Dinitrogen fixation in a unicellular chlorophyll d

0 downloads 0 Views 733KB Size Report
Feb 6, 2011 - 21 fixation / microbial diversity / nitrogenase. 22. 23 .... (ASN0) by centrifugation in a sterilized tube at 6,000× g for 20 min, washed once with. 9 .... samples) at 3,750 rpm (Allegra 6 KR, Beckman Coulter, Krefeld, Germany). ... °C, 40 x (3 s at 95 °C, 30 s at 60 °C)) was used for the preliminary analysis of the.
1 1 2

Dinitrogen fixation in a unicellular chlorophyll d-containing

3

cyanobacterium

4 5

Ulrike Pfreundt1, Lucas J. Stal2,3, Björn Voß1, Wolfgang R. Hess1

6 7

1

8

Germany

9

2

University of Freiburg, Faculty of Biology, Schänzlestr. 1, D-79104 Freiburg,

Department of Marine Microbiology, Centre for Estuarine and Marine Ecology

10

(NIOO-KNAW), Yerseke, The Netherlands

11

3

12

Netherlands

Department of Aquatic Microbiology, IBED, University of Amsterdam, The

13 14 15

Running title: Nitrogenase in Acaryochloris spp. HICR111A

16 17

*Corresponding author: Wolfgang R. Hess; Phone: +49-761-203-2796

18

Fax: +49-761-203-2745; Email: [email protected]

19 20

ISMEJ-11-00621OAR In press, NOV 27, 2011

2 1

ABSTRACT

2

Marine cyanobacteria of the genus Acaryochloris are the only known organisms that

3

use chlorophyll d as a photosynthetic pigment. However, based on chemical

4

sediment analyses, chlorophyll d has been recognized to be widespread in oceanic

5

and lacustrine environments. Therefore it is highly relevant to understand the genetic

6

basis for different physiologies and possible niche adaptation in this genus. Here we

7

show that unlike all other known isolates of Acaryochloris the strain HICR111A,

8

isolated from waters around Heron Island, Great Barrier Reef, possesses a unique

9

genomic region containing all the genes for the structural and enzymatically active

10

proteins of nitrogen fixation and cofactor biosynthesis. Their phylogenetic analysis

11

suggests a close relation to nitrogen fixation genes from certain other marine

12

cyanobacteria. We show that nitrogen fixation in Acaryochloris sp. HICR111A is

13

regulated in a light-dark dependent fashion. We conclude that nitrogen fixation, one

14

of the most complex physiological traits known in bacteria, might be transferred

15

among oceanic microbes by horizontal gene transfer more often than anticipated so

16

far. Our data show that the two powerful processes of oxygenic photosynthesis and

17

nitrogen fixation co-occur in one and the same cell also in this branch of marine

18

microbes and characterize Acaryochloris as a physiologically versatile inhabitant of

19

an ecological niche which is primarily driven by the absorption of far-red light.

20 21

Keywords: Acaryochloris / chlorophyll d / coral reef / cyanobacteria / dinitrogen

22

fixation / microbial diversity / nitrogenase

23

3 1

INTRODUCTION

2

The fixation of dinitrogen is a major and crucial biogeochemical process in the ocean

3

where it replaces the losses of bioavailable combined nitrogen due to denitrification

4

and anaerobic ammonium oxidation (ANAMMOX) and allows new primary production

5

(Zehr & Kudela 2011). The capacity of fixing dinitrogen is widely distributed among

6

Bacteria and a few Archaea, but not among Eukarya and it gives the organisms that

7

possess this capacity access to the largest pool of nitrogen, atmospheric N2.

8

Nitrogenase, the enzyme complex that is responsible for the fixation of dinitrogen is

9

extremely oxygen sensitive and therefore most of the N2-fixing microorganisms are

10

obligate or facultative anaerobes. A notable exception is the oxygenic phototrophic

11

cyanobacteria, many of which possess the capacity of fixing dinitrogen (Zehr 2011).

12

Therefore, it is generally assumed that cyanobacteria are the main group of

13

organisms that are responsible for the fixation of N2 in the ocean, although this

14

viewpoint has been challenged (Riemann et al. 2010). Obviously, cyanobacteria

15

need adaptations and follow certain strategies to provide nitrogenase with an

16

anaerobic environment (Bergman et al. 1997; Berman-Frank et al. 2003).

17

Heterocystous cyanobacteria confine nitrogenase to special differentiated anoxic

18

cells that lack photosystem II, thereby separating it spatially from oxygenic

19

photosynthesis. Many marine unicellular diazotrophic cyanobacteria, such as

20

Crocosphaera or Cyanothece follow another strategy and separate the two

21

incompatible processes temporally, confining the fixation of dinitrogen to the dark

22

period. A notable exception is the filamentous non-heterocystous cyanobacterium

23

Trichodesmium, which is abundant in the tropical oceans and one of the most

24

important diazotrophs (Karl et al. 2002). This organism may exhibit a combination of

25

spatial and temporal separation in which cells temporally switch off oxygenic

26

photosynthesis and fix N2 (Berman-Frank et al. 2001). Another exception is the

4 1

hitherto uncultivated putative unicellular diazotrophic cyanobacteria UCYN-A. The

2

analysis of the metagenome of flow cytometrically sorted cells has shown that the

3

genes for the oxygenic photosystem II and for CO2 fixation are absent (Tripp et al.

4

2010). The nitrogen fixation (nif) genes of this organism are expressed during the

5

daytime and N2 fixation also occurs during the day but further nothing is known about

6

the identity of this enigmatic cyanobacterium.

7 8

Acaryochloris is a unicellular marine cyanobacterium that is atypical in using

9

chlorophyll d for light harvesting (Miyashita et al. 1996). Based on chemical analyses,

10

chlorophyll d has only recently been recognized to be widespread in oceanic and

11

lacustrine environments (Kashiyama et al. 2008). Therefore it is highly relevant to

12

understand the genetic basis for different physiologies and possible niche adaptation

13

in this genus. At present, genome sequences are available of two Acaryochloris

14

strains. Acaryochloris marina MBIC 11017 (from here A. marina) was isolated from

15

waters around the Palau islands in the Pacific Ocean (Miyashita et al. 1996),

16

Acaryochloris sp. CCMEE 5410 from the Salton Sea lake in California (Miller et al.

