IJSEM Papers in Press. Published May 8, 2014 as doi:10.1099/ijs.0.064576-0

1

Roseiarcus fermentans gen. nov. sp. nov., a Bacteriochlorophyll a - Containing

2

Fermentative Bacterium Phylogenetically Related to Alphaproteobacterial

3

Methanotrophs and Description of the Family Roseiarcaceae fam. nov.

4 5

Irina S. Kulichevskaya1, Olga V. Danilova1, Vera M. Tereshina1, Vadim V.

6

Kevbrin1, and Svetlana N. Dedysh1

7 8

1

9

Russia

S.N. Winogradsky Institute of Microbiology, Russian Academy of Sciences, Moscow 117312,

10 11

Author for correspondence: Svetlana N. Dedysh

12

Tel: 7 (499) 135 0591. Fax: 7 (499) 135 6530. Email: [email protected]

13 14 15

Journal: International Journal of Systematic and Evolutionary Microbiology (IJSEM)

16

Contents Category: New Taxa - Proteobacteria

17 18

Running title: Roseiarcus fermentans gen. nov., sp. nov.

19

The GenBank/EMBL/DDBJ accession numbers for the 16S rRNA gene sequence and the partial

20

sequence of the nifH gene of Roseiarcus fermentans Pf56T are KJ406703 and KJ406704,

21

respectively.

1

1

ABSTRACT

2

A light-pink-pigmented, microaerophilic bacterium was obtained from a methanotrophic

3

consortium enriched from acidic Sphagnum peat and designated strain Pf56T. Cells of this

4

bacterium are Gram-negative, non-motile, thick curved rods, which contain a vesicular

5

intracytoplasmic membrane system characteristic for some purple non-sulfur

6

Alphaproteobacteria. Absorption spectrum of the acetone: methanol extracts of cells grown

7

in the light shows maxima at 363, 475, 505, 601 and 770 nm; the peaks 363 and 770 nm are

8

characteristic for bacteriochlorophyll a. However, in contrast to purple non-sulfur

9

bacteria, strain Pf56T is unable to grow phototrophically under anoxic conditions in the

10

light. Best growth occurs on some sugars and organic acids under micro-oxic conditions by

11

means of fermentation. The fermentation products are propionate, acetate, and hydrogen.

12

Slow chemo-organotrophic growth is also observed under fully oxic conditions. Light

13

stimulates growth. C1 substrates are not utilized. Strain Pf56T grows in the pH range 4.0-

14

7.0 (optimum 5.5-6.5) and at 15- 30 C (optimum 22–28 °C). The major cellular fatty acids

15

are 19:0 CYCLO ω8c and 18:1ω7c; quinones are represented by ubiquinone Q10. The

16

G+C content of the DNA is 70.0 mol%. Strain Pf56 displays 93.6-94.7 and 92.7-93.7% 16S

17

rRNA gene sequence similarity to members of the families Methylocystaceae and

18

Beijerinckiaceae, respectively, and belongs to a large cluster of environmental sequences

19

retrieved in cultivation-independent studies from various wetlands and forest soils.

20

Phenotypic, genotypic and chemotaxonomic characteristics of strain Pf56 T suggest that it

21

represents a novel genus and species of bacteriochlorophyll a-containing fermentative

22

bacteria, for which the name Roseiarcus fermentans is proposed. Strain Pf56T (=DSM

23

24875T = VKM B-2876T) is the type strain, which is also the first characterized member of

24

a novel family within the class Alphaproteobacteria, Roseiarcaceae fam. nov.

o

2

1

Keywords: Roseiarcus fermentans gen. nov., sp. nov., Roseiarcaceae fam. nov.,

2

bacteriochlorophyll a-containing bacteria, fermentative growth, acidic wetlands.

3 4

The class Alphaproteobacteria accommodates a wide variety of microorganisms with the

5

phototrophic lifestyle, including the purple non-sulfur bacteria and the aerobic

6

bacteriochlorophyll a-containing bacteria (Imhoff, 2006; Swingley et al., 2009). Purple non-

7

sulfur bacteria are metabolically versatile anoxygenic phototrophs that grow phototrophically

8

under anoxic conditions in the light or chemotrophically under microoxic to oxic conditions in

9

the dark (Imhoff, 1995, 2001). Many purple non-sulfur bacteria are also capable of fermentative

10

growth under anoxic conditions in the dark (Uffen & Wolfe, 1970; Madigan et al., 1980; Schultz

11

& Weaver, 1982). By contrast, aerobic bacteriochlorophyll a-containing bacteria are incapable of

12

anaerobic photosynthesis and light stimulates a transient enhancement of aerobic growth after a

13

shift from the dark to illumination (Yurkov & Beatty, 1998; Yurkov, 2001). In our study, we

14

describe an unusual, microaerophilic, bacteriochlorophyll a-possessing isolate, strain Pf56T,

15

which prefers to grow by means of fermentation and is phenotypically different from both the

16

purple non-sulfur bacteria and the earlier described aerobic bacteriochlorophyll a-containing

17

bacteria.

