Sei sulla pagina 1di 11

Phytoplankton composition and biomass across the southern Indian Ocean

Louise Schl uter


a,n
, Peter Henriksen
b
, Torkel Gissel Nielsen
b,c
, Hans H. Jakobsen
c,b
a
DHI Water Environment & Health, Agern Alle 5, DK-2970 Hrsholm, Denmark
b
National Environmental Research Institute, Aarhus University, Department of Marine Ecology, Frederiksborgvej 399, P.O. Box 358, DK-4000 Roskilde, Denmark
c
National Institute of Aquatic Resources, DTU Aqua, Section for Oceanecology and Climate, Technical University of Denmark, Jgersborg Alle 1, DK-2920 Charlottenlund, Denmark
a r t i c l e i n f o
Article history:
Received 7 September 2010
Received in revised form
9 February 2011
Accepted 14 February 2011
Available online 2 March 2011
Keywords:
Phytoplankton
Pigments
Indian Ocean
HPLC
CHEMTAX
Galathea3
a b s t r a c t
Phytoplankton composition and biomass was investigated across the southern Indian Ocean. Phyto-
plankton composition was determined from pigment analysis with subsequent calculations of group
contributions to total chlorophyll a (Chl a) using CHEMTAX and, in addition, by examination in the
microscope. The different plankton communities detected reected the different water masses along a
transect from Cape Town, South Africa, to Broome, Australia. The rst station was inuenced by the
Agulhas Current with a very deep mixed surface layer. Based on pigment analysis this station was
dominated by haptophytes, pelagophytes, cyanobacteria, and prasinophytes. Sub-Antarctic waters of
the Southern Ocean were encountered at the next station, where new nutrients were intruded to the
surface layer and the total Chl a concentration reached high concentrations of 1.7 mg Chl a L
1
with
increased proportions of diatoms and dinoagellates. The third station was also inuenced by Southern
Ocean waters, but located in a transition area on the boundary to subtropical water. Prochlorophytes
appeared in the samples and Chl a was low, i.e., 0.3 mg L
1
in the surface with prevalence of
haptophytes, pelagophytes, and cyanobacteria. The next two stations were located in the subtropical
gyre with little mixing and general oligotrophic conditions where prochlorophytes, haptophytes and
pelagophytes dominated. The last two stations were located in tropical waters inuenced by down-
welling of the Leeuwin Current and particularly prochlorophytes dominated at these two stations, but
also pelagophytes, haptophytes and cyanobacteria were abundant. Haptophytes Type 6 (sensu Zapata
et al., 2004), most likely Emiliania huxleyi, and pelagophytes were the dominating eucaryotes in the
southern Indian Ocean. Prochlorophytes dominated in the subtrophic and oligotrophic eastern Indian
Ocean where Chl a was low, i.e., 0.0430.086 mg total Chl a L
1
in the surface, and up to 0.4 mg Chl a L
1
at deep Chl a maximum. From the pigment analyses it was found that the dinoagellates of unknown
trophy enumerated in the microscope at the oligotrophic stations were possibly heterotrophic or
mixotrophic. Presence of zeaxanthin containing heterotrophic bacteria may have increased the
abundance of cyanobacteria determined by CHEMTAX.
& 2011 Elsevier Ltd. All rights reserved.
1. Introduction
The Indian Ocean is one of the largest yet the least studied
oceans in the world. In October 2006, the Danish expedition,
Galathea 3, crossed the southern Indian Ocean from Cape Town in
South Africa to Broome in north-western Australia. The cruise
passed the nutrient-rich Agulhas Current, which runs along the
east coast of South Africa, encountered sub-Antarctic waters of
the Southern Ocean, passed the oligotrophic subtropical mid part
of the Indian Ocean, and approached the western coast of
Australia where down-welling occurs due to inuence of the
Leeuwin Current. This gave a unique opportunity to compare the
various phytoplankton communities inuenced by the oceano-
graphy in the different areas of this large ocean.
Nano- and picoplankton are important components of phyto-
plankton in the oceans, and in oligotrophic oceans 80% of the
phytoplankton communities are composed by cells smaller than
3 mm (Goericke, 1998). Classical microscopy is insufcient for
identication and quantication of these prominent phytoplank-
ton components in the large oceanic regions. Methods have been
developed especially suited for detecting cells that were pre-
viously overlooked, e.g., epiuorescence microscopy, electron
microscopy, and cell-ow cytometry (reviewed by Jeffrey et al.,
1999), which have greatly improved the knowledge of especially
picoplankton and documented their key role in the oligotrophic
ecosystems (Zapata, 2005). More recently molecular techniques
have been applied (e.g., Moon-van der Staay et al., 2001; Dez
et al., 2001). However, these methods cannot yet be used on a
Contents lists available at ScienceDirect
journal homepage: www.elsevier.com/locate/dsri
Deep-Sea Research I
0967-0637/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.dsr.2011.02.007
n
Corresponding author. Tel.: 45 4516 9557; fax: 45 4516 9292.
E-mail address: lsc@dhigroup.com (L. Schl uter).
Deep-Sea Research I 58 (2011) 546556
routine basis for measuring the whole size range of phytoplank-
ton. Instead the fast and automated chemotaxonomic method,
pigment analysis by High Performance Liquid Chromatography
(HPLC), is increasingly used as a method for measuring the
composition and chlorophyll a (Chl a) biomass of phytoplankton
groups (e.g., Wright and Jeffrey, 2006). Following the HPLC
analysis of pigments the Chl a biomass of the individual phyto-
plankton groups can be calculated by using the CHEMTAX
program developed by Mackey et al. (1996). CHEMTAX calculates
the contribution from different phytoplankton groups to Chl a
based on ratios between accessory pigments and Chl a, which are
loaded into the program together with the eld measurements of
pigment concentrations. Knowledge on the phytoplankton com-
munities in the area sampled is important to achieve trustworthy
results of the CHEMTAX calculated biomasses (Wright et al.,
1996; Ansotegui et al., 2003; Irigoien et al., 2004). General
agreement between results of microscopy and pigment analyses
including the CHEMTAX calculation of the biomass of the phyto-
plankton groups has been found in samples from many different
environments, i.e., freshwater, estuaries and coastal waters, and
oceanic regions (Andersen et al., 1996; Schl uter et al., 2000,
2006; Henriksen et al., 2002; Havskum et al., 2004).
Monitoring of phytoplankton by the pigment method to
describe the general structure and composition of phytoplankton
groups has been conducted in oligotrophic oceans (e.g., Gibb et al.,
2000; Marty et al., 2002; Sakamoto et al., 2004; Veldhuis and
Kraay, 2004). Knowledge on phytoplankton assemblages in ocea-
nic regions of the southern hemisphere is, however, still limited.
In the Indian Ocean only a few investigations on phytoplankton
communities encompassing the whole size range of phytoplank-
ton have been published. Barlow et al. (2007) analyzed phyto-
plankton by HPLC in surface samples only on transects in the
different oceans of the southern hemisphere, and found relatively
low biomass and prokaryote dominance in the Indian Ocean. Not
et al. (2008) used pigment analysis and subsequent CHEMTAX
analysis to measure phytoplankton in two regions of the Indian
Ocean with special emphasis on the picoeucaryotes.
The water column structure has impact on the nutrient supply
to the euphotic zone. In general well-mixed nutrient-rich oceans
will support classic food chains with large phytoplankton and
large copepods, whereas oligotrophic stratied waters are domi-
nated by small phytoplankton where the productivity is based on
nutrients regenerated from microbial driven food webs. Knowl-
edge about the phytoplankton communities is essential to
improve our understanding of the plankton community structure
and productivity of the southern Indian Ocean. In order to get a
better understanding of the structure and function of the phyto-
plankton in the different regions in the large Indian Ocean, which
are inuenced by different water masses each with different
physical characteristics (Visser et al., submitted), more investiga-
tions on the vertical and horizontal distribution, abundance and
composition of the whole size range of the phytoplankton com-
munities are needed. The aim of the present paper is to analyze
and characterize the phytoplankton communities across the
Indian Ocean in relation to the different oceanographic regimes
by HPLC supplemented by classic microscopy to gain new
information on the primary producers in this poorly studied area.
