13.11.2014 Views

Live (Rose Bengal stained) and dead benthic foraminifera from the ...

Live (Rose Bengal stained) and dead benthic foraminifera from the ...

Live (Rose Bengal stained) and dead benthic foraminifera from the ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Marine Micropaleontology 62 (2007) 45–73<br />

www.elsevier.com/locate/marmicro<br />

<strong>Live</strong> (<strong>Rose</strong> <strong>Bengal</strong> <strong>stained</strong>) <strong>and</strong> <strong>dead</strong> <strong>benthic</strong> <strong>foraminifera</strong><br />

<strong>from</strong> <strong>the</strong> oxygen minimum zone of <strong>the</strong> Pakistan<br />

continental margin (Arabian Sea)<br />

Stefanie Schumacher a, ⁎ , Frans J. Jorissen a , Delphine Dissard a,1 ,<br />

Kate E. Larkin b , Andrew J. Gooday b<br />

a Laboratory of Recent <strong>and</strong> Fossil Bio-Indicators (BIAF), Angers University, 2 Bd Lavoisier, 49045 Angers Cedex 01,<br />

France, <strong>and</strong> Laboratory of Marine Bio-Indicators (LEBIM), Ile d'Yeu, Ker Chalon, France<br />

b National Oceanography Centre, Southampton, DEEPSEAS Benthic Biology Group, University of Southampton<br />

Waterfront Campus, European Way, Southampton SO14 3ZH, United Kingdom<br />

Received 29 July 2005; received in revised form 10 July 2006; accepted 12 July 2006<br />

Abstract<br />

<strong>Live</strong> (<strong>Rose</strong> <strong>Bengal</strong> <strong>stained</strong>) <strong>and</strong> <strong>dead</strong> <strong>benthic</strong> <strong>foraminifera</strong>l communities (hard-shelled species only) <strong>from</strong> <strong>the</strong> Pakistan<br />

continental margin oxygen minimum zone (OMZ) have been studied in order to determine <strong>the</strong> relation between faunal composition<br />

<strong>and</strong> <strong>the</strong> oxygenation of bottom waters. Samples were taken <strong>from</strong> 136 m to 1870 m water depth during <strong>the</strong> intermonsoon season of<br />

2003 (March–April). <strong>Live</strong> <strong>foraminifera</strong>l densities show a clear maximum in <strong>the</strong> first half centimetre of <strong>the</strong> sediment only few<br />

specimens are found down to 4 cm depth. The faunas exhibit a clear zonation across <strong>the</strong> Pakistan margin OMZ. Down to 500 m<br />

water depth, Uvigerina ex gr. U. semiornata <strong>and</strong> Bolivina aff. B. dilatata dominate <strong>the</strong> assemblages. These taxa are largely<br />

restricted to <strong>the</strong> upper cm of <strong>the</strong> sediment. They are adapted to <strong>the</strong> very low bottom-water oxygen values (≈ 0.1 ml/l in <strong>the</strong> OMZ<br />

core) <strong>and</strong> <strong>the</strong> extremely high input of organic carbon on <strong>the</strong> upper continental slope. The lower part of <strong>the</strong> OMZ is characterised by<br />

cosmopolitan faunas, containing also some taxa that in o<strong>the</strong>r areas have been described in deep infaunal microhabitats. The contrast<br />

between faunas typical for <strong>the</strong> upper part of <strong>the</strong> OMZ, <strong>and</strong> cosmopolitan faunas in <strong>the</strong> lower part of <strong>the</strong> OMZ, may be explained by<br />

a difference in <strong>the</strong> stability of dysoxic conditions over geological time periods. The core of <strong>the</strong> OMZ has been characterised by<br />

prolonged periods of stable, strongly dysoxic conditions. The lower part of <strong>the</strong> OMZ, on <strong>the</strong> contrary, has been much more variable<br />

over time-scales of 1000s <strong>and</strong> 10,000 years because of changes in surface productivity <strong>and</strong> a fluctuating intensity of NADW<br />

circulation. We suggest that, as a consequence, well-adapted, shallow infaunal taxa occupy <strong>the</strong> upper part of <strong>the</strong> OMZ, whereas in<br />

<strong>the</strong> lower part of <strong>the</strong> OMZ, cosmopolitan deep infaunal taxa have repeatedly colonised <strong>the</strong>se more intermittent low oxygen<br />

environments.<br />

© 2006 Elsevier B.V. All rights reserved.<br />

Keywords: live (<strong>Rose</strong> <strong>Bengal</strong> <strong>stained</strong>); <strong>dead</strong> <strong>benthic</strong> <strong>foraminifera</strong>; oxygen minimum zone; Arabian Sea<br />

⁎ Corresponding author. Present address: Alfred Wegener Institute for Polar <strong>and</strong> Marine Research, Alter Hafen 26, 27568 Bremerhaven, Germany.<br />

Fax: +49 471 4831 1923.<br />

E-mail addresses: sschumacher@awi-bremerhaven.de (S. Schumacher), frans.jorissen@univ-angers.fr (F.J. Jorissen),<br />

ddissard@awi-bremerhaven.de (D. Dissard), kel1@noc.soton.ac.uk (K.E. Larkin), ang@noc.soton.ac.uk (A.J. Gooday).<br />

1 Now at Alfred Wegener Institute Wegener Institute for Polar <strong>and</strong> Marine Research, Am H<strong>and</strong>elshafen 12, 27570 Bremerhaven, Germany.<br />

0377-8398/$ - see front matter © 2006 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.marmicro.2006.07.004


46 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

47<br />

1. Introduction<br />

The surface circulation of <strong>the</strong> Arabian Sea is driven<br />

by southwest <strong>and</strong> nor<strong>the</strong>ast monsoonal winds, which<br />

cause a strong seasonality (Wyrtki, 1973; Burkill et al.,<br />

1993). During <strong>the</strong> SW monsoon (May to September), an<br />

anticyclonic surface circulation (Fig. 1, black arrows)<br />

causes an intense coastal upwelling off Somalia <strong>and</strong><br />

Oman, <strong>and</strong> off <strong>the</strong> southwestern part of India. During <strong>the</strong><br />

more gentle NE monsoon (November through March) a<br />

cyclonic surface circulation prevails (Fig. 1, gray arrows),<br />

<strong>and</strong> most of <strong>the</strong> upwelling phenomena disappear.<br />

Only off Pakistan, local <strong>and</strong> sporadical upwelling events<br />

are described during <strong>the</strong> NE monsoon (Wyrtki, 1973).<br />

The upwelling of nutrient-rich surface water leads to a<br />

high biological production during <strong>the</strong> SW monsoon<br />

(Ry<strong>the</strong>r <strong>and</strong> Menzel, 1965; Qasim, 1977). A second<br />

maximum in biological productivity occurs during <strong>the</strong><br />

NE monsoon (e.g. Caron <strong>and</strong> Dennett, 1999; Rixen<br />

et al., 2000). This second maximum is caused by convective<br />

mixing due to winter cooling <strong>and</strong> <strong>the</strong> injection of<br />

nutrients into <strong>the</strong> euphotic zone (Banse <strong>and</strong> McClain,<br />

1986; Madhupratap et al., 1996). In contrast to <strong>the</strong><br />

western Arabian Sea, where <strong>the</strong> organic matter fluxes<br />

are much higher during <strong>the</strong> SW monsoon than during <strong>the</strong><br />

NE monsoon, <strong>the</strong> fluxes in <strong>the</strong> central <strong>and</strong> eastern<br />

Arabian Sea are nearly <strong>the</strong> same during both monsoon<br />

seasons (Rixen et al., 2000). According to satellite observations<br />

by Antoine et al. (1996), annual primary<br />

productivity in <strong>the</strong> NE Arabian Sea is very high, <strong>and</strong><br />

amounts to 400 gC m − 2 a − 1 .<br />

In <strong>the</strong> Arabian Sea, <strong>the</strong> combination of high primary<br />

productivity <strong>and</strong> a pronounced <strong>the</strong>rmohaline stratification<br />

leads to a stable <strong>and</strong> exp<strong>and</strong>ed mid-water (150–<br />

1400 m) oxygen minimum zone (OMZ) (You <strong>and</strong><br />

Tomczak, 1993; Olson et al., 1993; Wyrtki, 1973; Helly<br />

<strong>and</strong> Levin, 2004), defined as <strong>the</strong> zone with oxygen<br />

concentrations below 0.5 ml/l (Levin, 2003a). The OMZ<br />

results <strong>from</strong> different factors including oxygen consumption<br />

by organic matter decay, supply of oxygendepleted<br />

intermediate water-masses <strong>from</strong> <strong>the</strong> south <strong>and</strong><br />

west (Swallow, 1984; Olson et al., 1993) <strong>and</strong> <strong>the</strong> semienclosed<br />

nature of <strong>the</strong> nor<strong>the</strong>rn Arabian Sea (Wyrtki,<br />

1973; Shetye et al., 1994).<br />

Benthic <strong>foraminifera</strong> provide a valuable tool for<br />

reconstructing paleoceanographic parameters (Gooday,<br />

Table 1<br />

Locations, water depth <strong>and</strong> approximated bottom water oxygen values<br />

<strong>from</strong> CTD profiles for <strong>the</strong> stations studied here<br />

Station Water depth (m) Latitude (N) Longitude (E) Oxygen (ml/l)<br />

55901#11 136 23°16.61' 66°42.30' 1.4<br />

55808#3 150 23°16.57' 66°39.41' 0.5<br />

55803#5 306 23°12.53' 66°34.07' 0.1<br />

55818#4 512 23°08.24' 66°30.12' 0.105<br />

55916#1 598 23°01.69' 66°42.07' 0.11<br />

55922#2 738 22°57.57' 66°41.63' 0.12<br />

55921#1 844 22°57.61' 66°37.66' 0.13<br />

55918#7 944 22°53.58' 66°36.65' 0.175<br />

55907#1 1000 22°54.73' 66°34.93' 0.18<br />

55802#4 1201 22°59.99' 66°24.44' 0.4<br />

55830#3 1870 22°51.32' 66°00.01' 1.7<br />

2003), including paleo-oxygenation (Kaiho, 1994;<br />

Bernhard et al., 1997; Jorissen, 1999; Den Dulk et al.,<br />

2000; Gooday, 2003; Schmiedl et al., 2003) as well as<br />

paleoproductivity conditions (Herguera <strong>and</strong> Berger,<br />

1991; Loubere, 1999; Loubere et al., 2003). The input<br />

of organic matter <strong>and</strong> <strong>the</strong> oxygenation of bottom- <strong>and</strong><br />

sediment pore-water regulates faunal density, composition<br />

<strong>and</strong> vertical distribution in <strong>the</strong> sediment (Jorissen<br />

et al., 1995, Fontanier et al., 2002). High densities of<br />

<strong>benthic</strong> <strong>foraminifera</strong> in oxygen-deficient environments<br />

have been described by several authors (Phleger <strong>and</strong><br />

Soutar, 1973; Bernhard, 1993; Sen Gupta <strong>and</strong> Machain-<br />

Castillo, 1993; Alve, 1994, 1995; Bernhard et al., 1997;<br />

Bernhard <strong>and</strong> Sen Gupta, 1999; Gooday et al., 2000).<br />

More particularly, high abundances of deep infaunal<br />

taxa such as Bolivina with a high length/width ratio,<br />

Chilostomella <strong>and</strong> Globobulimina are typical for most<br />

of <strong>the</strong>se low oxygen settings (Mackensen <strong>and</strong> Douglas,<br />

1989; Alve, 1994, 1995; Bernhard et al., 1997; Jorissen<br />

et al., 1998; Fontanier et al., 2002). However, shallow<br />

infaunal species also may be common (Hermelin <strong>and</strong><br />

Shimmield, 1990; Jannink et al., 1998; Maas, 2000). In<br />

order to better constrain <strong>the</strong> correlation between faunal<br />

density <strong>and</strong> composition, <strong>and</strong> bottom water oxygenation<br />

in <strong>the</strong>se dysoxic environments, more systematical investigations<br />

are necessary.<br />

The composition of <strong>the</strong> <strong>benthic</strong> <strong>foraminifera</strong>l faunas<br />

of <strong>the</strong> Arabian Sea <strong>and</strong> Indian Ocean OMZ is still poorly<br />

known. Only few publications describe living (<strong>Rose</strong><br />

<strong>Bengal</strong> <strong>stained</strong>) <strong>foraminifera</strong>l faunas <strong>from</strong> <strong>the</strong> OMZ.<br />

Gooday et al. (2000) <strong>and</strong> Hermelin <strong>and</strong> Shimmield<br />

Fig. 1. Surface hydrography in <strong>the</strong> Arabian Sea with <strong>the</strong> investigated area (Square) <strong>and</strong> previous studies on Recent <strong>benthic</strong> <strong>foraminifera</strong> (stars).<br />

During <strong>the</strong> SW monsoon (May through September), a strong anticyclonic surface circulation (black arrows), following <strong>the</strong> Findlater Jet (FJ) (after<br />

Rixen et al., 2000), dominates <strong>the</strong> surface circulation. The NE monsoon (November through March) causes a cyclonic circulation (light gray arrows),<br />

which is weaker than <strong>the</strong> SW monsoonal circulation. Site locations, investigated during April <strong>and</strong> May 2003, are given in <strong>the</strong> lower map. Bathymetry<br />

of <strong>the</strong> sampling area after Bett (2004a).


48 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

(1990) described <strong>the</strong> faunal distribution at stations <strong>from</strong><br />

<strong>the</strong> Northwest Arabian Sea <strong>from</strong> water depths below<br />

400 m. In <strong>the</strong> Nor<strong>the</strong>ast Arabian Sea, Jannink et al.<br />

(1998) described two transects <strong>from</strong> 500 to 3000 m<br />

water depth. Until now, only one site in <strong>the</strong> upper part of<br />

<strong>the</strong> OMZ (233 m water depth) has been described in <strong>the</strong><br />

Nor<strong>the</strong>ast Arabian Sea (Maas, 2000). Kurbjeweit et al.<br />

(2000) <strong>and</strong> Heinz <strong>and</strong> Hemleben (2003) provide an<br />

account of <strong>foraminifera</strong>l assemblages at greater water<br />

depths (N1900 m) in more central regions of <strong>the</strong> Arabian<br />

Sea.<br />

In this study we describe <strong>the</strong> hard-shelled <strong>foraminifera</strong>l<br />

assemblages sampled during <strong>the</strong> intermonsoon<br />

period (spring 2003) along a composite sampling profile<br />

(based on two transects) <strong>from</strong> 136 to 1870 m water<br />

depth. Our main aims are to improve knowledge of: (1)<br />

<strong>the</strong> faunal composition of live (<strong>Rose</strong> <strong>Bengal</strong> <strong>stained</strong>)<br />

<strong>and</strong> <strong>dead</strong> <strong>foraminifera</strong>l faunas, <strong>the</strong>ir bathymetrical<br />

distribution <strong>and</strong> microhabitats, (2) <strong>the</strong> relation between<br />

low oxygen environments <strong>and</strong> <strong>foraminifera</strong>l distribution<br />

<strong>and</strong> diversity indices <strong>and</strong> (3) <strong>the</strong> long-term adaptations<br />

that allow specific <strong>foraminifera</strong>l communities to survive<br />

in low oxygen environments. Preliminary observations<br />

on <strong>benthic</strong> <strong>foraminifera</strong> at <strong>the</strong> same sampling sites,<br />

based on ship-board sorting <strong>and</strong> including both hard<strong>and</strong><br />

soft-shelled taxa, were reported by Larkin et al.<br />

(2003).<br />

2. Material <strong>and</strong> methods<br />

During R.R.S. Charles Darwin Cruises 145 <strong>and</strong> 146<br />

(12 March to 28 May 2003), 11 multicores were taken on<br />

<strong>the</strong> continental margin off Karachi, Pakistan (Fig. 1,<br />

Table 1). Two transects were sampled, constituting a<br />

composite bathymetric profile <strong>from</strong> 136 m (above <strong>the</strong><br />

OMZ in spring 2003) down to 1870 m water depth<br />

(Fig. 2). Cores (surface area 25.5 cm 2 ) were processed as<br />

follows: for stations situated above, <strong>and</strong> in <strong>the</strong> upper part<br />

Fig. 2. Profiles over <strong>the</strong> Pakistan continental margin with locations of sampling sites <strong>and</strong> oxygen content in <strong>the</strong> water column (Bett, 2004a).