17

2005). Compared to other cyanobacteria, large and complex genomes were found for

18

both isolates. A. marina harbors a total of 8.3 million base pairs with more than 25%

19

of its 8462 genes being located on nine different plasmids, eight of which are larger

20

than 100 kb (Swingley et al. 2008). A similarly complex genome was also reported for

21

the Salton Sea strain (Miller et al. 2011). Despite this impressive coding capacity, no

22

genes for dinitrogen fixation were found in either of these two well-studied isolates.

23 24

We recently described Acaryochloris sp. HICR111A, a new member of the

25

Acaryochloris clade (Mohr et al. 2010). The strain was isolated from microbial mat-

26

like material on a dead coral skeleton in shallow water on the southern reef flat of

5 1

Heron Island, Great Barrier Reef (23° 26' 31.2'' S, 151° 54' 50.4'' E). The cells are rod

2

shaped, between 1 and 2 μm long with a diameter of 0.75–1 μm, which is somewhat

3

smaller than the 2–3 μm in length and the diameter of 1.5–2.0 μm reported for A.

4

marina (Mohr et al. 2010). The presence of Chl d was monitored spectro-

5

photometrically (Qy maxima of Chl d detected at 707 nm and a single fluorescence

6

maximum at 750 nm after Chl d excitation at 470 nm) and by high performance liquid

7

chromatography repeatedly from the early cultivation phase onwards. Purified

8

cultures use Chl d as main photopigment, as indicated by a 1:50 ratio of Chl a / Chl d,

9

which changed only little with increasing light intensity (from 5–10 µmol to 75–100

10

µmol photons m-2 s-1) (Mohr et al. 2010). Acaryochloris sp. HICR111A is, according

11

to 16S rRNA-based phylogeny, an early branching member of the genus

12

Acaryochloris (Mohr et al. 2010). Its 16S ribosomal RNA is 2% divergent from the two

13

previously described A. marina strains, which, together with several distinct genetic,

14

morphological and physiological differences, led us to conclude that cyanobacteria of

15

the genus Acaryochloris occur as distinct ecotypes. A preliminary sequence analysis

16

of its genome based on Illumina technology yielded a genome with an accumulated

17

length of 8,371,965 nt (Mohr et al. 2010). Here we report the existence of a complete

18

nitrogenase gene cluster in Acaryochloris sp. HICR111A, we demonstrate that the nif

19

genes are strongly induced upon removal of combined nitrogen, and we detected

20

nitrogenase activity.

21 22

MATERIALS AND METHODS

23

Culture conditions

24

Cultures of Acaryochloris sp. HICR111A (Mohr et al. 2010) were grown in ASN

25

medium (artificial seawater + BG-11 nutrients) (Rippka et al. 1979). For preliminary

26

analysis of dinitrogen fixation genes by qPCR, cultures were grown at room

6 1

temperature (summer: ~25 °C) under naturally changing light conditions (from here:

2

“summer cultures”) and later that year moved to an incubator at 26 °C under a

3

12h|12h light|dark cycle with warm white light (Sanyo fluorescent tubes, 40 watt) of

4

35 µmol quanta m-2 s-1. For measurement of nitrogenase activity two cultures of

5

Acaryochloris HICR111A were grown and used as biological replicates (the assays

6

were done on subsequent days). Both cultures were initially grown under a constant

7

12h|12h light|dark cycle at 26°C and 35 µmol quanta m-2 s-1. Nine days prior to the

8

nitrogenase assay, half of one culture was shifted to nitrogen depleted medium

9

(ASN0) by centrifugation in a sterilized tube at 6,000× g for 20 min, washed once with

10

ASN0 (a third of the original volume), and resuspended in the original volume of

11

ASN0. The other half of this culture served as control and was also centrifuged, but

12

resuspended in an equal volume of fresh ASN medium. Subsequently, both were

13

incubated under a light|dark regime of 12h|12h at 50-60 µmol quanta m-2 s-1. The

14

replicate culture had been incubated at 50-60 µmol quanta m-2 s-1 for 22 days and

15

half of it was subsequently shifted to ASN0 7 days prior to the assay as described

16

above. One day prior to the nitrogenase assay, all cultures were shifted to 150 µmol

17

quanta m-2 s-1 and 29 °C. For the nitrogenase assay six 15-ml samples of the ASN0

18

culture and one 15-ml sample of ASN culture were filtered onto GF-F filters and

19

placed on a perforated support in contact with ASN0 (or ASN in case of the control) in

20

a custom made, temperature controlled incubator. Each of the seven chambers was

21

closed by a glass window, leaving 3 ml headspace. The cultures were incubated

22

under a constant flow of N2 + 0.03 % CO2. The light/dark cycle, light intensity and

23

temperature were kept the same.

24 25

Determination of nitrogenase activity

7 1

Nitrogenase activity was measured in custom-made incubators with separate

2

chambers to hold filters of 45 mm diameter (one incubator containing 6 chambers,

3

the other containing a single chamber which served as the control). The chambers

4

were identical to the one described in (Staal et al. 2001). The incubator was

5

connected to a circulating cryostat water bath in order to thermostat the incubation

6

chambers. The single chamber was temperature controlled by a computer-controlled

7

Peltier element. The chambers were connected in series to a gas line of N2 with

8

0.03% CO2 premixed (Westfalen Gases, Münster, Germany) at 3 L h-1, controlled by

9

a mass flow controller (5850S) and control unit (0152, both Brooks Instrument, Ede,

10

The Netherlands). The incubator was covered with a hood containing the light source

11

(two cool white fluorescent tubes) that was turned on and off by a timer.

12 13

Nitrogenase activity was measured using the acetylene reduction assay (ARA)

14

(Steward et al. 1967). Nitrogenase activity was measured at six time points: (1) -2 – 0

15

h, (2) 0 – 2 h, (3) 3 – 5 h, (4) 6 – 8 h, (5) 10 – 12 h, and (6) 12 – 14 h (where the

16

beginning of the dark phase is taken as 0 h). The ASN culture (nitrate-supplemented)

17

was taken as the control and measured along with the second and the fifth time

18

points in the two replicate experiments, respectively. The samples had been

19

incubated under anaerobic conditions for 7-8 h before the acetylene reduction assay

20

was started to prevent a false negative result due to oxygen inactivation.

