Download pdf

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Genomic library wikipedia , lookup

Copy-number variation wikipedia , lookup

Gene therapy wikipedia , lookup

Primary transcript wikipedia , lookup

Genomic imprinting wikipedia , lookup

Cre-Lox recombination wikipedia , lookup

Biology and consumer behaviour wikipedia , lookup

Genetic engineering wikipedia , lookup

Gene nomenclature wikipedia , lookup

Minimal genome wikipedia , lookup

Gene expression programming wikipedia , lookup

Nutriepigenomics wikipedia , lookup

Transposable element wikipedia , lookup

Ridge (biology) wikipedia , lookup

Genomics wikipedia , lookup

Epigenetics of human development wikipedia , lookup

Genome (book) wikipedia , lookup

Pathogenomics wikipedia , lookup

Human genome wikipedia , lookup

Genome evolution wikipedia , lookup

Sequence alignment wikipedia , lookup

Vectors in gene therapy wikipedia , lookup

History of genetic engineering wikipedia , lookup

Non-coding DNA wikipedia , lookup

Gene desert wikipedia , lookup

Microsatellite wikipedia , lookup

Gene wikipedia , lookup

Therapeutic gene modulation wikipedia , lookup

Computational phylogenetics wikipedia , lookup

Gene expression profiling wikipedia , lookup

Genome editing wikipedia , lookup

RNA-Seq wikipedia , lookup

Point mutation wikipedia , lookup

Microevolution wikipedia , lookup

Designer baby wikipedia , lookup

Helitron (biology) wikipedia , lookup

Site-specific recombinase technology wikipedia , lookup

Metagenomics wikipedia , lookup

Artificial gene synthesis wikipedia , lookup

Transcript
Distribution and Relative Quantification of Key Genes Involved in Fixed
Nitrogen Loss From the Arabian Sea Oxygen Minimum Zone
Amal Jayakumar
Department of Geosciences, Princeton University, Princeton, New Jersey, USA
S. Wajih A. Naqvi
National Institute of Oceanography, Council of Scientific and Industrial Research, Dona Paula, India
Bess B. Ward
Department of Geosciences, Princeton University, Princeton, New Jersey, USA
The Arabian Sea (AS) oxygen minimum zone (OMZ) is one of the largest
pelagic low-oxygen environments in the open ocean. It is responsible for the
removal of up to 60 Tg of nitrogen annually, roughly an eighth of the global fixed
nitrogen sink. Although denitrification has long been believed to be the major
process responsible for fixed nitrogen loss from the oceans, recent studies show
that anaerobic ammonium oxidation (anammox) is potentially a more important
process involved. We have investigated the phylogeny of both anammox and
denitrifying microbes in the AS, and here we report on their diversity in terms of
their characteristic genes. Denitrifiers were targeted using nirS and nirK genes and
anammox bacteria with the 16S rRNA gene. The nirK gene was amplified from
all the samples, but nirS gene could only be detected when nitrite was present.
The distribution of phylotypes was related to the concentration of nitrite and the
apparent stage of denitrification. Most nirK or nirS genes from the AS had low
identities with other published sequences. The closest identities were to sequences
from other water column denitrifying environments rather than sedimentary or
terrestrial environments. Phylogenetic analysis of the nirS and nirK genes revealed
overall lower diversity than the very high diversities reported from estuarine and
sedimentary environments. 16S rRNA partial gene sequences revealed very limited
diversity among anammox sequences. All the anammox sequences were ≥98%
identical to each other and were similar to sequences from other marine and water
column environments, which are distantly related to Scalindua sp. Quantification
Indian Ocean Biogeochemical Processes and Ecological Variability
Geophysical Monograph Series 185
Copyright 2009 by the American Geophysical Union.
10.1029/2008GM000730
187
188 DISTRIBUTION AND RELATIVE QUANTIFICATION OF KEY GENES
using quantitative polymerase chain reaction assays showed that both nirS genes
and anammox 16S rRNA genes were more abundant at depths with higher nitrite
concentrations. The presence and abundance of genes indicative of both processes
suggest that both canonical denitrification and anammox likely occur in the OMZ
of the AS. However, the abundance of the nirS gene, one of the genes responsible
for canonical denitrification, was an order of magnitude higher than the abundance
of the anammox 16S gene, indicating that denitrifiers are numerically dominant in
this environment. The greater diversity of nirS and nirK genes relative to anammox
genes both among and within stations also suggests that the denitrifier assemblages
are more dynamic in response to environmental conditions.
1. Introduction
Nitrogen is a key element in biological processes and is
often limiting in marine ecosystems. This essential nutrient
is efficiently recycled within ecosystems, and fresh input occurs mainly through nitrogen fixation. Oceans lose >400 Tg
N of fixed nitrogen every year, mainly in low-oxygen environments [Codispoti, 2007]. Low-oxygen pelagic environments are responsible for ~35% of the marine-fixed nitrogen
loss through microbially catalyzed reductive processes [Codispoti, 2007] and are hot spots for intense turnover of nitrous oxide, a potent greenhouse gas. The oxygen minimum
zones (OMZ) contain a large volume of water with oxygen
levels marginally above 1 µM, the threshold value thought
to induce dissimilatory nitrate reduction [Naqvi et al., 2003;
Codispoti et al., 2005]. Minor perturbations of the delicate
balance between oxygen consumption and supply caused by,
for example, an increase in organic matter loading may result
in large changes in the fixed N loss. Such shifts might occur
through global warming-induced changes in circulation patterns or via input of excess nitrogen or organic carbon resulting from fertilizer runoff or atmospheric deposition [Duce et
al., 2008].
The Arabian Sea (AS) OMZ is one of the largest OMZs
in the open ocean and alone could be responsible for the removal of up to 60 Tg of nitrogen annually [Codispoti, 2007].
The region is distinguished by very low dissolved oxygen
(DO) concentration at depths between 150 and 1200 m, especially east of 60oE and north of 12oN. Nitrogen transformations appear to be most intense between 150 and 600 m.
In this zone, net loss of combined nitrogen occurs on a
perennial basis, as indicated by the persistent presence of
a broad secondary nitrite maximum (SNM) and large fixed
N deficits [Naqvi, 1991]. Microorganisms capable of denitrification and anaerobic ammonium oxidation (anammox)
both convert oxidized forms of nitrogen to gaseous species
and thereby cause net fixed N loss. The relative importance
of anammox and conventional denitrification in the removal
of fixed nitrogen in the AS is not known. There have been
no systematic studies in this region targeting the distribution, diversity, and abundance of organisms capable of these
conversions.
Knowledge of the diversity, distribution, and relative
abundance of denitrifiers and anammox bacteria will help
to determine what mechanisms are responsible for the observed net nitrogen deficits. Such knowledge is also essential for understanding microbial community structure and its
dynamics in natural assemblages, as well as the response of