2. Material and methods
2.1. Sampling strategy
Water samples were collected in 30-L Niskin bottles mounted
on a rosette with a Seabird 9/11 CTD on a transect with
seven stations from Cape Town in South Africa to Broome in
north-western Australia (Fig. 1) during late spring, October
18November 5, 2006. A detailed description of the sampling
and the oceanography of the transect can be found in Visser et al.
(submitted). Briey, water samples from 10, 30, 60, 100, and
200 m depth were taken at all stations. Furthermore, in between
these casts discrete surface samples were taken along the transect
by a seawater intake system positioned approximately 5 m below
the ocean surface, which enabled water sampling of surface water
while cruising.
2.2. HPLC analyses
A Shimadzu LC-10A HPLC, composed of one pump (LC-
10ADVP), photodiode array detector (SPD-M10A VP), SCL-10ADVP
System controller with Class-VP software, temperature-controlled
auto sampler (set at 4 1C), column oven (CTO-10ASVP) and a
degasser, was installed in one of the six laboratory containers
mounted on the quarterdeck of the vessel. In order to hamper the
noise on the power supply generated by the vessel, the electricity
in all the laboratory containers was EMC secured (electromag-
netic compatibility) according to military standards and supplied
from a UPS (universal power supply) grounded to a single node in
each container. The baseline noise of the HPLC was only slightly
increased compared to land based HPLC analyses.
The samples for HPLC analysis were ltered in dim light onto
25 mm Whatman GF/F lters and immediately frozen in liquid
Fig. 1. Cruise track with station locations superimposed on map of surface temperature.
L. Schl uter et al. / Deep-Sea Research I 58 (2011) 546556 547
nitrogen. The samples were analyzed onboard within 3 days after
sampling. The lters were placed in a syringe mounted with a
0.2 mm Teon syringe lter and 2 mL 95% acetone containing
0.025 mg mL
1
vitamin E acetate as internal standard was added.
The samples were sonicated in the syringes for 10 s on ice with a
Sonics Vibra Cell Ultrasonic processor, and ltered directly into
HPLC vials. The vials were placed in the cooling rack (4 1C) of the
HPLC together with a parallel set of vials with injection buffer
(90:10, 28 mM aqueous tetrabutyl ammonium acetate (TBA), pH
6.5; methanol). The samples were mixed with buffer using the
auto injector by programming it to make a mix in the loop of
buffer and sample in the ratio 5:2. A total volume of 500 mL was
injected. The method used for HPLC analysis was the Van
Heukelem and Thomas (2001) method, with an Eclipse XDB C8,
4.6 mm150 mm column (Agilent Technologies). Solvent A:
(70:30) methanol: 28 mM aqueous TBA (hydroxide titrant, JT
Baker HPLC reagent V365-07), pH 6.4, solvent B: 100% methanol.
Solvents were mixed using linear gradients along the following
time program: 0 min: 95% A, 5% B, 22 min: 5% A, 95% B, 30 min:
95% A, 5% B, 31 min: 100% A, 0% B, 34 min: 100% A, 0% B, 35 min
5% A, 95% B, 41 min: Stop. The ow rate was 1.1 mL min
1
and
the temperature of the column oven was set at 60 1C. The HPLC
was calibrated with pigment standards from DHI Lab Products,
Denmark. The internal standard was detected at 222 nm, while
the rest of the pigments were detected at 450 nm. Peak identities
were routinely conrmed by on-line PDA analysis.
A QA threshold procedure, application of limit of quantitation
(LOQ) and limit of detection (LOD), was applied to the pigment
data as described by Hooker et al. (2005) to reduce the uncer-
tainty of pigments found either in low concentrations or not
detected at all, causing false positives or false negatives, which
frequently occur when pigments are quantied near the
detection limit.
To investigate if all phytoplankton cells were collected on the
lters used in this study to collect the whole phytoplankton
community (Whatman GF/F, nominal pore size 0.7 mm), subsam-
ples of ltrates (the volumes were not recorded) of two surface
samples from station 6 were collected and ltered onto GE
Osmonics polycarbonate 0.2 mm lters, extracted and analyzed
by HPLC as described above.
2.3. Enumeration, identication, and biomass estimates of
autotrophic protists
Samples from 10 and 60 m were collected from the rosette
immediately after it landed on deck, xed in acidic Lugols (nal
conc. 5%) and stored in 300 mL brown bottles at 5 1C. Samples
were subsequently counted within 3 months from the sampling
date. Sample aliquots were allowed to settle in 100 mL Uterm ohl
settling chambers for 24 h and analyzed in an inverted micro-
scope. Cellso20 mm were observed using an objective of 40X,
while larger cells were analyzed by a 10X objective. Depending on
the abundance of the species in consideration, a fraction of or the
entire sample was counted. Each counted cell was assigned to a
morphological group and size class. The size classes were made of
10 mm equivalent spherical diameter (ESD) intervals and a
volume was assigned using the appropriate morphology volume
relationship equations. Except for a few rare species most of the
thecate dinoagellates could be identied to genera. Hence,
nutritional mode could be deduced from the literature. The naked
dinoagellates presented a challenging taxonomical problem, and
only Gymnodinium spirale and members of the genus Cochlodinium
spp. were identied as heterotrophic (data on heterotrophic
protozoans will be presented in a later publication, Jo nasdo ttir
et al., in preparation). The cellular carbon content of protists was
estimated from the taxon-specic ESD:carbon relationships
(Menden-Deuer and Lessard, 2000).
2.4. CHEMTAX analyses
The pigment concentrations were loaded into the CHEMTAX
program to calculate the Chl a biomass of the individual phyto-
plankton groups (Mackey et al., 1996). The pigment data set was
divided in two oceanographic regions: the samples taken during
the rst part of the cruise, where the Agulhas Bank encounters
sub-Antarctic waters of the Southern Ocean until 861E longitude
(south-western (SW) Indian Ocean), and samples taken from
911151E longitude (south-eastern (SE) Indian Ocean) (Fig. 1),
where total Chl a was below 0.1 mg L
1
. These two data sets were
further divided into two sets: surface samples and samples from
and below the vertical Chl a maximum (Chl a
max
).
The pigment ratios used as input values for the CHEMTAX
calculations were from Schl uter et al. (2000), Higgins and Mackey
(2000), Gibb et al. (2001), Rodrguez et al. (2002), and Higgins
et al. (in press). The CHEMTAX program version 1.95 was used to
construct 60 different ratio matrices from the initial ratios for
each of the four data sets. Monovinyl (MV) Chl a was used for
calculating the biomass of all other groups than prochlorophytes
for which divinyl (DV) Chl a was used. 10% (n6) of the ratios
creating the lowest residual root mean square were averaged and
run repeatedly until the ratios became stable.
3. Results
3.1. The oceanography of the southern Indian Ocean
The oceanography of the cruise track (Fig. 1) is described in
details in Visser et al. (submitted). Briey, the rst station was
characterized by the conuence of the circumpolar circulation
and the water masses of the Agulhas Bank, which is a region of
elevated biological production. Station 2 was located where the
waters masses of the Agulhas Bank meet the converging sub-
tropical and sub-Antarctic fronts just north of the Crozet Plateau
in the Southern Ocean. Station 3 was situated on the subtropical
front where relatively cold and fresh Antarctic intermediate
waters are subducted under the salty-warm waters of the sub-
tropical gyre. Stations 4 and 5 were located in the subtropical gyre
with little mixing from winds or from meso-scale eddies and
oligotrophic conditions. Stations 6 and 7 were situated in tropical
waters inuenced by the Indonesian through ow where station
7 was on the shelf break of the North-West Australian shelf and
inuenced by coastal condition (Fig. 1).
3.2. Phytoplankton biomass: results of pigment analyses
The different oceanic regimes crossed were reected in the
phytoplankton biomass and diversity. The phytoplankton biomass
in the surface of the SW Indian Ocean was between 0.2 and
0.4 mg Chl a L
1
with an increase to 1.4 mg Chl a L
1
in the area
with inuence of Southern Ocean water (Fig. 2). In the SE Indian
Ocean the Chl a biomass dropped at the surface to concentrations
of 0.040.09 mg Chl a L
1
(Fig. 2). At the rst stations in the SW
Indian Ocean the depth prole showed a Chl a
max
at around
3040 m, while the Chl a
max
was located deeper, i.e., at around
100 m, in the middle part of the Indian Ocean (Fig. 3). Towards
the western coast of Australia, the Chl a
max
was again located
higher up in the water column at around 60 m (Fig. 3).