21 22

For each time point one chamber was disconnected from the gas flow and 1 ml of

23

acetylene was injected into the chamber (~30 % of gas phase) while the gas was

24

displaced into a second syringe connected to the gas outlet of the incubator. The gas

25

in the headspace was mixed by ‘pumping’ using the two syringes. After 2 h of

26

incubation the gas was mixed again, two 0.5 ml gas samples (technical duplicates)

8 1

were withdrawn from the headspace, and the filter with the culture snap-frozen in

2

liquid nitrogen for later RNA extraction. Acetylene and ethylene in the gas samples

3

were determined gas chromatographically (Chrompack CP9001, Middelburg, The

4

Netherlands). The gas chromatograph was equipped with a flame ionization detector

5

(FID) and a wide-bore silica fused (0.53 mm internal diameter) Porapak U column

6

(Chrompack). The carrier gas was N2 at 10 ml min-1 and the flows of H2 and air for

7

the FID were 30 and 300 ml min-1, respectively. The temperatures for injector,

8

detector and oven were 90, 120, and 55 °C, respectively. The total amount of

9

ethylene produced by the culture was calculated using acetylene as an internal

10

standard (Stal 1988) and normalized to chlorophyll d (measured in a separate 15-ml-

11

sample of the original culture).

12 13

Determination of chlorophyll d

14

Fifteen ml of culture were transferred to a 50-ml Falcon tube and centrifuged for 15

15

min at 5,000 g. Five ml of 96 % ethanol was added to the cell pellet, vortexed, and

16

the chlorophyll was extracted overnight in the dark at room temperature. The extract

17

was centrifuged for 20 min at 6,000 g and the absorption was read

18

spectrophotometrically at 665 and 696 nm. The concentration of chlorophyll d was

19

calculated using the formula Chl d [µg/ml] = A665nm (-0.2006) + A696nm (12.0995)

20

(Ritchie 2006).

21 22

RNA extraction

23

Acaryochloris cells were collected by rapid filtration on hydrophilic polyethersulfone

24

filters (Supor 800 Filter, 0.8 µm, Pall, NewYork, USA) and transferred to 15-ml

25

Sarstedt tubes containing 2 ml of PGTX RNA extraction solution (Pinto et al. 2009).

26

Alternatively, the frozen GF-F filters with the samples from the acetylene reduction

9 1

assays were quickly thawed, the cells removed with forceps and immediately placed

2

into a 15-ml Sarstedt plastic tube containing 2 ml of PGTX extraction solution.

3

Samples in Sarstedt tubes were vortexed together with the PGTX, frozen in liquid N2

4

and stored at -80 °C until extraction. Frozen Acaryochloris cells in PGTX were

5

transferred from -80 °C immediately to a 95 °C water bath for 5 min with occasional

6

vortexing and subsequently put on ice for 5 min. For phase extraction, 700 µl of

7

chloroform/isoamylalcohol (IAA) (v/v 24:1) per 1 ml PGTX was added, the mixture

8

shaken vigorously and left at 24°C for 10 min with occasional shaking and vortexing.

9

The phases were separated by centrifugation for 15 min (20 min for GF-F filter

10

samples) at 3,750 rpm (Allegra 6 KR, Beckman Coulter, Krefeld, Germany). The

11

aqueous phase containing the RNA was transferred to a new reaction tube and 1

12

volume chloroform/IAA (v/v 24:1) was added, followed by a second centrifugation

13

step (for GF-F filter samples, the aqueous phase was first re-extracted with phenol

14

(pH 4.5-5)/chloroform/IAA (v/v/v 25:24:1)). The aqueous phase was transferred to a

15

2-ml reaction tube (1 ml per tube) and the RNA precipitated overnight with 1 volume

16

of isopropanol at -20 °C. Precipitated RNA was pelleted by centrifugation for 25 min

17

at 13,000 g at 4 °C, the pellet washed with 1 ml 70 % ethanol and centrifuged for 10

18

min at 13,000 g at 4 °C. The pellet was subsequently air-dried and dissolved in 100

19

µl (30 µl for GF-F filter samples) nuclease-free Milli-Q H2O. RNA concentration and

20

grade of contamination of a sample were measured with a spectrophotometer

21

(NanoDrop ND-1000 or ND-2000, Peqlab, Erlangen, Germany). Quality of RNA was

22

verified on 1.5% formaldehyde-agarose gels.

23 24

Northern blot analysis

25

Blots for the detection of nifD mRNA were prepared from the separation of 1 µg of

26

total RNA on 1.5 % denaturing formaldehyde-agarose gels. Separated RNA was

10 1

transferred to Amersham Hybond-N nylon membranes (GE Healthcare, Munich,

2

Germany) by capillary blotting with 20x SSC buffer (3 M NaCl, 0.3 M C6H5Na3O7, pH

3

7.0). The membranes were hybridized with [α-32P]UTP-labeled transcript probes

4

overnight at 62oC in 50 % deionized formamide, 7 % SDS, 250 mM NaCl, 120 mM

5

phosphate buffer (sodium salts), pH 7.2 in screw-cap hybridization tubes. The

6

membranes were washed in 2× SSC, 1 % SDS for 10 min; 1× SSC, 0.5 % SDS for

7

10 min; and in 0.1× SSC, 0.1 % SDS for 10 min at 57 °C. Signals were detected and

8

analyzed on a Personal Molecular Imager FX system with Quantity One software

9

(Bio-Rad, Munich, Germany).

10 11

Quantitative real-time reverse transcription PCR (RT-qPCR)

12

For elimination of contamination with genomic DNA, 2 µg or 8 µg (for first and second

13

qPCR experiment, respectively) of each RNA sample were incubated twice with 2

14

units of TURBO DNase (TURBO DNA-freeTM Kit, Ambion, Austin, TX, USA) for 20

15

min at 37 °C. The DNase was inactivated following the manufacturer’s instructions.

16

RNA was re-extracted by phenol (pH 4.5 - 5)/chloroform and precipitated with 3

17

volumes of NaOAc/ethanol. RNA samples were tested for residual contamination with

18

DNA by PCR (30 amplification cycles) using the nifD primers of the qPCR reactions.