microbes to environmental factors. It is also not known how
the denitrifying and anammox communities in the OMZ of
the AS are related to their counterparts in similar environments elsewhere, such as OMZs in the eastern tropical North
and South Pacific. If each OMZ harbors a unique microbial
assemblage, community composition might be related to
differences in N transformations and fluxes in the different
OMZs.
The key step in the denitrification process is the reduction of nitrite, which is the rate-limiting step in the transformation of nitrate to dinitrogen gas [Zumft, 1997], since
the downstream products are gases that cannot be readily
assimilated. Nitrite reduction is carried out by two enzymes,
one that requires copper and is encoded by the gene nirK,
and one that requires iron and is encoded by the gene nirS.
The two enzymes apparently are functionally and physiologically equivalent [Zumft, 1997]. The nirK gene from
Pseudomonas auerofaciens can be expressed in a mutant of
Pseudomonas stutzeri that lacks the nirS gene [Glockner et
al., 1993], but the two enzymes do not usually occur in the
same organism.
Denitrifying bacteria are taxonomically diverse and are
found in over 50 genera. The nitrite reductase genes appear
to have spread by horizontal gene transfer, as their phylogeny is not congruent with that of the 16S rRNA genes [Song
and Ward, 2003]. Therefore, a 16S rRNA-based approach
to studying the diversity of microbial assemblages does not
provide reliable information on their denitrification capabil-
JAYAKUMAR et al. 189
ity. The nirS gene is found nearly exclusively in proteobacteria, but recently has been reported from planctomycetes
[Strous et al., 2006], while the nirK gene has been reported
from many diverse and unrelated taxa [Zumft, 1997]. There
is minimal information available on functional genes for
anammox, but 16S rRNA reliably differentiates this unique
functional group [Schmid et al., 2005]. Here we present
data on the diversity, distribution, and abundance of the two
marker genes, nirS and nirK, for denitrifying organisms and
the 16S rRNA genes for anammox, in order to evaluate their
involvement in the major pathways for the removal of fixed
nitrogen from the OMZ of the AS.
2. Materials and Methods
Samples were collected on board ORV Sagar Kanya from
five stations during two cruises, SK208 and SK209, in August–
September 2004 (Figure 1). The maximum water column
depth at stations 1, 2, and 23 was 2000–4000 m, while station
K03 was 40 m and station 17 was 65 m deep, respectively.
Samples for nutrients and DNA analysis were collected using a Sea-Bird Electronics conductivity-temperature-depth
(CTD)-rosette sampling system, Model SBE 9, fitted with
Niskin/Go-flo bottles. Nutrients were analyzed using a SKALAR autoanalyzer following standard methods [Grasshoff et
al., 1983] and DO was estimated using the Winkler method
[Carpenter, 1965]. For DNA analysis, 5–10 L of seawater
Figure 1. Station locations for Sagar Kanya cruise SK208 and
SK209 September–October 2004. K03 is offshore of Karwar, G8
and V4 are station locations from the previously published coastal
study [Jayakumar et al., 2004]. Contour line encloses the region in
the Arabian Sea (AS) where nitrite concentrations exceed 0.5 µM
within the OMZ region [Naqvi, 1991].
was filtered onto Sterivex capsules (0.2 µm pore size, Millipore, Inc., Bedford, MA, USA) with a peristaltic pump, quick
frozen in liquid nitrogen and stored at −80°C until the DNA
could be extracted. DNA analyses are reported for samples
from the SNM for the open-ocean locations. The sampling
depth was determined after a quick analysis of the depth distribution of nitrite concentration in the water column from an
initial cast. For the coastal stations, DNA samples were collected from the depth at which the DO was low; nitrite was
not present at the sample depth at either coastal location.
DNA extraction, polymerase chain reaction (PCR) amplification, and clone library construction methods for nirS
gene [Jayakumar et al., 2004] and nirK gene [Santoro et
al., 2006] have been described previously. Anammox 16S
rRNA genes were amplified using primers AMX368F and
AMX820R [Schmid et al., 2003]. Clones were selected
randomly from the clone libraries and amplified using T7M13 primers. The products were used as templates for cycle
sequencing from both ends with Big DyeTM V.3.1 terminator chemistry (Applied Biosystems) with the same primers
using an ABI 310 or ABI3100 Genetic Analyzer (Applied
Biosystems). Sequences were edited using Sequencher
4.1TM (Genetic Codes) and assembled. Assemblies of 16S
rRNA anammox genes and translated amino acid sequences
for nirS and nirK genes were then aligned using ClustalX
[Thompson et al., 1997]. Neighbor-joining trees were produced using distance matrix methods (PAUP 4.0, Sinauer
Associates), and the reliability of the phylogenetic reconstructions (1000 replicates) was evaluated by bootstrapping.
For descriptive purposes, nirS and nirK gene clusters were
identified as groups containing sequences with >90% identity at DNA level, although some clusters contained slightly
more divergent sequences, and cluster designations are
therefore somewhat subjective at a few branch points. The
clone numbering includes the station number as the prefix,
and depth is not specified in the clone name because data
from only one depth per station are reported.
Analytical rarefaction 1.3TM software [Holland, 2001]
was used to compute the rarefaction statistics [Heck et al.,
1975]. nirS and nirK operational taxonomic units (OTUs)
were defined as DNA sequences >90% identical. nirS data
from stations G8 and V4 (Figure 1) from a previous study
in the coastal AS [Jayakumar et al., 2004] have been included for comparison with data from the open-ocean environment. OTUs for 16S rRNA sequences were defined at
2% difference in identity. Anammox 16S rRNA partial gene
sequences reported previously [Jayakumar et al., 2009] are
included in the larger environmental data set analyzed here.
The sequences reported in this study have been deposited
in GenBank under the following accession numbers: nirK,
FJ596039–FJ596098; nirS, FJ596099–FJ596141.
190 DISTRIBUTION AND RELATIVE QUANTIFICATION OF KEY GENES
DNA was quantified using PicoGreen (Molecular Probes),
and quantitative PCR (QPCR) analyses were performed on an
Mx3000p real-time PCR system (Stratagene) using a QuantiTect SYBR Green PCR master mix kit (Qiagen). Three gene
fragments were quantified: total nirS (Tot nirS), a nirS gene
sequence that was dominant in the station 1 clone library (Dom
nirS) [Jayakumar et al., 2009], and a fragment of the anammox16S rRNA gene. DNA from the environmental samples,
negative controls and standards were all run in triplicate in the
same QPCR assay. Standards were prepared by amplifying a
constructed plasmid containing the respective gene. Both Tot
nirS and Dom nirS regions are present in the cloned fragment
obtained with primers nirS1F and nirS6R [Braker et al., 1998].
The same anammox 16S rRNA gene fragment used for phylogenetic analysis, obtained with the primers AMX 368F and
AMX820R, was used to construct the standard plasmid for
anammox. Standard curves were used to determine the copy
number of Tot nirS/Dom nirS/anammox 16S rRNA genes.