The HPLC results revealed presence of DV Chl a, the
diagnostic pigment of prochlorophytes, in samples from station
3 and onwards. Zeaxanthin, another important pigment in
L. Schl uter et al. / Deep-Sea Research I 58 (2011) 546556 548
prochlorophytes, was detected in the samples from the rst part
of the cruise where DV Chl a was absent, indicating presence of,
e.g., cyanobacteria, prasinophytes, and/or chlorophytes. Crypto-
phytes were detected in practically all samples by the diagnostic
pigment alloxanthin. The presence of MV Chl c
3
as well as
19
0
-butanoyloxyfucoxanthin (19
0
-but) and 19
0
-hexanoyloxyfu-
coxanthin (19
0
-hex) in almost all samples across the Indian Ocean
indicated presence of at least haptophytes Type 6; Emiliana
huxleyi and Gephyrocapsa oceanica according to Zapata et al.
(2004), and probably also pelagophytes (Andersen et al., 1993).
Peridinin, the diagnostic marker of dinoagellates, prasinoxanthin
as well as other pigments present in prasinophytes and/or
chlorophytes (Chl b, lutein, violaxanthin, neoxanthin) were
detected in many samples. Calculation of the pigment ratios:
prasinoxanthin/Chl b and lutein/Chl b (Schl uter and Mhlenberg,
2003) for SW Indian Ocean, where DV Chl b was absent and
therefore did not interfere these calculations due to coelution of
DV and MV Chl b, indicated that both prasinophytes with
prasinoxanthin (prasinophytes Type 3, Higgins et al., in press,
and prasinophytes without prasinoxanthin and including chlor-
ophytes (prasinophytes Type 1, Higgins et al., in press) were
present in the samples.
The output ratios from the CHEMTAX analyses are shown in
Table 1. The ratios of peridinin/Chl a in dinoagellates and
prasinoxanthin/Chl a in prasinophytes were both horizontally
and vertically relatively constant, while fucoxanthin/Chl a in
diatom ratios were constant across different water masses, but
differed vertically with higher ratios in the deeper part of the
water column (Table 1). The fucoxanthin/Chl a ratios in E. huxleyi-
Type 6 haptophytes were variable around 0.1 showing no clear
difference for the different data sets, but the ratio of the
diagnostic pigment 19
0
-hex to Chl a was higher in the SE Indian
Ocean and increased in the deeper part of the water column
(Table 1). 19
0
-but/Chl a in pelagophytes and alloxanthin/Chl a
ratios in cryptophytes were lower in the deeper parts of the water
column in the SW Indian Ocean, but tended to increase in the SE
Indian Ocean. Such different responses were also obvious in the
two subtypes of prasinophytes Types 1 and 3; Chl b/Chl a ratios of
prasinophytes with prasinoxanthin were relatively stable across
the southern Indian Ocean, while those of prasinophytes Type 1
(incl. chlorophytes) were approx. twice as high in the SW Indian
Ocean (Table 1). While zeaxanthin/Chl a ratios in the upper part
of the water column were more than twice as high as in the
deeper waters for cyanobacteria, they were more than 5 times
higher for prochlorophytes (Table 1). Both groups had higher
ratios in the eastern part of the Indian Ocean.
The phytoplankton groups calculated by CHEMTAX revealed
that the increase in Chl a at station 2 were caused by a general
increase in the Chl a biomass of most phytoplankton groups
(Fig. 3), but particularly dinoagellates and diatoms were abun-
dant (Table 2). The phytoplankton populations in the surface
waters after station 2 were dominated by haptophytes (Fig. 2). At
station 4 the phytoplankton biomass in the surface dropped to a
total Chl a biomass less than 0.05 mg Chl a L
1
, and became
dominated by prochlorophytes, but haptophytes, cyanobacteria,
and pelagophytes also constituted an important part of the
phytoplankton population (Fig. 2, Table 2).
The CHEMTAX calculations showed that diatoms as well as
dinoagellates were only sporadically present when prochloro-
phytes became dominating in the SE Indian Ocean. Both vertically
and horizontally, haptophytes and pelagophytes constituted a
signicant part of the phytoplankton biomass, and pelagophytes
even exceeded the biomass of prochlorophytes at Chl a
max
at
station 5 (Fig. 3, Table 2). Except at station 2 cyanobacteria also
appeared to constitute an important part of the phytoplankton
population in the Indian Ocean (Fig. 3).
The ltrates of the GF/F ltered samples collected onto 0.2 mm
lters showed that all phytoplankton cells were collected on
the GF/F lters in this study, since the Chl a concentrations in the
0.2 mm samples were below the limit of detection. However, the
pigment zeaxanthin was detected on the 0.2 mm lters as the only
accessory pigment indicating that non-autotrophic cells with a size
less than the nominal size of GF/F lters of approx. 0.7 mm contain-
ing this photo-protective pigment were present in the samples.
3.3. Phytoplankton biomass: results of microscopy
The phototrophic protists 45 mm identied by inverted micro-
scopy showed a dominance of unidentied agellates at most
stations except station 2, where the highest biomass of
280 mg C L
1
was measured at 60 m depth and diatoms and
dinoagellates dominated, and at station 7 off the Australian west
coast (Table 3). At this westernmost station pennate diatoms
o20 mm and dinoagellates of unknown trophy dominated. The
autotrophic thecate dinoagellates constituted an insignicant
biomass except at station 2, where species of the genera Gonyau-
lax, Heterocapsa, Prorocentrum, and Torodinium were observed.
0
0.2
0.4
0.6
0.8
1
1.2
1.4
21.10
St. 1
22.10 23.10
St. 2
24.1025.1026.10 27.10
St. 3
27.10 28.10 29.10

g

C
h
l

a

L
-
1
0
0.02
0.04
0.06
0.08
0.1
30.10
St. 4
31.10
St. 5
2.11
St. 6
4.11
St. 7
Prochlorophytes
Diatoms
Cyanobacteria
Prasinophytes type 1
Pelagococcus
Haptophytes type 6
Cryptophytes
Dinoflagellates
Prasinophytes type 3
Fig. 2. Biomass of the phytoplankton population calculated by CHEMTAX in the surface in the southern Indian Ocean. The x-axis shows the sampling dates and the
stations. The dates sampled in between stations are sampled with the seawater intake system described in Material and methods.
L. Schl uter et al. / Deep-Sea Research I 58 (2011) 546556 549
Naked dinoagellates of unknown trophy were present in most
samples (Table 3). Assigning nutrition strategy to this group is
challenging. Regression between the biomass of dinoagellate
estimated in microscope averaged over 10 and 60 m against the
biomass of dinoagellates determined by HPLC from all stations
except station 2 revealed that that the slope of the regression line
was not different from 0 (N6; P0.31, when including dino-
agellates of unknown trophy and P0.79 when excluding dino-
agellates of unknown trophy). Station 2 was not included in this
analysis as the food web at this station appeared very different
from the remaining stations. As the dinoagellates of unknown
trophy did not add any signicance to the relationship between
the biomass estimated by HPLC and microscopy, respectively, the
dinoagellates of unknown trophy were most likely either mixo-
trophic with a low pigment content or strictly heterotrophic.
4. Discussion
4.1. Phytoplankton in relation to the oceanography of the southern
Indian Ocean
The different phytoplankton communities measured were reect-
ing the different water masses of the southern Indian Ocean. The rst
station sampled, station 1, was inuenced by the Agulhas Current
with a very deep mixed surface layer, down to 300 m (Visser et al.,
submitted), where total Chl a was 0.4 mg L
1
at the surface. Flagel-
lates and athecate dinoagellates were detected by microscopy
(Table 3). HPLC measurements suggested that phytoplankton was
in fact dominated by haptophytes, pelagophytes, and cyanobacteria.