19

All tests were negative. cDNA synthesis was done with 350 ng RNA of each sample

20

using the Quantitect kit (Qiagen, Hilden, Germany) with a mix of 9 primers (the

21

reverse primer for each gene to be tested), each at a final concentration of 0.5 µM,

22

and following manufacturer’s instructions. A 7500 Fast system (Applied Biosystems,

23

Foster City, CA, USA) was used for qPCR. The modified “Fast Mode” (10 min at 95

24

°C, 40 x (3 s at 95 °C, 30 s at 60 °C)) was used for the preliminary analysis of the

25

“summer culture” samples. For analysis of the samples from the acetylene reduction

26

assays, the longer “Standard Mode” (2 min at 50 °C, 10 min at 95 °C, 40 x (15 s at 95

11 1

°C, 60 s at 60 °C)) was used. Each 15-µl reaction contained 7.5 µl of Power SYBR

2

Green Mix (Applied Biosystems), 4.5 µl of diluted cDNA (1:1280 for nifB, nifD, fdxH,

3

psbA, psaA and coxA; 1:80 for pdhB and 1: 5,242,880 for 16S rRNA), 1.5 µl each of

4

forward and reverse primer. Primers were designed manually to contain similar GC

5

content and have an equal length of 24 nt. The “PrimersList” tool on

6

http://primerdigital.com/tools/ was used for all calculations; for melting temperature

7

calculation the option “using Allawi’s thermodynamics parameters” was chosen

8

(Kalendar et al. 2011). Prior to the final experiment, a primer-matrix was tested

9

(amplifying genomic DNA) to determine the most efficient primer concentration for

10

each primer pair and to confirm primer efficiencies to be similar. Primer sequences,

11

concentrations, and efficiencies are shown in Table 1. Primers were synthesized by

12

Sigma-Aldrich, Hamburg, Germany. All reactions were done in triplicate and for each

13

reaction a ‘no RT’ control was included. 16S rRNA was amplified as reference. Melt

14

curves for all amplifications showed only one product (no product for “no RT” control).

15

Data were analyzed with the 7500 Fast System SDS software (Applied Biosystems),

16

with manual Ct and automatic baseline setting (followed by manual control). Relative

17

transcript quantities were calculated with the ΔΔCt method (Livak & Schmittgen

18

2001). As calibrator conditions, samples of the nitrate-supplemented culture were

19

used, namely the one harvested at 12:00 and 22:00 for the 1st and 2nd experiment

20

(replicate), respectively.

21 22

Software

23

The here described contig was fully annotated and submitted to Genbank (accession

24

JN585763). Artemis Release 12 (Rutherford et al. 2000) was used for visualization

25

and annotation of genome information stored in Genbank or text files. For multiple

26

sequence alignments and BLAST searches in local databases, BioEdit version

12 1

7.0.5.3 (Hall 1999) was used. Phylogenetic analyses were done with MEGA version 5

2

(Tamura et al. 2011) and with Geneious Pro version 5.4.6 (Drummond et al. 2011)

3

using the Mr. Bayes PlugIn (Huelsenbeck & Ronquist 2001). The BLAST algorithm

4

(Altschul et al. 1990) was used with the BLOSUM62 matrix to find putative homologs

5

of genes (BLASTN) or proteins (BLASTP) in the NCBI database and to determine

6

conservation between homologous sequences.

7 8 9

Results and Discussion

10

A cassette for nitrogen fixation genes in Acaryochloris sp. HICR111A

11

During annotation of the Acaryochloris sp. HICR111A genome we encountered a

12

possible cassette for nitrogen fixation (nif) genes on a 63005 bp long contig

13

(GenBank accession JN585763). This contig harbours 60 protein-coding genes, from

14

which 22 physically clustered genes are related to dinitrogen fixation. With a few

15

exceptions like cysteine synthase A and serine-O-acetyltransferase, these have no

16

homologs in the genomes of the two other sequenced Acaryochloris (Swingley et al.

17

2008; Miller et al. 2011) but several flanking genes on both sides of the nif gene

18

cluster have homologs in A. marina (Figure 1), albeit at different positions in the

19

genome. It should be noted that the genome was assembled from DNA of a non-

20

axenic culture containing one strain of an α-proteobacterium and possibly another

21

bacterium. However, these were the only nif genes found in the whole metagenome.

22

Besides the nitrogen fixation cassette, the genome of Acaryochloris sp. HICR111A

23

contains genes for several other nitrogen assimilation pathways, including

24

nitrate/nitrite-, ammonia-, urea-, and cyanate assimilation. It is missing the

25

nitrate/nitrite transporters encoded by nrtABCD and focA, but contains the nrtP gene

13 1

which also codes for a nitrate/nitrite transporter. A similar situation can be found in

2

the reference strain A. marina.

3 4

Genome assemblies solely based on Illumina technology may be erroneous in

5

particular in case of the Acaryochloris genomes which are among the biggest

6

(Swingley et al. 2008; Miller et al. 2011) and most complex genomes in the phylum

7

cyanobacteria (Hess 2011). Therefore, the sequence assembly linking this

8

nitrogenase gene cluster to the neighbouring genes with homologs in A. marina was

9

verified by PCR yielding three amplicons of 365, 2802 and 2182 nt (Figure 2). Among

10

the 22 N2 fixation-related genes are the nifHDK genes encoding the three different

11

nitrogenase subunits, genes related to molybdenum cofactor biosynthesis or

12

transport (hesAB, nifBENX, modABC), Fe-S cluster assembly (nifSU), nif-specific

13

ferredoxins (fdxHN) and a flavodoxin. On the left side, this cluster is flanked by a glgP

14

gene for a putative glycogen/starch/alpha-glucan phosphorylase, closely related to A.

15

marina gene AM1_2114, on the right side by several genes of different phylogenetic

16

matches, including four restriction/modification (RM) genes, followed by 12 homologs

17

of A. marina genes AM1_4104 to AM1_4121 (Figure 1). Very surprisingly, the RM

18

genes are interspersed by a gene encoding a protein of the XisH superfamily. The

19

only known function of these proteins is the cell-type specific excision of a 55 kb

20

element interrupting the fdxN gene during heterocyst differentiation in nitrogen-fixing

21

filamentous cyanobacteria of the Nostocales group (Ramaswamy et al. 1997). The

22

function of this XisH homolog in a unicellular non-heterocystous cyanobacterium

23

appears enigmatic: the fdxN gene in Acaryochloris sp. HICR111A is not interrupted

24

by a genetic element, moreover, the sequence comparison of XisH yielded mainly

25

matches in the Nostocales, but also in some unicellular dinitrogen fixers such as

14 1

Cyanothece and Crocosphaera, as well as in some cyanobacteria lacking this ability,

2

such as Microcystis aeruginosa and, most interestingly, A. marina (gene AM1_4785).