18

Strain Pf56T was isolated from a methanotrophic enrichment culture that was established

19

from an acidic peat soil (pH 3.8) sampled at a depth of 10 cm of Sphagnum peat bog Staroselsky

20

moss (56o34’ N, 32o46’ E), Tver region, Russia, in August 2008. A methanotrophic enrichment

21

culture was obtained from this peat sample using liquid mineral medium M2 of pH 5.0 (Dedysh

22

et al., 2000) and incubation with 30% (vol/vol) methane. A cell suspension of this enrichment

23

culture was serially diluted and used to inoculate deep-agar cultures prepared with semi-liquid

3

1

agar (0.1%, w/v) medium M2 in 25-ml test tubes. The inoculated tubes were sealed with

2

Parafilm M and incubated in jars with 30% (vol/vol) methane under light. After 1.5 months of

3

incubation, intensive development of a cream-colored methanotrophic biofilm was observed in

4

the surface (0-3 mm depth) agar layer of these tubes. A methanotrophic bacterium, strain S284,

5

isolated this biofilm was further described as a member of a novel species of the genus

6

Methylocystis, Methylocystis bryophila (Belova et al., 2013). At a depth of 1-3 cm below this

7

biofilm, however, we noticed the development of brownish-red-colored colonies of 2-3 mm in

8

diameter. Microscopic examination of a cell material taken from these colonies revealed the

9

presence of thick curved rods, which occurred singly or in pairs (Fig. 1A). These cells showed

10

bright probe-conferred signal when hybridized to the probe M-450 with reported group

11

specificity for Methylosinus/Methylocystis-like methanotrophs (Eller et al., 2001) and, therefore,

12

were preliminary identified as belonging to the family Methylocystaceae. Given these

13

identification results, the routine procedure for methanotroph isolation was employed. The

14

respective cell material was spread plated onto the medium M2 solidified with Phytagel and the

15

plates were incubated at 20°C in desiccators under a methane/air (50 : 50) gas mixture. The

16

unpigmented colonies that developed on these plates after 6 weeks of incubation were again

17

examined by means of phase-contrast microscopy and whole-cell hybridization with the probe

18

M-450. The selected colonies were picked and purified by successive re-streaking until the target

19

organism, designated strain Pf56T, was obtained in a pure culture. Partial sequencing of the 16S

20

rRNA gene from this isolate showed that it is affiliated with the Alphaproteobacteria and is

21

equally divergent (92-94% sequence similarity) from members of the two methanotroph-

22

accommodating families, Methylocystaceae and Beijerinckiaceae.

4

1

Unexpectedly, our further attempts to maintain this isolate as a methanotrophic

2

bacterium, in screw-cap 500 ml serum bottles containing 100 ml liquid medium M2 and 10-20%

3

(v/v) CH4 in the headspace, were unsuccessful. No growth was observed also when strain Pf56T

4

was re-streaked on plates with medium M2 solidified with agar and incubated in a desiccator

5

under a methane/air gas mixture. Apparently, development of this bacterium on the isolation

6

medium could only be explained by Phytagel utilization. This was further confirmed by

7

cultivating strain Pf56T in 160 ml screw-cap serum bottles filled with 20 ml liquid medium M2

8

and supplemented with 0.05% (wt/vol) Phytagel as a carbon source. These growth conditions,

9

however, were not fully optimal for strain Pf56T since only a relatively poor growth was

10

observed. Further search for the optimal growth conditions revealed that strain Pf56T can be

11

maintained in flasks filled by 4/5th - 9/10th with a liquid medium M3 containing (g per liter of

12

distilled water): KH2PO4, 0.1; NH4Cl, 0.2; MgSO4 × 7H2O, 0.1; CaCl2× 2H2O, 0.02; yeast

13

extract, 0.1; glucose (or malate) 0.5; and 1 ml of trace element solution ‘SLA’ (Imhoff, 2006),

14

water 1L; pH 5.8-6.0. Growth in this medium was observed after 15-20 days of incubation in

15

static conditions. Incubation in the light caused apparent (approximately by 20-30%) stimulation

16

of growth.

17

Liquid cultures of strain Pf56T grown in the light displayed homogenous turbidity and

18

were slightly pinkish in color. This color intensity was not comparable to that in known purple

19

bacteria, which form intensively colored, purplish red cell suspensions when incubated in the

20

light under anoxic conditions. Nonetheless, cell pellets of strain Pf56T collected from the cultures

21

incubated in the light were red and, therefore, were examined for the presence of pigments. The

22

absorption spectra of living cells were recorded with a LOMO SPh-56 spectrophotometer. For

23

these measurements, the cells were suspended in 50% glycerol. In addition, pigments were

5

1

extracted with acetone/methanol (7:1, v/v) and the absorption spectra of these extracts were

2

recorded. Carotenoids were analyzed by HPLC as described previously (Kulichevskaya et al.,

3

2006). The absorption spectrum of the acetone: methanol extracts of cells grown in the light

4

showed maxima at 363, 475, 505, 601 and 770 nm (Fig. 2). The peaks 363 and 770 nm are

5

characteristic for bacteriochlorophyll a. Absorption maxima of living cells were at 455, 489, 528,

6

593, 806 and 865 nm. Carotenoid analysis by HPLC identified spirilloxanthin (35.3% of total

7

carotenoids), rhodopin (34.8%), and 3,4- didehydrorhodopin (15.5%) as the major carotenoids of

8

strain Pf56T. Lycopene (8.1%) and anhydrorhodovibrin (3.3%) were also detected. Cell pellets of

9

strain Pf56T collected from the cultures incubated in the dark were light-cream;

10

bacteriochlorophyll a was not detected in these cells.