Prasinophytes Types 1 and 3 were also important at this station. This
is comparable to the results of Not et al. (2008), who also found
cyanobacteria, i.e., Synechococcus, and picoeucaryotes (pelagophytes,
haptophytes, and chlorophytes/prasinophytes) to dominate at the
more coastal, nutrient-rich stations in the Indian Ocean. At station
2 sub-Antarctic waters of the Southern Ocean were encountered,
reducing the surface temperatures and the salinity (Visser et al.,
submitted). The Southern Ocean is one of the high-nutrient low-
chlorophyll (HNLC) regions, but at this station the total Chl a
concentration reached relatively high values for oceanic regions of
1.4 mg Chl a L
1
in the surface (Fig. 2) with a Chl a
max
of 1.7 mg L
1
in
30 m depth. The high Chl a concentration was accompanied by
increased proportions of diatoms and dinoagellates in the phyto-
plankton population detected by both methods (Tables 2 and 3). This
is commonly found in areas with a supply of new nutrients since
these opportunistic taxa are particularly well suited to take advantage
of excess nutrients (Fogg, 1991; Claustre, 1994). High nutrient
concentrations were indeed measured from the surface throughout
Table 1
Output ratios of pigment/chlorophyll a from the CHEMTAX calculations for the four different data sets analyzed (see text for details).
Chl c
3
Chl c
2
Chl c
1
MV Chl c
3
Peri 19
0
-but Fuco Neox Pras Viol 19
0
-hex Allo Zeax Lut Chl b
South-western Indian Ocean, surface
Prasinophytes Type 3 0.111 0.377 0.050 0.020 0.784
Dinoagellates 0.456 0.698
Cryptophytes 0.065 0.307
Haptophytes Type 6 0.195 0.084 0.036 0.006 0.032 0.781
Pelagophytes 0.131 0.512 1.108 0.222
Prasinophytes Type 1 0.091 0.111 0.038 0.054 0.793
Cyanobacteria 1.378
Diatoms 0.032 0.026 0.307
Prochlorophytes 0.466 0.147
South-western Indian Ocean, from and below chlorophyll a maximum
Prasinophytes Type 3 0.113 0.458 0.079 0.018 0.679
Dinoagellates 0.289 0.711
Cryptophytes 0.060 0.172
Haptophytes Type 6 0.140 0.162 0.007 0.008 0.184 1.900
Pelagophytes 0.365 0.176 0.471 0.085
Prasinophytes Type 1 0.068 0.063 0.005 0.005 0.539
Cyanobacteria 0.650
Diatoms 0.140 0.012 0.809
Prochlorophytes 0.086 0.338
South-eastern Indian Ocean, surface
Prasinophytes Type 3 0.123 0.488 0.074 0.016 0.905
Dinoagellates 0.271 0.735
Cryptophytes 0.069 0.184
Haptophytes Type 6 0.289 0.199 0.079 0.014 0.212 1.483
Pelagophytes 0.188 0.334 0.821 0.095
Prasinophytes Type 1 0.058 0.157 0.045 0.124 0.350
Cyanobacteria 2.692
Diatoms 0.105 0.014 0.370
Prochlorophytes 0.732 0.096
South-eastern Indian Ocean, from and below chlorophyll a maximum
Prasinophytes Type 3 0.065 0.402 0.071 0.014 0.601
Dinoagellates 0.247 0.748
Cryptophytes 0.105 0.265
Haptophytes Type 6 0.106 0.132 0.006 0.009 0.102 1.949
Pelagophytes 0.769 0.344 1.059 0.095
Prasinophytes Type 1 0.236 0.219 0.014 0.011 0.313
Cyanobacteria 0.781
Diatoms 0.163 0.029 0.719
Prochlorophytes 0.156 1.184
Abbreviations: Chl, chlorophyll; MV, monovinyl; peri, peridinin; 19
0
-but, 19
0
-butanoyloxyfucoxanthin; fuco, fucoxanthin; neox, neoxanthin; pras, prasinoxanthin; viol,
violaxanthin; 19
0
-hex, 19
0
-hexanoyloxyfucoxanthin; allo, alloxanthin; zeax, zeaxanthin; lut: lutein.
L. Schl uter et al. / Deep-Sea Research I 58 (2011) 546556 550
the mixed layer, and preceding days of rough weather conditions had
intruded newnutrients, and possibly iron enrichment fromthe Crozet
Plateau (Pollard et al., 2007) to the area (Visser et al., submitted).
Station 3 was also inuenced by the Southern Ocean waters and at
this station prochlorophytes appeared for the rst time in the samples
and Chl a was relatively low, i.e., 0.3 mg L
1
in the surface waters,
with prevalence of haptophytes particularly in the surface. Further-
more, pelagophytes and cyanobacteria were abundant (Fig. 3), indi-
cating that this station was located in a transition area on the
boundary to subtropical water. Stations 4 and 5 were located within
the subtropical gyre with little mixing and general oligotrophic
conditions. This was reected in the composition and biomass of
phytoplankton with low surface concentrations of Chl a and a
maximum of 0.2 mg Chl a L
1
at 100 m depth at both stations,
showing dominance of prochlorophytes, cyanobacteria, and small
agellates (Fig. 3) typical for oligotrophic areas where regenerated
nutrients are the only nutrient source. The last two stations, 6 and 7,
were located in tropical waters inuenced by down-welling of the
Leeuwin Current. However, the presence of pennate diatoms at
station 7 suggests a complex exchange of ocean water and coastal
water masses that inject Si to the water column. Particularly
prochlorophytes dominated at these two stations (Figs. 2 and 3),
but also pelagophytes, haptophytes, and cyanobacteria were abun-
dant. These cells were not detected by the microscopy method used,
by which only larger cells were identied. The Chl a
max
at station
7 located on the shelf break inuenced by coastal conditions was
situated higher up in the water column at 60 m and picoprocaryotes
(cyanobacteria and prochlorophytes) contributed 4567% to the total
Chl a biomass in and above Chl a
max
(Table 2). This is comparable to
the range of 5565% found by Hanson et al. (2007) for picoprocar-
yotes in deep Chl a maximum (DCM) in the Leeuwin Current in
coastal waters of western Australia in close vicinity of our station 7,
with haptophytes as the other primary contributor (2132%). How-
ever, contrary to the present study the phytoplankton population in
0
50
100
150
200
0
m
g Chl a L
-1
St. 1
0
50
100
150
200
0
m
g Chl a L
-1
St. 2
0
50
100
150
200
0
m
St. 3
0
50
100
150
200
0
m
St. 4
0
50
100
150
200
0
m
St. 5
0
50
100
150
200
0
m
St. 6
0
50
100
150
200
0
m
Prasinophytes type 3
Dinoflagellates
Cryptophytes
Haptophytes type 6
Pelagophytes
Prasinophytes type 1
Cyanobacteria
Diatoms
Prochlorophytes
St. 7
0.05 0.1 0.15 0.1 0.2 0.3 0.4 0.5 0.6
0.02 0.04 0.06 0.08 0.04 0.08 0.12 0.16
0.01 0.02 0.03 0.04 0.05 0.06 0.07
0.04 0.08 0.12 0.16
0.05 0.1 0.15 0.2
Fig. 3. Depth distribution of the biomass of the individual phytoplankton groups as Chl a concentration calculated by CHEMTAX at the stations sampled.
L. Schl uter et al. / Deep-Sea Research I 58 (2011) 546556 551
the surface waters of the study of Hanson et al. (2007) was dominated
by cyanobacteria and haptophytes, while prochlorophytes were
virtually absent. Hanson et al. (2007) did, however, not determine
the diagnostic pigment of prochlorophytes, DV Chl a, by HPLC. Instead
this group was calculated by CHEMTAX using the pigments zeax-
anthin and Chl b. Nevertheless, Barlow et al. (2007) did indeed
measure DV Chl a, indicative of prochlorophytes, and the DV
Chl a/total Chl a ratio was 0.4 in the surface transect of the Indian
Ocean at 201S, the station closest to the Australian coast placed at
approx. 1131E, close to our station 7 (1151E, 201S). This is comparable
to the present study where this ratio was 0.47 at station 7. Barlow
et al. (2007) sampled only surface water and found dominance by
prokaryotes and low total Chl a biomass down to 0.02 mg
total Chl a L
1
at a transect more northerly (201S) than ours. In the
oligotrophic SE Indian Ocean we measured total Chl a concentrations
from 0.043 to 0.086 mg Chl a L
1
in the surface (Fig. 2). Barlow et al.
Table 2
Contribution in percentage of the different phytoplankton groups to chlorophyll a biomass obtained by CHEMTAX.