3 4

When the 22 core nif genes were compared against sequence databases, we noticed

5

a remarkable co-linearity with a similar gene cluster in Synechococcus PCC 7335

6

including the same peculiar drop in GC content upstream of nifB (Figure 1). The

7

average blastP amino acid identity of all nif genes is 80%, showing the highest

8

similarity between the homologs of the structural genes nifHDK (90% average

9

identity) and the lowest similarity between the homologs of the nif related genes from

10

cysE to nifT (73% average identity, compare Figure 1). Of all proteins in the NCBI

11

database, Synechococcus PCC 7335 homologs had the highest similarity to those of

12

Acaryochloris sp. HICR111A in all cases (Table 2). Phylogenetic analyses of

13

Acaryochloris sp. HICR111A Nif proteins confirmed a very high sequence similarity,

14

locating the respective orthologs from both strains in the same clade. For these

15

analyses, we concatenated the NifHDK sequences from 19 different cyanobacteria

16

and used the homologs from Rhodospirillum rubrum as outgroup (Figure 3b).

17

Remarkable is not only the tight clustering of Synechococcus PCC 7335 and

18

Acaryochloris sp. HICR111A Nif proteins in a single clade but also the totally

19

unexpected closer relatedness to the respective orthologs in Trichodesmium

20

erythraeum (Figure 3b). Furthermore, a discrepancy to the general phylogeny of

21

these species is evident when comparing NifHDK phylogeny to the 16S rRNA-based

22

phylogeny (Figure 3a). This pattern of presence/absence of nif genes throughout

23

different clades can only be explained by events of selective gene gains or losses

24

and horizontal gene transfer must have played an important role in these processes.

25

Of special interest was also the fact that Acaryochloris sp. HICR111A expresses a

26

fused version of the genes nifE and nifN (Figure 1). Of all fully sequenced

15 1

diazotrophic cyanobacteria, only four strains have such a fused version of nifEN

2

(Figure 3c). Remarkably, besides Acaryochloris sp. HICR111A, these are

3

Synechococcus PCC 7335 (GenBank accession: EDX83984), Trichodesmium

4

erythraeum (NCBI reference sequence: YP_723620.1), and Anabaena variabilis

5

(nifEN2, GenBank accession: AAA93024) which has one fused and one separate

6

version in the nif2 and nif1 clusters, respectively. The expression of nifEN was

7

verified experimentally for Trichodesmium (Dominic et al. 2000) and is evident at the

8

DNA sequence level in the other three organisms. The phylogenetic analysis (Figure

9

3c) of the NifEN protein shows that the occurrence of fused nifEN coincides with

10

higher sequence similarity, clearly separating the Acaryochloris nif genes from those

11

of typical unicellular diazotrophic cyanobacteria. This strengthens the hypothesis that

12

the Acaryochloris nif cluster had originally been acquired by horizontal gene transfer.

13 14

Expression of Acaryochloris sp. HICR111A nif genes is controlled by the light/dark

15

cycle

16

Horizontally transferred genes are not necessarily actively expressed genes. They

17

might just have been acquired and now be in the process of decline. The

18

incompatibility between photosynthetic oxygen production and the well-known

19

oxygen sensitivity of the nitrogenase enzyme should also be considered. Therefore,

20

we set out to measure the expression of these genes in two different ways.

21 22

First, we measured expression of nifD in N-deplete (-N) versus N-replete (+N)

23

conditions in cultures growing under natural light/dark cycles (German summer,

24

varying light intensities and length of day/night). Using quantitative RT-PCR we

25

observed a several thousand fold induction of gene expression in the –N cultures

26

during the dark, starting already towards the end of the light phase (Figure 4a) and

16 1

detected the appearance of a strongly accumulating transcript of about 2.3 kb in

2

Northern blot hybridization in -N cultures only (Figure 4b). The large size of the nifD

3

transcript characterizes it as a likely polycistronic mRNA. Similar results were

4

obtained for other probed nif genes (not shown). The fact that nif genes were already

5

induced before the onset of the dark phase indicates that the circadian clock

6

probably played a role in the timing of the expression, the key components of which

7

are encoded by the kai genes present in this genome. Follow-up experiments

8

including hoaxing experiments with shifted light and dark phases should give further

9

insight.

10

To prove that the induction of gene expression detected at the mRNA level actually

11

results in enzymatically active nitrogenase, we measured nitrogenase activity during

12

six 2h-periods and in parallel analyzed the expression of nif genes in each sample.

13

This experiment was done twice with two different cultures. Figure 5a shows that

14

both cultures, despite different periods of adaptation to the final light and temperature

15

conditions (9 and 22 days, respectively) and differing time points of their shift to N-

16

deplete medium (9 and 7 days, respectively), were performing dinitrogen fixation in

17

the second half of the 12h dark period. In contrast, the expression of three different

18

genes, nifB, nifD and fdxH, measured at the RNA level, started to increase with the

19

begin of the dark period, whereas a gene for a typical photosynthetic protein, psbA,

20

chosen for control, showed an inverse behavior with high expression during the light

21

period and a minimum at night (Figure 5b,c). Thus, compared to the induction of nifB,

22

nifD, and fdxH expression at the RNA level, the peaks of dinitrogen fixation are offset

23

by about 6 hours. The maximum observed nitrogenase activity rate was 485 nmol

24

C2H4 / mg Chl d /2h-1. This relatively low N2 fixation rate corresponds to the slow

25

growth of Acaryochloris sp. HICR111A observed in culture. Our measurements were

17 1

done under micro-aerobic conditions. In an aerobic environment the small diameter

2

of the cells would make it impossible for Acaryochloris sp. HICR111A to maintain an

3

anaerobic intracellular environment for nitrogenase as the diffusion of oxygen would

4

be higher than the scavenging of O2 by respiration. However, Acaryochloris sp.

5

HICR111A forms aggregates in culture (Mohr et al. 2010). Such aggregates may

6

enhance anoxic conditions in the dark and hence support the formation of active

7

nitrogenase. The energy required for N2 fixation in Acaryochloris sp. HICR111A is

8

most likely obtained from the aerobic respiration of the intracellular glycogen reserve,

9

possibly involving the enzymatic activity of the putative glycogen phosphorylase

10

encoded by the glgP gene next to the nif gene cluster. However, we can also not

11

exclude the possibility of anaerobic fermentation, when aggregates of biofilms of this

12

organism turn anoxic at night as observed for other cases (Stal & Moezelaar 1997).

13 14

Conclusions

15

Acaryochloris sp. HICR111A differs from other chlorophyll d-containing Acaryochloris

16

strains in that it possesses a complete and fully functional cluster of nitrogenase

17

genes that seems to be acquired through horizontal gene transfer from other

18

unicellular

19

nitrogenase activity was detected in the second half of the dark phase. Nitrogenase

20

activity was comparatively low but in line with the slow growth rate of this organism.