Standard curves were quite reproducible, and results were included only from replicate assays that yielded reproducible
standard curves. All assays used the same annealing temperature, 62 oC, and were performed on the same plates to maximize
reproducibility. To quantify Tot nirS gene, primers nirS1F and
nirS3R (100 pM) [Braker et al., 1998] were used to target a
260-bp conserved region within the nirS1F-nirS6R region. For
Dom nirS, primers (50 pM) targeted a different 280-bp region
that was conserved only within the dominant nirS clone sequences (forward primer nirSASDom30F 5′-CGATGGTCTTACGCTCGAGCC-3′; reverse primer nirSASDom310R
5′-GAGATCGACGATCACCATCG-3′). Clones representing
all clusters from total nirS phylogeny were tested to verify the
specificity of the Dom nirS primers.
Primer concentrations for the QPCR assays were optimized, and specificity of the QPCR for the production of
only one product was assured by performing a melting curve
analysis at the end of 40 cycles. Gel electrophoresis of the
QPCR products further confirmed that only one product of
the right size was amplified. nirS, nirK, and anammox amplifications for clone library construction and QPCR were all
done from the same DNA samples for each location. Clone
library analysis using primers AMX 368F and AMX820R
detected only anammox 16S rRNA gene sequences, confirming that the QPCR analysis was also specific for anammox bacteria.
3. Results and Discussion
3.1. Hydrography of the Region
The AS is enclosed by land on three sides. Its unusual circulation pattern and large supply of organic material from
the productive surface waters combine to produce an intense
and perennial OMZ. High primary productivity in the AS
occurs during both the NE and SW monsoon seasons [Barber et al., 2001]. The sinking and decomposition of the locally produced and laterally advected organic carbon from
the highly productive upwelling regions culminates in excessive consumption of DO and its near depletion at midwater depths (Figure 2) [Naqvi et al., 2003]. The classical
Winkler method is not sufficiently sensitive to determine the
exact concentration of DO in these waters. However, a more
sensitive colorimetric method suggests that DO concentrations are <1 µM when sequential reduction of nitrate is induced, and nitrite accumulates in OMZ waters [Morrison et
al., 1999; Naqvi et al., 2003].
Based on our measurements, DO concentrations were
<10 µM between 125 and 1000 m in the northern locations
(stations 1 and 2) and between 200 and 800 m in the southern location (station 23, Figure 2). The SNM occurred within
a narrow band within this zone (≥0.5 µM nitrite) [Naqvi,
1991]. The SNM was much thicker (125–400 m) in the
northeastern region than toward the west and south, where it
was centered around 250 m. Within the core of the OMZ, the
concentration of nitrite varied from traces to a maximum of
5.2 µM at 250 m at station 1. Direct measurements of denitrification rates in the OMZs are sparse [Devol et al., 2006],
and hence, the concentration of nitrite in these environments
is often considered as an indicator of high microbial activity
[Jayakumar et al., 2009]. Although the SNM is a perennial
feature of the NE AS, it is not clear how the organic substrate demand is met during the less productive seasons.
The nearly complete DO depletion coupled with high concentrations of nitrate provides an ideal niche for organisms
that can respire nitrate. High rates of decomposition of organic matter sinking from the productive surface waters lead
to high bacterial abundances, not only in the surface waters
[Ducklow et al., 2001] but also within the OMZ. During the
monsoon seasons, nitrite concentrations increase in the SNM
of the AS [Morrison et al., 1999], and the highest bacterial
cell numbers in the SNM are comparable to those in surface
waters [Jayakumar et al., 2009]. We suggest that this pattern
represents a bloom of denitrifying organisms.
The hydrographic characteristics of individual stations are
reflected in the composition of the clone libraries described
below. The sample from station 1 came from the core of
the OMZ, where denitrifying conditions prevail throughout the year, while stations 2 and 23 were located toward
the periphery of the perennial OMZ as mapped by Naqvi
[1991]. Station 23 may experience suboxia seasonally, similar in intensity to station 1, due to the phytoplankton blooms
caused by the offshore advection of water upwelling off
Oman, which reach this location during the SW monsoon
JAYAKUMAR et al. 191
Figure 2. Distribution of dissolved nitrite ( µM), nitrate (µM), and DO (µM) from SK209 four stations (1, 2, 17, 23), SK208
one station (K03), and Sagar Sukti-1 one station (G8) [Jayakumar et al., 2004]. Locations: station 1, 17°N 68°E; station 2,
19.75°N 64.62°E; station 17, 18.3°N 57.5°E; station 23, 15°N 64°E; station K03, 14.4°N 74°E; station G8, 15.5°N 73.4°E.
192 DISTRIBUTION AND RELATIVE QUANTIFICATION OF KEY GENES
(S. W. A. Naqvi, unpublished data, 2004). Station 2 may
sometimes experience nondenitrifying conditions when the
Persian Gulf water penetrates the depth of the core of
the OMZ. Although this high salinity water originating in
the shallow coastal Persian Gulf quickly loses its initially
high DO, it is nevertheless a significant source of variability
of denitrification in the northwestern AS, as even the small
amounts of oxygen introduced by it can inhibit the process
[Naqvi and Jayakumar, 2000]. Due to the complex circulation pattern in the region, which is dominated by mesoscale
eddies [Kim et al., 2001], the supply of Persian Gulf Water
near the location of station 2 appears to be patchy, and its
age variable, such that the salinity maximum that characterizes this water mass is sometimes associated with the SNM,
while at other times, it is not (K. Banse et al., manuscript in
preparation 2009). Thus, waters in the open AS represent different stages of denitrification at different times, as indicated
by varying concentrations of nitrate, nitrite, and DO [Naqvi,
1991; Morrison et al., 1999]. Individual stations likely undergo temporal changes in the intensity of denitrification, and
when sampled on one cruise, they represent both spatial and
temporal variability in the apparent stage of denitrification
[Jayakumar et al., 2009].
3.2. nirS Phylogeny
nirS gene was amplified from open-ocean stations 1, 2,
and 23, where nitrite was present. nirS gene could not be
amplified from coastal stations 17 and K03, where nitrite
was absent. Previously, nirS gene was also amplified from
coastal locations, but only where nitrite was present [Jayakumar et al., 2004]. The NirS amino acid sequences from the
AS (Plate 1) show between 44% and 99% identity with previously published database sequences. Exclusion of the most
divergent sequences from this analysis (45 sequences in
cluster V) narrows the range to 70–99%. Phylogenetic analysis placed the 175 NirS sequences into seven major clusters
(Plate 1), four of which (clusters I, II, V, VIII) included sequences from all three stations. All of the clusters contained
sequences from at least one of the open ocean stations. The
nirS denitrifying assemblages in the open ocean were quite
different, however, from those in the coastal AS [Jayakumar
et al., 2004], where diversity was also much higher. Most