11

Morphological observations and cell-size measurements were made with a Zeiss Axioplan 2

12

microscope and Axiovision 4.2 software (Zeiss). Cells of strain Pf56T were Gram-negative, non-

13

motile, non-spore-forming, thick curved rods with an outer diameter of 2.7-4.0 µm and a width

14

of 0.6-1.2 µm, that occurred singly or in pairs (Fig. 1A). For preparation of ultrathin sections,

15

cells of the exponentially growing cultures were collected by centrifugation and pre-fixed with

16

1.5% (w/v) glutaraldehyde in 0.05 M cacodylate buffer (pH 6.5) for 1 h at 4oC and then fixed in

17

1% (w/v) OsO4 in the same buffer for 4 h at 20oC. After dehydration in an ethanol series, the

18

samples were embedded in a Spurr epoxy resin. Thin sections were cut on an LKB-4800

19

microtome, stained with 3% (w/v) uranyl acetate in 70% (v/v) ethanol. The specimen samples

20

were examined with a JEM-100C transmission electron microscope. Electron microscopy of

21

ultrathin sections revealed the presence of a vesicular intracytoplasmic membrane system (Fig.

22

1B), which is characteristic for some purple non-sulfur alphaproteobacteria (Imhoff, 2006).

23

Granules of polyhydroxybutyrate and polyphosphate were also observed in cells of strain Pf56T.

6

1

Tests for photo-organotrophic growth in the light under anoxic conditions were performed in

2

screw-cap 120 ml serum bottles containing 100 ml of medium M3 with 0.05% (w/v) acetate,

3

malate, succinate, glucose, fructose or pyruvate. Before autoclaving, these flasks were flushed

4

with N2 for 10 min. No growth of strain Pf56T was observed under these conditions. Negative

5

results were also obtained in tests for photo-lithotrophic growth, which were performed in screw-

6

cap 120 ml serum bottles containing 30 ml medium M3 without a carbon source and a mixture of

7

H2 and CO2 (8 : 2, v/v) in a gas phase. Apparently, strain Pf56T was incapable of phototrophic

8

growth under anoxic conditions.

9

Tests for aerobic growth in the dark were performed in screw-cap 120 ml serum bottles

10

containing 10 ml of medium M3 with 0.05% (w/v) concentrations of the following carbon

11

sources: glucose, fructose, xylose, sucrose, cellobiose, trehalose, acetate, malate, citrate,

12

succinate, pyruvate, α-oxoglutarate, propionate, galacturonic and glucuronic acids, ethanol,

13

methanol, formate, and Phytagel. The bottles were incubated on a shaker (120 r.p.m.) and the

14

growth was monitored by measuring OD600 and comparing to a negative control. Under these

15

conditions, only a slow growth (generation time of 40-60 h) of strain Pf56T was observed on

16

glucose, fructose, xylose, malate, pyruvate, galacturonic acid and Phytagel. The cultures grown

17

under fully aerobic conditions contained forked and misshapen cells indicating that these

18

conditions are not optimal for this bacterium (Suppl. Fig. S1).

19

The same range of growth substrates was examined during incubations in the dark in micro-

20

oxic conditions (bottles filled with a liquid medium by 9/10th of total volume with ambient air in

21

the headspace). Best growth (doubling time 25-30 h) occurred with glucose, fructose, xylose, and

22

galacturonic acid, while strain Pf56T grew also with trehalose, malate, succinate, pyruvate, and

23

Phytagel. Under these conditions, the growth occurred by means of fermentation. Fermentation

7

1

products (organic acids, alcohols) and substrates (sugars) were assayed using a Stayer HPLC

2

chromatograph (Aquilon, Russia) equipped with a refractometric detector (Knauer, Germany)

3

and an Aminex HPX-87H column (Bio-Rad, USA), operated isocratically with 5 mM H2SO4 as

4

eluent at 0.6 ml/min. The products of fructose fermentation by strain Pf56T were acetate,

5

propionate, and hydrogen (Suppl. Fig. S2). The same fermentation products were detected in

6

micro-oxic cultures of strain Pf56T incubated in the light. Notably, the growth yield in the light

7

was somewhat (approximately 20-30%) higher than that in the dark.

8 9

Nitrogen sources were tested by replacing NH4Cl in a medium M3 with 0.01% (w/v) KNO3, NaNO2, urea, alanine, asparagine, histidine, lysine, glutamate or yeast extract. Strain Pf56T

10

utilized ammonium salts, histidine, glutamate and yeast extract as nitrogen sources. It was also

11

capable of growth in a liquid nitrogen-free medium. Partial fragment of the nifH gene (encoding

12

dinitrogenase reductase) was amplified using primers and reaction conditions described by

13

Dedysh et al. (2004b). The sequences of nifH gene fragments from this bacterium displayed

14

highest similarity (90-92% nucleotide sequence similarity and 95-96% derived amino acid

15

sequence identity) to the corresponding gene fragments from various strains of

16

Rhodopseudomonas palustris.

