Next Article in Journal
Myo-D-inositol Trisphosphate Signalling in Oomycetes
Previous Article in Journal
Virulence Potential and Antibiotic Susceptibility of S. aureus Strains Isolated from Food Handlers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genomic Analysis and Characterization of Pseudotabrizicola formosa sp. nov., a Novel Aerobic Anoxygenic Phototrophic Bacterium, Isolated from Sayram Lake Water

1
Marine College, Shandong University, Weihai 264209, China
2
The Walter and Eliza Hall Institute of Medical Research, Parkville, VIC 3052, Australia
3
Infection Program and Department of Microbiology, Monash Biomedicine Discovery Institute, Monash University, Clayton, VIC 3800, Australia
4
Weihai Research Institute of Industrial Technology of Shandong University, Weihai 264209, China
*
Authors to whom correspondence should be addressed.
Microorganisms 2022, 10(11), 2154; https://doi.org/10.3390/microorganisms10112154
Submission received: 12 October 2022 / Revised: 27 October 2022 / Accepted: 28 October 2022 / Published: 30 October 2022
(This article belongs to the Section Environmental Microbiology)

Abstract

:
Aerobic anoxygenic photosynthetic bacteria (AAPB) are a kind of heterotrophic prokaryote that can use bacteriochlorophyll (BChl) for photosynthesis without oxygen production and they are widely distributed in aquatic environments, including oceans, lakes, and rivers. A novel aerobic anoxygenic photosynthetic bacterium strain XJSPT was isolated during a study of water microbial diversity in Sayram Lake, Xinjiang Province, China. Strain XJSPT was found to grow optimally at 33 °C, pH 7.5 with 1.0% (w/v) NaCl, and to produce bacteriochlorophyll a and carotenoids. Phylogenetic analysis based on 16S rRNA gene sequence and concatenated alignment sequences of 120 ubiquitous single-copy proteins both supported that strain XJSPT belonged to the genus Pseudotabrizicola. Both average nucleotide identity (ANI) and DNA–DNA hybridization (DDH) values were below the species delineation threshold. The primary polar lipids were phosphatidylcholine, phosphatidylglycerol, phosphatidylethanolamine, one unknown lipid, and one unidentified phospholipid. Based on the results of polyphasic analyses performed in this study, strain XJSPT represents a new member of the genus Pseudotabrizicola, for which the name Pseudotabrizicola formosa sp. nov. is proposed. The type strain is XJSPT (= KCTC 52636T = MCCC 1H00184T = SDUM 107003T). Comparative genomic analysis showed that four species of the genus Pseudotabrizicola shared 2570 core genes and possessed a complete anoxygenic photosystem II.

1. Introduction

Aerobic anaerobic photosynthetic bacteria (AAPB) are widely distributed in aquatic environments (oceans, lakes, and rivers). As a heterotrophic group, they can use reduced organic matter as electron donor under aerobic conditions to carry out non-oxygen-producing photosynthesis, relying on unique bacteriochlorophyll (BChl) and light reaction center [1,2]. The ATP produced by photosynthesis replenishes the energy required for growth, not only reducing the consumption of organic carbon, but also increasing the amount of dissolved organic carbon (DOC) entering the cells, which is vital for biogeochemical cycles [3,4,5].
The genus Tabrizicola, belonging to the family Rhodobacteraceae in the class Alphaproteobacteria, was firstly proposed by Vahideh et al. in 2013 and consists of eight validly published species and three effectively described species at the time of writing [6]. In 2022, Ma et al. reclassified Tabrizicola sediminis, Tabrizicola alkalilacus, and Tabrizicola algicola into a novel genus, Pseudotabrizicola gen. nov., as Pseudotabrizicola sediminis comb. nov., Pseudotabrizicola alkalilacus comb. nov., and Pseudotabrizicola algicola comb. nov. according to the results of polyphasic investigations [7]. Existing members of the genus Pseudotabrizicola are Gram-strain-negative, catalase- and oxidase-positive, and have Q-10 as the main respiratory quinone. Moreover, P. sediminis KCTC 72015T and P. algicola KCTC 72206T were reported to belong to AAPB, a kind of heterotrophic bacteria which have a photosynthetic gene and can produce BChl a in aerobic condition but cannot grow photoautotrophically under anaerobic conditions [8,9,10].
During our research of bacterial diversity at Sayram Lake, a cream-colored bacterium designated XJSPT was isolated from a lake water sample using a dilution-plating procedure and conventional isolation techniques. Polyphasic taxonomic investigations, including phenotypic characterizations, chemotaxonomic properties, and phylogenetic analysis, showed that strain XJSPT was a novel aerobic anoxygenic phototrophic bacterium species affiliated to the genus Pseudotabrizicola.