sequences from the open AS (>80%) had <75% identity at
the DNA level to the coastal AS sequences, suggesting that
the two environments harbored different communities, despite their geographical proximity. Twenty-seven nirS gene
sequences, all from Station 2, from the open-ocean OMZ
had ≥94% identity to some coastal AS sequences (in clusters
VI and VII). Some nirS gene sequences from the open AS
were ≥94% identical to sequences from the Baltic Sea [Han-
nig et al., 2006], Black Sea [Oakley et al., 2007], the OMZ
of the eastern South Pacific (ETSP) [Castro-Gonzalez et al.,
2005], sediment of the NW Pacific [Braker et al., 2000], and
River Colne Estuary in the United Kingdom [Nogales et al.,
2002].
The lowest nirS gene diversity was found at station 1,
where the nitrite concentration was the highest, and denitrification was presumably more advanced. The sequences
from this location were distributed only in four of the seven
clusters. For sequence types in clusters I and II, the closest
identities (>70% at amino acid level) found in the database
were other published AS sequences from the coastal environment and one mRNA clone sequence from Colne Estuary, Anis-62 [Nogales et al., 2002]. Cluster II contains 29%
of the clones from the three open-ocean nirS gene libraries
and represents a major nirS phylotype in the open ocean.
These two clusters also contained sequences from the Baltic Sea, indicating that some of these types are cosmopolitan. The sequences of cluster I were the dominant sequence
type found in the coastal AS, but accounted for <15% of the
clones from all three libraries from the open ocean. Thus,
different sequence types apparently attain dominance in different environments.
Three of the seven clusters, (III, IV, and VI) contained
sequences present only in clone libraries from the edges of
the permanent OMZ locations, and these groups were not
detected in the core of the OMZ at station 1. It is possible
that these denitrifying phylotypes are overgrown by other
denitrifiers that are better adapted to suboxic conditions, as
at station 1, and their relative abundance decreases, making
it rare in the clone library.
Cluster V contained sequences from all the open-ocean
stations of the AS, coastal AS, Baltic Sea, and also sequences from the continental margin sediments of the NW
Pacific. The sequences in this cluster are quite distinct from
sequences in other clusters. The amplicons in cluster V were
only ~830 bp DNA, while the rest are around ~890 bp. The
identity of the cluster V sequences to other clusters was
only ~40% at the amino acid level. This particular cluster
is abundant in the AS clone library, but these short amplicons were detected only rarely in libraries from Washington
Coast sediment [Braker et al., 2000]. For example, clone
WA20 is most similar to our cluster V (81% identity). Sequences in this dominant cluster are very different from the
other database NirS sequences, and the alignments with the
other 890 bp gene sequences have many gaps. Cluster V is
the second major group in the AS open ocean libraries, containing 25% of the sequences. At station 1, however, these
sequences dominate and account for 51% of the clones from
this location. The sequences from this study that groups in
Cluster V were very similar to each other, varying by <1%
JAYAKUMAR et al. 193
Plate 1. Phylogenetic relationships among translated nirS sequences (270 amino acids) from the water column of the AS
with Paracoccus denitrificans and Paracoccus pantotrophus as outgroups. Bootstrap values (>50%) are indicated at the
branch points. Sequence names of nirS clones from the present study are in color. Station 1, green; Station 2, blue; station
23, red. Database sequences are in black. Abbreviations: CAS, coastal AS; ETSP, eastern tropical South Pacific; PN,
Pacific northwest, CEst, Colne Estuary.
at the amino acid level. These sequences were identified as
the dominant nirS gene sequence at station 1 and used for the
development of clade-specific QPCR primers (see below).
The nirS gene of the anammox organism, Kuenenia stuttgartiensis, also falls into cluster V. The identity of K. stuttgartiensis NirS amino acid sequence is no more than 44%
(66% positive) to any NirS sequence from the AS OMZ. At
the DNA level, there is no significant similarity between the
nirS gene sequence from K. Stuttgartiensis and the nirS gene
sequences from the AS. Despite their evident divergence
from the rest of the NirS tree, the deduced amino acid alignments of cluster V sequences revealed that the Histidine
352 (Pseudomonas aeruginosa numbering) was conserved
among all the cluster V sequences from the AS. This supports
194 DISTRIBUTION AND RELATIVE QUANTIFICATION OF KEY GENES
Plate 1. (continued)
the conclusion that His 352 is the heme d1 ligand [Braker et
al., 2000]. His 352 is conserved in K. stuttgartiensis as well,
suggesting that this organism might be capable of carrying
out the nitrite-reduction step in the canonical denitrification
pathway. His 302 is also conserved in K. stuttgartiensis but
not in all the other sequences obtained in this study.
Cluster VI contained sequences only from station 2, representing the most abundant group from this location (~33%
of sequences). These sequences had close identities with
sequences from the ETSP. Some sequences in this group
were 99% identical to Marinobacter aquaeolei, a halophilic,
hydrocarbon degrading Gammaproteobacterium, which is
also a facultative mixotrophic iron oxidizer. It was isolated
from a Vietnamese oil-producing well [Nguyen et al., 1999]
and is also known as Marinobacter hydrocarbonoclasticus
[Márquez and Ventosa, 2005]. This cluster also contained
more distantly related sequences from the coastal AS, Baltic Sea, and isolate C10-1 from Pacific northwest sediment.
This dominant group at station 2 was not represented in the
libraries from the more intensely suboxic stations 1 and 23.
Cluster VII is a cosmopolitan group, as it contains sequences from all three locations as well as sequences from
JAYAKUMAR et al. 195
the coastal AS and ETSP. The closest isolate sequence
to this group is Azoarcus tolulyticus, with low bootstrap
support.
3.3. nirK Phylogeny
Unlike nirS, nirK gene fragments were amplified from all
the samples, regardless of whether or not nitrite was present.
The community dynamics of nirK gene phylotypes were dis-
tinct from those of nirS gene phylotypes. Similar to nirS,
high nitrite waters exhibited obvious dominance among nirK
types. However, even in waters without nitrite, nirK gene
also exhibited dominance, instead of the very high diversity
characteristic of nirS gene phylogeny in low nitrite waters.
Such dominance in nirK gene phylotypes was also reported
from the Black Sea suboxic zone [Oakley et al., 2007].
The identities between NirK sequences from the open AS
with sequences in the database ranged between 45% and
Plate 2. Phylogenetic relationships among translated nirK sequences (157 amino acids) from the water column of the
AS (both coastal zone and open ocean), with Neisseria gonorrhoeae AniA as the outgroup. Bootstrap values (>50%) are
indicated at the branch points. Phylogenetic positions of nirK clones from the present study are in color. Station 1, green;