17

Despite its isolation from a methanotrophic enrichment culture, strain Pf56T was unable to

18

grow on methane. The presence of a key enzyme of aerobic methanotrophs, the particulate

19

methane monooxygenase (pMMO), in this bacterium could not be confirmed by any of the

20

commonly used approaches. Cells of this isolate did not contain intracytoplasmic membranes

21

characteristic of all pMMO-possessing methanotrophic proteobacteria. Our attempts to amplify a

22

pmoA gene fragment from DNA of strain Pf56T using any of the two primer sets for this gene,

23

i.e. A189/A682 (Holmes et al., 1995) and A189/Mb661r (Costello & Lidstrom, 1999), were

8

1

unsuccessful. The mmoX gene, coding for the soluble methane monooxygenase (sMMO), also

2

could not be amplified from DNA of our novel isolate with any of the previously described

3

mmoX-targeted primers (McDonald et al., 1995; Miguez et al., 1997; Shigematsu et al., 1999;

4

Auman et al., 2000; Vorobev et al., 2011). Methanol and formate also did not support the growth

5

of strain Pf56T. In summary, we did not find any evidence for the presence of C1 metabolic

6

capabilities in this bacterium.

7

Physiological tests were performed in the liquid medium M3 with glucose under micro-

8

oxic conditions in the dark. Growth of strain Pf56T was monitored for 10-14 days under a variety

9

of conditions, including temperatures of 10-37oC, pH 3.5-7.5 and NaCl concentrations of 0.01-

10

3.0 % (w/v). Variations in the acidity level were achieved by mixing 0.1M solutions of HCl and

11

KOH. Strain Pf56T grew in the pH range 4.0-7.0 with a pH optimum of 5.5-6.5 (Fig. 3). The

12

temperature range for growth was 15- 30 C, with an optimum at 22–28 °C. Growth inhibition

13

was observed in the presence of NaCl in the medium above 0.5% (w/v).

o

14

The presence of catalase in strain Pf56T was tested by using the Method 1 described by

15

Gerhardt et al. (1981). Oxidase was tested using the commercial kit (bioMérieux). Strain Pf56T

16

was cytochrome oxidase-negative and catalase-possitive.

17

Cell biomass for cellular fatty acid, polar lipid, isoprenoid quinone analyses and for DNA

18

extraction was obtained from batch cultures grown in liquid medium M3 with glucose under

19

micro-oxic conditions in the light at 20ºC for 2 weeks. This corresponds to the late exponential

20

growth phase. The fatty acid profiles were analyzed at the Identification Service of the Deutsche

21

Sammlung von Mikroorganismen und Zellkulturen (DSMZ, Braunschweig, Germany) as

22

described by Kämpfer & Kroppenstedt (1996), using an Agilent 6890N GC, Sherlock MIS

23

version 6.0 and the library TSBA40 4.10. The cellular fatty acid composition of strain Pf56T is

9

1

shown in Table 1. Similar to many alphaproteobacteria, strain Pf56T contained significant

2

amounts (26.3%) of 11-cys-octadecenoic acid (18:1ω7c), which is also typical for members of

3

the families Methylocystaceae and Beijerinckiaceae. However, the major component of the fatty

4

acid profile (39% of total fatty acids) in our novel isolate was 19:0 CYCLO ω8c, which is

5

absent from methanotrophs of the Methylocystaceae (Bodelier et al., 2009) and was detected in

6

significant amounts (up to 13%) only in a very few members of the Beijerinckiaceae, such as

7

Methylocella tundrae (Dedysh et al. 2004a).

8 9

For polar lipid analysis, the cells were washed with distilled water, re-suspended in isopropanol and incubated for 30 min at 70°C. The biomass residue was removed by filtration

10

and extracted twice with isopropanol–chloroform (1 : 1) (Nichols, 1963). The combined extract

11

was dried on a rotor evaporator; the residue was dissolved in 3 ml of chloroform–methanol (1:

12

1), and the water-soluble compounds were removed by addition of 2.5% sodium chloride (4 ml).

13

The chloroform layer was separated, dehydrated by passing through anhydrous sodium sulfate,

14

dried in a rotor evaporator, and vacuum-dried to a constant weight. The obtained residue was

15

dissolved in chloroform–methanol (1 : 1) and stored at –21°C. The separation of phospho- and

16

sphingolipids was performed by two-dimensional TLC on glass plates (Merck, Germany) using

17

chloroform–methanol–water (65 : 25 : 4) in the first direction and chloroform–acetone–

18

methanol–acetic acid–water (50 : 20 : 10 : 10 : 5) in the second direction (Benning et al., 1995).

19

Several plates with increased amount of the analyzed material were used in order to ensure the

20

detection of lipids present in minor amounts. The chromatograms were developed by spraying

21

with 5% sulfuric acid in ethanol with subsequent heating at 180°C. The lipids were identified

22

using the individual markers and qualitative reactions with ninhydrin (for the presence of amino

23

groups), the Dragendorff reagent (for choline), and α-naphthol (for glycolipids). The

10

1

sphingolipid nature of glycolipids was determined by the saponification method (Kates, 1972).

2

Lipid analyses were carried out using the following standards: phosphatidylcholine (Sigma,

3

USA) for phospholipids and glycoceramide mixture (Larodan, Sweden) for sphingolipids. The

4

polar lipids of strain Pf56T consisted of phosphatidylcholine, phosphatidylethanolamine,

5

phosphatidylglycerol, cardiolipin and sphingolipids (Suppl. Fig. S3). This polar lipid

6

composition is characteristic for phototrophic bacteria of the family Rhodospirillaceae (Imhoff et

7

al., 1982).