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

49<br />

of <strong>the</strong> OMZ, sediment slices were taken for <strong>the</strong> 0–0.5 <strong>and</strong><br />

0.5–1 cm intervals, <strong>and</strong> <strong>the</strong>n in 1 cm intervals down to<br />

10 cm. For <strong>the</strong> lower part of <strong>the</strong> OMZ, <strong>the</strong> second<br />

centimetre was also sliced in half-centimetre intervals.<br />

Each sample was stored in 10% borax-buffered formalin<br />

for fur<strong>the</strong>r processing. Onshore, <strong>the</strong> samples were wet<br />

sieved over 63 μm, 150 μm <strong>and</strong> 300 μm sieves <strong>and</strong> <strong>the</strong><br />

residues were <strong>stained</strong> for one week in ethanol with <strong>Rose</strong><br />

<strong>Bengal</strong> (Walton, 1952). After staining, <strong>the</strong> residue was<br />

washed again. The <strong>stained</strong> faunas were picked wet in<br />

three granulometric fractions (63–150 μm, 150–300 μm<br />

<strong>and</strong> N300 μm), down to 10 cm depth. To gain more<br />

insight into <strong>the</strong> population dynamics (compare Jorissen<br />

<strong>and</strong> Wittling, 1999), we investigated <strong>the</strong> <strong>dead</strong> (un<strong>stained</strong>)<br />

<strong>foraminifera</strong> in <strong>the</strong> 2–3 cm level for <strong>the</strong> fractions<br />

150–300 μm <strong>and</strong> N300 μm. At this sediment depth we<br />

expected a strong decrease of taxa with rapidly<br />

decomposing tests (especially some arenaceous taxa),<br />

<strong>and</strong> also a more correct representation of deeper living<br />

infaunal species, that may be under-represented in<br />

superficial <strong>dead</strong> faunas (Loubere, 1989). The smaller<br />

fraction of <strong>the</strong> <strong>dead</strong> fauna has not been studied, since it<br />

may be seriously affected by post-mortem transport<br />

(Murray, 1991). When possible, a minimum of 200–300<br />

specimens was collected; if necessary samples were<br />

splitted with an Otto microsplitter. To enable a<br />

comparison of <strong>the</strong> sites, <strong>the</strong> 1–2 cm interval is treated<br />

as a single unit, even when it was sampled in halfcentimetre<br />

slices. The fractions N300 μm <strong>and</strong> 150–<br />

300 μm show nearly <strong>the</strong> same faunal distribution <strong>and</strong><br />

<strong>the</strong>refore <strong>the</strong> results are presented here for both fractions<br />

combined (i.e. <strong>the</strong> N150 μm fraction). Data are available<br />

on doi:10.1594/PANGAEA.475995. The recognition of<br />

<strong>stained</strong> <strong>foraminifera</strong> may be somewhat subjective; for<br />

problems <strong>and</strong> discussion about <strong>the</strong> use of <strong>Rose</strong> <strong>Bengal</strong><br />

see Bernhard (1988). We applied our staining criteria<br />

very strictly <strong>and</strong> only specimens with all chambers except<br />

<strong>the</strong> last one <strong>stained</strong> bright pink were counted as<br />

<strong>stained</strong>. Species determinations are based on <strong>the</strong> taxonomic<br />

references given in <strong>the</strong> Appendix. For our quantitative<br />

analysis we used all individuals of <strong>the</strong> total fauna<br />

with <strong>the</strong> exception of fragments of tubular <strong>foraminifera</strong><br />

(e.g. Hyperammina, Rhizammina <strong>and</strong> Rhabdammina).<br />

In order to obtain a “potential fossil” fauna (Mackensen<br />

et al., 1990) we considered separately <strong>the</strong> “calcareous<br />

fauna”, containing only calcareous-perforate <strong>and</strong> porcellaneous<br />

taxa, for living as well as <strong>dead</strong> faunas. The<br />

numbers of fossilizable arenaceous taxa were negligible.<br />

Fig. 3. Faunal densities (counts/core 25.5 cm 2 ) of live (<strong>stained</strong>) <strong>foraminifera</strong> against water depth <strong>and</strong> oxygen content in <strong>the</strong> water column.


50 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

For <strong>the</strong> description of <strong>the</strong> vertical distribution, we use <strong>the</strong><br />

Average Living Depth (ALD x , Jorissen et al., 1995).<br />

ALD x ¼ X i¼0;xðn i D i Þ=N<br />

x<br />

n i<br />

D i<br />

N<br />

lower boundary of <strong>the</strong> deepest sample<br />

number of individuals in interval i<br />

midpoint of sample interval i<br />

total number of individuals for all levels<br />

For all stations, ALD 10 was calculated for <strong>the</strong> entire<br />

hard-shelled (calcareous +arenaceous) live fauna <strong>and</strong> for<br />

<strong>the</strong> calcareous component of <strong>the</strong> fauna. We also present<br />

<strong>the</strong> species numbers <strong>and</strong> <strong>the</strong> diversities, expressed by <strong>the</strong><br />

Shannon–Wiener Index (H(S ln )) (Shannon, 1948) <strong>and</strong><br />

by <strong>the</strong> Fisher Alpha Index (Fisher et al., 1943). Calculations<br />

were based on <strong>the</strong> entire hard-shelled fauna as<br />

well as on calcareous taxa alone.<br />

Oxygen profiles of bottom water were obtained with a<br />

CTD (Fig. 2) at all sites <strong>and</strong> oxygen concentrations in<br />

Fig. 4. Vertical distribution of live (<strong>stained</strong>) <strong>foraminifera</strong> in <strong>the</strong> sediment for each core.


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

51<br />

core–top water were determined by Winkler titration<br />

(Cowie, 2003; Bett, 2004a). Down to about 150 m water<br />

depth, bottom waters were well oxygenated during<br />

March/April 2003. Oxygen values dropped to around<br />

0.1 ml/l between 150 <strong>and</strong> 500 m. Between 500 <strong>and</strong><br />

1000 m <strong>the</strong>y increased gradually <strong>and</strong> reach values of about<br />

0.2 ml/l. This is followed by a strong increase of oxygen in<br />

<strong>the</strong> lowest part of <strong>the</strong> OMZ <strong>and</strong> oxygen values N0.5 ml/<br />

l below 1300 m water depth. CTD profiles obtained later<br />

in <strong>the</strong> year (September/October 2003) showed an upward<br />

shift of <strong>the</strong> upper OMZ boundary (0.5 ml/l oxygen level)<br />

<strong>from</strong> c. 180 m to c. 80 m (Bett, 2004b).<br />

Sediments are well bioturbated at 136 m water depth,<br />

in <strong>the</strong> lowest part of <strong>the</strong> OMZ (1000–1300 m) <strong>and</strong> below<br />

<strong>the</strong> OMZ (Levin, 2003b). The sediments within <strong>the</strong> core<br />

of <strong>the</strong> OMZ are laminated; <strong>the</strong> laminations are irregular at<br />

<strong>the</strong> 300-m site but are more regularly formed between 500<br />

<strong>and</strong> 700 m. The laminations are overprinted by vertical<br />

burrows at <strong>the</strong> 850-m <strong>and</strong> 950-m sites (Levin, 2003b).<br />

3. Results<br />

3.1. Foraminiferal densities<br />

Total abundances of live (<strong>Rose</strong> <strong>Bengal</strong> <strong>stained</strong>) <strong>foraminifera</strong><br />

are shown for each core in Fig. 3. Faunal densities<br />

in <strong>the</strong> N150 μm fraction clearly decreased with increasing<br />

water depth. At 136 m, <strong>the</strong>re were N620 individuals/core<br />

(25.5 cm 2 surface area), compared with only 13 per core at<br />

1000 m <strong>and</strong> 50 per core below <strong>the</strong> OMZ. The densities of<br />

calcareous taxa showed <strong>the</strong> same trend with a strong<br />

decrease below 600 m water depth. Densities in <strong>the</strong><br />

smaller size fraction showed ra<strong>the</strong>r constant values in <strong>the</strong><br />

upper 600 m, a sudden maximum at <strong>the</strong> 738-m site<br />

followed by a decrease at deeper sites. Total densities in<br />

<strong>the</strong> 63–150 μm fraction varied <strong>from</strong> 50 to about 2000<br />

individuals/core. Densities of calcareous taxa reached a<br />

maximum at 150 m water depth with a second maximum<br />

at 598 to 944 m. Densities in <strong>the</strong> two size fractions (N150<br />

<strong>and</strong> 63–150 μm) were very similar, except between 512<br />

<strong>and</strong> 844 m water depth, where <strong>the</strong> density in <strong>the</strong> smaller<br />

size fractions could be 10 times higher than in <strong>the</strong> larger<br />

size fraction. The overall (N63 μm) abundance of live<br />

<strong>foraminifera</strong> was highest <strong>from</strong> 138 to 306 m <strong>and</strong> 598 to<br />

844 m, with very low values below 944 m depth.<br />

3.2. Vertical distribution<br />

<strong>Live</strong> <strong>foraminifera</strong> were found at a maximum sediment<br />

depth of 2 cm in <strong>the</strong> OMZ <strong>and</strong> down to 3 cm <strong>and</strong> 4 cm<br />

depth above <strong>and</strong> below <strong>the</strong> OMZ, respectively. Most<br />

stations showed a clear maximum faunal density in <strong>the</strong><br />

first half centimetre with an abrupt decrease in deeper<br />

layers (Fig. 4). At most stations more than 75% of <strong>the</strong><br />

fauna occurred in <strong>the</strong> first half centimetre for both size<br />

fractions. This superficial life position was confirmed by<br />

Fig. 5. Average Living Depth for <strong>the</strong> entire hard-shelled fauna <strong>and</strong> calcareous fauna against water depth <strong>and</strong> oxygen content in <strong>the</strong> water column.


52 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

Fig. 6. Species number, Shannon–Wiener Index (H(S ln )) <strong>and</strong> Fisher Alpha Index for <strong>the</strong> living fauna N150 μm, 63–150 μm <strong>and</strong> N63 μm/core against<br />

water depth.


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

53<br />

Fig. 6 (continued).<br />

<strong>the</strong> fact that Average Living Depth of <strong>the</strong> entire hardshelled<br />

faunas (ALD 10 ) was mostly b0.4 cm for <strong>the</strong><br />

N150 μm fraction (Fig. 5). Values were slightly higher<br />

below 700 m <strong>and</strong> a maximum value of about 1.4 cm was<br />

found at <strong>the</strong> deepest (1850-m) site. For <strong>the</strong> smaller size<br />

fraction, <strong>the</strong> ALD 10 was comparable to those derived<br />

<strong>from</strong> <strong>the</strong> N150 μm fraction. Between 512 <strong>and</strong> 844 m<br />

water depth <strong>the</strong> ALD 10 was minimal.<br />

In <strong>the</strong> N150 μm fraction, <strong>the</strong> ALD 10 for calcareous<br />

taxa was similar to <strong>the</strong> ALD 10 for <strong>the</strong> entire hard-shelled<br />

fauna down to 598 m water depth, whereas at deeper<br />

sites, <strong>the</strong> ALD 10 of calcareous taxa was systematically<br />

higher than <strong>the</strong> ALD 10 of <strong>the</strong> entire hard-shelled fauna.<br />

This pattern showed that more non-fossilizing arenaceous<br />

taxa lived in very superficial niches, whereas<br />

some calcareous taxa may have lived slightly deeper in<br />

<strong>the</strong> sediment. For <strong>the</strong> smaller size fraction, <strong>the</strong> ALD 10 of<br />

calcareous taxa was always comparable to that of <strong>the</strong><br />

entire hard-shelled fauna (Fig. 5).<br />

3.3. Diversity<br />

3.3.1. Living faunas<br />

Between 8 <strong>and</strong> 17 species were observed in <strong>the</strong><br />

N150 μm size fraction, between 13 <strong>and</strong> 33 species in<br />

<strong>the</strong> 63–150 μm size fraction, <strong>and</strong> between 17 <strong>and</strong> 37<br />

species in <strong>the</strong> N63 μm fraction (Fig. 6). The species<br />

number in <strong>the</strong> larger fraction did not vary significantly<br />

with water depth, whereas <strong>the</strong> species number in <strong>the</strong><br />

small fraction was fairly well correlated with <strong>the</strong> faunal<br />

densities <strong>and</strong> showed maximum values at 138 to 150 m<br />

<strong>and</strong> at 737 m. The number of calcareous species in <strong>the</strong><br />

larger size fraction decreased with increasing water<br />

depth, but <strong>the</strong>re was no clear trend in <strong>the</strong> case of <strong>the</strong><br />

smaller size fraction. In <strong>the</strong> larger size fraction, <strong>the</strong><br />

Shannon–Wiener index (H(S ln )) showed a trend towards<br />

higher values at greater water depth. The Fisher<br />

Alpha index showed a similar tendency but it was<br />

weaker <strong>and</strong> started deeper. Both indices were positively<br />

correlated with bottom-water oxygen concentration in<br />

<strong>the</strong> 63–150 μm fraction. Shannon–Wiener <strong>and</strong> <strong>the</strong><br />

Fischer Alpha values for <strong>the</strong> calcareous part of <strong>the</strong> fauna<br />

(large size fraction) were highly variable <strong>and</strong> followed<br />

<strong>the</strong> diversity of <strong>the</strong> entire hard-shelled fauna but with<br />

lower values. The difference between entire hard-shelled<br />

<strong>and</strong> calcareous values increased with increasing water<br />

depth, due to <strong>the</strong> higher proportion of arenaceous<br />

<strong>foraminifera</strong> at deeper sites. In <strong>the</strong> smaller size fraction,<br />

diversity indices for <strong>the</strong> calcareous taxa followed <strong>the</strong><br />

same pattern with water depth as <strong>the</strong> entire hard-shelled


54 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

Fig. 7. Species number, Shannon–Wiener Index (H(S ln )) <strong>and</strong> Fisher Alpha Index for <strong>the</strong> <strong>dead</strong> fauna N150 μm in2–3 cm depth against water depth.<br />

fauna, but with slightly lower values. Diversity trends<br />

for <strong>the</strong> N63 μm fraction were similar to those for <strong>the</strong> 63–<br />

150 μm fraction.<br />

3.3.2. Dead fauna<br />

Between 13 to 49 species were encountered (Fig. 7)<br />

The species number showed a clear positive correlation<br />

with bottom-water oxygen values, with minimum numbers<br />

in <strong>the</strong> OMZ. The two diversity indices also showed a<br />

good correlation with oxygen. For <strong>the</strong> calcareous fauna,<br />

species numbers were lower, but showed <strong>the</strong> same trend<br />

as <strong>the</strong> entire hard-shelled fauna. Similar observations<br />

could be made for <strong>the</strong> two diversity indices.<br />

3.4. Faunal composition<br />

The live <strong>and</strong> <strong>dead</strong> faunas were mainly represented by<br />

calcareous-perforate <strong>and</strong> arenaceous <strong>foraminifera</strong>. Miliolids<br />

were very rare. The percentage of arenaceous<br />

foraminifers in <strong>the</strong> living fauna increased with increasing<br />

water depth. At <strong>the</strong> 136 <strong>and</strong> 306 m sites, <strong>the</strong> percentage<br />

of arenaceous <strong>foraminifera</strong> was about 20% in<br />

both size fractions. In deeper water, <strong>the</strong> proportion of<br />

arenaceous <strong>foraminifera</strong> in <strong>the</strong> N150 μm fraction rose to<br />

90% at 1200 m <strong>and</strong> 100% at 1850 m. In <strong>the</strong> 63–150 μm<br />

fraction, a maximum of 72% arenaceous taxa at 738 m<br />

water depth coincided with a strong abundance maximum<br />

(Fig. 3). In <strong>the</strong> o<strong>the</strong>r OMZ samples, arenaceous<br />

taxa accounted for 12 to 63% of <strong>the</strong> fauna. Most <strong>foraminifera</strong><br />

lived in <strong>the</strong> first half centimetre of <strong>the</strong> sediment.<br />

There was no clear microhabitat differentiation;<br />

<strong>the</strong>re were no species restricted to <strong>the</strong> uppermost level,<br />

nor infaunal taxa showing subsurface density maxima.<br />

The <strong>dead</strong> fauna was dominated by calcareous <strong>foraminifera</strong><br />

(60–100%), with <strong>the</strong> exception of a maximum<br />

(N60%) of arenaceous <strong>foraminifera</strong> at <strong>the</strong> 598- <strong>and</strong><br />

738-m sites. The latter were mainly represented by two<br />

species of Ammodiscus (see Appendix) <strong>and</strong> various<br />

species of Reophax <strong>and</strong> <strong>the</strong> Trochamminacea. The tests<br />

of all <strong>the</strong>se taxa normally are strongly affected by early<br />

diagenesis <strong>and</strong> are not preserved in fossil faunas (Mackensen<br />

et al., 1990, 1993). Because we want to use our<br />

description of faunal patterns in <strong>the</strong> OMZ as a basis for<br />

paleoceanographic methods, we will now chiefly focus<br />

on <strong>the</strong> calcareous part of <strong>the</strong> fauna, which was usually<br />

dominated by <strong>the</strong> group of bi- <strong>and</strong> triserial species. For<br />

<strong>the</strong> live fauna N150 μm, this group reached a maximum<br />

density between 136 <strong>and</strong> 598 m water depth (more than


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

55<br />

Fig. 8. Counts of bi- <strong>and</strong> triserial species <strong>and</strong> <strong>the</strong>ir relative abundance in <strong>the</strong> entire calcareous fauna for <strong>the</strong> living fauna N150 μm <strong>and</strong> 63–150 μm, <strong>and</strong><br />

for <strong>the</strong> <strong>dead</strong> fauna N150 μm against water depth.<br />

100 specimens/25.5 cm 2 core) (Fig. 8). In this depth<br />

interval, bi- <strong>and</strong> triserial taxa always accounted for more<br />

than 85% of <strong>the</strong> calcareous fauna. At deeper sites, <strong>the</strong>se<br />

became much less abundant, although relative abundances<br />

remained elevated (50–100%). In <strong>the</strong> smaller<br />

size fractions, this group also strongly dominated <strong>the</strong><br />

faunas (N75%) down to 844 m water depth. Among <strong>the</strong><br />

<strong>dead</strong> faunas, all sites within <strong>the</strong> OMZ were strongly<br />

dominated (48–100%) by biserial <strong>and</strong> triserial taxa.<br />

Planoconvex species were infrequent in <strong>the</strong><br />

N150 μm fraction <strong>and</strong> occurred in <strong>the</strong> live as well as<br />

in <strong>the</strong> <strong>dead</strong> fauna mainly above 306 m <strong>and</strong> below 800 m<br />

water depth (Fig. 9). Among live assemblages, absolute<br />

numbers were highest between 136 <strong>and</strong> 306 m water<br />

depth for <strong>the</strong> larger fraction <strong>and</strong> between 136 <strong>and</strong> 150 m<br />