station 2, blue; station 17, brown; station 23, red; station K03, pink. Database sequences are in black.
196 DISTRIBUTION AND RELATIVE QUANTIFICATION OF KEY GENES
Plate 2. (continued)
97% at the amino acid level (Plate 2). The closest identities
at the amino acid level were with coastal aquifer sequences
at 85% [Santoro et al., 2006] and with sequences from the
Black Sea suboxic zone at 74% [Oakley et al., 2007]. Interestingly, all the database sequences that had close identities
with the environmental sequences from the AS, Black Sea
suboxic zone, and coastal aquifer were from contaminated
sites or wastewater. Overall, the diversity of nirK gene sequences from the AS was comparable to the level of diversity in the coastal aquifer [Santoro et al., 2006] and Black
Sea suboxic zone [Oakley et al., 2007], based on the number
and divergence of phylogenetic groupings. This level of
identity between AS and database nirK gene sequences is
similar to the level of identities observed for nirS gene (see
above), although the database contains many more nirS than
nirK gene sequences from similar environments.
A total of 220 nirK clones were sequenced from the three
open-ocean stations and the two coastal stations off Oman
(station 17) and India (station KO3). Phylogenetic analysis
of NirK sequences from the five samples clustered the sequences into seven groups. None of these sequences were
closely related to sequences from cultured organisms, but
a few of the sequences had up to 97% identity to an environmental sequence, suggesting that natural assemblages are
poorly represented in culture collections. A majority of NirK
sequences from cultured denitrifying organisms grouped
JAYAKUMAR et al. 197
together, mostly with sequences obtained from waste­water.
Two AS sequences from station 23 were very different from
the rest of the sequences and had only 45% identity at the
amino acid level to any NirK sequences in the database.
These did not group with any other sequence and had long
branch lengths. All the clusters had strong bootstrap support within clusters but weak bootstrap support between
clusters.
Cluster I contained sequences only from the open-ocean
stations and accounted for 35% of the sequences from these
locations. This cluster included the nirK gene from the cultivated nitrifiers, Nitrosomonas sp. C113a and Nitrosomonas
sp. NO3W [Casciotti and Ward, 2001], suggesting that
these sequences might represent nitrifiers; however, there
is no strong bootstrap support between the cultivated nitrifier and the environmental sequences. Cluster I split into two
clades IA and IB with weak bootstrap support between them.
Clade IA contained sequences from all three open-ocean
locations evenly distributed, while clade IB had sequences
predominantly from stations 1 and 23, where more nitrite
was present. Sequences in clade IB clustered with sequences
from a coastal aquifer (78%identity), from a location where
the salinity was 34.5, and ammonium was present in a high
concentration [Santoro et al., 2006].
Cluster II mainly contained sequences from the station off
Karwar (K03), and one representative each from stations 2
and 23. The open-ocean sequences branched separately from
coastal sequences, and the sequence from station 23 was 82%
identical at the amino acid level to Roseovarius sp.217, an
Alphaproteobacterium of the Rhodobacterales family. Both
open-ocean sequences in this group clustered also with nirK
gene sequences from Baltic Sea water and sediment (O.-S.
Kim et al., unpublished Genbank sequences, 2005).
Cluster III, a very small group, contained two sequences
from station 23 and one from station 17 (coastal Oman).
This cluster contained sequences from Baltic Sea sediment
(O.-S. Kim et al., unpublished Genbank sequences, 2005)
and a brackish groundwater aquifer [Santoro et al., 2006].
Cluster IV contained sequences mainly from the two coastal
stations and a few sequences from stations 1 and 2. These
clusters did not contain any other database sequences and
groups separated out as open ocean and coastal stations subgroups, without strong bootstrap support.
The majority of the sequences (68%) from the Oman
coast station 17 grouped in cluster V. This cluster also contained sequences from open-ocean stations and, again, did
not contain any representatives from the database. Cluster
VI had sequences only from off the coast of Karwar. One
sequence that belonged to cluster VI was 97% identical to
a sequence obtained from wastewater sludge. Apart from a
few wastewater-derived environmental sequences, this clus-
ter also included sequences from Pseudomonas mendocina,
Enterococcus sp., and Alcaligenes sp., organisms related to
sewage, suggesting anthropogenic impacts in these coastal
waters.
Cluster VII is a cosmopolitan group, further branching out
into three smaller subgroups A, B, and C, where VIIA and
VIIC contain sequences from all locations. Cluster VIIA is
predominantly coastal and VIIC predominantly open ocean.
Cluster VIIB contained sequences from station 23 and the
Oman coast only, which clustered with the cultured bacterium Rhodobacterales bacterium (75% identity). This cluster also contained sequences from Baltic Sea water column
and sediment (O.-S. Kim et al., unpublished Genbank sequences, 2005) and the cultured strain Mesorhizobium sp.
4FB11 [Song and Ward, 2003].
3.4. Anammox Phylogeny
Anammox 16S rRNA gene fragments were amplified
from all three open-ocean stations, but this analysis was not
attempted for the coastal samples. Ladderane lipids, specific biomarkers for anammox organisms [Jaeschke et al.,
2007], and the presence of anammox 16S rRNA gene fragments [Woebken et al., 2008] have been detected in the open
AS OMZ waters. The 16S rRNA partial gene sequences
revealed very low diversity of anammox phylotypes in
the open AS (Figure 3). All 58 sequences obtained from the
three clone libraries were ≥98% identical to each other. The
forward primer AMX368F targets all anammox organisms
(genera Brocadia, Kuenenia, Scalindua, and Anammoxoglobus) [Schmid et al., 2001, 2003], while the reverse primer
AMX820R was designed to target only the Candidatus brocadia and Candidatus kuenenia groups. Using various sets
of primers, several studies have amplified monophyletic sequences that were ≥95% identical to Scalindua type anammox bacteria from several natural aquatic habitats [Kuypers
et al., 2003; Risgaard-Petersen et al., 2004; Kuypers et al.,
2005]. As in most of these previous studies, the closest data­
base match (96%) to the open AS sequence was also the
enriched anammox organism, Candidatus scalindua sp. As
the primers used in the present study were more specific for
Brocadia and Kuenenia types, it is likely that these two types
were not present in this environment. Scalindua sp., is apparently the major type that colonizes the OMZ habitat
(Benguela OMZ zones [Schmid et al., 2007]; ETSP OMZ
[Stevens and Ulloa, 2008]; Lake Tanganyika [Shubert et
al., 2006]). The sequences from the AS OMZ were 99%
identical to 16S rRNA anammox sequences from Lake Tanganyika [Schubert et al., 2006]. It is possible that the primers used in this study do not target all anammox bacteria
in this environment. Many studies in similar marine and
198 DISTRIBUTION AND RELATIVE QUANTIFICATION OF KEY GENES