Station Depth
(m)
Prasinophytes
Type 3
Dinoagellates Cryptophytes Haptophytes
Type 6
Pelagophytes Prasinophytes
Type 1
Cyanobacteria Diatoms Prochlorophytes
1 10 9 0 5 38 12 11 9 15 0
30 8 0 5 18 28 11 24 6 0
60 9 0 7 17 29 11 21 5 0
100 9 0 8 16 30 10 21 6 0
2 10 1 16 3 27 9 7 2 36 0
30 3 13 5 15 29 11 3 23 0
60 5 3 3 13 32 6 1 35 0
100 5 13 6 18 22 12 3 18 0
3 10 0 1 1 62 10 0 9 5 13
30 0 0 1 26 22 1 25 13 11
60 0 0 0 13 26 0 32 0 29
100 4 0 1 12 56 4 8 0 15
4 10 3 2 8 12 10 9 10 7 39
30 3 2 6 17 16 9 7 6 36
60 3 2 6 18 19 9 4 1 39
100 1 1 2 15 18 5 19 0 40
5 10 3 3 10 29 16 12 6 0 20
30 3 2 8 23 21 11 8 0 25
60 3 1 5 28 26 9 2 0 26
100 2 1 2 21 31 6 11 5 22
6 10 3 2 7 15 11 9 13 3 38
30 3 3 7 19 9 9 13 2 36
60 3 3 8 21 14 8 11 3 29
100 1 1 2 10 16 2 20 1 47
7 10 2 3 3 9 4 7 17 8 47
30 2 3 2 9 5 6 13 6 54
60 7 3 4 11 16 5 11 8 34
80 4 3 3 11 25 4 10 11 29
100 2 2 5 16 33 4 3 15 20
Average 3 3 5 19 20 7 12 8 31
Table 3
Biomasses of autotrophic protists across the Indian Ocean and the sum of all groups at 10 and 60 m depths. Units mg C L
1
.
Diatoms Flagellates45 lm Autotrophic thecate
dinoagellates
Autotrophic athecate
dinoagellates
Athecate dinoagellates of
unknown trophy
Sum of all
groups
Station 10 m 60 m 10 m 60 m 10 m 60 m 10 m 60 m 10 m 60 m 10 m 60 m
1 0.09 0.07 0.03 1.00 0.47 0.46 0.05 0.05 1.44 1.73 2.08 3.30
2 6.74 133.92 9.15 34.35 31.91 57.17 0.82 1.31 56.06 53.41 104.68 280.16
3 0.23 0.12 7.75 3.25 0.93 1.35 0.12 0.13 4.31 3.05 13.34 7.90
4 0.02 0.01 2.22 10.97 0.57 1.09 0.09 0.11 2.53 3.81 5.44 16.00
5 0.00 0.08 4.02 3.31 0.65 1.27 0.06 0.03 2.49 2.93 7.23 7.63
6 0.09 0.14 3.15 0.02 0.16 0.64 0.06 0.23 1.04 6.95 4.51 7.99
7 0.44 2.55 0.02 0.10 0.77 1.00 0.07 0.11 2.79 3.04 4.08 6.81
L. Schl uter et al. / Deep-Sea Research I 58 (2011) 546556 552
(2007) found total Chl a of 0.09 mg L
1
, which is comparable to this
study. Prochlorophytes were, however, even more concentrated in
the DCM (Fig. 3) where total Chl a reached 0.4 mg Chl a L
1
.
4.2. The use of CHEMTAX for determining phytoplankton
composition and biomass
Analyzing phytoplankton pigments by HPLC gives the advan-
tage of getting information on the whole phytoplankton commu-
nity by one method, which cannot be achieved in one single step
by other methods. However, the subsequent application of pro-
grams like CHEMTAX requires subjective interpretations on
which phytoplankton groups and pigment/Chl a ratios to load
into CHEMTAX. In order to diminish the uncertainty on these
taxonomical interpretations, Higgins et al. (in press) have made a
guide for quantitative chemotaxonomic interpretation of pigment
data. Briey, it is important to obtain as much information as
possible on which phytoplankton groups to expect in the samples,
i.e., also by using alternative methods to pigment analysis, since
the results of the CHEMTAX analyses rely on the input to the
program. Furthermore, the pigment ratios and phytoplankton
groups chosen to load into CHEMTAX should reect the phyto-
plankton communities sampled. It is recommended to divide the
pigment data into subsamples of populations with equal environ-
mental conditions (light adaptations, water mass properties, etc.),
and carefully select the initial pigment/Chl a ratios, using multiple
starting pigment ratios. Then to run the CHEMTAX calculations
repeatedly by optimizing the input ratios in order to minimize the
residual root mean square error (Higgins et al., in press). These
approaches were used in this study. The pigment data set was
divided to represent two oceanographic regions, SE and SW
Indian Ocean. Although the phytoplankton populations, detected
at the different stations sampled indicated, that even more sub-
grouping of this large ocean could be considered, the number of
samples was limited and we found it more important to divide
the dataset vertically.
The choice of which phytoplankton groups to load into
CHEMTAX was supported by microscopic enumerations made
by inverted microscope, showing presence of diatoms, dinoa-
gellates, and unidentiable agellates (Table 3). Peridinin is a
diagnostic marker of dinoagellates, but the pigment method
only found peridinin containing dinoagellates to be a signicant
part of the phytoplankton population at station 2, and absent or
sporadically present in the rest of the samples (Fig. 3, Table 2).
Some dinoagellates have acquired their chloroplast and pig-
ments from other taxa and contain, e.g., fucoxanthin and its
derivates (De Salas et al., 2003). If present, these organisms will
have been included in the haptophytes by the CHEMTAX analyses.
The dinoagellate genera Ornithocercus, Histioneis, Parahistioneis
and Citharistes, Amphisolenia and Triposolenia observed by micro-
scopy had ectosymbiotic cyanobacteria, while the latter two also
had endosymbionts of eukaryotic origin (Farnelid et al., 2010;
Tarangkoon et al., 2010). Although, ecto- and endosymbiotic
mixotrophy is found in a wide range of oceanic dinoagellate
species, their abundance is o2 cells L
1
(Tarangkoon et al., 2010)
and thus of minor importance. The discrepancy between the
HPLC pigment method and microscopic analysis may be
explained by heterotrophic nutrition in the dinoagellates, which
may be dominant yet invisible to pigment analysis, except for
what they have consumed (Higgins et al., in press). Hence,
we suggest that most of the dinoagellates enumerated in the
microscope as dinoagellates of unknown trophy were most
likely heterotrophic.
Prochlorophytes, cryptophytes, and prasinophytes Type 3
were detected by their unique diagnostic pigments DV Chl a,
alloxanthin, and prasinoxanthin, respectively. Cyanobacteria are
usually detected by the non-specic pigment zeaxanthin, which
was present at all stations. Since Synechococcus has long been
documented by ow cytometry as an important constituent of the
prokaryotic algal community in many oceanic regions, including
the Indian Ocean (Not et al., 2008), this group was included in the
CHEMTAX calculations. The presence of prasinophytes
Type 1 (including chlorophytes) could be identied by pig-
ment ratios as mentioned in the results.
The presence of 19
0
-hex indicated presence of haptophytes.
Zapata et al. (2004) investigated pigments of haptophytes and
found 8 different types based on their pigment content, where
3 types contained 19
0
-hex. One of the pigments, MV Chl c
3
which
was detected in this study, has been found to be strongly
associated with the globally important species Emiliania huxleyi
and one other species, Gephyrocapsa oceanica, grouped as hapto-
phytes Type 6 in Zapata et al. (2004). MV Chl c
3
was detected in
most samples in this study along with 19
0
-hex and 19
0
-but, and
haptophytes Type 6 was consequently included in CHEMTAX
(Table 1). This group was found to constitute an important
part of the phytoplankton population in the Indian Ocean, and
particularly in the SW Indian Ocean haptophytes Type 6 tended
to dominate in the surface waters, but were also abundant
in the deeper parts of the water column in the SE Indian Ocean
(Fig. 3). E. huxleyi has been found to dominate the coccolithophore
populations in various regions of the worlds oceans (e.g., Boeckel
et al., 2006; Lipsen et al., 2007; Siegel et al., 2007; Gravalosa
et al., 2008), and this study shows that haptophytes Type 6,
i.e., E. huxleyi and G. oceanica, are important in the Indian Ocean
too. Flagellates were grouped as unknown agellates by micro-
scopy. Since acidic Lugols iodine was used to x the samples, the
coccoliths were most likely dissolved (Sournia, 1978), which
made the identication of coccolithophorids impossible.