<strong>and</strong> below 844 m for <strong>the</strong> smaller fraction. Planoconvex<br />

taxa were absent <strong>from</strong> <strong>the</strong> live assemblages N150 μm<br />

below 1000 m depth, although <strong>the</strong>y persisted in <strong>the</strong><br />

<strong>dead</strong> fauna.<br />

We could also recognise a third group containing<br />

species that were concentrated in <strong>the</strong> topmost sediment<br />

in our study area but have been described elsewhere as<br />

deep infaunal (e.g. Jorissen et al., 1998; Fontanier et al.,<br />

2002; Licari et al., 2003). This group includes Bulimina<br />

exilis, Fursenkoina spp. Globobulimina spp. Chilostomella<br />

spp. <strong>and</strong> Praeglobobulimina spp. All <strong>the</strong>se taxa,<br />

except Chilostomella, are also included in <strong>the</strong> group of<br />

bi- <strong>and</strong> triserial taxa. Significant numbers of <strong>the</strong>se “deep<br />

infaunal” taxa occurred mainly in <strong>the</strong> lower part of <strong>the</strong><br />

OMZ (Fig. 10) <strong>and</strong> especially in <strong>the</strong> <strong>dead</strong> fauna. For <strong>the</strong><br />

smaller size fraction, a clear maximum (63–72%) could<br />

be observed <strong>from</strong> 306 to 738 m water depth, coinciding<br />

with <strong>the</strong> area of lowest oxygen concentrations.<br />

3.5. Single species distribution<br />

A limited number of dominant species characterised<br />

<strong>the</strong> fauna. In <strong>the</strong> upper part of <strong>the</strong> OMZ <strong>the</strong>se were<br />

Uvigerina ex gr. U. semiornata, Bolivina aff. B. dilatata<br />

<strong>and</strong> Cassidulina laevigata in <strong>the</strong> N150 μm fraction <strong>and</strong><br />

Bolivina aff. B. dilatata, B. exilis, Bolivina dilatata <strong>and</strong><br />

Alliatina primitiva in <strong>the</strong> smaller fraction. Cancris auriculus<br />

also made a significant contribution to <strong>the</strong> <strong>dead</strong><br />

assemblages at <strong>the</strong> shallowest stations. In <strong>the</strong> lower part<br />

of <strong>the</strong> OMZ, Uvigerina peregrina replaced Uvigerina<br />

ex gr. U. semiornata <strong>and</strong> o<strong>the</strong>r widely distributed species<br />

such as Pullenia bulloides, P. quinqueloba, <strong>and</strong><br />

Bulimina aculeata also occurred. Among <strong>the</strong> arenaceous<br />

<strong>foraminifera</strong>, Ammodiscus spp. <strong>and</strong> several species of<br />

Reophax were common.


56 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

Fig. 9. Counts of planoconvex species <strong>and</strong> <strong>the</strong>ir relative abundance in <strong>the</strong> entire calcareous fauna for <strong>the</strong> living fauna N150 μm <strong>and</strong> 63–150 μm, <strong>and</strong><br />

for <strong>the</strong> <strong>dead</strong> fauna N150 μm against water depth.<br />

Uvigerina ex gr. U. semiornata (see Appendix) was<br />

common in <strong>the</strong> OMZ core <strong>and</strong> between 136 <strong>and</strong> 306 m<br />

water depth, where it represented up to 90% of <strong>the</strong><br />

calcareous assemblage N150 μm. With increasing water<br />

depth, <strong>the</strong> density of Uvigerina ex gr. U. semiornata<br />

decreased down to 512 m, where it was replaced by<br />

U. peregrina which was present down to 1000 m water<br />

depth <strong>and</strong> represented up to 89% of <strong>the</strong> live <strong>and</strong> <strong>dead</strong><br />

calcareous fauna. In <strong>the</strong> smaller living fraction, Uvigerina<br />

ex gr. U. semiornata was less numerous, whereas<br />

U. peregrina sometimes still represented more than 60%<br />

of calcareous <strong>foraminifera</strong> (Figs. 11 <strong>and</strong> 12). In <strong>the</strong> <strong>dead</strong><br />

fauna, Uvigerina ex gr. U. semiornata was again common<br />

down to 512 m water depth, where U. peregrina<br />

occurred for <strong>the</strong> first time. The latter species was common<br />

down to 1000 m water depth (Fig. 13). The absence<br />

of Uvigerina ex gr. U. semiornata <strong>and</strong> U. peregrina in<br />

<strong>the</strong> <strong>dead</strong> fauna of deeper stations was consistent with <strong>the</strong><br />

absence of <strong>the</strong>se species in <strong>the</strong> living fauna N150 μm.<br />

Bolivina aff. B. dilatata (see Appendix) was a dominant<br />

species in <strong>the</strong> larger size fraction at <strong>the</strong> shallower<br />

stations, while in <strong>the</strong> smaller fraction B. dilatata was<br />

also common. In <strong>the</strong> living fauna N150 μm an absolute<br />

density maximum of Bolivina aff. B. dilatata was present<br />

at 306 m water depth (229 individuals/core, 45%),<br />

while <strong>the</strong> highest percentage (88%) was found at <strong>the</strong><br />

598 m site. For <strong>the</strong> <strong>dead</strong> fraction, Bolivina aff. B. dilatata<br />

reached a maximum relative abundance at 306 m<br />

(58%), correlating with <strong>the</strong> absolute density maximum<br />

observed in <strong>the</strong> living fauna. In <strong>the</strong> smaller size fraction,<br />

Bolivina aff. B. dilatata occurred between 136 <strong>and</strong><br />

598 m water depth, while B. dilatata occurred in significant<br />

numbers only between 136 <strong>and</strong> 306 m. Where <strong>the</strong><br />

two species occurred toge<strong>the</strong>r, B. dilatata was more<br />

abundant than Bolivina aff. B. dilatata. The total dominance<br />

of <strong>the</strong>se taxa seen in <strong>the</strong> larger size fraction was<br />

not observed in <strong>the</strong> finer fractions, where <strong>the</strong>ir relative<br />

abundance was b60%.<br />

The above-mentioned Uvigerina <strong>and</strong> Bolivina species<br />

strongly dominated <strong>the</strong> living <strong>and</strong> <strong>dead</strong> faunas in<br />

<strong>the</strong> N150 μm assemblages in <strong>the</strong> upper part of <strong>the</strong> OMZ.<br />

Low numbers of C. laevigata, C. auriculus, <strong>and</strong> o<strong>the</strong>r<br />

planoconvex species occurred here as well (Figs. 8 <strong>and</strong><br />

11). In <strong>the</strong> smaller size fraction, B. exilis <strong>and</strong> A. primitiva<br />

also occurred in significant numbers in <strong>the</strong> live<br />

fauna. B. exilis occurred <strong>from</strong> 150 to 944 m water depth,<br />

with its maximum abundance between 306 <strong>and</strong> 738 m,<br />

where it made up N60% of <strong>the</strong> assemblage. This species


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

57<br />

Fig. 10. Counts of “deep-infaunal” species <strong>and</strong> <strong>the</strong>ir relative abundance in <strong>the</strong> entire calcareous fauna for <strong>the</strong> living fauna N150 μm <strong>and</strong> 63–150 μm,<br />

<strong>and</strong> for <strong>the</strong> <strong>dead</strong> fauna N150 μm against water depth.<br />

seemed to replace Bolivina aff. B. dilatata <strong>and</strong> B. dilatata,<br />

<strong>the</strong> dominant species in <strong>the</strong> upper region of <strong>the</strong><br />

OMZ (150–306 m). B. exilis was nearly absent in <strong>the</strong><br />

larger live <strong>and</strong> <strong>dead</strong> fraction except at <strong>the</strong> 600 m site<br />

where it accounted for 10% of <strong>the</strong> <strong>dead</strong> assemblage.<br />

A. primitiva was only represented in <strong>the</strong> smaller living<br />

fraction <strong>and</strong> occurred between 136 <strong>and</strong> 944 m water<br />

depth. Maximum values were found between 512 <strong>and</strong><br />

738 m (Fig. 12). Some planoconvex species such as<br />

Gyroidina orbicularis <strong>and</strong> G. altiformis were a minor<br />

faunal element in <strong>the</strong> upper part of <strong>the</strong> OMZ (Abb. 8).<br />

The lower part of <strong>the</strong> OMZ had a more diverse fauna<br />

(compare Figs. 6 <strong>and</strong> 7) <strong>and</strong> contained more cosmopolitan<br />

species. Members of <strong>the</strong> Globobulimina group<br />

(G. cf. G. pyrula, Globobulimina sp. 3, Praeglobobulimina<br />

sp. 1 <strong>and</strong> Praeglobobulimina pupoides) were<br />

represented over <strong>the</strong> total OMZ with low numbers in<br />

<strong>the</strong> living fraction N 150 μm, except at <strong>the</strong> 512 m site<br />

where <strong>the</strong>y accounted for more than 40% of <strong>the</strong> fauna.<br />

These species were virtually absent in <strong>the</strong> smaller size<br />

fraction. These were well represented in <strong>the</strong> <strong>dead</strong> fraction,<br />

with a maximum relative abundance of 21% at<br />

598 m water depth (Figs. 11 <strong>and</strong> 13). It appears that<br />

G. cf. G. pyrula was present across <strong>the</strong> OMZ, while<br />

Praeglobobulimina sp. 1 was more common below<br />

512 m. A significant difference could be observed in<br />

<strong>the</strong> relative abundances of this group between <strong>the</strong> living<br />

<strong>and</strong> <strong>dead</strong> faunas, with generally higher percentages in<br />

<strong>the</strong> <strong>dead</strong> faunas.<br />

Between 512 <strong>and</strong> 1201 m water depth, Reophax spp.<br />

<strong>and</strong> Ammodiscus spp. accounted for more than 20% of<br />

<strong>the</strong> entire hard-shelled fauna. This area was below <strong>the</strong><br />

main occurrence of Uvigerina ex gr. U. semiornata,<br />

Bolivina aff. B. dilatata, <strong>and</strong> B. dilatata. Down to 738 m,<br />

<strong>the</strong> small sized B. exilis <strong>and</strong> A. primitiva occurred toge<strong>the</strong>r<br />

with significant numbers of Reophax spp. <strong>and</strong><br />

Ammodiscus spp. (Fig. 14).<br />

4. Discussion<br />

The assemblages described above present a clear<br />

zonation across <strong>the</strong> Pakistan margin OMZ in samples<br />

collected during March/April 2003 (Fig. 15). The main<br />

features of this pattern are as follows.<br />

1. At 136 m, above <strong>the</strong> OMZ, where <strong>the</strong> bottom water<br />

was relatively well oxygenated (N1.5 ml/l) during<br />

March/May 2003 (although not later in <strong>the</strong> year


58 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

Fig. 11. Single species distribution over <strong>the</strong> OMZ for <strong>the</strong> calcareous live faunaN150 μm.<br />

during <strong>the</strong> monsoonal season when this site becomes<br />

dysoxic), living <strong>foraminifera</strong> are present down to<br />

3 cm depth in <strong>the</strong> sediment. Faunal densities <strong>and</strong><br />

diversities are high, especially for <strong>the</strong> smaller (63–<br />

150 μm) living <strong>and</strong> larger (N150 μm) <strong>dead</strong> fractions.<br />

The dominant species in <strong>the</strong> living <strong>and</strong> <strong>dead</strong> faunas of<br />

<strong>the</strong> larger size fraction are Uvigerina ex gr. U. semiornata<br />

<strong>and</strong> Bolivina aff. B. dilatata; in <strong>the</strong> 63–<br />

150 μm fraction B. dilatata is a major element <strong>and</strong><br />

Uvigerina ex gr. U. semiornata less frequent.


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

59<br />

Fig. 12. Single species distribution over <strong>the</strong> OMZ for <strong>the</strong> calcareous live fauna 63–150 μm.<br />

2. At <strong>the</strong> upper boundary of <strong>the</strong> OMZ (150 m, oxygen<br />

0.5 ml/l during March/May 2003) <strong>and</strong> in <strong>the</strong> core of <strong>the</strong><br />

OMZ (below 150 to about 500 m water depth), where<br />

bottom water oxygenation is very low (≈ 0.1 ml/l),<br />

living <strong>foraminifera</strong> are only found in <strong>the</strong> upper 2 cm of<br />

<strong>the</strong> sediment. High faunal densities contrast<br />

with relatively low diversity values. Uvigerina ex gr.<br />

U. semiornata <strong>and</strong> Bolivina aff. B. dilatata dominate<br />

<strong>the</strong> larger size fraction of <strong>the</strong> live as well as <strong>dead</strong><br />

faunas, whereas B. dilatata <strong>and</strong> B. exilis are <strong>the</strong>


60 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

Fig. 13. Single species distribution over <strong>the</strong> OMZ for <strong>the</strong> calcareous <strong>dead</strong> fauna N150 μm.<br />

dominant taxa in <strong>the</strong> 63–150 μm fraction, where<br />

Uvigerina ex gr. U. semiornata is less frequent.<br />

3. In <strong>the</strong> lower part of <strong>the</strong> OMZ (about 600 to 1200 m<br />

depth) bottom water oxygenation shows a slight increase<br />

(values of 0.1 to 0.2 ml/l), <strong>and</strong> live <strong>foraminifera</strong><br />

are found down to 1 or 2 cm depth in <strong>the</strong> sediment. In<br />

<strong>the</strong> larger fraction (N150 μm) faunal densities decrease<br />

significantly with increasing water depth. In <strong>the</strong><br />

smaller size fraction (63–150 μm) a density maximum<br />

is observed at 738 m water depth. The dominant taxa<br />

are U. peregrina, Bolivina aff. B. dilatata (only at<br />

598 m), B. exilis, Reophax spp. <strong>and</strong> Ammodiscus spp.


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

61<br />

Fig. 14. Relative abundance of Ammodiscus sp. 1 <strong>and</strong> Reophax spp. in <strong>the</strong> entire hard-shelled living faunaN150 μm <strong>and</strong> 63–150 μm.<br />

Fig. 15. Scheme of species zonation across <strong>the</strong> OMZ. Given are <strong>the</strong> oxygen profile, <strong>the</strong> sediment depth until which <strong>stained</strong> <strong>foraminifera</strong> are found (0–<br />

4 cm), <strong>the</strong> densities <strong>and</strong> diversity (L=low, H=high), dominant taxa for <strong>the</strong> large size fraction (<strong>dead</strong> as well as live fauna) <strong>and</strong> for <strong>the</strong> 63–150 μm (live<br />

fauna). The hatched field indicates deep infaunal taxa, that are more prominent in <strong>the</strong> <strong>dead</strong>, than in <strong>the</strong> live faunas.


62 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

63<br />

4. Below <strong>the</strong> OMZ (1870 m water depth, bottom water<br />

oxygen concentration≈1.75 ml/l), living <strong>foraminifera</strong><br />

are found down to 4 cm in <strong>the</strong> sediment, faunal<br />

diversities are high <strong>and</strong> <strong>the</strong> fauna constitutes a typical,<br />

cosmopolitan deep-sea assemblage.<br />

4.1. Identification of dominant taxa<br />

Many of <strong>the</strong> dominant taxa found in <strong>the</strong> upper part of<br />

<strong>the</strong> OMZ (above 500 m) were difficult to identify. Earlier<br />

authors (Jannink et al., 1998; Maas, 2000) working with<br />

Indian Ocean faunas ei<strong>the</strong>r incorrectly attributed <strong>the</strong>se<br />

abundant species to morphologically very different<br />

cosmopolitan taxa, or placed <strong>the</strong>m in <strong>the</strong> open<br />

nomenclature.<br />

4.1.1. Uvigerina ex gr. U. semiornata<br />

The test is normally triserial, rounded in outline <strong>and</strong><br />

with a relatively low length/width ratio. Chambers are<br />

strongly inflated <strong>and</strong> overlapping. The short neck is<br />

positioned in a prominent depression, a tooth is present<br />

in <strong>the</strong> aperture. Except on <strong>the</strong> last chamber, where <strong>the</strong>y<br />

may be absent, <strong>the</strong> test is ornamented with costae which<br />

can cross <strong>the</strong> sutures. Large pores are situated between<br />

<strong>the</strong> costae. Some specimens are more elongated <strong>and</strong> in<br />

adults <strong>the</strong> last chambers may become biserial. The position<br />

of <strong>the</strong> neck, <strong>the</strong> length/width ratio of <strong>the</strong> test <strong>and</strong><br />

<strong>the</strong> inflated chambers, strongly overlapping <strong>the</strong> older<br />

chambers, clearly shows that this species belongs in <strong>the</strong><br />

U. semiornata/U. mediterranea (Van der Zwaan et al.,<br />

1986) group <strong>and</strong> is very different <strong>from</strong> U. peregrina,<br />

which is also present in our material. Uvigerina ex gr.<br />

U. semiornata differs <strong>from</strong> U. peregrina by <strong>the</strong> position<br />

of <strong>the</strong> short neck in a depression. U. mediterranea has<br />

more closely spaced costae, lacks a tendency towards<br />

biserial or uniserial chamber arrangement in later<br />

ontogentic stages, <strong>and</strong> does never become very<br />

elongated. We refrain <strong>from</strong> putting our material in <strong>the</strong><br />

synonymy of U. semiornata because this taxon has<br />

mainly been described <strong>from</strong> Miocene <strong>and</strong> older<br />

deposits.<br />

Uvigerina ex gr. U. semiornata was reported previously<br />

on <strong>the</strong> Indo-Pakistan Continental Margin by<br />

Maas (2000). He described abundant populations of this<br />

species at 233 m water depth, where U. peregrina is<br />

absent. In two transects off Karachi, Jannink et al. (1998)<br />

reported Uvigerina phlegeri (=Rectuvigerina phlegeri<br />

Le Calvez, 1995) occurring toge<strong>the</strong>r with U. peregrina<br />

at <strong>the</strong> shallowest station (500 m). We strongly suspect<br />

that U. phlegeri of Jannink et al. (1998) is identical to<br />

Uvigerina ex gr. U. semiornata of <strong>the</strong> present study<br />

<strong>and</strong> of Maas (2000). Adult specimens are sometimes<br />

very elongated, <strong>and</strong> have, like <strong>the</strong> much smaller U.<br />

phlegeri, a tendency to become biserial or even uniserial<br />

in <strong>the</strong> later stages, leading to a confusion between <strong>the</strong> two<br />

species. However, our assemblages always contain many<br />

specimens that are triserial throughout. Gooday (2003)<br />

misidentified this species <strong>from</strong> <strong>the</strong> Oman margin of <strong>the</strong><br />