Figure 3. Phylogenetic relationships among anammox 16S rRNA partial gene sequences (390 bases) from the water
column of open AS (stations 1, 2, and 23) with an environmental clone CTD005-50B-02 from a ridge flank crustal fluid
as the outgroup. Bootstrap values (>50%) are indicated at the branch points. Names of 16S rRNA clones from the present
study are in bigger font and marked with asterisks, Station 1, *; station 2, **; station 23, ***; and those sequences from
other oceanic environments and published databases are in smaller font.
JAYAKUMAR et al. 199
freshwater environments using various primer combinations
have failed to amplify other novel anammox organisms. The
sequences presented here and from other marine and fresh
water environments [Schmid et al., 2007] show the ubiquitous presence of anammox bacteria, but suggest low diversity
of these organisms in suboxic/anoxic marine ecosystems.
3.5. Rarefaction Analysis
Rarefaction analysis showed that nirS gene richness was
similar in all three open-ocean clone libraries (Figure 4a)
and the libraries saturated at about nine phylotypes. nirK
gene rarefaction curves were quite different among stations
(Figure 4b). The richness at stations 1 and K03 appear to
be saturated, while at stations 2, 17, and 23, more sequencing effort will likely detect more novel sequence types. To
compare the richness between the open-ocean and coastal
AS, the sequences for each gene type were grouped as either coastal or open ocean. By this analysis, nirS gene communities were much richer in the coastal environment than
in the open ocean (Figure 4c). Total richness was not as
dramatically different between open-ocean and coastal stations for nirK gene (Figure 4d), but the extrapolation of the
curves implies greater richness in libraries from the coastal
environment. Because very little diversity was detected in
the anammox 16S rRNA gene sequences of the open AS, a
1% cutoff was used to investigate the degree of small-scale
variability in the sequence distribution (Figure 4e). If 98%
similarity for the 16SrRNA gene delineates a genus, then the
data shown here indicate that there is some species level diversity in the AS anammox assemblage. However, there was
no definite pattern with respect to their distribution among
the locations, and richness was more or less saturated at all
three locations.
3.6. Quantification of nirS and Anammox Genes
Total nirS gene copy numbers ranged from 1.9 to 5.6 ×
105 ml−1 of seawater, with highest numbers at station 1
(Figure 5), assuming 100% efficiency in DNA extraction.
Total nirS, Dom nirS, and anammox 16S rRNA gene copy
Figure 4. (opposite) Rarefaction curves. Ten percent difference in
DNA sequences was the cutoff limit for defining an OTU for nirS
and nirK; 1% for anammox 16S rRNA OTUs. (a) nirS sequences
from the open AS (three stations). (b) nirK sequences from AS (five
stations). (c) nirS sequences grouped by open ocean (three stations)
and coastal ocean (two stations). (d) nirK sequences grouped by
open ocean (three stations) and coastal ocean (two stations). (e)
Anammox 16S rRNA sequences from the open AS (three stations).
200 DISTRIBUTION AND RELATIVE QUANTIFICATION OF KEY GENES
tuted between 0.4% and 1.5% of the total cell count. The cell
numbers and relative abundance of anammox as a percent
of the total appear to be higher (10–12%) in the AS than in
the other OMZ environments for which comparable data are
available.
3.7. Ecology of Denitrification and Anammox
Figure 5. Abundance of nirS and anammox genes. Total nirS
genes, Dom1 nirS genes, and anammox 16S rRNA genes at three
stations from cruise SK209 (October 2004) determined by quantitative polymerase chain reaction. Error bars represent standard
deviations of three to six replicate assays for each gene.
numbers all varied in similar patterns among stations: all three
were more abundant, by a factor of ~2, at the two stations
(1, 23) with highest nitrite concentrations. Anammox 16S
rRNA gene copy number was highest at station 1 (84,398 ml−1)
but always was <15% of the total nirS gene copy number,
while Dom nirS was about 20–30% of total nirS. Both anammox 16S rRNA and nirS genes are assumed to be present at
one copy per cell [Strous et al., 2006; Zumft, 1997], so even
without absolute normalization to volume, it is clear that
nirS type denitrifiers outnumbered anammox cells by 7- to
11-fold. The single phylotype represented by the Dom nirS
primers exceeded anammox copy numbers by 3.3- to 4.6fold. nirK gene copy numbers were not investigated.
Total cell abundance in the OMZ of the AS ranges from
about 105 to 106 cells ml−1 [Ducklow et al., 2001], with the
highest cell numbers associated with highest nitrite concentrations [Jayakumar et al., 2009]. High nitrite concentrations
have been interpreted as representing intense suboxia and
implying classical denitrification. Even if nirK-type denitrifier abundance is only a fraction of the abundance of nirStype denitrifiers, the overwhelming majority of cells in the
OMZ are either denitrifiers or anammox bacteria. Both nirS
and anammox 16S rRNA copy numbers increased with increasing nitrite concentration, suggesting increased activity
of both groups with increased suboxia.
In Lake Tanganyika [Schubert et al., 2006], the Black Sea
[Kuypers et al., 2003], and the Benguela upwelling regions
[Woebken et al., 2007], anammox cell abundance consti-
Using data from incubation experiments designed to measure denitrification and anammox rates [Ward et al., 2008;
Jayakumar et al., 2009], and in situ sampling described here,
we detected a relationship between denitrifying community
dynamics, as deduced from nirS and nirK gene diversity, and
the stage of denitrification [Jayakumar et al., 2009]. As denitrification progresses from initially low-oxygen-high-nitrate
waters, to low-oxygen-high-nitrite conditions, denitrifier
diversity decreases, while dominance increases, analogous
to a typical bloom situation [Jayakumar et al., 2009].
Anammox bacteria, on the other hand, were represented by
a single phylotype (C. scalindua). Although anammox cell
number varied by threefold (stations 1 and 23 versus 2; Figure 5), the anammox assemblage composition does not appear to vary among stations. Both anammox and denitrifying
organisms are probably always present in this environment,
the perennial SNM, but their activities and abundances appear to increase with increased organic deposition (and the
accompanying increase in nitrite concentration in the SNM)
during the monsoon periods.
In the Namibian upwelling system, Kuypers et al. [2003]
detected anammox activity in the presence of up to 9 µM
DO. Anammox cells of the Scalindua sp. type were associated with particles that also contained assemblages of proteobacteria [Woebken et al., 2007]. It was suggested that
heterotrophic bacteria consumed the oxygen within the particles, producing conditions that favored anammox activity.
It seems unlikely that this scenario applies to the open AS.
If heterotrophic bacteria were growing by respiring oxygen,
their cell numbers should decrease with increasing suboxia,
but the opposite trend, increasing total cell number with increasing suboxia and increasing nitrite concentration, is observed in OMZ waters [Ward et al., 2008]. Moreover, the
particle maxima in the OMZ are composed predominantly of