2. Materials and Methods

2.1. Bacterial Isolation and Culture

Samples from various habitats were gathered for bacterial enrichment and isolation as a part of the study about bacterial resource diversity in our lab [11]. A water sample, collected from Sayram Lake, Xinjiang Province, China (44° 30′ 30.41″ N, 81° 12′ 39.55″ E), was diluted stepwise using sterile distilled water and each diluted sample was spread evenly on marine agar 2216 (MA; Becton Dickinson, Franklin Lakes, NJ, USA). The strain XJSPT was isolated from the coated medium, which was incubated at 25 °C for 10 days. Pure cultures were preserved for long-term in sterile 15% (v/v) glycerol supplemented with 1% (w/v) NaCl at –80 °C. The type of strain P. sediminis KCTC 72015T was purchased as an experiment control strain from Korean Collection for Type Cultures center (KCTC).

2.2. S rRNA Gene Sequencing and Phylogenetic Analysis

The 16S rRNA genes of strain XJSPT were amplified using polymerase chain reaction (PCR) technology with two universal primers for bacteria (27F and 1492R) and a purified gene product was cloned using the method described previously to obtain almost complete 16S rRNA gene sequence [12]. The 16S rRNA gene similarities between strain XJSPT and closely related species were calculated using the NCBI BLAST service and EzBioCloud database. The 16S rRNA gene sequence of strain XJSPT and those of relevant strains were aligned by MUSCLE service [13] and phylogenetic trees were reconstructed with 1000 bootstrap replicates based on neighbor-joining (NJ), minimum-evolution (ME), and maximum-likelihood (ML) algorithms in MEGA X software [14,15]. The integrated method T92 + G + I was calculated as the best-fit substitution pattern for reconstructing the ML tree.

2.3. Whole-Genome Sequencing and Genome Annotation

Purified genomic DNA was obtained employing the SteadyPure bacterial genomic DNA extraction kit (Accurate Biotechnology Co., Ltd., Hunan Province, China) following the user guide. The draft genome of strain XJSPT was sequenced by Novogen (Tianjin, China) using Illumina Hiseq platform with the sequencing protocol of paired-end 150 bp fragment libraries and genome assembly was carried out with the Velvet software (v. 1.2.10) [16]. The genome sequences of related strains used in this paper were downloaded from the NCBI genomes repository. Gene prediction and annotation were carried out by Prodigal server [17] and the prokaryotic genome annotation pipeline (PGAP) implemented in NCBI [18].

2.4. Phylogenomic and Comparative Genomic Analysis

Genome similarity indexes, average nucleotide identity (ANI), and DNA–DNA hybridization (DDH) were calculated employing the JSpeciesWS online service offered by Ribocon (https://jspecies.ribohost.com/jspeciesws/ (accessed on 13 July 2022)) [19] and genomes comparison calculator v. 3.0 (http://ggdc.dsmz.de/ggdc.php) [20], respectively. The IQ-TREE based on concatenated alignment sequences of 120 ubiquitous single-copy proteins was reconstructed by GTDB-Tk v. 1.3.0 with the LG+F+I+G4 pattern and 1000 bootstrap replicates [21,22,23]. To further investigate gene and protein differences among members of the genus Pseudotabrizicola, comparative genomic analysis was achieved by the ultra-fast bacterial pan-genome analysis tool (BPGA) with default parameters [24]. The analysis of metabolic pathways and the search of putative secondary metabolite biosynthetic gene clusters were accomplished by BlastKOALA service (v. 2.2) in Kyoto Encyclopedia of Genes and Genomes (KEGG, https://www.kegg.jp/blastkoala/ (accessed on 26 October 2022)) [25] and antiSMASH 6.0 (https://antismash.secondarymetabolites.org/ (accessed on 31 August 2022)) [26]. Photosynthetic genes pufML encoding for M and L subunit of core photosynthetic reaction center were detected in the genome based on the primer pufL-67F (5′-TTC GAC TTY TGG RTN GGNCC-3′) and pufM-781R (5′-CCA KSG TCC AGC GCC AGAANA-3′) using the software SnapGene v. 4.1.9 [27].