Arabian Sea as Rectuvigerina cylindrica.<br />

4.1.2. Bolivina dilatata Reuss, 1850<br />

Our morphotype correspond very well to <strong>the</strong> typical<br />

forms, originally described by Reuss (1850), <strong>and</strong> later<br />

figured by Von Daniels (1970) <strong>and</strong> Barmawidjaja et al.<br />

(1992). They have an elongated, slowly tapering test with<br />

<strong>the</strong> axis sometimes slightly twisted. Only <strong>the</strong> outer parts of<br />

<strong>the</strong> test are slightly compressed. The periphery is rounded.<br />

There are 16 to 20 chambers in adult specimens. The later<br />

chambers become strongly inflated, <strong>and</strong> are often<br />

imperforate in <strong>the</strong>ir upper parts. The sutures are not<br />

limbate, <strong>and</strong> are strongly curved downwards towards <strong>the</strong><br />

center of <strong>the</strong> test.<br />

Plate I. Scale bar=100 μm.<br />

1–6. Uvigerina ex gr. U. semiornata. 1. 55901#11, 0–0.5 cm, 150–300 μm; 2. 55901#11, 0–0.5 cm, 150–300 μm; 3. 55901#11, 0–0.5 cm, 150–<br />

300 μm; 4. 55818#4, 0–0.5 cm, 150–300 μm; 5. 55808#3, 0–0.5 cm. 63–150 μm; 6. 55901#11, 0–0.5 cm, 63–150 μm.<br />

7. Uvigerina peregrina Cushman 55922#2, 0–0.5 cm, 150–300 μm.<br />

8–12. Bolivina aff. B. dilatata. 8. 55818#4, 0–0.5 cm, 150–300 μm; 9. 55901#11, 0–0.5 cm, 150–300 μm; 10. 55803#5, 0–0.5 cm, 150–300 μm;<br />

11. 55818#4, 0–0.5 cm, 150–300 μm; 12. 55901#11, 0–0.5 cm, 63–150 μm.<br />

13–14. Bolivina dilatata Reuss. 13. 55901#11, 0–0.5 cm, 63–150 μm; 14. 55901#11, 0–0.5 cm, 63–150 μm.<br />

15. Bulimina exilis, Brady. 5818#4, 0–0.5 cm, 63–150 μm.<br />

16. Alliatina primitiva (Cushmann <strong>and</strong> McCulloch). 55803#5, 0–0.5 cm, 63–150 μm.<br />

17. Praeglobobulimina sp. 1. 55922#2, 0–0.5 cm, 150–300 μm.<br />

18. Praeglobobulimina pupoides (d'Orbigny). 55922#2, 0–0.5 cm, 150–300 μm.<br />

19. Globobulimina sp. 3. 55922#2, 0–0.5 cm, N300 μm.<br />

20. Globobulimia cf. G. pyrula. 55921#1, 2–3 cm, N300 μm.<br />

21. Ammodiscus sp. 1 (Brady). 55916#1, 0–0.5 cm, 150–300 μm.<br />

22. Reophax bilocularis Flint. 55921#1, 0–0.5 cm, 150–300 μm.<br />

23. Reophax scorpiurus Montfort. 55921#1, 0–0.5 cm, 150–300 μm.


64 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

4.1.3. Bolivina aff. B. dilatata<br />

The test of Bolivina aff. B. dilatata is strongly tapering,<br />

strongly flattened, with large pores. The periphery<br />

is more or less rounded, without a keel, very often<br />

provided with a striate ornamentation. Adult specimens<br />

consist of about 10 chambers, rectangular in shape,<br />

making an angle of about 45° with <strong>the</strong> periphery. The<br />

sutures are limbate, straight until <strong>the</strong> very center of <strong>the</strong><br />

test where <strong>the</strong>y may be strongly inflated. Megalospheric<br />

forms have a very large <strong>and</strong> inflated first chamber (see<br />

Plate 1, fig. 12). Sometimes, a faint striate ornamentation<br />

on <strong>the</strong> earlier part of <strong>the</strong> test can be observed. This<br />

species was determined as B. dilatata by Maas (2000)<br />

<strong>and</strong> Jannink et al. (1998) but we are convinced that it is a<br />

different species. It is slightly resembles Longinelli's<br />

(1956) ra<strong>the</strong>r unclear figure of B. dilatata Reuss var.<br />

abbreviata, <strong>and</strong> correspond to his description. Unfortunately<br />

no type material is available for this taxon, so that<br />

a relationship between our species, which is very frequent<br />

in modern Indian Ocean faunas, <strong>and</strong> Longinelli's<br />

Miocene–Pleistocene variety, cannot be confirmed.<br />

Faunas dominated by morphotypes resembling Bolivina<br />

aff. B. dilatata have been documented <strong>from</strong> <strong>the</strong><br />

Nor<strong>the</strong>astern Arabian Sea at 230 m (Maas, 2000, as<br />

B. dilatata) <strong>and</strong> 550 m water depth (Jannink et al., 1998,<br />

as B. dilatata). This species was identified as Bolivina<br />

seminuda by Gooday et al. (2000), at 412 m on <strong>the</strong><br />

Oman margin, where was <strong>the</strong> dominant taxon.<br />

4.2. Comparison with previous studies in <strong>the</strong> Indian<br />

Ocean<br />

Because we had a larger number of sampling points,<br />

we were able to determine <strong>the</strong> distribution of <strong>foraminifera</strong>l<br />

species in relation to water depth <strong>and</strong> bottomwater<br />

oxygen concentrations on <strong>the</strong> Indo-Pakistan Continental<br />

Margin in greater detail than in <strong>the</strong> study of Maas<br />

(2000). He described assemblages <strong>from</strong> 233, 658 <strong>and</strong><br />

902 m water depth in an area where <strong>the</strong> bottom water<br />

oxygenation (∼ 0.2 ml/l at 230 <strong>and</strong> 650 m) was slightly<br />

higher than in <strong>the</strong> upper part of <strong>the</strong> OMZ in our study<br />

area. As in <strong>the</strong> present study, <strong>the</strong> faunas described by<br />

Maas (2000) at 233 m were dominated by B. dilatata<br />

(our Bolivina aff. B. dilatata), Uvigerina ex. gr. U.<br />

semiornata <strong>and</strong> C. auriculus, whereas arenaceous taxa<br />

were most abundant at 902 m depth. However, <strong>the</strong><br />

assemblages described by Maas (2002) generally contained<br />

more “deep-infaunal” species (Globobulimina),<br />

especially at 650 m where Globobulimina affinis (our P.<br />

pupoides) were dominant.<br />

The work of Jannink et al. (1998) was based on two<br />

transects <strong>from</strong> 500 to 2000 m water depth off Karachi.<br />

As in our study area, <strong>the</strong> faunas described by Jannink<br />

et al. (1998) <strong>from</strong> 500 m were dominated by B. dilatata<br />

(our Bolivina aff. B. dilatata), C. laevigata <strong>and</strong> Ammodiscus<br />

sp., <strong>and</strong>, in <strong>the</strong> 63–150 μm size fraction, by<br />

B. exilis. From 1000 to 1254 m water depth, U. peregrina,<br />

Rotaliatinopsis semiinvoluta <strong>and</strong> Ammodiscus sp. were<br />

frequent, while below <strong>the</strong> OMZ B. aculeata <strong>and</strong> Lagenammina<br />

sp. were abundant. Although <strong>the</strong>se taxa are not<br />

exactly <strong>the</strong> same as those found in our study, <strong>the</strong> decrease<br />

in faunal density <strong>and</strong> <strong>the</strong> increase in diversity <strong>and</strong> ALD 10<br />

with increasing water depth is very similar.<br />

Hermelin <strong>and</strong> Shimmield (1990) described faunas<br />

<strong>from</strong> <strong>the</strong> northwestern Arabian Sea (Oman margin). The<br />

upper bathyal (440–640 m) assemblages were characterised<br />

by high abundances of various Bulimina <strong>and</strong><br />

Bolivina species, whereas below 530 m U. peregrina<br />

is present in low numbers. The authors consider <strong>the</strong>se<br />

faunas to reflect <strong>the</strong> high organic carbon content of <strong>the</strong><br />

sediment. In our samples, Uvigerina ex gr. U. semiornata<br />

is always a conspicuous faunal element in <strong>the</strong><br />

upper part of <strong>the</strong> OMZ, between 136 <strong>and</strong> 500 m, although<br />

it decreases significantly with water depth.<br />

Hermelin <strong>and</strong> Shimmield (1990) do not mention this<br />

species. However, it is very abundant in samples <strong>from</strong><br />

100 m water depth on <strong>the</strong> Oman margin analysed by<br />

Ar<strong>and</strong>a da Silva (pers. comm. 2005). A specimen of<br />

Uvigerina ex gr. U. semiornata <strong>from</strong> this site was<br />

illustrated by Gooday (2003, Fig. 3F) as R. cylindrica (a<br />

misidentification). In <strong>the</strong> deeper part of <strong>the</strong> northwestern<br />

Arabian Sea (600–1000), oxygen depletion is slightly<br />

less severe, as in our study area. Here, Hermelin <strong>and</strong><br />

Shimmield (1990) described a fauna dominated by<br />

U. peregrina, Ehrenbergina trigona, Hyalinea balthica<br />

<strong>and</strong> Tritaxis sp. 1. Although not all <strong>the</strong>se species are<br />

found in our study area, <strong>the</strong> increase in diversity towards<br />

deeper sites is very similar in both areas.<br />

4.3. Comparison with o<strong>the</strong>r oxygen-depleted areas<br />

There are differences between <strong>the</strong> taxonomic composition<br />

of faunas <strong>from</strong> <strong>the</strong> Arabian Sea OMZ <strong>and</strong> those<br />

reported <strong>from</strong> o<strong>the</strong>r oxygen-depleted areas. In <strong>the</strong><br />

dysoxic Santa Catalina Basin (bottom-water oxygen<br />

0.2 to 0.5 ml/l) faunas are dominated by Bolivina spissa<br />

<strong>and</strong> Globobulmina pacifica, while in <strong>the</strong> Santa Monica<br />

Basin (oxygen b 0.2 ml/l) <strong>the</strong> fauna is more diverse <strong>and</strong><br />

includes Cancris inaequalis, Globobulimina pacifica,<br />

Rosalina columbiensis <strong>and</strong> Planulina ariminensis<br />

(Mackensen <strong>and</strong> Douglas, 1989). In <strong>the</strong> Santa Barbara<br />

Basin (oxygen 0.02–0.5 ml/l), B. seminuda, B. argentea,<br />

Bulimina tenuata, Chilostomella ovoidea <strong>and</strong> Nonionella<br />

stella are common taxa (Bernhard et al., 1997).


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

65<br />

Small, thin-shelled forms of Bolivina <strong>and</strong> Bolivina-like<br />

species are also abundant within <strong>the</strong> sou<strong>the</strong>rn East<br />

Pacific margin OMZ (Sen Gupta <strong>and</strong> Machain-Castillo,<br />

1993). However, species of Uvigerina are almost absent<br />

in most previously described low-oxygen faunas,<br />

whereas this genus is a dominant faunal element in <strong>the</strong><br />

Arabian Sea OMZ. All Bolivina species encountered in<br />

o<strong>the</strong>r low-oxygen areas tend to be very elongated, with a<br />

high length/width ratio, whereas our dominant taxon,<br />

Bolivina aff. B. dilatata, is strongly tapering <strong>and</strong> has a<br />

low length/width ratio. Low-oxygen faunas <strong>from</strong><br />

Mediterranean sapropels exhibit very low diversity <strong>and</strong><br />

are mainly dominated by deep infaunal taxa, e.g. Chilostomella<br />

spp. <strong>and</strong> Globobulimina spp. (Jorissen, 1999;<br />

Schmiedl et al., 2003). B. dilatata has been described as<br />

part of <strong>the</strong> repopulating faunas found in <strong>the</strong> upper parts of<br />

some sapropels (Schmiedl et al., 2003).<br />

4.4. Distribution of species across <strong>the</strong> OMZ<br />

In this section, we discuss species distributions on <strong>the</strong><br />

Pakistan margin, using water depth as a proxy for bottom-water<br />

oxygenation. There is a major break at about<br />

500 m water depth in our study area. At this depth a live<br />

fauna dominated by Uvigerina ex gr. U. semiornata,<br />

Bolivina aff. B. dilatata, <strong>and</strong> in <strong>the</strong> 63–150 μm fraction<br />

also by B. dilatata <strong>and</strong> B. exilis, is replaced by a fauna<br />

with abundant U. peregrina, Reophax spp. <strong>and</strong> Ammodiscus<br />

spp., with B. exilis again present in <strong>the</strong> smaller<br />

fraction. The <strong>dead</strong> fauna below 500 m also includes<br />

Globobulimina spp., B. aculeata <strong>and</strong> H. balthica.<br />

Most species found below 600 m depth are well<br />

known in <strong>the</strong> literature <strong>and</strong> can be considered as cosmopolitan.<br />

U. peregrina is described in all ocean basins,<br />

including <strong>the</strong> North Atlantic (Lutze, 1986; Loubere<br />

et al., 1995; Jorissen et al., 1998; Fontanier et al., 2002,<br />

2003), South Atlantic (Mackensen et al., 1995; Licari<br />

et al., 2003), Pacific Ocean (Loubere, 1994; Kitazato<br />

et al., 2000), Indian Ocean below <strong>the</strong> OMZ (Hermelin<br />

<strong>and</strong> Shimmield, 1990; Jannink et al., 1998) <strong>and</strong> Mediterranean<br />

Sea (De Rijk et al., 2000). This species is also<br />

common in slightly oxygen depleted areas, e.g. off California<br />

(Bernhard, 1992), in <strong>the</strong> Angola Basin (Van<br />

Leeuwen, 1989) <strong>and</strong> is found associated with cold methane<br />

seeps (Rathburn et al., 2000). In <strong>the</strong> Arabian Sea<br />

OMZ, U. peregrina is described below 500 m water<br />

depth by several authors, in <strong>the</strong> northwestern (Hermelin<br />

<strong>and</strong> Shimmield, 1990), as well as in <strong>the</strong> nor<strong>the</strong>astern part<br />

(Jannink et al., 1998; Maas, 2000).<br />

In our study area, <strong>the</strong> dominant Reophax species are<br />

R. dentaliniformis, R. bilocularis (with a varying test<br />

morphology), <strong>and</strong> morphotypes close to R. scorpiurus.<br />

These taxa have been described in many o<strong>the</strong>r ocean<br />

basins, for example, <strong>the</strong> North Atlantic (Schröder, 1986;<br />

Wollenburg <strong>and</strong> Mackensen, 1998; Gooday <strong>and</strong> Hughes,<br />

2002), South Atlantic (Mackensen et al., 1990, 1993;<br />

Licari et al., 2003), Pacific Ocean (Jonasson et al., 1995)<br />

<strong>and</strong> <strong>the</strong> Indian Ocean below <strong>the</strong> OMZ (Gooday et al.,<br />

2000). In <strong>the</strong> northwestern Arabian Sea, Hermelin <strong>and</strong><br />

Shimmield (1990) described Reophax species around <strong>the</strong><br />

lower boundary of <strong>the</strong> OMZ (at 1048 m depth), whereas<br />

Gooday et al. (2000) described high abundances at<br />

3500 m, well below <strong>the</strong> OMZ. In <strong>the</strong> nor<strong>the</strong>astern<br />

Arabian Sea, Reophax species occur with moderate<br />

abundances below 600 m water depth (Maas, 2000).<br />

The high densities of Ammodiscus spp. found in our<br />

study area between 300 m <strong>and</strong> 850 m are more unusual.<br />

The most abundant of <strong>the</strong>se species, Ammodiscus sp. 1<br />

(see Appendix), was described as A. cretaceus <strong>from</strong><br />

233 m <strong>and</strong> at 902 m water depth by Maas (2000).<br />

Jannink et al. (1998) described it as Ammodiscus sp.,<br />

with occurrences between 500 <strong>and</strong> 2000 m depth on<br />

both her transects off Karachi. Similar species have been<br />

described on <strong>the</strong> Antarctic Shelf by Igarashi et al. (2001,<br />

as Ammodiscus sp.) <strong>and</strong> in <strong>the</strong> North Atlantic (Schröder,<br />

1986 as Ammodiscus incertus).<br />

We found B. exilis mainly in <strong>the</strong> 63–150 μm fraction<br />

between 150 <strong>and</strong> 944 m water depth. This usually very<br />

fusiform buliminid is known <strong>from</strong> upwelling <strong>and</strong> lowoxygen<br />

environments <strong>and</strong> is common between 700 <strong>and</strong><br />