8

Isoprenoid quinones were extracted according to Collins (1985) and analyzed using a

9

tandem-type MS LCQ ADVANTAGE MAX and ionization MS Finnigan Mat 8430 with

10

atmospheric pressure chemical ionization. The mass spectra were first recorded in MS mode and

11

then analyzed using MS/MS mode. Similarly to many phototrophic representatives of the

12

Alphaproteobacteria, cells of strain Pf56T contained ubiquinone Q10.

13

The DNA base composition of strain Pf56T was determined by thermal denaturation using a

14

Cary 100Bio spectrophotometer (Varian, USA) at a heating rate of 0.5°C min-1. The mol % G+C

15

value was calculated according to Owen et al. (1969). The G+C content of the DNA of strain

16

Pf56T was 70.0 mol%, which is higher than in all currently characterized members of the

17

Methylocystaceae and Beijerinckiaceae (GC content in the range of 55-65 mol%). At the same

18

time, it is within the range of high GC content characteristic for DNA from purple nonsulfur

19

bacteria (GC content in the range of 59-72 mol%) (Imhoff, 1995).

20

PCR-mediated amplification of the 16S rRNA gene was performed using primers 9f and

21

1492r and reaction conditions described by Weisburg et al. (1991). Phylogenetic analysis was

22

carried out using the ARB program package (Ludwig et al., 2004). The trees were constructed

23

using distance-based (neighbor-joining), maximum-likelihood (DNAml), and maximum-

11

1

parsimony methods. The significance levels of interior branch points obtained in neighbor-

2

joining analysis were determined by bootstrap analysis (1000 data re-samplings) using PHYLIP

3

(Felsenstein, 1989). A phylogenetic tree constructed on the basis of 16S rRNA gene sequences

4

(Fig. 4) indicated that strain Pf56T belongs to the class Alphaproteobacteria and forms a lineage,

5

which is most closely related to the family Methylocystaceae (93.6-94.7% sequence similarity).

6

Interestingly, the probe M-450 with reported group specificity for Methylosinus/Methylocystis-

7

like methanotrophs (Eller et al., 2001) matched the corresponding target region of 16S rRNA

8

gene of strain Pf56T, and the cells of this bacterium showed bright probe-conferred signal when

9

hybridized to this probe. This may cause an overestimation of methanotroph abundance in

10

environments colonized with strain Pf56T-like bacteria. The 16S rRNA gene sequence similarity

11

of strain Pf56T with members of the family Beijerinckiaceae was between 92.7 and 93.7%. The

12

most closely related phototrophs which, nonetheless, belong to a separate lineage, are acidophilic

13

purple bacteria Rhodoblastus acidophilus and Rhodoblastus sphagnicola (93.4-93.7% 16S rRNA

14

gene similarity). At the same time, strain Pf56T displayed high (96.3-98.6%) 16S rRNA gene

15

similarity to a large number of environmental sequences that were retrieved in cultivation-

16

independent studies from various boreal and tropical peatlands (GenBank accession numbers

17

FR720624, FR720611, AM162440, AM162436, JX504960, AY163571, GQ402716, GQ402766,

18

GQ402819), forest soil (FJ624894) and some other environments such as volcanic deposits

19

(AY425766). The phylogenetic cluster formed by these environmental 16S rRNA gene

20

sequences and the corresponding gene sequence from strain Pf56T was stable independently of

21

the algorithm used for the tree construction and was supported by a bootstrap value of 100%

22

(Fig. 4).

12

1

In summary, we showed that strain Pf56T is phenotypically distinct from all phylogenetically

2

related alphaproteobacteria. It does not possess methanotrophic capabilities and, therefore,

3

cannot be placed in the family Methylocystaceae, which is phenotypically uniform and

4

accommodates aerobic, pMMO-containing methanotrophs (Bowman, 2000). The family

5

Beijerinckiaceae is more heterogeneous and contains both organotrophs and methano-

6

/methylotrophs (Tamas et al., 2014) but all these organisms are strict aerobes. Finally, strain

7

Pf56T is different from anaerobic phototrophs of the genus Rhodoblastus in possessing

8

fermentative metabolism and inability to grow phototrophically under anoxic conditions. We

9

therefore suggest that strain Pf56T (=DSM 24875T=VKM B-2876T) should be classified as a

10

novel genus and species of bacteriochlorophyll a-possessing, fermentative bacteria, for which the

11

name Roseiarcus fermentans is proposed. We also propose a novel family to accommodate this

12

novel genus, Roseiarcaceae fam. nov.

13 14

Description of Roseiarcus gen. nov.

15

Roseiarcus (Ro.se.i.ar’cus. L. adj. roseus, rose, pink; L. masc. n. arcus, arch, bow; N. L. masc. n.

16

Roseiarcus, pink bow).

17

Gram-negative, non-motile, non-spore-forming, pink-pigmented, thick curved rods that occur

18

singly or in pairs. Reproduce by normal cell division. Cells contain bacteriochlorophyll a and a

19

vesicular intracytoplasmic membrane system when grown in the light. Grow best in micro-oxic

20

conditions by means of fermentation. Growth substrates are some sugars and organic acids.

21

Unable to grow phototrophically under anoxic conditions. Capable of slow aerobic

22

chemoorganotrophic growth. Incubation in the light stimulates growth. C1 compounds are not

23

utilized. Capable of dinitrogen fixation. Moderately acidophilic and mesophilic. The major

13

1

cellular fatty acids are 19:0 CYCLO ω8c and 18:1ω7c. The polar lipids consist of

2

phosphatidylcholine, phosphatidylethanolamine, phosphatidylglycerol, cardiolipin and

3

sphingolipids. Quinones are represented by ubiquinone Q10. Belong to the family

4

Roseiarcaceae, class Alphaproteobacteria. Found in peatlands and soils. The type species is

5

Roseiarcus fermentans.

