophysical Research 105, 23,895–23,926.
12, 577–585.

You might also like