2000 m water depth in sediments off Cape Blanc (Caralp,<br />

1984, 1989), off Northwest Portugal (Caralp,<br />

1984), <strong>and</strong> on <strong>the</strong> continental slope off Southwestern<br />

Africa, (Schmiedl et al., 1997). In <strong>the</strong> Arabian Sea it<br />

commonly has a shorter, less fusiform test, with a lower<br />

length/width ratio. Maas (2000) described its main<br />

occurrence (as Buliminella exilis) at 658 <strong>and</strong> 902 m<br />

water depth. In <strong>the</strong> profiles off Karachi, Jannink et al.<br />

(1998) described our shorter morphotype between 500<br />

<strong>and</strong> 1555 m water depth, with highest abundances in <strong>the</strong><br />

63–150 μm fraction between 500 <strong>and</strong> 1000 m.<br />

4.5. Low-oxygen tolerance of Arabian Sea OMZ species<br />

As indicated in Section 4.1, <strong>the</strong>re are no reports of<br />

Uvigerina ex gr. U. semiornata <strong>and</strong> Bolivina aff. B.<br />

dilatata, <strong>the</strong> dominant species in <strong>the</strong> upper part of <strong>the</strong><br />

Arabian Sea OMZ, in o<strong>the</strong>r oceans. They are largely<br />

restricted to <strong>the</strong> upper cm of <strong>the</strong> sediment, similar to <strong>the</strong><br />

usual shallow infaunal microhabitat occupied by related<br />

species, such as U. mediterranea (De Stigter et al., 1998;<br />

Fontanier et al., 2002, 2003; Schmield et al., 2004), <strong>and</strong><br />

B. dilatata/spathulata (Barmawidjaja et al., 1992; De<br />

Stigter et al., 1998). Apparently, in <strong>the</strong> Indian Ocean, <strong>the</strong>se


66 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

shallow infaunal taxa are extremely tolerant of <strong>the</strong> persistent<br />

low-oxygen conditions encountered at <strong>the</strong> sediment–water<br />

interface within <strong>the</strong> upper OMZ. The faunas<br />

of <strong>the</strong> lower part of <strong>the</strong> OMZ, on <strong>the</strong> contrary, contain<br />

various elements known <strong>from</strong> deep infaunal microhabitats<br />

in o<strong>the</strong>r, better oxygenated areas. They include Globobulimina<br />

spp. (Jorissen et al., 1998; Kitazato, 1994;<br />

Rathburn et al., 2000; Fontanier et al., 2002; Licari et al.,<br />

2003), Praeglobobulimina sp., Chilostomella (Corliss<br />

<strong>and</strong> Emmerson, 1990; Kitazato, 1994; Fontanier et al.,<br />

2002; Licari et al., 2003), Fursenkoina (Jorissen et al.,<br />

1998; Licari et al., 2003) <strong>and</strong>B. exilis (Caralp, 1989;<br />

Jannink et al., 1998; Maas, 2000). These taxa, which<br />

occupy deep infaunal niches elsewhere, are found in<br />

superficial sediments in <strong>the</strong> lower part of <strong>the</strong> OMZ of our<br />

study area. This observation are consistent with <strong>the</strong> TROX<br />

model of Jorissen et al. (1995) that predicts a migration of<br />

deep infaunal taxa to <strong>the</strong> sediment surface under very low<br />

oxygen conditions. In Late Quaternary Mediterranean<br />

sapropels, <strong>the</strong> same deep infaunal taxa, that are well<br />

adapted to low oxygen conditions, dominated <strong>the</strong> <strong>benthic</strong><br />

<strong>foraminifera</strong>l faunas immediately before <strong>the</strong> onset of<br />

sapropel formation (Jorissen, 1999; Casford et al., 2003;<br />

Schmiedl et al., 2003).<br />

An interesting question is why <strong>the</strong>se deep infaunal<br />

taxa, that are very resistant to low oxygen conditions (as<br />

is shown by <strong>the</strong>ir usual microhabitat around <strong>the</strong> zero<br />

oxygen boundary), are poorly represented in <strong>the</strong> superficial<br />

low oxygen environments in <strong>the</strong> upper part of <strong>the</strong><br />

OMZ. In o<strong>the</strong>r words, why did a ra<strong>the</strong>r specific fauna,<br />

with species that we do not find in o<strong>the</strong>r oceans, dominate<br />

in <strong>the</strong> upper part of <strong>the</strong> OMZ, but not <strong>the</strong> lower<br />

part? We suggest that this important difference between<br />

upper <strong>and</strong> lower OMZ faunas may be explained by a<br />

difference in <strong>the</strong> stability of dysoxia over geological<br />

time.<br />

Various geochemical <strong>and</strong> micropaleontological studies<br />

describe variations in Arabian Sea OMZ intensity<br />

during <strong>the</strong> Late Quaternary (Reichart et al., 1997, 1998,<br />

2002; Schulte et al., 1999; Von Rad et al., 1999; Den<br />

Dulk et al., 2000). Changes in OMZ intensity are<br />

mainly inferred <strong>from</strong> reconstructed primary production<br />

rates (e.g. Reichart et al., 1997; Schulte et al., 1999;<br />

Agnihotri et al., 2003), that are correlated with<br />

millennium-scale orbital forcing <strong>and</strong> have been linked<br />

to climate at high latitudes in <strong>the</strong> nor<strong>the</strong>rn hemisphere<br />

(Leuschner <strong>and</strong> Sirocko, 2000, 2003; Reichart et al.,<br />

2002). Increased primary production has been associated<br />

with maximum summer monsoon intensity in<br />

glacial periods, whereas interglacial periods are supposed<br />

to have been characterised by lower productivity<br />

(Reichart et al., 1998; Schulte et al., 1999; Von Rad et<br />

al., 1999; Agnihotri et al., 2003). High glacial surfacewater<br />

productivity probably caused higher organic<br />

carbon flux rates, a vertical expansion of <strong>the</strong> OMZ, <strong>and</strong><br />

lower O 2 levels over <strong>the</strong> whole OMZ. In interglacial<br />

periods, on <strong>the</strong> contrary, reduced productivity should<br />

cause a weakened, shallower, <strong>and</strong> more instable OMZ<br />

(Von Rad et al., 1999; Den Dulk et al., 2000).<br />

Bottom water circulation is a second factor influencing<br />

<strong>the</strong> lower boundary of <strong>the</strong> OMZ (Schulte et al.,<br />

1999; Den Dulk, 2000; Schmiedl <strong>and</strong> Leuschner, 2005).<br />

Today, <strong>the</strong> OMZ is positioned below <strong>the</strong> Arabian Sea<br />

High Salinity Water, formed during winter monsoons in<br />

<strong>the</strong> Nor<strong>the</strong>rn Arabian Sea. The waters residing in <strong>the</strong><br />

OMZ are formed by a mixture of intermediate water<br />

masses, <strong>the</strong> Red Sea <strong>and</strong> <strong>the</strong> Persian Sea Outflow Water,<br />

<strong>and</strong> <strong>the</strong> North Indian Intermediate Water formed in <strong>the</strong><br />

sou<strong>the</strong>rn Indian Ocean. The very low (northward) transport<br />

rates of this water mass amplify its oxygen-depleted<br />

character (Prasanna Kumar <strong>and</strong> Prasa, 1999; Schott <strong>and</strong><br />

McCreary, 2001). Below <strong>the</strong> OMZ, <strong>the</strong> Arabian Sea is<br />

ba<strong>the</strong>d by well-oxygenated North Indian Deep Water<br />

(NIDW), which is formed mainly by North Atlantic<br />

Deep Water (NADW) <strong>and</strong> Circumpolar Deep Water<br />

(CDW) (Schott <strong>and</strong> McCreary, 2001; Van Aken et al.,<br />

2004).<br />

According to Reichart et al. (2002), in <strong>the</strong> Late<br />

Quaternary, <strong>the</strong> upper part of <strong>the</strong> OMZ was only<br />

ventilated during <strong>the</strong> coldest phases of <strong>the</strong> Dansgaard–<br />

Oeschger cycles, <strong>the</strong> so-called Heinrich events, as a<br />

consequence of local intermediate water formation.<br />

Both during glacial periods, <strong>and</strong> during normal interglacial<br />

conditions, <strong>the</strong> upper part of <strong>the</strong> OMZ was<br />

always oxygen depleted. The lower part of <strong>the</strong> OMZ,<br />

on <strong>the</strong> contrary, has probably been much more<br />

sensitive to short-term changes in primary production<br />

(<strong>and</strong> resulting export production), <strong>and</strong> will probably<br />

have shifted up <strong>and</strong> down over time (Den Dulk, 2000;<br />

Schmiedl <strong>and</strong> Leuschner, 2005). The result is that <strong>the</strong><br />

core region of <strong>the</strong> OMZ has been stable over periods<br />

of 1000s to 10,000s of years, despite evidence for<br />

seasonal fluctuations of <strong>the</strong> upper boundary (Bett,<br />

2004b). Such prolonged periods of severe dysoxia will<br />

have enabled faunal elements living in <strong>the</strong> OMZ core<br />

to acquire <strong>the</strong> evolutionary adaptations needed to<br />

invade <strong>the</strong>se low oxygen environments (Rogers, 2000),<br />

that are attractive because of <strong>the</strong>ir extremely high food<br />

availability. In <strong>the</strong> lower part of <strong>the</strong> OMZ, on <strong>the</strong><br />

contrary, conditions have probably been much more<br />

variable over geological time, because of changes in<br />

surface productivity <strong>and</strong> <strong>the</strong> intensity of <strong>the</strong> influence<br />

of NADW (Den Dulk, 2000; Schmiedl <strong>and</strong> Leuschner,<br />

2005). As a consequence, <strong>the</strong> cosmopolitan, deep-


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

67<br />

infaunal taxa were most successful in colonizing <strong>the</strong>se<br />

environments that fluctuated between oxic <strong>and</strong> dysoxic<br />

conditions. The fact that <strong>the</strong> same or similar deep<br />

infaunal taxa also dominate <strong>the</strong> Mediterranean presapropel<br />

faunas for very short periods of time (50–<br />

100 years) when <strong>the</strong> system shifts <strong>from</strong> welloxygenated<br />

(N4 ml/l) to anoxic conditions (Jorissen,<br />

1999; Schmiedl et al., 2003), fur<strong>the</strong>r illustrates <strong>the</strong><br />

need for prolonged periods of time for <strong>the</strong> adaptation<br />

of taxa to severe low-oxygen conditions. As in <strong>the</strong><br />

case of <strong>the</strong> onset of sapropels in <strong>the</strong> Late Quaternary<br />

Mediterranean, <strong>the</strong> lower part of <strong>the</strong> OMZ in <strong>the</strong><br />

Indian Ocean apparently has not been sufficiently<br />

stable over geological time scales to allow <strong>the</strong><br />

adaptation of a highly specialized, shallow infaunal<br />

fauna.<br />

5. Conclusion<br />

There is a clear zonation of <strong>benthic</strong> <strong>foraminifera</strong>l<br />

faunas in <strong>the</strong> Nor<strong>the</strong>ast Arabian Sea OMZ. Our composite<br />

profile <strong>from</strong> 136 to 1870 m water depth shows<br />

major faunal breaks at 300, 500 <strong>and</strong> 1000 m water depth.<br />

In <strong>the</strong> core of <strong>the</strong> OMZ <strong>and</strong> above <strong>the</strong> OMZ (136–<br />

500 m) a very specific fauna, unknown in o<strong>the</strong>r oceans,<br />

is found, dominated by shallow infaunal taxa such<br />

as Uvigerina ex gr. U. semiornata <strong>and</strong> Bolivina aff.<br />

B. dilatata; deep infaunal taxa are poorly represented. In<br />

<strong>the</strong> lower part of <strong>the</strong> OMZ (below 500 m), <strong>the</strong> fauna has a<br />

more cosmopolitan character. It includes various taxa,<br />

such as Globobulimina spp. Praeglobobulimina sp. B.<br />

exilis, which occupy deep infaunal niches in better oxygenated<br />

areas but are found in superficial sediment in our<br />

Pakistan margin samples.<br />

The contrasting nature of <strong>the</strong> faunas found in <strong>the</strong><br />

core region <strong>and</strong> <strong>the</strong> lower part of <strong>the</strong> OMZ may be<br />

explained by a difference in <strong>the</strong> stability of dysoxia<br />

over geological time. During glacial <strong>and</strong> interglacial<br />

periods, <strong>the</strong> core of <strong>the</strong> OMZ was always oxygen<br />

depleted, except during <strong>the</strong> coldest phases, <strong>the</strong> Heinrich<br />

events, when it was ventilated as a consequence of local<br />

intermediate water formation. The core region has<br />

<strong>the</strong>refore been stable over long time periods. This<br />

enabled <strong>the</strong> adaptation of shallow infaunal taxa to<br />

extremely low oxygen conditions with a very high food<br />

availability. The lower part of <strong>the</strong> OMZ, on <strong>the</strong><br />

contrary, has been much more unstable over time,<br />

reflecting a higher sensitivity to short-term changes in<br />

primary productivity <strong>and</strong> <strong>the</strong> intensity of NADW flow.<br />

As a consequence, cosmopolitan deep infaunal taxa<br />

were most successful in colonizing <strong>the</strong> temporarily<br />

available low oxygen environments.<br />

Acknowledgements<br />

We thank Greg Cowie for <strong>the</strong> chance to work on<br />

Arabian Sea material <strong>and</strong> for providing <strong>the</strong> opportunity<br />

for two of us (KEL <strong>and</strong> AJG) to sample on <strong>the</strong> Pakistan<br />

margin. We have special <strong>and</strong> kind thoughts for <strong>the</strong> crews<br />

<strong>and</strong> captains of <strong>the</strong> RRS Charles Darwin. We are grateful<br />

to members of <strong>the</strong> scientific parties on RRS Charles<br />

Darwin Cruises 145 <strong>and</strong> 146, particularly Rachel Jeffreys,<br />

Lisa Levin, Matt Schwartz, Christine Whitcraft <strong>and</strong> Clare<br />

Woulds, who helped in numerous ways. We are very<br />

grateful to Barun Sen Gupta <strong>and</strong> an anonymous reviewer<br />

for <strong>the</strong>ir valuable comments. SSch was supported by <strong>the</strong><br />

Conseil Général of <strong>the</strong> Vendée, France. KEL <strong>and</strong> AJG<br />

were supported by <strong>the</strong> Natural Environment Research<br />

Council grant number NER/A/S/2000/01383.<br />

Appendix A<br />

Species of <strong>benthic</strong> <strong>foraminifera</strong> of <strong>the</strong> Pakistan Continental<br />

margin, with reference to plates <strong>and</strong> figures in<br />

<strong>the</strong> literature.<br />

Perforate species<br />

Alliatina primitiva<br />

(Cushmann <strong>and</strong><br />

MaCulloch,<br />

1939) Plate 1,<br />

fig. 16<br />

Amphicoryna scalaris<br />

(Batsch, 1791)<br />

Amphicoryna sublineata<br />

(Brady, 1884)<br />

Anomalinoides<br />

globulosus Chapman<br />

<strong>and</strong> Parr, 1937<br />

Astrononion echolsi<br />

Kennett, 1967<br />

Astrononion sp. 1<br />

Bolivina alata<br />

(Seguenza, 1862)<br />

Bolivina dilatata Reuss,<br />

1850 Plate 1,<br />

figs. 13–14<br />

Bolivina pacifica<br />

Cushmann <strong>and</strong><br />

MaCulloch, 1942<br />

Bolivina seminuda<br />

Cushman, 1911<br />

Bolivina spathulata<br />

(Williamson, 1858)<br />

Bolivina aff. B. dilatata<br />

Plate 1, figs. 8–12<br />

References<br />

Schiebel (1992), pl. 5,<br />

fig. 11<br />

Jones (1994), pl. 63,<br />

figs. 19–22<br />

Jones (1994), pl. 63,<br />

figs. 28–31<br />

Jones (1994) Cibicidoides<br />

globulosus, pl. 94, figs. 4–5<br />

Corliss (1979), pl. 3,<br />

figs. 16–17<br />

It differs <strong>from</strong> A. echolsi Kennett,<br />

1967, by a less significant<br />

ornamentation.<br />

Jones (1994) Brizalina alata,<br />

pl. 53, figs. 2–4<br />

Schiebel (1992), pl. 1,<br />

fig. 4a<br />

Schiebel (1992), pl. 1,<br />

figs. 6b,c<br />

Den Dulk (2001), pl. 1,<br />

fig. 4<br />

Jones (1994) Brizalina<br />

spathulata, pl. 52, figs. 20–21<br />

Test strongly tapering, strongly<br />

flattened, with large pores.<br />

(continued on next page)