6 7

Description of Roseiarcus fermentans sp. nov.

8

Roseiarcus fermentans (fer.men’tans. L. part. adj. fermentans fermenting).

9

The description is as for the genus but with the following additional traits. Cells are 2.7-4.0

10

µm in lenght and 0.6-1.2 µm in wight. Liquid cultures are slightly pinkish or colorless when

11

grown in the light and in the dark, respectively. Absorption maxima of living cells grown in the

12

light are at 455, 489, 528, 593, 806 and 865 nm. Bacteriochlorophyll a is produced only in the

13

light. Spirilloxanthin, rhodopin and 3,4- didehydrorhodopin are the major carotenoids. Preferred

14

mode of growth is fermentation of some sugars and organic acids under micro-oxic conditions.

15

Carbon sources utilized include glucose, fructose, xylose, trehalose, malate, succinate, pyruvate,

16

galacturonic acid, and Phytagel. Methane, methanol and formate are not utilized. Nitrogen

17

sources utilized include ammonium salts, histidine, glutamate, yeast extract and N2. Oxidase-

18

negative and catalase-positive. Grows in the pH range 4.0-7.0 (optimum 5.5-6.5) and at 15- 30 C

19

(optimum 22–28 °C). NaCl inhibits growth at concentrations above 0.5% (w/v). The DNA G+C

20

content is 70.0 mol%. The type strain, Pf56T (=DSM 24875T = VKM B-2876T), was isolated

21

from the acidic Sphagnum peat bog Staroselsky moss, Tver region, Russia.

o

22 23

Description of Roseiarcaceae fam. nov.

14

1

Roseiarcaceae (Ro.se.i.ar.ca.ce’ae. N.L. masc. n. Roseiarcus type genus of the family; -aceae

2

ending to denote a family; N.L. fem. pl. n. Roseiarcaceae the Roseiarcus family).

3

Gram-negative, non-spore-forming bacteria. Aerobes capable of fermentation under micro-oxic

4

conditions. Mesophilic and mildly acidophilic. The family Roseiarcaceae belongs to the class

5

Alphaproteobacteria, order Rhizobiales. The type genus is Roseiarcus.

6 7 8

ACKNOWLEDGMENTS

9

This research was supported by the Program “Molecular and Cell Biology” and the Russian Fund

10

of Basic Research (project No 12-04-00768). The authors thank Vladimir M. Gorlenko for

11

helpful discussions and valuable advices, Natalia E. Suzina for electron microscopy analysis and

12

E.N. Detkova for DNA G+C content analysis.

15

1

REFERENCES

2

Auman, A.J., Stolyar, S., Costello, A.M. & Lidstrom, M.E. (2000). Molecular characterization

3

of methanotrophic isolates from freshwater lake sediment. Appl Environ Microbiol 66, 5259-

4

5266.

5

Belova, S.E., Kulichevskaya, I.S., Bodelier, P.L.E. & Dedysh, S.N. (2013). Methylocystis

6

bryophila sp. nov., a novel, facultatively methanotrophic bacterium from acidic Sphagnum peat,

7

and emended description of the genus Methylocystis (ex Whittenbury et al. 1970) Bowman et al

8

1993. Int J Syst Evol Microbiol 63, 1096-1104.

9

Benning, C., Huang, Z.H., & Gage, D.A. (1995). Accumulation of a novel glycolipid and a

10

betaine lipid in cells of Rhodobacter sphaeroides grown under phosphate limitation. Arch

11

Biochem Biophys 317, 103–111.

12

Bodelier, P.L.E., Gillisen, M.-J.B., Hordijk, K., Sinninghe Damsté, J.S., Rijpstra, W.I.C.,

13

Geenevasen, J.A.J. & Dunfield, P.F. (2009). A reanalysis of phospholipid fatty acid as

14

ecological biomarkers for methanotrophic bacteria. The ISME Journal 3, 606-617.

15

Bowman, J. (2000). The methanotrophs – the families Methylococcaceae and Methylocystaceae.

16

In The prokaryotes: en evolving electronic resource for the microbiological community, 3rd ed.,

17

release 3.1. Dworkin, M. et al. (eds.) [Online] http://link.springer-

18

ny.com/link/service/books/10125/.

19

Collins, M.D. (1985). Analysis of isoprenoid quinones. Methods Microbiol, 18, 329-366.

20

Costello, A. M. & Lidstrom, M. E. (1999). Molecular characterization of functional and

21

phylogenetic genes from natural populations of methanotrophs in lake sediments. Appl Environ

22

Microbiol 65, 5066-5074.

23

Dedysh, S. N., Liesack, W., Khmelenina, V. N., Suzina, N. E., Trotsenko, Y. A., Semrau, J.

24

D., Bares, A. M., Panikov, N. S. & Tiedje, J. M. (2000). Methylocella palustris gen. nov., sp.

25

nov., a new methane-oxidizing acidophilic bacterium from peat bogs, representing a novel

26

subtype of serine-pathway methanotrophs. Int J Syst Evol Microbiol 50, 955-969.