2.5. Phenotypic Characteristics

Strain XJSPT was incubated on MA medium at 33 °C for the implementation of phenotypic characteristics investigations. After culturing for four days, the Gram staining reaction was checked with the Gram-stain kit produced by bioMérieux company, and cell morphology was observed by employing light microscopy (E600, Nikon, Tokyo, Japan) and scanning electron microscopy (model Nova NanoSEM450, FEI). The growth of strains at various pH ranges (pH 5.5–9.5, at intervals of 0.5) was tested in marine broth 2216 (MB; Becton Dickinson, Franklin Lakes, NJ, USA) with various pH values, and growth status was quantified using a microplate reader at 600 nm. The pH of mediums was adjusted using commercial additional buffers at a concentration of 20 mM: MES (pH 5.5 and 6.0), PIPES (pH 6.5 and 7.0), HEPES (pH 7.5 and 8.0), Tricine (pH 8.5), and CAPSO (pH 9.0 and 9.5). Temperature conditions for growth were tested at 0, 4, 10, 15, 20, 25, 28, 30, 33, 37, 40, 42, and 45 °C for approximately 7 days on MA medium (growth was recorded every 12 h). Salt tolerance was assayed using modified MA (prepared according to the MA formula, but without NaCl) with different NaCl concentrations (0, 0.5, 1, 2, 3, 4, 5, 6, 7, 8, 9, and 10%, w/v).
Oxidase test was examined by employing the commercial bioMérieux oxidase test kit, and catalase activity was detected through bubbles production after adding 3% (v/v) H2O2 to plate with fresh cultures. Hydrolysis tests of starch (0.2%, w/v), CM-cellulose (0.5%, w/v), alginate (2%, w/v), and Tweens (20, 40, 60, and 80, 1%, v/v) were determined based on the previous methods [28]. The commercial bioMérieux API 50CH and API ZYM reagent strips were used to test acid production and enzyme activities, respectively. Other biochemical analyses were performed applying the BIOLOG GEN III MicroPlates and API 50CH and all reagent strip tests were implemented following the user guide, except for adjusting the NaCl concentration to the optimum. The antimicrobial susceptibility test was investigated using the disc diffusion method under optimum conditions for a week [29].
Photoheterotrophic growth was tested under light exposure (2400 lx) and anaerobic conditions in the following liquid medium (per liter: 3 g sodium pyruvate, 1.2 g NH4Cl or 1g KNO3), prepared with modified artificial seawater (per liter of distilled water: 3.3 g MgSO4, 2.3 g MgCl2, 1.2 g CaCl2, 0.7 g KCl, 10 g NaCl) at 33 °C for 14 days. Photoautotrophic growth was determined by anaerobically incubating strain XJSPT under light condition (2400 lx) with the following liquid medium (0.5 mM Na2S, 0.5 mM Na2S2O3 and 0.1% (w/v) NaHCO3), prepared with modified artificial seawater as described above [9]. Anaerobic conditions were achieved by boiling the liquid medium and adding sterilized liquid paraffin. Additionally, the presence of pigments was detected by in vitro spectrometric methods described by Biebl et al. [30]. Cells cultivated aerobically in MB medium for four days were collected, washed twice, and suspended in a mixture of acetone and methanol (7:2, v/v) to extract the pigments. Absorption spectra were measured using a spectrophotometer.