68 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

Appendix A (continued)<br />

Appendix A (continued)<br />

Perforate species<br />

References<br />

Perforate species<br />

References<br />

Bolivina aff. B. dilatata<br />

Plate 1, figs. 8–12<br />

Bolivina sp. 2<br />

Bolivina subspinescens<br />

Chusman, 1922<br />

Bulimina aculeata<br />

d'Orbigny, 1826<br />

Bulimina alazanensis<br />

Cushman, 1972<br />

Bulimina exilis Brady,<br />

1884 Plate 1, fig. 15<br />

Bulimina marginata<br />

d'Orbigny, 1826<br />

Cancris auriculus (Fichtel<br />

<strong>and</strong> Moll, 1798)<br />

Cancris oblongus<br />

(Williamson, 1858)<br />

Cassidulina laevigata var.<br />

carinata Silvestri 1896<br />

Cassidulina laevigata<br />

d'Orbigny, 1826<br />

Cassidulina obtusa<br />

Williamson, 1858<br />

Cassidulinoides bradyi<br />

(Norman, 1881)<br />

Chilostomella oolina<br />

Schwager, 1878<br />

Chilostomella ovoidea<br />

Reuss, 1850<br />

Cibicides lobatulus<br />

(Walker <strong>and</strong> Jacob,<br />

1798)<br />

Periphery more or less rounded,<br />

without a keel, very often<br />

provided with a striate<br />

ornamentation. Adult specimens<br />

with about 10 chambers,<br />

rectangular in shape, making an<br />

angle of about 45° with <strong>the</strong><br />

periphery. Sutures limbate,<br />

straight until <strong>the</strong> very centre of<br />

<strong>the</strong> test where <strong>the</strong>ir end may be<br />

strongly inflated. Megalospheric<br />

forms with a very large <strong>and</strong><br />

inflated first chamber.<br />

Sometimes with a faint striate<br />

ornamentation on <strong>the</strong> earlier<br />

part of <strong>the</strong> test.<br />

Maas (2000) Bolivina dilatata,<br />

pl. 2, fig. 5<br />

Jannink et al. (1998)<br />

Bolivina dilatata, pl. 1, fig. 1<br />

A biserial species with a Bolivina<br />

shape. The chambers are inflated<br />

<strong>and</strong> may be spinose at <strong>the</strong>ir lower<br />

end.<br />

Jones (1994) Brizalina<br />

subspinescens, pl. 52, figs. 24–25<br />

Jones (1994), pl. 51, figs. 7–9<br />

Schmiedl (1995), pl. 2, fig. 7<br />

Jannink et al. (1998), pl. 1, fig. 3.<br />

Our specimens are close to <strong>the</strong>se<br />

relatively short specimens.<br />

Van Morkhoven et al. (1986),<br />

pl. 4, figs. 1−2. This figure shows<br />

<strong>the</strong> typical, more elongated<br />

morphotype.<br />

Jones (1994), pl. 51, figs. 5–6<br />

Jones (1994), pl. 106, fig. 4<br />

Jones (1994), pl. 106, fig. 5<br />

Jones (1994), pl. 54, figs. 2–3<br />

Den Dulk (2000), pl. 3, fig. 7<br />

Jones (1994), pl. 54, fig. 5<br />

Jones (1994), pl. 54, figs. 6–9<br />

Jones (1994), pl. 55, figs. 12–14,<br />

figs. 17–18<br />

Jones (1994), pl. 55, figs. 15–16,<br />

figs. 19–23<br />

Jones (1994), pl. 92, fig. 10,<br />

pl. 115, figs. 4 <strong>and</strong> 5<br />

Cibicidoides bradyi<br />

(Trauth, 1981)<br />

Cibicidoides kullenbergi<br />

(Parker, 1953)<br />

Cibicidoides<br />

robertsonianus<br />

(Brady, 1881)<br />

Cibicidoides spp.<br />

Cibicidoides wuellerstorfi<br />

(Schwager, 1866)<br />

Dentalina filiformis<br />

(d'Orbigny, 1826)<br />

Dentalina spp.<br />

Ehrenbergina spp.<br />

Epistominella exigua<br />

(Brady, 1884)<br />

Eponides pusillus Parr, 1950<br />

Fissurina spp.<br />

Fursenkoina mexicana<br />

(Cushman, 1922)<br />

Fursenkoina rotundata<br />

(Parr, 1950)<br />

Gavelinopsis lobatulus<br />

(Parr, 1950)<br />

Globobulimina cf. G.<br />

pyrula (d'Orbigny,<br />

1846) Plate 1, fig. 20<br />

Globobulimina sp. 3<br />

Plate 1, fig. 19<br />

Globocassidulina<br />

subglobosa<br />

(Brady, 1881)<br />

Gyroidina altiformis (R.E.<br />

<strong>and</strong> K.C. Stewart, 1930)<br />

Gyroidina orbicularis<br />

(d'Orbigny, 1826)<br />

Gyroidinoides polius<br />

(Phleger <strong>and</strong> Parker,<br />

1951)<br />

Gyroidioides soldanii<br />

(d'Orbigny, 1826)<br />

Hanzawaia boueana<br />

(d'Orbigny, 1846)<br />

Hyalinea balthica<br />

(Schroeter, 1783)<br />

Ioanella tumidula<br />

(Brady, 1884)<br />

Lagena spp.<br />

Laticarinina spp.<br />

Lenticulina calcar<br />

(Linné, 1767)<br />

Lenticulina spp.<br />

Marginulina obesa<br />

(Chushman, 1923)<br />

Den Dulk (2000), pl. 6, fig. 2<br />

Den Dulk (2000), pl. 6, fig. 4<br />

Jones (1994), pl. 95, fig. 4<br />

Jones (1994), pl. 98, figs. 8–9<br />

Jones (1994), pl. 63, figs. 3–5<br />

Schmiedl et al. (1997),<br />

pl. 2, figs. 7–9<br />

Van Leeuwen (1989) Nuttallides<br />

pusillus pusillus, pl. 14,<br />

figs. 10–14<br />

Schmiedl (1995), pl. 2,<br />

figs. 14–15<br />

Jones (1994), pl. 52, figs. 10–11<br />

Den Dulk (2000), pl. 4, fig. 6<br />

Fontanier (2003), pl. 1,<br />

fig. A–E (p. 245),<br />

Maas (2000), G. turgida<br />

(Bailey, 1851), pl. 2, fig. 7<br />

This species shows more whirls,<br />

than Globobulimina cf. G. pyrula<br />

Jones (1994), pl. 54, fig. 17<br />

Fontanier (2003), pl. 7,<br />

figs. I, J, K<br />

Den Dulk (2000), pl. 8, fig. 1<br />

Van Leeuwen (1989), pl. 12,<br />

figs. 10–12<br />

Den Dulk (2000), pl. 8, fig. 2<br />

Abu-Zied (2001), pl. 11, figs. 3–4<br />

Jones (1994), pl. 112, figs. 1–2<br />

Jones (1994), pl. 95, figs. 6–7<br />

Jones (1994), pl. 70, figs. 8–12<br />

Maas (2000), L. articulata<br />

(Terquem, 1862), pl. 2, fig. 10<br />

Jones (1994), pl. 65, figs. 5–6


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

69<br />

Appendix A (continued)<br />

Perforate species<br />

Melonis pompilioides<br />

(Fichtel <strong>and</strong> Moll, 1798)<br />

Melonis za<strong>and</strong>amae<br />

(Van Voorthuysen, 1952)<br />

Neolenticulina variabilis<br />

(Reuss, 1850)<br />

Nonion fabum<br />

(Fichtel <strong>and</strong> Moll, 1798)<br />

Oridorsalis umbonatus<br />

(Reuss, 1851)<br />

Osangularia culter<br />

(Parker <strong>and</strong> Jones, 1865)<br />

Planodiscorbis sp.<br />

Praeglobobulimina<br />

pupoides (d'Orbigny)<br />

Plate 1, fig. 18<br />

Praeglobobulimina sp. 1<br />

Plate 1, fig. 17<br />

Pullenia bulloides<br />

(d'Orbigny, 1846)<br />

Pullenia quinqueloba<br />

(Reuss, 1851)<br />

Reussella spinulosa<br />

(Reuss, 1850)<br />

Robertina chapmani<br />

(Heron-Allen <strong>and</strong><br />

Earl<strong>and</strong>, 1922)<br />

Rosalina vilardeboana<br />

(d'Orbigny, 1839)<br />

Rotaliatinopsis<br />

semiinvoluta<br />

(Germeraad, 1946)<br />

Saracenaria latifrons<br />

(Brady, 1884)<br />

Sphaeroidina bulloides<br />

(Deshayes, 1832)<br />

Spirilina vivipara<br />

(Ehrenberg, 1843)<br />

Uvigerina ex gr. U.<br />

semiornata Plate 1, figs.<br />

1−6<br />

References<br />

Jones (1994), pl. 109, figs. 10–12<br />

Schmiedl et al. (1997), pl. 2,<br />

figs. 12–13<br />

Jones (1994), pl. 68, figs. 11–16<br />

Jones (1994), pl. 109, figs. 12–13<br />

Jones (1994), pl. 95, fig. 11<br />

Schmiedl et al. (1997), pl. 2,<br />

figs. 4–6<br />

Jones (1994), pl. 50, figs. 14–15<br />

Den Dulk (2000), Globobulimina<br />

affinis (d'Orbigny, 1846), pl. 1,<br />

fig. 11<br />

Maas (2000), Protoglobobulimina<br />

pupoides (d'Orbigny, 1846),<br />

pl. 3, fig. 12<br />

Jones (1994), pl. 84, figs. 12–13<br />

Den Dulk (2000), pl. 4, figs. 2–3<br />

Jones (1994), pl. 47, figs. 1–3<br />

Wiesner (1931), pl. 50, fig. 18<br />

Jones (1994), pl. 86, fig. 9<br />

Loeblich <strong>and</strong> Tappan (1988),<br />

pl. 714, figs. 7 <strong>and</strong> 11<br />

Jones (1994), pl. 113, fig. 11<br />

Jones (1994), pl. 85, Figs. 1–5<br />

Jones (1994), pl. 85, figs. 1–4<br />

Test normally triserial, rounded in<br />

outline <strong>and</strong> with a relatively low<br />

length/width ratio. Chambers<br />

strongly inflated <strong>and</strong> overlapping.<br />

The short neck is positioned in a<br />

prominent depression, a tooth is<br />

present in <strong>the</strong> aperture. Test is<br />

ornamented with costae, which<br />

can cross <strong>the</strong> sutures. On <strong>the</strong> last<br />

chambers, <strong>the</strong> costae can miss. Large<br />

pores between <strong>the</strong> costae. Some<br />

specimens become more elongated<br />

<strong>and</strong> in adult specimens <strong>the</strong> last<br />

chambers may become biserial.<br />

Maas (2000), U. ex,<br />

gr. U. semiornata, pl. 2, figs. 1–3<br />

Appendix A (continued )<br />

Perforate species<br />

Uvigerina peregrina<br />

Cushman, 1923 Plate 1,fig.7<br />

Uvigerina probiscidea<br />

Schwager, 1866<br />

Valvulineria araucana<br />

(d'Orbigny, 1839)<br />

Valvulineria minuta<br />

(Schubert, 1904)<br />

Porcellaneous species<br />

Pyrgo spp.<br />

Sigmoilopsis schlumbergeri<br />

(Silvestri, 1904)<br />

Spiroloculina communis<br />

Cushman <strong>and</strong> Todd, 1944<br />

Spiroloculina rotunda<br />

d'Orbigny, 1826<br />

Spirosigmoilina tenuis<br />

(Czjzek 1848)<br />

Triloculina spp.<br />

References<br />

Schmiedl et al. (1997),<br />

pl. 3, fig. 3<br />

Van Morkhoven et al.<br />

(1986), pl. 6<br />

Phleger et al. (1953), pl. 8,<br />

figs. 29, 30<br />

Jones (1994), pl. 91, fig. 4<br />

Jones (1994), pl. 8, figs. 1–4<br />

Jones (1994), pl. 9, figs. 5–6<br />

Jones (1994), pl. 5, figs. 15–16<br />

Jones (1994), pl. 10, figs. 7,8,11<br />

Arenaceous species<br />

Adercotryma glomerata Jones (1994), pl. 34, figs. 15–18<br />

(Brady, 1878)<br />

Ammobaculites filiformis Jones (1994), pl. 32,<br />

(Earl<strong>and</strong>, 1934)<br />

Figs. 21?, 22–23<br />

Ammodiscus sp. 1 (Brady, Maas (2000) A. cretaceus<br />

1881) Plate 1, fig. 21 (Reuss, 1845), pl. 1, fig. 2<br />

Jannink et al. (1998)<br />

Ammodiscus sp., pl. 1, fig. 2<br />

Ammodiscus sp. 2<br />

A single-celled species with<br />

vaulted, coiled, test.<br />

Ammomarginulina foliacea Jones (1994) Eratidus foliaceus,<br />

(Brady, 1881)<br />

pl. 33, figs. 20–25<br />

Cribrostomoides jeffreysii Jones (1994) Veleroninoides<br />

(Williamson, 1858)<br />

jeffreysii, pl. 35, figs. 1–2<br />

Maas (2000) Labrospira sp. A1,<br />

pl. 1, fig. 1<br />

Cribrostomoides cf. C.<br />

Hess (1998), pl. 7, fig. 9<br />

nitidum (Goes, 1896)<br />

Cribrostomoides scitulus Jones (1994) Veleroninoides<br />

(Brady, 1881)<br />

scitulus, pl. 34, figs. 11–13<br />

Cribrostomoides subglobosus Jones (1994), pl. 34, figs. 8–10<br />

(Cushman, 1910)<br />

Cribrostomoides wiesneri Jones (1994) Veleroninoides<br />

(Parr, 1950)<br />

wiesneri, pl. 40, figs. 14–15<br />

Cyclammina cancellata Jones (1994), pl. 37, figs. 8–16<br />

(Brady, 1879)<br />

Deuterammina grahami Wollenburg <strong>and</strong> Mackensen<br />

Brönnimann <strong>and</strong><br />

(1998), pl. 2, figs. 6–8<br />

Whittaker, 1988<br />

Deuterammina spp.<br />

Egerella bradyi (Cushman, 1911) Jones (1994), pl. 47, figs. 4–7<br />

Eggerella sp. 1<br />

A triserial test, with a high length/<br />

width ratio. Chambers are less<br />

inflated than by E. bradyi. The<br />

test material is coarse grained.<br />

(continued on next page)


70 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

Appendix A (continued)<br />

Perforate species<br />

Eggerelloides scabra<br />

(Williamson, 1858)<br />

Globotextularia anceps<br />

(Bradyi, 1884)<br />

Glomospira charoides<br />

(Jones <strong>and</strong> Parker, 1860)<br />

Glomospira gordialis<br />

(Jones <strong>and</strong> Parker, 1860)<br />

Haplophragmoides<br />

sphaeriloculus<br />

(Cushman, 1910)<br />

Karrerulina conversa<br />

(Grzybowski, 1901)<br />

Karrerulina sp. 1<br />

Lagenammina apulacea<br />

(Brady, 1881)<br />

Lagenammina difflugiformis<br />

(Brady, 1879)<br />

Martinotiella communis<br />

(d'Orbigny, 1846)<br />

Paratrochammina challengeri<br />

Brönnimann <strong>and</strong><br />

Whittaker, 1988<br />

Portatrochammina bipolaris<br />

Brönnimann <strong>and</strong><br />

Whittaker, 1988<br />

Recurvoides contortus<br />

Earl<strong>and</strong>, 1934<br />

Recurvoides turbinatus<br />

(Brady, 1881)<br />

Reophax bilocularis Flint,<br />

1899 Plate 1, fig. 22<br />

Reophax dentaliniformis<br />

Brady, 1881<br />

Reophax fusiformis<br />

(Williamson, 1858)<br />

Reophax gracilis<br />

(Kiaer, 1900)<br />

Reophax guttifera Brady, 1881<br />

Reophax micaceus Earl<strong>and</strong>,<br />

1934<br />

Reophax mortenseni<br />

Hogker, 1972<br />

Reophax pilulifera<br />

Brady, 1884<br />

Reophax scorpiurus<br />

Montfort, 1808<br />

Plate 1, fig. 23<br />

Reophax spp.<br />

Saccammina sphaerica<br />

Brady, 1871<br />

Siphotextularia spp.<br />

Textularia spp.<br />

Trochammina globulosa<br />

Cushman, 1920<br />

References<br />

Jones (1994) pl. 47, figs. 15–17<br />

Jones (1994) pl. 35, figs. 12–15<br />

Jones (1994) Usbekistania<br />

charoides, pl. 38, figs. 10–16<br />

Jones (1994), pl. 38, figs. 7–8<br />

Schröder (1986), pl. 18, figs. 5–7<br />

Jones (1994), pl. 46, figs. 17–19<br />

Jones (1994), pl. 30, fig. 6<br />

Mackensen et al. (1990), pl. 6,<br />

fig. 9<br />

Jones (1994), pl. 48,<br />

figs. 1–2, 4–8, 3?<br />

Brönnimann <strong>and</strong> Whittaker (1988),<br />

pl. 16, figs. h–k<br />

Brönnimann <strong>and</strong> Whittaker (1988),<br />

pl. 20–31<br />

Loeblich <strong>and</strong> Tappan (1988),<br />

pl. 68, figs. 7–14<br />

Jones (1994), pl. 35, fig. 9<br />

Schmiedl et al. (1997),<br />

pl. 1, figs. 3–4<br />

Jones (1994) pl. 30, figs. 21–22<br />

Jones (1994), pl. 30, figs. 7–10<br />

Maas (2000) Nouria<br />

polymorphoides Heron-Allen <strong>and</strong><br />

Earl<strong>and</strong>, 1914, pl. 1, fig. 3<br />

Hess (1998), pl. 3, fig. 12<br />

Jones (1994) Hormosinella<br />

guttifera, pl. 31, figs. 10–15<br />

Timm (1992), pl. 2, fig. 6<br />

Jones (1994), pl. 31, figs. 3–4<br />

Jones (1994) Hormosina<br />

pilulifera, pl. 30, figs. 18–20<br />

Schröder (1986), pl. 14,<br />

figs. 1–5, pl. 23<br />

Jones (1994), pl. 18,<br />

figs. 11–15, ?17<br />

Schröder (1986), pl. 19,<br />

figs. 9–11<br />

Appendix A (continued)<br />

Perforate species<br />

Trochammina spp.<br />

Verneuilinulla propinqua<br />

(Brady, 1848)<br />

References<br />

References<br />

Jones (1994), pl. 47, figs. 8–12<br />

Abu-Zied, 2001. High resolution LGM—present paleoceanography of<br />

<strong>the</strong> NE Mediterranean: a <strong>benthic</strong> perspective. PhD Thesis, University<br />

of Southampton, 195 pp.<br />

Agnihotri, R., Sarin, M.M., Somayalulu, B.L.K., Jull, A.J.T., Burr,<br />

G.S., 2003. Late-Quaternary biogenic productivity <strong>and</strong> organic<br />

carbon deposition in <strong>the</strong> eastern Arabian Sea. Palaeogeography,<br />