Haptophytes types 15 do not contain the characteristic
pigments 19
0
-hex and 19
0
-but (Zapata et al., 2004) and if present,
they were included in the groups of diatoms by CHEMTAX. An
examination of the diatoms carbon/Chl a (C/Chl a) ratios showed
that these were varying with an average of 132 and when
excluding station 2, 60 m, the C/Chl a ratio was in average 27
(data not shown). The very high C/Chl a ratio at station 2, 60 m
together with a generally low photosynthetic activity (Visser
et al., submitted), indicates a decline of the bloom encountered
at this station. Although uncertainty exists on counting and
biomass estimations made by microscopy and particularly dia-
toms of varying C/Chl a ratios (Schl uter and Mhlenberg, 2003),
the potential inclusion of haptophytes Types 15 in the group
diatoms by the CHEMTAX calculations did not lead to low C/Chl
a ratios of diatoms. Thus haptophytes types 15 seem to have
been of minor importance in the Indian Ocean. Furthermore,
except at station 2 diatoms determined by pigment analysis
usually only constituted a few percentages and 15% at maximum
(Table 2), which also indicates that any haptophytes types
15 included in diatoms by CHEMTAX were of insignicant
importance.
The 19
0
-hex/19
0
-but-ratios of all data in the present data set
were 2.471.1 (average7standard deviation, n223). E. huxleyi
and G. oceanica contain no or very little 19
0
-but and according
to Zapata et al. (2004) the Type 6 haptophytes, which were found
to contain 19
0
-but, had 19
0
-hex/19
0
-but-ratios of at least 50. This
indicates that other 19
0
-but containing algae were present in the
samples of this study. In the study of Andersen et al. (1996)
pelagophytes were found to be an important phytoplankton
group both by pigments and electron microscopy in the Atlantic
and Pacic Oceans, and subsequently, the pelagophytes have been
identied by 19
0
-but in various oligotrophic waters (e.g., Bidigare
and Ondrusek, 1996; Steinberg et al., 2001; Suzuki et al., 2002;
L. Schl uter et al. / Deep-Sea Research I 58 (2011) 546556 553
Marty et al., 2008). Consequently, pelagophytes were included in
the CHEMTAX calculations, and were found to comprise an
important part, on average 20%, of the phytoplankton population
in the Indian Ocean, particularly in the deeper parts of the ocean
(Table 2). This agreed well with results of Not et al. (2008) from a
more northerly transect in the Indian Ocean, where pelagophytes
were found by HPLC to constitute a similar part of the phyto-
plankton population. As found in other oceans (e.g., Andersen
et al., 1996; Steinberg et al., 2001) haptophytes and pelagophytes
were the most abundant eucaryotes in the Indian Ocean con-
tributing profoundly to the Chl a
max
(Fig. 3, Table 2). In the
present study pelagophytes tended to be placed even lower in the
water column than the haptophytes, a pattern also found in a few
other studies, e.g., in the Atlantic Ocean (Veldhuis and Kraay,
2004) and in the Mediterranean (Marty et al., 2008).
4.3. Effects of applying QA threshold
The QA threshold procedure applied to the data set in this
study, where the baseline noise was slightly increased due to the
noise from the power supply generated by the vessel, resulted in
an improved CHEMTAX solution. This is apparent from an
evaluation of the residuals (pigment content unexplained by the
CHEMTAX solution as root mean square, data not shown). The
residuals from the CHEMTAX analyses, carried out after the QA
threshold procedure was applied, were up to 15 times lower
when compared to the residuals achieved before applying LOD
values to the results, thus proving a better data t when LOD
values were applied. The reason for this is that the QA procedure
reduced the uncertainties contributed by false negatives and
false positives, which occurs when pigments are quantied near
the method detection limit (Hooker et al., 2005). In the present
data set these pigments were secondary pigments like neox-
anthin, violaxanthin, MV Chl c
3
, and lutein. Instead of zero values
in the spread sheet, when such pigments could not be detected
and quantied, the LOD values applied caused that CHEMTAX
included these pigments in the calculations. For example when
low concentrations of prasinoxanthin were detected the acces-
sory pigments neoxanthin and lutein of prasinophytes Type
3 were often below the detection limit. Applying LOD values
instead of zeroes for these secondary pigments improved the
CHEMTAX calculations, thus causing a better data t.
4.4. Presence of other zeaxanthin containing organisms in the
southern Indian Ocean
Analyses of organisms not retained by the GF/F lters used to
lter the samples revealed that all autotrophic pico-sized algae
were in fact collected by the lters, since no Chl a was detected on
0.2 mm lters after passage of the GF/F lters. The 0.2 mm lters
did, nevertheless, retain some zeaxanthin-containing organisms,
probably a marine bacterium such as Paracoccus zeaxanthinifaciens
(formerly Flavobacterium; Berry et al., 2003). In a recent study in
Antarctic waters (Wright et al., 2009) high zeaxanthin concentra-
tions caused an unrealistic high contribution of cyanobacteria by
the CHEMTAX calculations, and bacteria rather than cyanobac-
teria were the most likely source to zeaxanthin. However, in the
Antarctic waters the bacteria were retained by GF/F lters
(Wright et al., 2009), indicating that the size of such pigmented
marine bacteria is variable or that the Antarctic bacteria were
attached to aggregates such as mucilage. If such larger sized
bacteria also were present in the Indian Ocean they would have
inuenced the CHEMTAX calculations and increased particularly
the biomass of the cyanobacteria, which was calculated from the
zeaxanthin concentration (Table 1). The zeaxanthin/Chl a ratios of
cyanobacteria obtained by the CHEMTAX calculations were 1.38
in the SW Indian Ocean, but 2.69 in the SE Indian Ocean (Table 1),
and the latter value is higher than ratios of high light treated cells
of Synechococcus sp. cultures (Schl uter et al., 2000; Henriksen
et al., 2002). Although zeaxanthin is a photo-protective pigment
and zeaxanthin/Chl a ratios did show 2 and 5 times changes as
function of depth/light intensity for cyanobacteria and prochlor-
ophytes, respectively (Table 1), the high zeaxanthin/Chl a ratio of
cyanobacteria in the SE Indian Ocean indicated that other zeax-
anthin containing cells may have been retained on the GF/F lters
too. CHEMTAX calculated a variable contribution of cyanobacteria
with an average of 12% (Table 2). In the oligotrophic parts of the
Pacic Ocean the biomass of Synechococcus was found to always
constitute less than 10% by ow cytometry (Campbell et al., 1994,
1997; Blanchot et al., 2001). Of the few studies of picoplankton
conducted in the oligotrophic parts of the Indian Ocean Not et al.
(2008) used ow cytometry to enumerate cells of Prochlorococcus
and Synechococcus. Although no biomass estimations were made
the cell numbers appear comparable to the results of Blanchot
et al. (2001) from the Pacic Ocean. A seasonal succession
towards prokaryote dominance during high temperatures and
irradiance in summer has been demonstrated across the global
ocean basins in the subtropical southern hemisphere (Barlow
et al., 2007). While the study of Not et al. (2008) was carried out
during late fall, this study was carried out during late spring and a
higher prokaryotic proportion of phytoplankton should be
expected from station 4 and onwards where oligotrophic condi-
tions prevailed. Unfortunately, no ow cytometry measurements
were conducted, and in some occasions the cyanobacteria con-
stituted up to 20% (Table 2) of the phytoplankton population,
which might indicate that zeaxanthin from non-photosynthetic
bacteria could have biased the biomass of cyanobacteria deter-
mined by CHEMTAX. The presence and size range of such
zeaxanthin containing bacteria that might interfere with deter-
mination of cyanobacteria by CHEMTAX denitely need special
attention in future investigations.