cells (as reported for the ETSP OMZ) [Spinrad et al., 1989],
unlike the large flocs that were observed near the sediment
water interface in the Namibian system. Higher copy numbers of both anammox 16S rRNA and heterotrophic denitrification genes were found at later denitrification stages,
i.e., increasing cell number with increasing suboxia. If the
shift toward dominance by a few OTUs, in the later stages
of denitrification, which is exhibited by both nirS and nirK
genes, implies that both types of denitrifiers undergo bloom
JAYAKUMAR et al. 201
dynamics, then we expect the abundances of both types to be
greatest in the most intensely suboxic zones. Both types of
organisms presumably increase in abundance at the expense
of oxidized forms of nitrogen. It is possible that the growth
and biomass of denitrifying bacteria are limited by the supply of organic matter, while anammox is limited by the supply of ammonia that is produced from the decomposition
of organic matter by the denitrifiers and other heterotrophs.
Hydrographic data provide evidence that ammonia is very
rapidly recycled and is normally present at insignificant concentrations in these waters [Naqvi et al., 2006; http://usjgofs.
whoi.edu/datasys/mrg/mrgdir.html]. An inverse correlation
between C-deposition and the proportion of total N2 production that is due to anammox has been observed in sediment environments [Dalsgaard et al., 2005; Engstrom et al.,
2005]. Such a correlation in the OMZ would suggest that anam­
mox activity should decrease in importance compared to
denitrification during seasons of high organic matter input,
as encountered during the sampling period of this study.
The canonical denitrification process alone cannot fully
explain the excess N2 production observed in the AS OMZ
[Devol et al., 2006]. The excess N2, calculated from N2/Ar
ratios, is as large as 23 µg at N L−1, while the nitrate deficit
calculated from Redfield stoichiometry is ≤13 µg at N L−1
[Devol et al., 2006]. The presence and abundance dynamics
of genes representing denitrification and anammox suggest
that both processes are likely occurring in this environment.
The question is, “What is the relative contribution of these
two processes to the removal of fixed nitrogen?” The ratio of
nirS-type denitrifier abundance to that of anammox organisms is high in the open AS, but rates of N transformations
associated with the two groups of organisms have not been
reported. We suspect that the abundance ratio and the relative transformation rates vary in space and time, depending
on the availability of organic substrates, DO, and other environmental conditions.
4. Conclusions
The overall diversity of nitrite reductase genes was limited in the open AS compared to sedimentary and terrestrial
environments and the coastal AS. nirS-type denitrifiers were
more abundant than anammox bacteria in the OMZ of the
AS as quantified by QPCR. Denitrifiers also display greater
diversity and greater variability in community composition
across a range of different conditions of oxygen and nitrite
concentrations.
Most nirK or nirS genes from the AS are not closely related
to published sequences. The closest identities are with sequences from other water column denitrifying environments,
however, rather than sedimentary or terrestrial environ-
ments. Some wastewater-type denitrifying assemblages were
detected off the coast of Karwar. Some groups of nirK-type
denitrifiers were unique to the open ocean, and some were
unique to the coastal regions. Only two groups appeared to
be cosmopolitan. Some groups of nirS genes were common
to both open AS and ETSP. The open AS nirS assemblages
were quite different from those in the coastal AS.
Based on diversity dynamics and gene abundance data
alone, it appears that both canonical denitrifiers and anammox bacteria contribute to the removal of fixed nitrogen in
the OMZ of the AS. Denitrifiers are present in higher numbers in the nitrite maximum and show patterns of diversity
and dominance that suggest a dynamic bloom response to
denitrifying conditions. Thus, denitrifiers may play a larger
role in N removal than anammox during high carbon flux
seasons.
Acknowledgments. This work was funded by the US National
Science Foundation grants to BBW and AJ. We also thank P.V.
Narvekar, H. Naik, and A. Pratihary for assistance onboard ORV
Sagar Kanya with the nutrient analysis and Greg O’Mullan for his
help with the OTU analysis.
References
Barber, R. T., J. Marra, R. R. Bidigare, L. A. Codispoti, D. Halpern, Z. Johnson, M. Latasa, R. Goericke, and S. L. Smith (2001),
Primary productivity and its regulation in the Arabian Sea during
1995, Deep Sea Res., Part II, 48, 1127–1172.
Braker, G., A. Fesefeldt, and K.-P. Witzel (1998), Development of
PCR primer systems for amplification of nitrite reductase genes
(nirK and nirS) to detect denitrifying bacteria in environmental
samples, Appl. Environ. Microbiol., 64, 3769–3775.
Braker, G., J. Zhou, L. Wu, A. H. Devol, and J. M. Tiedje (2000),
Nitrite reductase genes (nirK and nirS) as functional markers to
investigate diversity of denitrifying bacteria in Pacific northwest
marine sediment communities, Appl. Environ. Microbiol., 66,
2096–2104.
Carpenter, J. H. (1965), The Chesapeake Bay Institute technique
for the Winkler dissolved oxygen method, Limnol. Oceanogr.,
10, 141–143.
Casciotti, K. L., and B. B. Ward (2001), Nitrite reductase genes
in ammonia-oxidizing bacteria, Appl. Environ. Microbiol., 67,
2213–2221.
Castro-González, M., G. Braker, L. Farías, and O. Ulloa (2005),
Communities of nirS-type denitrifiers in the water column of the
oxygen minimum zone in the eastern South Pacific, Environ.
Microbiol., 7, 1298–1306.
Codispoti, L. A. (2007), An oceanic fixed nitrogen sink exceeding
400 Tg N a−1 vs the concept of homeostasis in the fixed-nitrogen
inventory, Biogeosciences, 4, 233–253.
Codispoti, L. A., T. Yoshinari, and A. H. Devol (2005), Suboxic respiration in the oceanic water column, in Respiration in
202 DISTRIBUTION AND RELATIVE QUANTIFICATION OF KEY GENES
Aquatic Ecosystems, edited by P. A. del Giorgio and P. J. Le B.
Williams, pp. 225–247, Oxford Univ. Press.
Dalsgaard, T., B. Thamdrup, and D. E. Canfield (2005), Anaerobic
ammonium oxidation (anammox) in the marine environment,
Res. Microbiol., 156, 457–464.
Devol, A. H., A. G. Uhlenhopp, S. W. A. Naqvi, J. A. Brandes,
D. A. Jayakumar, H. Naik, S. Gaurin, L. A. Codispoti, and T.
Yoshinari (2006), Denitrification rates and excess nitrogen gas
concentrations in the Arabian Sea oxygen deficient zone, Deep
Sea Res., Part I, 53, 1533–1547.
Duce, R. A., et al. (2008), Impacts of atmospheric anthropogenic
nitrogen on the open ocean, Science, 320, 893–898.
Ducklow, H. W., D. C. Smith, L. Campbell, M. R. Landry, H. L.
Quinby, G. F. Steward, and F. Azam (2001), Heterotrophic bacterioplankton in the Arabian Sea: Basinwide response to yearround high primary productivity, Deep Sea Res., Part II, 48,
1303–1323.
Engstrom, P., T. Dalsgaard, S. Hulth, and R. C. Aller (2005),
Anaerobic ammonium oxidation by nitrite (anammox): Implications for N2 production in coastal marine sediments, Geochim.
Cosmochim. Acta, 69, 2057–2065.
Glockner, A. B., A. Jüngst, and W. G. Zumft (1993), Coppercontaining nitrite reductase from Pseudomonas aureofaciens is