Palaeoclimatology, Palaeoecology 197, 43–60.<br />

Antoine, D., André, J.-M., Morel, André, 1996. Oceanic primary production.<br />

2. Estimation at global scale <strong>from</strong> satellite (Coastal Zone<br />

Color Scanner) chlorophyll. Global Biogeochemical Cycles 10,<br />

57–69.<br />

Alve, E., 1994. Opportunistic features of <strong>the</strong> <strong>foraminifera</strong> Stainforthia<br />

fusiformis (Williamson): evidence <strong>from</strong> Frierfjord, Norway. Journal<br />

of Micropaleontology 13, 24.<br />

Alve, E., 1995. Benthic <strong>foraminifera</strong>l distribution <strong>and</strong> recolonization<br />

of formerly anoxic environments in Drammensfjord, sou<strong>the</strong>rn<br />

Norway. Marine Geology 25, 169–286.<br />

Barmawidjaja, D.M., Jorissen, F.J., Piskaric, S., Van der Zwaan, G.J.,<br />

1992. Microhabitat selection by <strong>benthic</strong> <strong>foraminifera</strong> in <strong>the</strong> nor<strong>the</strong>rn<br />

Adriatic Sea. Journal of Foraminiferal Research 22, 297–317.<br />

Banse, K., McClain, C.R., 1986. Winter blooms of phytoplankton in<br />

<strong>the</strong> Arabian Sea as observed by <strong>the</strong> Coastal Zone Colour Scanner.<br />

Marine Ecology Progress Series 34, 201–211.<br />

Bernhard, J.M., 1988. Postmortem vital staining in <strong>benthic</strong> <strong>foraminifera</strong>:<br />

duration <strong>and</strong> importance in population <strong>and</strong> distributional studies.<br />

Journal of Foraminiferal Research 18, 143–146.<br />

Bernhard, J.M., 1992. Benthic <strong>foraminifera</strong>l distribution <strong>and</strong> biomass<br />

related to porewater oxygen content: central California continental<br />

slope <strong>and</strong> rise. Deep-Sea Research 39, 585–605.<br />

Bernhard, J.M., 1993. Experimental <strong>and</strong> field evidence of Antarctic<br />

<strong>foraminifera</strong>l tolerance to anoxia <strong>and</strong> hydrogen sulfide. Marine<br />

Micropaleontology 20, 203–213.<br />

Bernhard, J.M., Sen Gupta, B.K., 1999. Foraminifera of oxygendepleted<br />

environments. In: Sen Gupta, B.K. (Ed.), Modern Foraminifera.<br />

Kluwer Academic Publisher, Dordrecht, pp. 201–216.<br />

Bernhard, J.M., Sen Gupta, B.K., Borne, P.F., 1997. Benthic <strong>foraminifera</strong>l<br />

proxy to estimate dysoxic bottom-water oxygen concentrations:<br />

Santa Barbara basin, U.S. Pacific continental margin.<br />

Journal of Foraminiferal Research 27, 301–310.<br />

Bett, B.J., 2004a. RRS Charles Darwin Cruise 145 12 Mar–09 Apr<br />

2003. Benthic ecology <strong>and</strong> biogeochemistry of <strong>the</strong> Pakistan margin.<br />

Southampton Oceanography Centre Cruise report, No. 50, 161 pp.<br />

Bett, B.J., 2004b. RRS Charles Darwin Cruise 150, 22 Aug–15 Sept<br />

2003. Benthic ecology <strong>and</strong> biogeochemistry of <strong>the</strong> Pakistan margin.<br />

Southampton Oceanography Centre Cruise report, No. 51,<br />

143 pp.<br />

Brönnimann, P., Whittaker, J.E., 1988. The Trochamminacea of <strong>the</strong><br />

Discovery Reports. British Museum (Natural History), London.<br />

152 pp.<br />

Burkill, P.H., Mantoura, R.F.C., Owens, N.J.P., 1993. Biogeochemical<br />

cycling in <strong>the</strong> northwestern Indian Ocean: a brief overview. Deep-<br />

Sea Research II 40, 643–649.


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

71<br />

Caralp, M.H., 1984. Impact de la matière organique dans des zones de<br />

forte productivité sur certains foraminifères benthiques. Oceanologica<br />

Acta 7, 509–515.<br />

Caralp, M.H., 1989. Abundance of Bulimina exilis <strong>and</strong> Melonis barleeanum:<br />

relationship to <strong>the</strong> quality of marine organic matter. Geomarine<br />

Letters 9, 37–43.<br />

Caron, D.A., Dennett, M.R., 1999. Phytoplankton growth <strong>and</strong> mortality<br />

during <strong>the</strong> 1995 nor<strong>the</strong>ast monsoon <strong>and</strong> spring intermonsoon<br />

in <strong>the</strong> Arabian Sea. Deep-Sea Research II 46, 1665–1690.<br />

Casford, J.S.L., Rohling, E.J., Abu-Zied, R.H., Fontanier, C., Jorissen,<br />

F.J., Leng, M.J., Schmiedl, G., Thomson, J., 2003. A dynamic<br />

concept for <strong>the</strong> eastern Mediterranean circulation <strong>and</strong> oxygenation<br />

during sapropel formation. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology 190, 103–119.<br />

Corliss, B.H., 1979. Taxonomy of recent deep-sea benthonic <strong>foraminifera</strong><br />

<strong>from</strong> <strong>the</strong> sou<strong>the</strong>ast Indian Ocean. Micropaleontology 25,<br />

1–19.<br />

Corliss, B.H., Emmerson, S., 1990. Distribution of <strong>Rose</strong> <strong>Bengal</strong> <strong>stained</strong><br />

deep-sea <strong>benthic</strong> <strong>foraminifera</strong> <strong>from</strong> <strong>the</strong> Nova Scotian continental<br />

margin <strong>and</strong> Gulf of Maine. Deep-Sea Research 37, 381–400.<br />

Cowie, G., 2003. RRS “Charles Darwin” Cruise RRS Charles Darwin<br />

Cruise 145 12 Mar–09 Apr 2003. Arabian Sea <strong>benthic</strong> processes<br />

study. University of Edinburgh, Cruise report, 133 pp.<br />

Den Dulk, M., 2000. Benthic <strong>foraminifera</strong>l response to Late Quaternary<br />

variations in surface water productivity <strong>and</strong> oxygenation in <strong>the</strong><br />

nor<strong>the</strong>rn Arabian Sea. Geologica Ultraiectina 188, 1–205.<br />

Den Dulk, M., Reichart, G.J., Van Heyst, S., Zachariasse, W.J., Van<br />

der Zwaan, G.J., 2000. Benthic <strong>foraminifera</strong> as proxies of organic<br />

matter flux <strong>and</strong> bottom water oxygenation? A case history <strong>from</strong><br />

<strong>the</strong> nor<strong>the</strong>rn Arabian Sea. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology 161, 337–359.<br />

De Rijk, S., Jorissen, F.J., Rohling, E.J., Troelstra, S.R., 2000. Organic<br />

flux control on bathymetric zonation of Mediterranean <strong>benthic</strong><br />

<strong>foraminifera</strong>. Marine Micropaleontology 40, 151–166.<br />

De Stigter, H.C., Jorissen, F.J., Van der Zwaan, G.J., 1998. Bathymetric<br />

distribution <strong>and</strong> microhabitat partitioning of live (<strong>Rose</strong><br />

<strong>Bengal</strong> <strong>stained</strong>) <strong>benthic</strong> <strong>foraminifera</strong> along a shelf to bathyal<br />

transect in <strong>the</strong> sou<strong>the</strong>rn Adriatic Sea. Journal of Foraminiferal<br />

Research 28, 40–65.<br />

Fisher, R.A., Corbet, A.S., Williams, C.B., 1943. The relation between<br />

<strong>the</strong> number of species <strong>and</strong> <strong>the</strong> number of individuals in a r<strong>and</strong>om<br />

sample of an animal population. Journal of Animal Ecology 12,<br />

4258.<br />

Fontanier, C., 2003. Écologie des foraminifères benthiques du Golf<br />

de Gascogne: Études de la variabilité spatiale et temporelle des<br />

faunes de foraminifères benthiques et de la composition isotopique<br />

(δ 18 O, δ 13 C) de leurs tests. PhD Thesis, University<br />

Bordeaux, 471 pp.<br />

Fontanier, C., Jorissen, F.J., Chaillou, G., David, C., Anschutz, P.,<br />

Lafon, V., 2003. Seasonal <strong>and</strong> interannual variability of <strong>benthic</strong><br />

<strong>foraminifera</strong>l faunas at 550 m depth in <strong>the</strong> Bay of Biscay. Deep-Sea<br />

Research I 50, 457–494.<br />

Fontanier, C., Jorissen, F.J., Licari, L., Alex<strong>and</strong>re, A., Anschutz, P.,<br />

Carbonel, P., 2002. <strong>Live</strong> <strong>benthic</strong> <strong>foraminifera</strong>l faunas <strong>from</strong> <strong>the</strong> Bay<br />

of Biscay: faunal density, composition, <strong>and</strong> microhabitats. Deep-<br />

Sea Research II 49, 751–785.<br />

Gooday, A.J., 2003. Benthic <strong>foraminifera</strong> (Protista) as tools in<br />

deep-water palaeoceanography: a review of environmental<br />

influences on faunal characteristics. Advances in Marine<br />

Biology 46, 1–90.<br />

Gooday, A.J., Hughes, J.A., 2002. Foraminifera associated with<br />

phytodetritus deposits at a bathyal site in <strong>the</strong> nor<strong>the</strong>rn Rockall<br />

Trough (NE Atlantic): seasonal contrasts <strong>and</strong> a comparison of <strong>stained</strong><br />

<strong>and</strong> <strong>dead</strong> assemblages. Marine Micropaleontology 46, 83–110.<br />

Gooday, A.J., Bernhard, J.M., Levin, L.A., Suhr, S.B., 2000. Foraminifera<br />

in <strong>the</strong> Arabian Sea oxygen minimum zone <strong>and</strong> o<strong>the</strong>r oxygendeficient<br />

settings: taxonomic composition, diversity, <strong>and</strong> relation<br />

to metazoan faunas. Deep-Sea Research II 47, 24–54.<br />

Heinz, P., Hemleben, C., 2003. Regional <strong>and</strong> seasonal variations of<br />

recent <strong>benthic</strong> deep-sea <strong>foraminifera</strong> in <strong>the</strong> Arabian Sea. Deep-Sea<br />

Research I 50, 435–447.<br />

Helly, J., Levin, L.A., 2004. Global distribution of naturally occurring<br />

marine hypoxia on continental margins. Deep Sea Research I 51,<br />

1159–1168.<br />

Herguera, J.C., Berger, W.H., 1991. Paleoproductivity <strong>from</strong> <strong>benthic</strong><br />

<strong>foraminifera</strong> abundance: glacial to postglacial change in <strong>the</strong> westequatorial<br />

Pacific. Geology 19, 1173–1176.<br />

Hermelin, J.O.R., Shimmield, G.B., 1990. The importance of <strong>the</strong><br />

oxygen minimum zone <strong>and</strong> sediment geochemistry in <strong>the</strong> distribution<br />

of Recent <strong>benthic</strong> <strong>foraminifera</strong> in <strong>the</strong> northwest Indian<br />

Ocean. Marine Geology 91, 1–29.<br />

Hess, S., 1998. Verbreitungsmuster rezenter benthischer Foraminiferen im<br />

Südchinesischen Meer. (Distribution patterns of recent <strong>benthic</strong> <strong>foraminifera</strong><br />

in <strong>the</strong> Sou<strong>the</strong>rn China Sea). Berichte — Reports. Geologisches<br />

und Paläontologisches Institut der Universität Kiel 59, 1–173.<br />

Igarashi, A., Numanami, H., Tsuchiya, Y., Fukuchi, M., 2001. Bathymetric<br />

distribution of fossil <strong>foraminifera</strong> within marine sediment<br />

cores <strong>from</strong> <strong>the</strong> eastern part of Lützow–Holm Bay, East Antarctica,<br />

<strong>and</strong> its paleoceanographic implications. Marine Micropaleontology<br />

42, 125–162.<br />

Jannink, N.T., Zachariasse, W.J., Van der Zwaan, G.J., 1998. Living<br />

(<strong>Rose</strong> <strong>Bengal</strong> <strong>stained</strong>) <strong>benthic</strong> <strong>foraminifera</strong> <strong>from</strong> <strong>the</strong> Pakistan<br />

continental margin (North Arabian Sea). Deep-Sea Research I 45,<br />

1283–1513.<br />

Jonasson, K.E., Schröder-Adams, C.J., Patterson, R.T., 1995. Benthic<br />

<strong>foraminifera</strong>l distribution at Middle Valley, Juan de Fuca Ridge, a<br />

nor<strong>the</strong>ast Pacific hydro<strong>the</strong>rmal venting site. Marine Micropaleontology<br />

25, 151–167.<br />

Jones, R.W., 1994. The Challenger Foraminifera. Oxford University<br />

Press, Oxford, New York. Tokyo, 149 pp.<br />

Jorissen, F.J., 1999. Benthic <strong>foraminifera</strong>l succession across Late Quaternary<br />

Mediterranean sapropels. Marine Geology 153, 91–101.<br />

Jorissen, F.J., Wittling, I., 1999. Ecological evidence <strong>from</strong> live-<strong>dead</strong><br />

comparison of <strong>benthic</strong> <strong>foraminifera</strong>l faunas off Cape Blanc (Northwest<br />

Africa). Palaeogeography, Palaeoclimatology, Palaeoecology<br />

149, 151–170.<br />

Jorissen, F.J., De Stigter, H.C., Widmark, J.G.V., 1995. A conceptual<br />

model explaining <strong>benthic</strong> <strong>foraminifera</strong>l microhabitats. Marine<br />

Micropaleontology 26, 3–15.<br />

Jorissen, F.J., Wittling, I., Peypouquet, J.P., Raboulle, C., Relexans,<br />

J.C., 1998. <strong>Live</strong> <strong>benthic</strong> <strong>foraminifera</strong>l faunas off Cape Blance,<br />

NW-Africa: community structure <strong>and</strong> microhabitates. Deep-Sea<br />

Research I 45, 2157–2188.<br />

Kaiho, K., 1994. Benthic <strong>foraminifera</strong>l dissolved-oxygen index <strong>and</strong><br />

dissolved-oxygen levels in <strong>the</strong> modern ocean. Geology 22,<br />

719–722.<br />

Kitazato, H., 1994. Foraminiferal microhabitats in four marine environments<br />

around Japan. Marine Micropaleontology 24, 29–41.<br />

Kitazato, H., Shirayama, Y., Nakatsuka, T., Jujiwara, S., Shimanaga,<br />

M., Kato, Y., Okada, Y., K<strong>and</strong>a, J., Yamaoka, A., Masuzawa, T.,<br />

Suzuki, K., 2000. Seasonal phytodetritus deposition <strong>and</strong> responses<br />

of bathyal <strong>benthic</strong> <strong>foraminifera</strong>l populations in Sagami Bay, Japan:<br />

preliminary results <strong>from</strong> “Project Sagami 1996–1999”. Marine<br />

Micropaleontology 40, 135–149.