21

We conclude that Acaryochloris sp. HICR111A may thrive diazotrophically in its

22

natural habitat by forming aggregates or biofilms at or close to the intertidal zone of

23

tropical reefs which are characterized by low oxygen levels during the night.

24

marine

cyanobacteria.

Nitrogenase

genes

were

expressed

and

18 1

REFERENCES

2 3 4 5 6

Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. (1990). Basic local alignment search tool. J Mol Biol 215:403-410. Bergman B, Gallon J., Rai A., Stal LJ. (1997). N2 Fixation by non-heterocystous cyanobacteria. FEMS Microbiol Rev 19:139-185.

7

Berman-Frank I, Lundgren P, Chen Y-B, Küpper H, Kolber Z, Bergman B, et al.

8

(2001). Segregation of nitrogen fixation and oxygenic photosynthesis in the

9

marine cyanobacterium Trichodesmium. Science 294:1534 -1537.

10

Berman-Frank I, Lundgren P, Falkowski P. (2003). Nitrogen fixation and

11

photosynthetic oxygen evolution in cyanobacteria. Res Microbiol 154:157-164.

12

Dominic B, Zani S, Chen Y-B, Mellon MT, Zehr JP. (2000). Organization of the nif

13

genes of the nonheterocystous cyanobacterium Trichodesmium sp. IMS101. J

14

Phycol 36:693-701.

15 16

Drummond A, Ashton B, Buxton S, Cheung M, Cooper A, Duran C, et al. (2011). Geneious v5.4. http://www.geneious.com.

17

Hall TA. (1999). BioEdit: a user-friendly biological sequence alignment editor and

18

analysis program for Windows 95/98/NT. Nucleic Acids Symp Ser 41:95-98.

19

Hess WR. (2011). Cyanobacterial genomics for ecology and biotechnology. Curr

20 21 22 23 24 25 26

Opin Microbiol 14. epub ahead of print Huelsenbeck JP, Ronquist F. (2001). MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics 17:754-755. Kalendar R, Lee D, Schulman AH. (2011). Java web tools for PCR, in silico PCR, and oligonucleotide assembly and analysis. Genomics 98:137-144. Karl D, Michaels A, Bergman B, Capone D, Carpenter E, Letelier R, et al. (2002). Dinitrogen fixation in the world’s oceans. Biogeochemistry 57-58:47-98.

19 1 2

Kashiyama Y, Miyashita H, Ohkubo S, Ogawa NO, Chikaraishi Y, Takano Y, et al. (2008). Evidence of global chlorophyll d. Science 321:658-658.

3

Livak KJ, Schmittgen TD. (2001). Analysis of relative gene expression data using

4

real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods

5

25:402-408.

6

Miller SR, Augustine S, Olson TL, Blankenship RE, Selker J, Wood AM. (2005).

7

Discovery of a free-living chlorophyll d-producing cyanobacterium with a hybrid

8

proteobacterial/cyanobacterial small-subunit rRNA gene. Proc Natl Acad Sci U S

9

A 102:850.

10

Miller SR, Wood AM, Blankenship RE, Kim M, Ferriera S. (2011). Dynamics of gene

11

duplication

12

implications for the ecological niche. Genome Biol Evol 3:601 -613.

13 14

in

the

genomes

of

chlorophyll

d-producing

cyanobacteria:

Miyashita H, Ikemoto H, Kurano N, Adachi K, Chihara M, Miyachi S. (1996). Chlorophyll d as a major pigment. Nature 383:402-402.

15

Mohr R, Voß B, Schliep M, Kurz T, Maldener I, Adams DG, et al. (2010). A new

16

chlorophyll d-containing cyanobacterium: evidence for niche adaptation in the

17

genus Acaryochloris. ISME J 4:1456-1469.

18

Pinto F, Thapper A, Sontheim W, Lindblad P. (2009). Analysis of current and

19

alternative phenol based RNA extraction methodologies for cyanobacteria. BMC

20

Mol Biol 10:79.

21

Ramaswamy KS, Carrasco CD, Fatma T, Golden JW. (1997). Cell-type specificity of

22

the Anabaena fdxN-element rearrangement requires xisH and xisI. Mol Microbiol

23

23:1241-1249.

24

Riemann L, Farnelid H, Steward GF. (2010). Nitrogenase genes in non-

25

cyanobacterial plankton: prevalence, diversity and regulation in marine waters.

26

Aquat Microb Ecol 61:235-247.

20 1

Rippka R, Deruelles J, Waterbury JB, Herdman M, Stanier RY. (1979). Generic

2

assignments, strain histories and properties of pure cultures of cyanobacteria.

3

Microbiology 111:1-61.

4 5 6 7

Ritchie RJ. (2006). Consistent sets of spectrophotometric chlorophyll equations for acetone, methanol and ethanol solvents. Photosynth Res 89:27-41. Rutherford K, Parkhill J, Crook J, Horsnell T, Rice P, Rajandream MA, et al. (2000). Artemis: sequence visualization and annotation. Bioinformatics 16:944-945.

8

Staal M, Lintel-Hekkert ST, Harren F, Stal LJ. (2001). Nitrogenase activity in

9

cyanobacteria measured by the acetylene reduction assay: a comparison

10

between batch incubation and on-line monitoring. Environ Microbiol 3:343-351.

11

Stal LJ. (1988). Nitrogen fixation in cyanobacterial mats. Meth Enzymol 167:474-484.

12

Stal LJ, Moezelaar R. (1997). Fermentation in cyanobacteria. FEMS Microbiol Rev

13 14 15

21:179-211. Steward WDP, Fitzgerald GP, Burris RH. (1967). In situ studies on N2 fixation using the acetylene reduction technique. PNAS 58:2071-2078.

16

Swingley WD, Chen M, Cheung PC, Conrad AL, Dejesa LC, Hao J, et al. (2008).

17

Niche adaptation and genome expansion in the chlorophyll d-producing

18

cyanobacterium Acaryochloris marina. Proc Natl Acad Sci U S A 105:2005.

19

Tamura K, Peterson D, Peterson N, Stecher G, Nei M, Kumar S. (2011). MEGA5:

20

molecular evolutionary genetics analysis using maximum likelihood, evolutionary

21

distance, and maximum parsimony methods. Mol Biol Evol msr121v1-msr121.

22

Tripp HJ, Bench SR, Turk KA, Foster RA, Desany BA, Niazi F, et al. (2010).

23

Metabolic streamlining in an open-ocean nitrogen-fixing cyanobacterium. Nature

24

464:90-94.

25 26

Zehr JP. (2011). Nitrogen fixation by marine cyanobacteria. Trends Microbiol 19:162173.