5. Conclusion
The difference of the pigment/Chl a ratios (Table 1) and the
large variety in the phytoplankton communities encountered
across the southern Indian Ocean reects the complexity of this
under-sampled ocean, which is inuenced by conuences of
water currents, upwelling and down-welling resulting in produc-
tive mixing areas to oligotrophic regions. The microscopy method
used in this study (inverted microscope) was providing only
limited, yet important, information on 45 mm algae. Valuable
information on the phytoplankton communities in the southern
Indian Ocean was obtained by combining the results from the two
phytoplankton identication methods. It could be deduced that
most of the dinoagellate community of unknown trophy in the
oligotrophic regions of the Indian Ocean were likely heterotrophic
species. For the different subtypes of haptophytes and pelago-
phytes the CHEMTAX setup could be justied by using the
information on diatom biomasses achieved by microscopy. In
order to obtain more information on the pico-sized phytoplank-
ton populations special microscopy methods are needed, which,
however, seldom are feasible when many samples need to be
analyzed. Since the pigment method certainly has limitations too
(e.g., Higgins et al., in press), and microscopic analyses also
introduce taxonomical misinterpretations even by competent
taxonomists (Culverhouse et al., 2003; Culverhouse, 2007), a
combination of several methods is warranted when examining
phytoplankton communities like those sampled in the southern
Indian Ocean.
L. Schl uter et al. / Deep-Sea Research I 58 (2011) 546556 554
Acknowledgments
The sampling was conducted during the third Danish Galathea
expedition. We thank the Captain of HMDS Vdderen, Carsten
Schmidt, and his crew for excellent assistance in connection with
our sampling. The project was supported by grants from Knud
Hjgaards Fond and the Danish Natural Sciences Research Coun-
cil. We are grateful to Simon Wright, Australian Antarctic Divi-
sion, for receiving CHEMTAX ver. 1.95, and to Jacob L. Hyer,
Danish Meteorological Institute for preparing Fig. 1. The present
work was carried out as part of the Galathea3 expedition under
the auspices of the Danish Expedition Foundation. This is
Galathea3 contribution no. P77.
References
Andersen, R.A., Saunders, G.W., Paskind, M.P., Sexton, J.P., 1993. Ultrastructure and
18S rRNA gene sequence for Pelagomonas Calceolata Gen. et sp. Nov. and the
description of a new algal class, the Pelagophyceae classis. Nov. J. Phycol. 29,
701715.
Andersen, R.A., Bidigare, R.R., Keller, M.D., Latasa, M., 1996. A comparison of
HPLC pigment signatures and electron microscopic observations for oligo-
trophic waters of the North Atlantic and Pacic Oceans. Deep-Sea Res. II 43,
517537.
Ansotegui, A., Sabrobe, A., Trigueros, J.M., Urrutxurtu, I., Orrive, E., 2003. Size
distribution of algal pigments, and phytoplankton assemblages in a coastal-
estuarine environment, contribution of small eukaryotic algae. J. Plankton Res.
25, 341355.
Barlow, R., Stuart, V., Lutz, V., Sessions, H., Sathyendranath, S., Platt, T., Kyewa-
lyanga, M., Clementson, L., Fukasawa, M., Watanabe, S., Devred, E., 2007.
Seasonal pigment patterns of surface phytoplankton in the subtrophical
southern hemisphere. Deep-Sea Res. I, 16871703.
Berry, A., Janssens, D., H+ umbelin, M., Jore, J.P.M., Hoste, B., Cleenwerck, I.,
Vancanneyt, M., Bretzel, W., Mayer, A.F., Lopez-Ulibarri, R., Shanmugam, B.,
Swings, J., Pasamontes, L., 2003. Paracoccus zeaxanthinifaciens sp. nov., a
zeaxanthin-producing bacterium. Int. J. Syst. Evol. Microbiol. 53, 231238.
Bidigare, R.R., Ondrusek, M.E., 1996. Spatial and temporal variability of phyto-
plankton pigment distribution in the central equatorial Pacic Ocean. Deep-
Sea Res. II 43, 809833.
Blanchot, J., Andre , J.-M., Navarette, C., Neveux, J., Radenac, M.-H., 2001. Picophy-
toplankton in the equatorial Pacic, vertical distributions in the warm pool
and in the high nutrient low chlorophyll conditions. Deep-Sea Res. 48,
297314.
Boeckel, B., Baumann, K., Henrich, R., Kinkel, H., 2006. Coccolith distribution
patterns in South Atlantic and Southern Ocean surface sediments in relation to
environmental gradients. Deep-Sea Res. I 53, 10731099.
Campbell, L., Nolla, H.A., Vaulot, D., 1994. The importance of Prochlorococcus to
community structure in the central North Pacic Ocean. Limnol. Oceanogr. 39,
954961.
Campbell, L., Liu, H., Nolla, H.A., Vaulot, D., 1997. Annual variability of phyto-
plankton and bacteria in the subtropical North Pacic Ocean at station ALOHA
during the 19911994 ENSO event. Deep-Sea Res. I 44, 167192.
Claustre, H., 1994. The trophic status of various oceanic provinces as revealed by
phytoplankton pigment signatures. Limnol. Oceanogr. 39, 12061210.
Culverhouse, P.F., Williams, R., Reguera, B., Herry, V., Gonzalez-Gil, S., 2003. Do
experts make mistakes? A comparison of human and machine identication of
dinoagellates. Mar. Ecol. Prog. Ser. 247, 1725.
Culverhouse, P.F., 2007. Human and machine factors in algae monitoring perfor-
mance. Ecol. Informatics 2 (4), 361366.
De Salas, M.F., Bolch, C.J.S., Botes, L., Nash, G., Wright, S.W., Hallegraeff, G.M., 2003.
Takayama gen. Nov. (Gymnodiniales, Dinophyceae), a new genus of unar-
mored dinoagellates with sigmoid apical grooves, including the description
of two new species. J. Phycol. 39, 12331246.
Dez, B., Pedros Alio , C., Massana, R., 2001. Study of genetic diversity of eukaryotic
picoplankton in different oceanic regions by small-subunit rRNA gene cloning
and sequenching. Appl. Environ. Microbiol. 67, 29322941.
Farnelid, H., Tarangkoon, W., Hansen, G., Hansen, P.J., Riemann, L., 2010. Putative
N-2-xing heterotrophic bacteria associated with dinoagellate-cyanobacteria
consortia in the low-nitrogen Indian Ocean. Aquat. Microb. Ecol. 61 (2),
105117.
Fogg, G.E., 1991. The phytoplanktonic ways of life. New Phytol. 118, 191232.
Gravalosa, J.M., Flores, J.-A., Sierro, F.J., Gersonde, R., 2008. Sea surface distribution
of coccolithophores in the eastern Pacic sector of the Southern Ocean
(Bellingshausen and Amundsen Seas) during the late austral summer of
2001. Marine Micropaleontol. 69, 1625.
Gibb, S.W., Barlow, R.G., Cummings, D.G., Rees, N.W., Trees, C.C., Holligan, P.,
Suggett, D., 2000. Surface phytoplankton pigment distribution in the Atlantic
Ocean, an assessment of basin scale variability between 501S. Prog. Oceanogr.
45, 339368.
Gibb, S.W., Cummings, D.G., Irigoien, X., Barlow, R.G., Mantoura, R.F.C., 2001.
Phytoplankton pigment chemotaxonomy of the northeastern Atlantic. Deep-
Sea Res. II 48, 795823.
Goericke, R., 1998. Response of phytoplankton community structure and taxon-
specic growth rates to seasonally varying physical forcing in the Sargasso Sea
off Bermuda. Limnol. Oceanogr. 43, 921935.
Hanson, C.E., Waite, A.W., Thompson, P.A., Pattiaratchi, C.B., 2007. Phytoplankton
community structure and nitrogen in Leeuwin Current and coastal waters of
the Gascoyne region of Western Australia. Deep-Sea Res. II 54, 902924.
Havskum, H., Schl uter, L., Scharek, R., Berdalet, E., Jacquet, S., 2004. Routine
quantication of phytoplankton groupsmicroscopy or pigment analyses?
Mar. Ecol. Prog. Ser. 273, 3142.
Henriksen, P., Riemann, B., Kaas, H., Srensen, H.M., Srensen, H.L., 2002. Effects
of nutrient-limitation and irradiance on marine phytoplankton pigments.
J. Plankton Res. 24, 835858.
Higgins, H.W., Mackey, D.J., 2000. Algal class abundances, estimated from
chlorophyll and carotenoid pigments, in the western Equatorial Pacic under
El Nino and non-El Nino conditions. Deep-Sea Res. I 47, 14611483.
Higgins, H., Wright, S., Schl uter, L., Quantitative interpretation of chemotaxonomic
pigment data. In: Roy, S., et al. (Eds.), Phytoplankton Pigments in Oceano-
graphy, Guidelines to Modern Methods. UNESCO Publishing, Paris (Chapter 5),
in press.