72 S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

Kurbjeweit, F., Schmiedl, G., Schiebel, R., Hemleben, Ch., Pfannkuche,<br />

O., Wallmann, K., Schäfer, P., 2000. Distribution, biomass <strong>and</strong><br />

diversity of <strong>benthic</strong> <strong>foraminifera</strong> in relation to sediment geochemistry<br />

in <strong>the</strong> Arabian Sea. Deep-Sea Research II 47, 2913–2955.<br />

Larkin, K.E., Gooday, A.J., Levin, L.A., 2003. Metazoan <strong>and</strong> protozoan<br />

assemblages. In: Cowie, G. (Ed.), RRS “Charles Darwin”<br />

Cruise 146: 12 April–30 May 2003. Benthic Ecology <strong>and</strong> Biogeochemistry<br />

of <strong>the</strong> Pakistan Margin. University of Edinburgh,<br />

Cruise report, pp. 64–71.<br />

Le Calvez, Y., 1995. Recent Travaux de l'Institut des Peches Maritimes,<br />

vol. 23, 263 pp.<br />

Levin, L.A., 2003a. Oxygen minimum zone benthos: adaptation <strong>and</strong><br />

community response to hypoxia. Oceanography <strong>and</strong> Marine Biology:<br />

an Annual Review 2003 41, 1–45.<br />

Levin, L.A., 2003b. Sediments — visual observations, photography, X-<br />

radiography. In: Cowie, G. (Ed.), RRS “Charles Darwin” Cruise 146:<br />

12 April–30 May 2003. Benthic Ecology <strong>and</strong> Biogeochemistry of <strong>the</strong><br />

Pakistan Margin. University of Edinburgh, Cruise report, pp. 53–59.<br />

Leuschner, D.C., Sirocko, F., 2000. The low-latitude monsoon climate<br />

during Dansgaard–Oeschger cycles <strong>and</strong> Heinrich events. Quaternary<br />

Science Reviews 19, 243–254.<br />

Leuschner, D.C., Sirocko, F., 2003. Orbital insolation forcing of <strong>the</strong><br />

Indian Monsoon — a motor for global climate changes? Palaeogeography,<br />

Palaeoclimatology, Palaeoecology 197, 83–95.<br />

Licari, L.N., Schumacher, S., Wenzhöfer, F., Zabel, M., Mackensen, A.,<br />

2003. Communities <strong>and</strong> microhabitats of living <strong>benthic</strong> <strong>foraminifera</strong><br />

<strong>from</strong> <strong>the</strong> tropical east Atlantic: impact of different productivity<br />

regimes. Journal of Foraminiferal Research 33, 10–31.<br />

Loeblich Jr., A.R., Tappan, H., 1988. Foraminiferal Genera <strong>and</strong> Their<br />

Classification. Van Nostr<strong>and</strong> Rheinhold Company, New York. 970 pp.<br />

Longinelli, A., 1956. Foraminiferi del Calabriano e Piacebziano di<br />

Rosignano Marittimo e della Val di Cecina. Paleontographica<br />

Italica 49, 1–159.<br />

Loubere, P., 1989. Bioturbation <strong>and</strong> sedimentation rate control of<br />

<strong>benthic</strong> microfossil taxon abundances in surface sediments: a<br />

<strong>the</strong>oretical approach to <strong>the</strong> analysis of species microhabitats. Marine<br />

Micropaleontology 14, 317–325.<br />

Loubere, P., 1994. Quantitative estimation of surface ocean productivity<br />

<strong>and</strong> bottom water oxygen concentration using <strong>benthic</strong> <strong>foraminifera</strong>.<br />

Paleoceanography 9, 723–737.<br />

Loubere, P., 1999. A multiproxy reconstruction of biological productivity<br />

<strong>and</strong> oceanography in <strong>the</strong> eastern equatorial Pacific for <strong>the</strong><br />

past 30,000 years. Marine Micropaleontology 37, 173–198.<br />

Loubere, P., Fariduddin, M., Murray, R.W., 2003. Patterns of export<br />

production in <strong>the</strong> eastern equatorial Pacific over <strong>the</strong> past<br />

130,000 years. Paleoceanography 18, 6-1–6-21.<br />

Loubere, P., Meyers, P., Gary, A., 1995. Benthic <strong>foraminifera</strong>l microhabitat<br />

selection, carbon isotope values, <strong>and</strong> association with larger<br />

animals: a test with Uvigerina peregrina. Journal of Foraminiferal<br />

Research 25, 83–95.<br />

Lutze, G.F., 1986. Uvigerina species of <strong>the</strong> eastern North Atlantic. In:<br />

Van der Zwaan, G.J., Jorissen, F.J., Verhalten, P.J.J.M., von Daniels,<br />

C.H. (Eds.), Atlantic–European Oligocene to Recent Uvigerina.<br />

Utrecht Micropaleontological Bulletins, vol. 35, pp. 21–46.<br />

Maas, M., 2000. Verbreitung lebendgefärbter benthischer Foraminiferen<br />

in einer intensivierten Sauerstoffminimumzone, Indo-Pakistanischer<br />

Kontinentalr<strong>and</strong>, nördliches Arabisches Meer (Distribution<br />

of <strong>Rose</strong> <strong>Bengal</strong> <strong>stained</strong> <strong>benthic</strong> <strong>foraminifera</strong> within an intensified<br />

oxygen minimum zone, Indo-Pakistan Continental Margin, Northwest<br />

Arabian Sea). Meyniana 52, 101–128.<br />

Mackensen, A., Douglas, R.G., 1989. Down-core distribution of live<br />

<strong>and</strong> <strong>dead</strong> deep-water <strong>benthic</strong> <strong>foraminifera</strong> in box cores <strong>from</strong> <strong>the</strong><br />

Weddell Sea <strong>and</strong> <strong>the</strong> California continental borderl<strong>and</strong>. Deep-Sea<br />

Research 36, 879–900.<br />

Mackensen, A., Fütterer, D.K., Grobe, H., Schmiedl, G., 1993. Benthic<br />

<strong>foraminifera</strong>l assemblages <strong>from</strong> <strong>the</strong> eastern South Atlantic Polar<br />

Front region between 35° <strong>and</strong> 57°S: distribution, ecology <strong>and</strong><br />

fossilization potential. Marine Micropaleontology 22, 33–69.<br />

Mackensen, A., Grobe, H., Kuhn, G., Fütterer, D.K., 1990. Benthic<br />

<strong>foraminifera</strong>l assemblages <strong>from</strong> <strong>the</strong> eastern Weddell Sea between<br />

68° <strong>and</strong> 73°S: distribution, ecology <strong>and</strong> fossilization potential.<br />

Marine Micropaleontology 16, 241–283.<br />

Mackensen, A., Schmiedl, G., Harloff, J., Giese, M., 1995. Deep-sea<br />

<strong>foraminifera</strong> in <strong>the</strong> Sou<strong>the</strong>rn Atlantic Ocean: ecology <strong>and</strong> assemblage<br />

generation. Micropaleontology 41, 342–358.<br />

Madhupratap, M., Prasanna Kumar, S., Bhattahiri, P.M.A., Dileep<br />

Kumar, M., Raghukumar, S., Nair, K.K.C., Ramaiah, N., 1996.<br />

Mechanisms of <strong>the</strong> biological response to winter cooling in <strong>the</strong><br />

nor<strong>the</strong>astern Arabian Sea. Nature 483, 549–552.<br />

Murray, J., 1991. Ecology <strong>and</strong> Paleoecology of Benthic Foraminifera.<br />

Longmann. 398 pp.<br />

Olson, D.B., Hitchcock, G.L., Fine, R.A., Warren, B.A., 1993. Maintenance<br />

of <strong>the</strong> low-oxygen layer in <strong>the</strong> central Arabian Sea. Deep-<br />

Sea Research II 40, 673–685.<br />

Phleger, F.B., Soutar, A., 1973. Production of <strong>benthic</strong> <strong>foraminifera</strong><br />

in three east Pacific oxygen minima. Micropaleontology 19,<br />

110–115.<br />

Phleger, F.B., Parker, F.L., Peirson, J.F., 1953. North Atlantic <strong>foraminifera</strong>.<br />

Reports of <strong>the</strong> Swedish Deep-Sea Expedition 7, sediment<br />

cores <strong>from</strong> <strong>the</strong> North Atlantic Ocean 1.<br />

Prasanna Kumar, S., Prasa, T.G., 1999. Formation <strong>and</strong> spreading of<br />

Arabian Sea high salinity water mass. Journal of Geophysical<br />

Research 104, 1255–1464.<br />

Qasim, S.Z., 1977. Biological productivity of <strong>the</strong> Indian Ocean. Indian<br />

Journal of Marine Science 6, 12–137.<br />

Rathburn, A.E., Levin, L.A., Held, Z., Lohmann, K.C., 2000. Benthic<br />

<strong>foraminifera</strong> associated with cold methane seeps on <strong>the</strong> nor<strong>the</strong>rn<br />

California margin: ecology <strong>and</strong> stable isotopic composition. Marine<br />

Micropaleontology 38, 247–266.<br />

Reichart, G.J., Den Dulk, M., Visser, H.J., Van der Weiden, C.H.,<br />

Zachariasse, W.J., 1997. A 225 kyr record of dust supply, paleoproductivity<br />

<strong>and</strong> <strong>the</strong> oxygen minimum zone <strong>from</strong> Murray Ridge<br />

(nor<strong>the</strong>rn Arabian Sea). Palaeogeography, Palaeoclimatology,<br />

Palaeoecology 134, 149–169.<br />

Reichart, G.J., Lourens, L.J., Zachariasse, W.J., 1998. Temporal variability<br />

in <strong>the</strong> nor<strong>the</strong>rn Arabian Sea oxygen minimum zone (OMZ)<br />

during <strong>the</strong> last 225,000 years. Paleoceanography 13, 607–621.<br />

Reichart, G.J., Schenau, S.J., de Lange, G.J., Zachariasse, W.J., 2002.<br />

Synchroneity of oxygen minimum zone intensity on <strong>the</strong> Oman <strong>and</strong><br />

Pakistan Margins at sub-Milankovitch time scales. Marine<br />

Geology 185, 403–415.<br />

Reuss, A.E., 1850. Neue Foraminiferen aus den Schichten des<br />

Österreischischen Tertiärbeckens. K. Akad. Wiss. Wien, Math.-<br />

Nat. Cl., Denkschr., 18, Bd. 1.<br />

Rixen, T., Ittekkot, V., Haake-Gaye, B., Schäfer, P., 2000. The influence<br />

of <strong>the</strong> SW monsoon on <strong>the</strong> deep-sea organic carbon cycle<br />

in <strong>the</strong> Holocene. Deep-Sea Research II 47, 2629–2651.<br />

Rogers, A.D., 2000. The role of oxygen minima in generating biodiversity<br />

in <strong>the</strong> deep sea. Deep-Sea Research II 47, 119–148.<br />

Ry<strong>the</strong>r, J.H., Menzel, D.W., 1965. On <strong>the</strong> production, composition <strong>and</strong><br />

distribution of organic matter in <strong>the</strong> western Arabian Sea. Deep-<br />

Sea Research I 12, 199–209.<br />

Schiebel, R., 1992. Rezente benthische Foraminiferen in Sedimenten<br />

des Schelfes und oberen Kontinentalhanges im Golf von Guinea


S. Schumacher et al. / Marine Micropaleontology 62 (2007) 45–73<br />

73<br />

(Westafika) (Recent <strong>benthic</strong> <strong>foraminifera</strong> in sediments of <strong>the</strong> shelf<br />

<strong>and</strong> upper continental slope <strong>from</strong> <strong>the</strong> Gulf of Guinea (West<br />

Afrika)). Berichte — Reports Geologisches und Paläontologisches<br />

Institut der Universität Kiel, vol. 59, pp. 1–179.<br />

Schmiedl, G., 1995. Rekonstruktion der spätquartären Tiefenwasserzirkulation<br />

und Produktivität im östlichen Südatlantik anh<strong>and</strong><br />

benthischer Foraminiferen (Late Quarternary <strong>benthic</strong> <strong>foraminifera</strong>l<br />

assemblages <strong>from</strong> <strong>the</strong> eastern South Atlantic Ocean: reconstruction<br />

of deep water circulation <strong>and</strong> productivity changes). Berichte zur<br />

Polarforschung 160, 1–207.<br />

Schmiedl, G., Leuschner, D.C., 2005. Oxygenation changes in <strong>the</strong><br />

deep western Arabian Sea during <strong>the</strong> last 190,000 years: productivity<br />

versus deepwater circulation. Paleoceanography 20,<br />

doi:10.1029/2004PA001044 PA2008.<br />

Schmiedl, G., Mackensen, A., Müller, P.J., 1997. Recent <strong>benthic</strong><br />

<strong>foraminifera</strong> <strong>from</strong> <strong>the</strong> eastern South Atlantic Ocean: dependence<br />

on food supply <strong>and</strong> water masses. Marine Micropaleontology 32,<br />

249–287.<br />

Schmiedl, G., Mitschele, A., Beck, S., Emeis, K.-C., Hemleben, C.,<br />

Schulz, H., Sperling, M., Weldeab, S., 2003. Benthic <strong>foraminifera</strong>l<br />

record of ecosystem variability in <strong>the</strong> eastern Mediterranean Sea<br />

during times of sapropel S5 <strong>and</strong> S6 deposition. Palaeogeography,<br />

Palaeoclimatology, Palaeoecology 190, 139–164.<br />

Schmield, G., Pfeilstricker, M., Hemleben, C., Mackensen, A., 2004.<br />

Environmental <strong>and</strong> biological effects on <strong>the</strong> stable isotope composition<br />

of recent deep-sea <strong>benthic</strong> <strong>foraminifera</strong> <strong>from</strong> <strong>the</strong> western<br />

Mediterranean Sea. Marine Micropaleontology 51, 129–152.<br />

Schott, F.A., McCreary Jr., J.P., 2001. The monsoon circulation of <strong>the</strong><br />

Indian Ocean. Progress in Oceanography 51, 1–123.<br />

Schröder, C.J., 1986. Deep-water arenaceous <strong>foraminifera</strong> in <strong>the</strong><br />

northwest Atlantic Ocean. Canadian Technical Report of Hydrography<br />

<strong>and</strong> Ocean Sciences 71, 1–188.<br />

Schulte, S., Rostek, F., Bard, E., Rullköter, J., Marchal, O., 1999. Variations<br />

of oxygen-minimum <strong>and</strong> primary productivity record in sediments of<br />

<strong>the</strong> Arabian Sea. Earth <strong>and</strong> Planetary Science Letters 173, 205–221.<br />

Sen Gupta, B.K., Machain-Castillo, M.L., 1993. Benthic <strong>foraminifera</strong><br />

in oxygen-poor habitats. Marine Micropaleontology 20, 18–201.<br />

Shannon, C.E., 1948. A ma<strong>the</strong>matical <strong>the</strong>ory of communication. Bell<br />

System Technical Journal 27, 379–423.<br />

Shetye, S.R., Gouveia, A.D., Shenoi, S.S.C., 1994. Circulation <strong>and</strong><br />

water masses of <strong>the</strong> Arabian Sea. In: Lal, D. (Ed.), Biogeochemistry<br />

of <strong>the</strong> Arabian Sea. Indian Academy of Sciences, Bangalore,<br />

pp. 9–25. 560 080, India.<br />

Swallow, J.C., 1984. Some aspects of <strong>the</strong> physical oceanography of <strong>the</strong><br />

Indian Ocean. Deep-Sea Research 31, 639–650.<br />

Timm, S., 1992. Rezente Tiefsee-Benthosforaminiferen aus Oberflächensedimenten<br />

des Golfes von Guinea (Westafrika) — Taxonomie,<br />

Verbreitung, Ökologie und Korngrößenfraktion. Berichte —<br />

Reports, Geologisches und Paläontologisches Institut der Universität<br />

Kiel 59, 1–192.<br />

Van Aken, H.M., Ridderinkhof, H., de Ruijter, W.P.M., 2004. North<br />

Atlantic deep water in <strong>the</strong> south-western Indian Ocean. Deep-Sea<br />

Research. Part 1. Oceanographic Research Papers 51, 755–776.<br />

Van Leeuwen, R.J.W., 1989. Sea-floor distribution <strong>and</strong> Late Quarternary<br />

faunal patterns of planktonic <strong>and</strong> <strong>benthic</strong> foraminifers in <strong>the</strong><br />

Angola Basin. Utrecht Micropleontological Bulletins 38, 1–288.<br />

Van der Zwaan, G.J., Jorissen, F.J., Verhallen, P.J.J.M., von Daniels,<br />

C.H., 1986. Uvigerina <strong>from</strong> <strong>the</strong> Eastern Atlantic, North Sea<br />

Basin, Paratethys <strong>and</strong> Mediterranean. In: Van der Zwaan, G.J.,<br />

Jorissen, F.J., Verhallen, P.J.J.M., von Daniels, C.H. (Eds.), Atlantic–European<br />

Oligocene to Recent Uvigerina. Utrecht Micropleontological<br />

Bulletins, vol. 35, pp. 7–19.<br />

Van Morkhoven, F.P.C.M., Berggren, W.A., Edwards, A.S., 1986.<br />

Cenozoic cosmopolitan deep-water <strong>benthic</strong> <strong>foraminifera</strong>. Bulletin<br />

des Centres de Recherches Exploration–Production Elf-Aquitain<br />

11 421 pp.<br />

Von Daniels, C.H., 1970. Quantitative ökologiosche Analyse der<br />

zeitlichen und räumlichen Verbreitung rezenter Foraminiferen im<br />

Limskikanal bei Rovinj (nördliche Adria). Göttinger Arbeiten zur<br />

Geologie und Paläontologie 8, 1–190.<br />

Von Rad, U., Schaaf, M., Michels, K.H., Schulz, H.D., Berger, W.H.,<br />

Sirocko, F., 1999. A 5000-yr record of climate change in varved<br />

sediments <strong>from</strong> <strong>the</strong> oxygen minimum zone off Pakistan, nor<strong>the</strong>astern<br />

Arabian Sea. Quarternary Research 51, 39–53.<br />

Walton, W.R., 1952. Techniques for recognition of living <strong>foraminifera</strong>.<br />

Contribution Cushman Foundation of Foraminiferal Research 3,<br />

56–60.<br />

Wiesner, H., 1931. Die Foraminiferen der Deutschen Südpolar-Expedition<br />

1901–1903. In: Drygalski, E.v. (ed.): Deutsche Südpolar-<br />

Expedition 1901–1903, Vol. 20, Zoologie 10, 49–169.<br />

Wollenburg, J.E., Mackensen, A., 1998. Living <strong>benthic</strong> foraminifers<br />

<strong>from</strong> <strong>the</strong> central Arctic Ocean. Faunal composition, st<strong>and</strong>ing stock<br />

<strong>and</strong> diversity. Marine Micropaleontology 34, 153–185.<br />

Wyrtki, K., 1973. Physical oceanography of <strong>the</strong> Indian Ocean. In:<br />

Zeitschel, B. (Ed.), The Biology of <strong>the</strong> Indian Ocean. Springer,<br />

Berlin, pp. 18–36.<br />

You, Y., Tomczak, M., 1993. Thermocline circulation <strong>and</strong> ventilation<br />

in <strong>the</strong> Indian Ocean derived <strong>from</strong> water mass analysis. Deep-Sea<br />

Research I 40, 13–56.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!