27

Dedysh, S. N., Berestovskaya, Y. Y., Vasylieva, L. V., Belova, S. E., Khmelenina, V. N.,

28

Suzina, N. E., Trotsenko, Y. A., Liesack, W. & Zavarzin, G. A. (2004a). Methylocella

16

1

tundrae sp. nov., a novel methanotrophic bacterium from acidic tundra peatlands. Int J Syst Evol

2

Microbiol 54, 151-156.

3

Dedysh, S.N., Ricke, P. & Liesack, W. (2004b). NifH and NifD phylogenies: a molecular basis

4

for understanding nitrogen fixation capabilities of methanotrophic bacteria. Microbiology 150,

5

1301-1313.

6

Eller, G., Stubner, S. & Frenzel, P. (2001). Group-specific 16S rRNA targeted probes for the

7

detection of type I and type II methanotrophs by fluorescence in situ hybridisation. FEMS

8

Microbiol Lett 198, 91-97.

9

Felsenstein, J. (1989). PHYLIP – phylogeny inference package (version 3.2). Cladistics 5, 164-

10

166.

11

Gerhardt, P. (1981). Manual of Methods for General Bacteriology. Washington, DC: American

12

Society for Microbiology.

13

Holmes, A. J., Costello, A., Lidstrom, M. E. & Murrell, J. C. (1995). Evidence that

14

particulate methane monooxygenase and ammonium monooxygenase may be evolutionarily

15

related. FEMS Microbiol Lett 132, 203-208.

16

Imhoff, J. F. (1995). Taxonomy and physiology of phototrophic purple bacteria and green sulfur

17

bacteria, p. 1–15. In R. E. Blankenship, M. T. Madigan, and C. E. Bauer (ed.), Anoxygenic

18

photosynthetic bacteria. Kluwer Academic Publishers, Dordrecht, The Netherlands.

19

Imhoff, J. F. (2001). The phototrophic alpha-Proteobacteria. In The Prokaryotes: an Evolving

20

Electronic Resource for the Microbiological Community, 3rd edn, release 3.6, 22 June 2001.

21

Edited by M. Dworkin & others. New York: Springer. http://141.150.157.117:8080/

22

prokPUB/index.htm.

23

Imhoff, J. F. (2006). The Phototrophic Alpha-Proteobacteria. In The Prokaryotes: a Handbook

24

on the Biology of Bacteria, 3rd edn, vol. 5, рp. 41-64. Edited by M. Dworkin, S. Falkow, E.

25

Rosenberg, K. H. Schleifer & E. Stackebrandt. New York: Springer.

26

Imhoff, J. F., Kushner, D.J., Kushwaha, S.C. & Kates, M. (1982). Polar lipids in phototrophic

27

bacteria of the Rhodospirillaceae and Chromatiaceae families. J Bacteriol 150, 1192-1201.

17

1

Kämpfer, P. & Kroppenstedt, R.M. (1996). Numerical analysis of fatty acid patterns of the

2

coryneform bacteria and related taxa. Can J Microbiol 42, 989-1005.

3

Kates, M. (1972). Techniques of Lipidology. New York: Elsevier.

4

Kulichevskaya, I.S., Guzev, V.S., Gorlenko, V.M., Liesack, W. & Dedysh, S.N. (2006).

5

Rhodoblastus sphagnicola sp. nov., a novel acidophilic purple non-sulfur bacterium from

6

Sphagnum peat bog. Int J Syst Evol Microbiol 56, 1397-1402.

7

Ludwig, W., Strunk, O., Westram, R., Richter, L., Meier, H., Yadhukumar, Buchner, A.,

8

Lai, T., Steppi, S., Jobb, G. & other authors (2004). ARB: a software environment for

9

sequence data. Nucleic Acids Res 32, 1363-1371.

10

Madigan, M.T., Cox, J.C., & Gest, H. (1980). Physiology of dark fermentative growth of

11

Rhodopseudomonas capsulata. J Bacteriol 142, 908-915.

12

McDonald, I.R., Kenna, E.M. & Murrell, J.C. (1995). Detection of methanotrophic bacteria in

13

environmental samples with the PCR. Appl Environ Microbiol 61, 116-121.

14

Miguez, C.B., Bourque, D., Sealy, J.A., Greer, C.W. & Groleau, D. (1997). Detection and

15

isolation of methanotrophic bacteria possessing soluble methane monooxygenase (sMMO) genes

16

using the polymerase chain reaction (PCR). Microb Ecol 33, 21-31.

17

Nichols, B.W. (1963). Separation of the lipids of photosynthetic tissues; improvement in

18

analysis by thin-layer chromatography, Biochim. Biophys. Acta 4145, 417–422.

19

Owen, R.J., Lapage, S.P. & Hill, L.R. (1969). Determination of base composition from melting

20

profiles in dilute buffers. Biopolymers 7, 503-516.

21

Schultz, J.E. & Weaver, P.F. (1982). Fermentation and anaerobic respiration by

22

Rhodospirillum rubrum and Rhodopseudomonas capsulata. J Bacteriol 149, 181-190.

23

Shigematsu, T., Hanada, S., Eguchi, M., Kamagata, Y., Kanagawa, T. & Kurane, R. (1999).

24

Soluble methane monooxygenase gene clusters fron trichloroethylene-degrading Methylomonas

25

sp. strains and detection of methanotrophs during in situ bioremediation. Appl Environ Microbiol

26

65, 5198-5206.