Hooker, S.B., Van Heukelem, L., Thomas, C.S., Claustre, H., Ras, J., Schl uter, L., Perl, J.,
Trees, C., Stuart, V., Head, E., Barlow, R., Sessions, H., Clementson, L., Fishwick, J.,
Llewellyn, C., Aiken, J., 2005. The Second Sea-WiFS HPLC Analysis Round-Robin
Experiment (SeaHARRE-2). NASA Tech. Memo. 2005212785, NASA Goddard
Space Flight Center, Greenbelt, Maryland.
Irigoien, X., Meyer, B., Harris, R., Harbour, D., 2004. Using HPLC pigment analysis to
investigate phytoplankton taxonomy, the importance of knowing your species.
Helgol. Mar. Res. 58, 7782.
Jo nasdo ttir, S.H., Satapoomin, S., Friis-Mller, E., Andersen, M.B., Jakobsen, H.H.,
Nielsen, T.G., Biological Oceanography across Southern Indian Ocean. The
micro- and mesozooplankton abundance, composition and production, in
preparation.
Jeffrey, S.W., Wright, S.W., Zapata, M., 1999. Recent advantages in HPLC pigment
analysis of phytoplankton. Mar. Freshwater Res. 50, 879896.
Lipsen, M.S., Crawford, D.W., Gower, J., Harrison, P.J., 2007. Spatial and temporal
variability in coccolithophore abundance and production of PIC and POC in the
NE subarctic Pacic during El Nin~ o (1998), La Nin~ a (1999) and 2000. Prog.
Oceanogr. 75, 304325.
Mackey, M.D., Mackey, D.J., Higgins, H.W., Wright, S.W., 1996. CHEMTAXa
program, for estimating class abundance from chemical markers, app-
lication to HPLC measurements of phytoplankton. Mar. Ecol. Prog. Ser. 144,
265283.
Marty, J.-C., Chiave rini, J., Pizay, M.-D., Avril, B., 2002. Seasonal and interannual
dynamics of nutrients and phytoplankton pigments in the western Mediter-
ranean Sea at the DYFAMED time-series station (19911999). Deep-Sea Res. II
49, 19651985.
Marty, J.-C., Garcia, N., Raimbault, P., 2008. Phytoplankton dynamics and primary
production under late summer conditions in the NW Mediterranean Sea.
Deep-Sea Res. I 55, 11311149.
Menden-Deuer, S., Lessard, E.J., 2000. Carbon to volume relationships for dino-
agellates, diatoms, and other protist plankton. Limnol. Oceanogr. 45 (3),
569579.
Moon-van der Staay, S.Y., De Wachter, R., Vaulot, D., 2001. Oceanic 18S rDNA
sequences from picoplankton reveal unsuspected eukaryotic diversity. Nature
409, 607610.
Not, F., Latasa, M., Scharek, R., Viprey, M., Karleskind, P., Balague , V., Ontoria-
Oviedo, I., Cumino, A., Goetze, E., Vaulot, D., Massana, R., 2008. Protistan
assemblages across the Indian Ocean, with specic emphasis on the picoeu-
caryotes. Deep-Sea Res. I 55, 14561473.
Pollard, R.T., Venables, J.T., Read, J.T., Allen, J.T., 2007. Large-scale circulation
around the Crozet Plateau controls an annual phytoplankton bloom in the
Crozet Basin. Deep-Sea Res. II 54, 19151921.
Rodrguez, F., Varela, M., Zapata, M., 2002. Phytoplankton assemblages in the
Gerlache and Branseld Straits (Antarctica Peninsula) determined by light
microscopy and CHEMTAX analysis of HPLC pigment data. Deep-Sea Res. II 49,
723747.
Sakamoto, C.M., Karl, D.M., Jannasch, H.W., Bidigare, R.R., Letelier, R.M., Walz, P.M.,
Ryan, J.P., Polito, P.S., Johnson, K.S., 2004. Inuence of Rossby waves on
nutrient dynamics and the plankton community structure in the North Pacic
subtropical gyre. J. Geophys. Res. 109, C05032. doi:10.1029/2003JC001976.
Schl uter, L., Mhlenberg, F., Havskum, H., Larsen, S., 2000. The use of phytoplank-
ton pigments for identifying and quantifying phytoplankton groups in coastal
areas, testing the inuence of light and nutrients on pigment/chlorophyll a
ratios. Mar. Ecol. Prog. Ser. 192, 4963.
Schl uter, L., Lauridsen, T.L., Krogh, G., Jrgensen, T., 2006. Identication and
quantication of phytoplankton groups in lakes using new pigment ratios
a comparison between pigment analysis by HPLC and microscopy. Freshwater
Biol. 51, 14741485.
Schl uter, L., Mhlenberg, 2003. Detecting presence of phytoplankton groups with
non-specic pigment signatures. J. Appl. Phycol. 15, 465476.
Siegel, H., Ohde, T., Gerth, M., Lavik, G., Leipe, T., 2007. Identication of
coccolithophore blooms in the SE Atlantic Ocean off Namibia by satellites
and in-situ methods. Continental Shelf Res. 27, 258274.
Sournia, A., 1978. Phytoplankton Manual. UNESCO Publishing, Paris.
L. Schl uter et al. / Deep-Sea Research I 58 (2011) 546556 555
Steinberg, D.K., Carlson, C.A., Bates, N.R., Johnson, R.J., Michaels, A.F., Knap, A.H.,
2001. Overview of the US JGOFS Bermuda Atlantic Time-series Study (BATS), a
decade-scale look at ocean biology and biogeochemistry. Deep-Sea Res. II 48,
14051447.
Suzuki, K., Minami, C., Liu, H., Saino, T., 2002. Temporal and spatial patterns of
chemotaxonomic algal pigments in the subarctic Pacic and the Bering Sea
during the early summer of 1999. Deep-Sea Res. II 49, 56855704.
Tarangkoon, W., Hansen, G., Hansen, P.J., 2010. Spatial distribution of symbiont-
bearing dinoagellates in the Indian Ocean in relation to oceanographic
regimes. Aquat. Microb. Ecol. 58 (2), 197213.
Van Heukelem, L., Thomas, C., 2001. Computer assisted high-performance liquid
chromatography method development with applications to the isolation and
analysis of phytoplankton pigments. J. Chromatogr. A 910, 3149.
Veldhuis, M.J.W., Kraay, G.W., 2004. Phytoplankton in the subtropical Atlantic
Ocean, towards a better assessment of the biomass and composition. Deep-Sea
Res. I 51, 507530.
Visser, A.W., Nielsen, T.G., Markager, S., Middelboe, M., Hyer, J.L., Biological
oceanography across the southern Indian Ocean: oceanography and the base
of the pelagic food web. Deep-Sea Res. I, submitted.
Wright, S.W., Thomas, D.P., Marchant, H.J., Higgins, H.W., Mackey, M.D., Mackey,
D.J., 1996. Analysis of phytoplankton in the Australian sector of the Southern
Ocean: comparisons of microscopy and size frequency data with interpreta-
tions of pigment HPLC data using the CHEMTAX matrix factorisation
program. Mar. Ecol. Prog. Ser. 144, 285298.
Wright, S.W., Jeffrey, S.W., 2006. Pigment markers for phytoplankton production.
In: Volkman, John K. (Ed.), Marine Organic Matter, Biomarkers, Isotopes and
DNA Series. The Handbook of Environmental Chemistry, vol. 2, Reactions and
Processes, Part 2N. Springer-Verlag.
Wright, S.W., Ishikawa, A., Marchant, H.J., Davidson, A.T., van den Enden, R.L., Nash,
G.V., 2009. Composition and signicance of picophytoplankton in Antarctic
waters. Polar Biol. 32, 797808.
Zapata, M., Jeffrey, S.W., Wright, S.W., Rodrguez, F., Garrido, J.L., Clementson, L.,
2004. Photosynthetic pigments in 37 species (65 strains) of Haptophyta,
implications for oceanography and chemotaxonomy. Mar. Ecol. Prog. Ser.
270, 83102.
Zapata, M., 2005. Recent advances in pigment analysis as applied to picophyto-
plankton. Vie et Milieu 55 (34), 233248.
L. Schl uter et al. / Deep-Sea Research I 58 (2011) 546556 556

Potrebbero piacerti anche