21 1 2 3 4 5

Zehr JP, Kudela RM. (2011). Nitrogen cycle of the open ocean: from genes to ecosystems. Annu Rev Marine Sci 3:197-225.

22 1

FIGURE LEGENDS

2 3

Figure 1

Comparison of the Acaryochloris HICR111A and Synechococcus PCC

4

7335 nif clusters. (a) GC content aligned with the scaled map of the Acaryochloris nif

5

cluster and surrounding genes. Numbers in the top right corner are %GC. Two

6

intergenic regions exhibiting a characteristic drop in GC content are boxed in red. (b)

7

Map and GC content of the Synechococcus PCC 7335 nif cluster. Orange block

8

arrows indicate genes associated with dinitrogen fixation, grey block arrows indicate

9

genes with known products, but not known to be associated with dinitrogen fixation,

10

white block arrows indicate genes encoding hypothetical proteins. Grey triangles

11

indicate stretches of sequence containing multiple genes not explicitly associated

12

with dinitrogen fixation.

13 14

Figure 2 Three linking PCRs were performed to validate the sequence assembly of

15

contig 62 of the Acaryochloris sp. HICR111A sequencing project containing the nif

16

gene cluster (for a detailed map of the cluster see Figure 1). The positions of forward

17

and reverse primers are indicated by red arrows. Genes with the closest orthologs

18

found in Acaryochloris marina are colored green; genes with the closest orthologs

19

found in the Synechococcus sp. PCC 7335 nif gene cluster in the same order are

20

colored orange; other genes in grey. Genes are clustered into boxes with only those

21

important for the PCRs shown as single arrows. Arrows and boxes are not drawn to

22

scale. The inset shows a 0.8% agarose gel with the three PCR products compared to

23

a λ (PstI) ladder.

24 25

Figure 3 Bayesian trees showing the phylogeny of 16S ribosomal RNA sequences

26

and of concatenated amino acid sequences of nitrogenase structural proteins NifHDK

23 1

and NifEN. (a) Phylogenetic analysis of 40 sequences of 16S ribosomal RNA from N2

2

fixing and other cyanobacteria. Rhodospirillum rubrum ATCC 11170 served as

3

outgroup;

4

(Prochlorococcus and Synechococcus) was collapsed for better overview. (b)

5

Phylogenetic analysis of 21 amino acid sequences of concatenated NifHDK proteins

6

from 19 different cyanobacteria and Rhodospirillum rubrum ATCC 11170 (outgroup).

7

For Anabaena variabilis, NifHDK protein sequences from nif clusters 1 (expressed in

8

the heterocysts) and 2 (expressed in the vegetative cells under anaerobic conditions)

9

were included. Variable C- and N-termini were cut out of the alignment (with a

10

maximum of 55 and 33 amino acids deleted at the NifH-NifD and NifD-NifK junctions,

11

respectively). (c) Phylogenetic analysis of the fused NifEN protein in N2 fixing

12

cyanobacteria and Rhodospirillum rubrum (outgroup). For all strains without the fused

13

version of these genes, amino acid sequences were concatenated for this analysis.

14

Strains which express the fused nifEN are indicated by a grey arrow. In all trees,

15

Acaryochloris strains are highlighted in orange, members of the Nostocales are

16

highlighted in grey. Synechococcus sp. 7335 is framed in orange for comparison with

17

Acaryochloris. All strains from (b) and (c) are included in (a).

one

subtree

showing

11

species

of

marine

picocyanobacteria

18 19

Figure 4 Transcript levels of nifD by quantitative real-time reverse transcription PCR

20

(RT-qPCR) and Northern blot analysis in two different experiments. a) Relative

21

transcript levels of nifD obtained by qPCR in Acaryochloris HICR111A grown with

22

(+N) and without (-N) combined nitrogen (nitrate) in summer under natural light

23

conditions. The shaded grey area shows the measured light intensity during the

24

sampling period. The data were normalized to 16S rRNA and calibrated to the 12:00

25

nitrate sample. (b) Northern blot analysis of nifD transcripts during the dark phase in

26

Acaryochloris HICR111A grown under controlled conditions with a defined 12h/12h

24 1

light/dark cycle. –N: grown without combined nitrogen; NO: grown with nitrate (NO3-).

2

5S rRNA was hybridized as internal standard.

3 4

Figure 5 Nitrogenase activity and real-time reverse transcription PCR (RT-qPCR)

5

analysis of relative transcript levels from genes involved in dinitrogen fixation and

6

photosynthesis. Shown are two different experiments with Acaryochloris HICR111A

7

cultures grown under a 12h/12h light-dark cycle. Nitrogenase activity (acetylene

8

reduction) and transcript levels were measured at six points in time during the

9

12h/12h light-dark cycle. (a) Nitrogenase activity. The culture used for measurements

10

on 02.06.2011 had adapted to the final temperature and light conditions for 9 days

11

and was shifted to nitrogen-deplete medium 9 days prior to the assay, the culture

12

used for measurements on 03.06.2011 had adapted to the final temperature and light

13

conditions for 22 days and was shifted to nitrogen-deplete medium 7 days prior to the

14

assay. Data points are set in the middle of the acetylene incubation time (total

15

incubation 2h). (b) and (c) relative transcript levels of nifB, nifD, and fdxH (nif cluster

16

genes) as well as psbA in nitrogen deplete culture samples. All transcript data were

17

normalized to 16S rRNA and calibrated to their respective mRNA levels in the one

18

sample of nitrate-supplemented culture harvested at 12:00 and 22:00 for the 1st (a)

19

and 2nd (b) experiment, respectively (relative expression for these samples is 1). The

20

black and white bars above each graph indicate light- and dark-periods.