27

Swingley, W.D., Blankenship, R.E. & Raymond, J. (2009). Evolutionary relationships among

28

purple photosynthetic bacteria and the origin of proteobacterial photosynthetic systems. Pp. 1718

1

29 In: The Purple Phototrophic Bacteria. Eds. Hunter, C.N., Thurnauer, M.C. & Beatty, J.T.

2

Springer Science + Business Media B.V.

3

Tamas, I., Smirnova, A.V., He, Z. & Dunfield, P.F. (2014). The (d)evolution of

4

methanotrophy in the Beijerinckiaceae – a comparative genomic analysis. The ISME Journal 8,

5

369-382.

6

Uffen, R.L., & Wolfe, R.S. (1970). Anaerobic growth of purple nonsulfur bacteria under dark

7

conditions. J Bacteriol 104, 462-472.

8

Vorobev, A.V., Baani, M., Doronina, N.V., Brady, A.L., Liesack, W., Dunfield, P.F. &

9

Dedysh, S.N. (2011). Methyloferula stellata gen. nov., sp. nov., an acidophilic, obligately

10

methanotrophic bacterium possessing only a soluble methane monooxygenase. Int J Syst Evol

11

Microbiol 61, 2456-2463.

12

Weisburg, W.G., Barns, S.M., Pelletier, D.A. & Lane, D.J. (1991). 16S ribosomal DNA

13

amplification for phylogenetic study. J Bacteriol 173, 697-703.

14

Yurkov, V.V. & Beatty, J.T. (1998). Aerobic anoxygenic phototrophic bacteria. Microbiol

15

Molec Biol Rev 62, 695-724.

16

Yurkov, V.V. (2001). Aerobic phototrophic proteobacteria. In The Prokaryotes: an Evolving

17

Electronic Resource for the Microbiological Community, 3rd edn, release 3.6, 22 June 2001.

18

Edited by M. Dworkin & others. New York: Springer. http://141.150.157.117:8080/

19

prokPUB/index.htm.

19

1

Table 1. Cellular fatty acid composition of strain Pf56T. Values are percentages of total fatty acids.

2

Major fatty acids are shown in bold.

Fatty acid

Strain Pf56T

12:0

0.3

13:0

0.4

14:0

0.5

15:0

0.7

16:1ω7c

0.3

16:0

9.3

17:1ω8c

1.4

17:0

6.5

18:1ω9c

0.8

18:1ω7c

26.3

18:1ω5c

0.6

11 methyl 18:1ω7c

0.9

17:0 3OH

0.6

19:0 CYCLO ω8c

38.9

18:0 3OH

2.5

20:2ω6,9c

0.5

20

1

FIGURE CAPTIONS

2

Figure 1. (A) Phase-contrast micrograph of cells of stain Pf56T, bar, 10 µm. (B) Electron

3

micrograph of an ultrathin section showing vesicular intracytoplasmic membranes, granules of

4

polyhydroxybutyrate and polyphosphate, bar, 1 µm.

5

Figure 2. Absorption spectrum of the acetone/ methanol extract of cells of phototrophically

6

grown strain Pf56T showing the characteristic peaks of BChl a at 363 and 770 nm.

7

Figure 3. Influence of medium pH on the growth of strain Pf56T.

8

Figure 4. 16S rRNA gene-based neighbor-joining tree showing the phylogenetic position of

9

strain Pf56T in relation to taxonomically characterized members of the families

10

Methylocystaceae and Beijerinckiaceae, as well as to some environmental clone sequences

11

retrieved in cultivation-independent studies from various wetlands and soils. Bootstrap values

12

(percentages of 1000 data resamplings) >50% are shown. Black circles indicate that the

13

corresponding nodes were also recovered in the maximum-likelihood and maximum-parsimony

14

trees. Four members of the Gammaproteobacteria, Methylomicrobium album (X72777),

15

Methylobacter luteus (AF304195), Methylomonas methanica S1 (AF304196) and Methylococcus

16

capsulatus Texas (AJ563935) were used as an outgroup. Bar, 0.05 substitutions per nucleotide

17

position.

18

Suppl. Figure S1. Phase-contrast micrographs of cells of stain Pf56T grown under fully oxic (a)

19

and micro-oxic conditions (b); bar, 10 µm.

20

Suppl. Figure S2. (A) Growth of strain Pf56 in anoxic conditions in the dark with (1) and

21

without (2) fructose. (B) Dynamics of fructose (1), propionate (2) and acetate (3) during

22

anaerobic fermentation of fructose by strain Pf56 T. 21

1

Suppl. Figure S3. Polar lipids of strain Pf56 T after separation by two dimensional TLC.

2

PC, phosphatidylcholine; PE, phosphatidylethanolamine; PG, phosphatidylglycerol;

3

CL, cardiolipin ; SL1 and SL2, sphingolipids; PL, unidentified phospholipid.

4

5 6

7

22

1 2 3

23

1

24

Descriptions of Roseiarcus fermentans gen. nov., sp. nov., a bacteriochlorophyll a-containing fermentative bacterium related phylogenetically to alphaproteobacterial methanotrophs, and of the family Roseiarcaceae fam. nov.

A light-pink-pigmented, microaerophilic bacterium was obtained from a methanotrophic consortium enriched from acidic Sphagnum peat and designated stra...
1MB Sizes 0 Downloads 3 Views

Recommend Documents