functional in a mutationally cytochrome cd1-free background
(NirS-) of Pseudomonas stutzeri, Arch. Microbiol., 160, 18–26.
Grasshoff, K., M. Ehrhardt, and K. Kremling (1983), Methods of
Seawater Analysis, Verlag Chemie, Weinheim.
Hannig, M., G. Braker, J. Dippner, and K. Jürgens (2006), Linking
denitrifier community structure and prevalent biogeochemical
parameters in the pelagial of the central Baltic Proper (Baltic
Sea), FEMS Microbiol. Ecol., 57, 260–271.
Heck, K. L., G. V. Belle, and D. Simberloff (1975), Explicit calculation of the rarefaction diversity measurement and the determination of sufficient sample size, Ecology, 56, 1459–1461.
Holland, S. M. (2001), Analytic Rarefaction 1.3, University of
Georgia, Athens, GA.
Jaeschke, A., E. C. Hopmans, S. G. Wakeham, S. Schouten, and J.
S. S. Damste (2007), The presence of ladderane lipids in the oxygen minimum zone of the Arabian Sea indicates nitrogen loss
through anammox, Limnol. Oceanogr., 52, 780–786.
Jayakumar, A., G. O’Mullan, S. W. A. Naqvi and B. B. Ward
(2009), Denitrifying bacterial community composition changes
associated with stages of denitrification in oxygen minimum
zones, Microb. Ecol., in press.
Jayakumar, D. A., C. A. Francis, S. W. A. Naqvi, and B. B. Ward
(2004), Diversity of nitrite reductase genes (nirS) in the denitrifying water column of the coastal Arabian Sea, Aquat. Microb.
Ecol., 34, 69–78.
Kim, H.-S., C. N. Flagg, and S. D. Howden (2001), Northern Arabian Sea variability from TOPEX/Poseidon altimetry data: An
extension of the US JGOFS shipboard ADCP study, Deep Sea
Res., Part II, 48, 1069–1096.
Kuypers, M. M. M., A. O. Sliekers, G. Lavik, M. Schmid, B. B.
Jorgensen, J. G. Kuenen, J. S. S. Damste, M. Strous, and M. S.
M. Jetten (2003), Anaerobic ammonium oxidation by anammox
bacteria in the Black Sea, Nature, 422, 608–611.
Kuypers, M. M. M., G. Lavik, D. Woebken, M. Schmid, B. M.
Fuchs, R. Amann, B. B. Jorgensen, and M. S. M. Jetten (2005),
Massive nitrogen loss from the Benguela upwelling system
through anaerobic ammonium oxidation, Proc. Natl. Acad. Sci.
U. S. A., 102, 6478–6483.
Márquez, M. C., and A. Ventosa (2005), Marinobacter hydrocarbonoclasticus Gauthier et al. 1992 and Marinobacter aquaeolei
Nguyen et al. 1999 are heterotypic synonyms, Int. J. Syst. Evol.
Microbiol., 55, 1349–1351.
Morrison, J. M., et al. (1999), The oxygen minimum zone in the Arabian Sea during 1995, Deep Sea Res., Part II, 46, 1903–1931.
Naqvi, S. W. A. (1991), Geographical extent of denitrification in
the Arabian Sea in relation to some physical processes, Oceanologia Acta, 14, 281–290.
Naqvi, S. W. A., and D. A. Jayakumar (2000), Ocean biogeochemistry and atmospheric composition: Significance of the Arabian
Sea, Curr. Sci., 78, 289–299.
Naqvi, S. W. A., H. S. Naik, and P. V. Narvekar (2003), The Arabian Sea, in Biogeochemistry of Marine Systems, edited by K.
Black and G. Shimmield, pp. 156–206, Blackwell, Oxford.
Naqvi, S. W. A., P. V. Narvekar, and E. Desa (2006), Coastal biogeochemical processes in the North Indian Ocean, in The Sea,
vol. 14, edited by A. Robinson and K. Brink, pp. 723–780, Harvard Univ. Press.
Nguyen, B. H., E. B. M. Denner, T. C. H. Dang, G. Wanner, and
H. Stan-Lotter (1999), Marinobacter aquaeolei sp. nov., a halophilic bacterium isolated from a Vietnamese oil-producing well,
Int. J. Syst. Bacteriol., 49, 367–375.
Nogales, B., K. N. Timmis, D. B. Nedwell, and A. M. Osborn
(2002), Detection and diversity of expressed denitrification genes
in estuarine sediments after reverse transcription-PCR amplification from mRNA, Appl. Environ. Microbiol., 68, 5017–5025.
Oakley, B. B., C. A. Francis, K. J. Roberts, C. A. Fuchsman, S.
Srinivasan, and J. T. Staley (2007), Analysis of nitrite reductase
(nirK and nirS) genes and cultivation reveal depauperate community of denitrifying bacteria in the Black Sea suboxic zone,
Environ. Microbiol., 9, 118–130.
Risgaard-Petersen, N., R. L. Meyer, M. Schmid, M. S. M. Jetten, A.
Enrich-Prast, S. Rysgaard, and N. P. Revsbech (2004), Anaerobic ammonium oxidation in an estuarine sediment, Aquat. Microb. Ecol., 36, 293–304.
Santoro, A. E., A. B. Boehm, and C. A. Francis (2006), Denitrifier community composition along a nitrate and salinity gradient in a coastal aquifer, Appl. Environ. Microbiol., 72, 2102–
2109.
Schmid, M., S. Schmitz-Esser, M. Jetten, and M. Wagner (2001),
16S-23S rDNA intergenic spacer and 23S rDNA of anaerobic
ammonium oxidizing bacteria: Implications for phylogeny and
in situ detection, Environ. Microbiol., 3, 450–459.
Schmid, M., et al. (2003), Candidatus “Scalindua brodae”, sp. nov.,
Candidatus “Scalindua wagneri”, sp. nov., two new species of
anaerobic ammonium oxidizing bacteria, Syst. Appl. Microbiol.,
26, 529–538.
Schmid, M., et al. (2005), Biomarkers for in situ detection of anaerobic ammonium-oxidizing (anammox) bacteria, Appl. Environ.
Microbiol., 71, 1677–1684.
JAYAKUMAR et al. 203
Schmid, M., et al. (2007), Anaerobic ammonium-oxidizing bacteria in marine environments: Widespread occurrence but low
diversity, Environ. Microbiol., 9, 1476–1484.
Schubert, C. J., E. Durisch-Kaiser, B. Wehrli, B. Thamdrup, P.
Lam, and M. M. Kuypers (2006), Anaerobic ammonium oxidation in the tropical freshwater system (Lake Tanganyika), Environ. Microbiol., 8, 1857–1863.
Song, B., and B. B. Ward (2003), Nitrite reductase genes in halobenzoate degrading denitrifying bacteria and related species, FEMS
Microbiol. Ecol., 43, 349–357.
Spinrad, R. W., H. E. Glover, B. B. Ward, L. A. Codispoti, and G.
Kullenberg (1989), Suspended particle and bacterial maxima in
Peruvian coastal waters during a cold water anomaly, Deep Sea
Res., Part A, 36, 715–733.
Stevens, H., and O. Ulloa (2008), Bacterial diversity in the oxygen
minimum zone of the eastern tropical South Pacific, Environ.
Microbiol., 10, 1244–1259.
Strous, M., et al. (2006), Deciphering the evolution and metabolism
of an anammox bacterium from a community genome, Nature,
440, 790–794.
Thompson, J. D., T. J. Gibson, F. Plewniak, F. Jeanmougin, and
D. G. Higgins (1997), The ClustalX windows interface: Flexible strategies for multiple sequence alignment aided by quality
analysis tools, Nucleic Acids Res., 24, 4876–4882.
Ward, B. B., C. B. Tuit, A. Jayakumar, J. J. Rich, J. Moffett, and
S. W. A. Naqvi (2008), Organic carbon, and not copper, controls
denitrification in oxygen minimum zones of the ocean, Deep Sea
Res., Part I, 55, 1672–1683.
Woebken, D., B. M. Fuchs, M. M. Kuypers, and R. Amann (2007),
Potential interactions of particle-associated anammox bacteria
with bacterial and archaeal partners in the Namibian upwelling
system, Appl. Environ. Microbiol., 73, 4648–4657.
Woebken, D., P. Lam, M. M. Kuypers, S. W. A. Naqvi, B. Kartal,
M. Strous, M. S. M. Jetten, B. M. Fuchs, and R. Amann (2008),
A microdiversity study of anammox bacteria reveals a novel
Candidatus scalindua phylotype in marine oxygen minimum
zones, Environ. Microbiol., 10, 3106–3119.
Zumft, W. G. (1997), Cell biology and molecular basis of denitrification, Microbiol. Mol. Biol. Rev., 61, 533–616.
A. Jayakumar and B. B. Ward, Department of Geosciences, Prince­
ton University, Princeton, NJ 08544, USA. (ajayakum@Princeton.
edu)
S. W. A. Naqvi, National Institute of Oceanography, Council
of Scientific and Industrial Research, Dona Paula, Goa 403 004,
India.