a

Acaryochloris sp. HICR111A ~ 20000 bp

orthologs to A. marina genes AM1_4121 to AM1_4104

1 kb

b ~ 10000 bp

~ 38800 bp

63.0

Synechococcus PCC 7335

31.0

1st amplicon365 bp

2nd amplicon2802 bp

3rd amplicon2182 bp

forward primer

modA - nifB

cysE – nifT 1 t 1st

reverse primer 4507 bp 2838 bp 2140 bp

805 bp 448 bp 339 bp

2 d 2nd

3d 3rd

genes of mixed origin

A. marina orthologs

a

Nostoc punctiforme PCC 73102 Anabaena variabilis ATCC 29413

100

100

Anabaena sp. PCC 7120 'Nostoc azollae' 0708

81

65

Raphidiopsis brookii D9

90 100

53

Cylindrospermopsis raciborskii CS-505

Microcoleus chthonoplastes PCC 7420 Trichodesmium erythraeum IMS101 100

Cyanothece sp. PCC 8802 Cyanothece sp. PCC 8801

77 100

Cyanothece sp. WH 8904 Gloeothece sp. KO68DGA

16S ribosomal RNA

98

Cyanothece sp. ATCC 51142

61

100

100

Crocosphaera watsonii WH 8501

UCYN-A Cyanothece sp. PCC 7822

100

100

b

Cyanothece sp. PCC 7424 70

Synechocystis sp. PCC 6803 99

concatenated nifHDK

Synechococcus sp. PCC 7002

Cyanothece sp. PCC 7425 73

Acaryochloris sp. HICR111A

97

100

Acaryochloris marina MBIC11017

Synechococcus sp. PCC 7335

Anabaena variabilis ATCC 29413 (nif1) 100

98

Anabaena variabilis ATCC 29413 (nif2) Anabaena sp. PCC 7120

Synechococcus elongatus PCC 7942

99

97

Cyanothece sp. PCC 7425

100

100

A Acaryochloris hl i sp. CCMEE5410

62

100

Picocyanobacteria (11 species)

Synechococcus sp. OS-B 100

Nostoc punctiforme PCC 73102

100

'Nostoc azollae' 0708

100

Synechococcus sp. OS-A

100

Cylindrospermopsis raciborskii CS-505

68

Gloeobacter violaceus PCC 7421

Gloeothece sp. KO68DGA

100

Cyanothece sp. ATCC 51142

100

Rhodospirillum rubrum ATCC 11170 100

UCYN-A

0.05

Crocosphaera watsonii WH 8501

Nostoc punctiforme PCC 73102

c

100 100

Cyanothece sp. PCC 7424

100

Anabaena variabilis ATCC 29413 nif1 100

100

98

Anabaena sp. PCC 7120

100

'Nostoc azollae' 0708 100

68

100

Cylindrospermopsis raciborskii CS-505 100

Trichodesmium erythraeum IMS101 51

Synechococcus sp. OS-B Acaryochloris sp. HICR111A 100

C Cyanothece th sp. PCC 7425 98

Synechococcus PCC 7335 100

Acaryochloris sp. HICR111A Synechococcus sp. OS-A

0.05

100

Synechococcus sp. OS-B Cyanothece sp. PCC 7822 Cyanothece sp. PCC 8801 100

Cyanothece sp. PCC 8802 UCYN-A 100 100

concatenated/fused nifEN

Gloeothece sp. KO68DGA Cyanothece sp. ATCC 51142 100

Crocosphaera watsonii WH 8501 Rhodospirillum rubrum ATCC 11170

0.05

Synechococcus sp sp. PCC 7335 Trichodesmium erythraeum IMS101

Rhodospirillum rubrum ATCC 11170

100

100

Cyanothece sp. PCC 8802

Synechococcus sp. OS-A

Anabaena variabilis ATCC 29413 nif2

100

Cyanothece sp. PCC 7822

Cyanothece sp. PCC 8801

a Nitrogena ase activity [nmol C2H4 / m mg Chl d /2h-1]

500.00 450 00 450.00 400.00 350.00 300.00 250.00

02.06.2011

200.00

03.06.2011

150.00 100.00 50.00 0 00 0.00 8:00

10:00

12:00

14:00

16:00

18:00

20:00

22:00

0:00

time of the day

b

c

relative e gene expressio on (log2 scale)

16384 0 16384.0 4096.0 4096.0 1024.0

1024.0

256.0

nifB

64.0

nifD fdxH

16.0

psbA

256.0 64.0 16.0

4.0

4.0

1.0

1.0

0.3

0.3 8:00 10:00 12:00 14:00 16:00 18:00 20:00 22:00 0:00 2:00

time of the day

8:00 10:00 12:00 14:00 16:00 18:00 20:00 22:00 0:00 2:00

time of the day

Table 1 Oligonucleotides used in this work

primer name

sequence in 5'3' direction

coxAfw coxArev fdxHfw fdxHrev nifBfw nifBrev nifDfw nifDrev pdhBfw pdhBrev psaAfw psaArev psbAfw psbArev 16S_fw 16S_rev

aagcacaccatctagacgaagtgg gaatcccaatcaccttatggtcgg atatcacgattcccgtggatgagg cttgccaacacagctagaacaagc aaacgaaggtcaagaagtctggtg gtagcagggatgcttttcaatccg agaaggcaccctgatttatgacga ggcaaagccattttctggaagaca agtactgttatttaacgccctccg tttgtaggagccaccatagtgtcc aacagctgaagttgacatcaaccc tgcaggttccaaatccaagttgtg tgacaacagtattgcaaagacgcg ccaaaccaaccgatgtataaacgg ttggaaacgactgctaataccgga actgatcgaagtcttggtaagcca

concentration in assay 0.6 µM 0.6 µM 0.6 µM 0.6 µM 0.6 µM 0.6 µM 0.6 µM 0.6 µM 1 µM 1 µM 1 µM 1 µM 0.6 µM 0.6 µM 0.6 µM 0.6 µM

primer PCR efficiency (Kalendar et al. 2011) 69 % 88 % 92 % 90 % 85 % 87 % 83 % 88 % 86 % 89 % 88 % 83 % 90 % 79 % 87 % 88 %

Table 2 Similarities between Acaryochloris HICR111A Nif proteins and their homologs in Synechococcus PCC 7335 (percentages of identical or similar residues measured by blastP). For Synechococcus PCC 7335, the sequence overlap was over 97% for 16 of the shown 22 proteins, and over 90% for the rest of the proteins except for NifB (80%) that is likely annotated too short in Synechococcus PCC 7335. An asterisk next to a number indicates that this is the top hit found in the NCBI database.

Gene or Gene Product

Synechococcus PCC 7335 blastP POSITIVES (%)

blastP IDENTITY (%)

modBC

78

68 *

cysteine synthase flavodoxin fdxH hesB hesA nifW ORF2 nifX nifEN nifZ nifK nifD nifH nifU nifS nifB cysE ferritin dps family nifV nifT/fixT parB

92 93 92 88 93 85 87 86 91 90 93 95 93 90 94 90 81 86 89 86 66

83 * 87 * 80 * 79 * 85 * 74 * 72 * 75 * 82 * 78 * 90 * 91 * 89 * 83 * 87 * 82 * 72 * 76 * 78 * 66 * 53 *