2.6. Chemotaxonomic Properties

Comparative analyses of chemotaxonomic property between Strain XJSPT and experiment control strain P. sediminis KCTC 72015T were performed using cells harvested in MB medium at the late stage of exponential growth phase. Lipids were obtained in the mixture system of chloroform, methanol, and water (2.5:5:2, v/v/v), and separated and identified by two-dimensional silica gel thin layer chromatography (TLC) plate [31,32]. Extracted fatty acids were separated and analyzed based on the TSBA40 database of the Sherlock Microbial Identification System (MIDI) by an Agilent gas chromatograph (product model 6890N), as used previously [33]. Respiratory quinones obtained from lyophilized thallus were separated by TLC plates and identified applying HPLC technology [34].

3. Results and Discussion

3.1. S rRNA Gene Sequence and Phylogenetic Analysis

Almost complete 16S rRNA gene sequence of strain XJSPT (1425 bp) was obtained in this study. The 16S rRNA gene sequence similarity values between Strain XJSPT and members of the genus Pseudotabrizicola showed 97.7–99.5% (Table 1). The NJ tree inferred from 16S rRNA gene sequence exhibited strain XJSPT located in the cluster of Pseudotabrizicola species, which supported strain XJSPT belonged to the genus Pseudotabrizicola (Figure 1). The topology of strain XJSPT and the genus Pseudotabrizicola was also obtained in the phylogenetic trees reconstructed with the ML and ME algorithm (Figure 1).

3.2. Genome Properties and Phylogenetic Analysis

The draft genome (strain XJSPT) of 3,702,758 bp in length was obtained after assembly with an average 300× coverage depth, producing 14 contigs, and the N50 value is 812,613 bp. All contigs were larger than 1595 bp, with the largest being 1,520,968 bp. The calculated G+C content was estimated to be 63.4 mol%. The 16S rRNA gene sequence of strain XJSPT detected from genome (1467 bp) covered that obtained by amplification (1425 bp). The PGAP results showed that a total of 3552 genes were predicted, including 52 RNA genes (3 rRNA genes, 3 ncRNA genes, and 46 tRNA genes) and 3470 potential protein-coding genes. Detailed comparison results of genome statistics of the Pseudotabrizicola are shown in Table 2.
The genome similarity indices ANI and DDH values between strain XJSPT and Pseudotabrizicola species were 81.4–87.3% and 23.9–32.5% respectively, with both being lower than the values for species demarcation [35,36] (Table 1), which indicated that strain XJSPT was a novel member belonging to the genus Pseudotabrizicola. The IQ-TREE built on concatenated alignment sequences of 120 ubiquitous single-copy proteins in bacteria showed the evolutionary relationships of strain XJSPT and the genus Pseudotabrizicola (Figure 2).

3.3. Pan-Genome Analysis of the Genus Pseudotabrizicola

Comparative genomic analysis of the genus Pseudotabrizicola was carried out to identify the consistency and difference of the members. As shown in Figure 3, 2570 core genes were shared by the four Pseudotabrizicola species, strain XJSPT, P. sediminis KCTC 72015T, P. alkalilacus KCTC 62173T, and P. algicola KCTC 72206T, which accounted for more than half (59.8–74.2%) of each genome. KEGG annotation was performed for core, accessory, and unique genes to analyze their distribution in different metabolic pathways. The results showed that the core genes were more involved in the metabolisms of amino acid, energy and nucleotide, translation and replication, and repair. The proportion of accessory genes was higher than that of core genes and unique genes in carbohydrate metabolism, cofactors and vitamins metabolism and xenobiotics biodegradation. However, unique genes contributed more to drug resistance, lipid metabolism, membrane transport, and signal transduction (Supplementary Figure S1).

3.4. Metabolic Pathways and Secondary Metabolites Analyses

The results of metabolic pathways analyzed by KEGG’s BlastKOALA service showed that most of carbohydrate metabolism pathways were intact, except for the incomplete glycolysis pathway (M00001) in strain XJSPT. The four species of the genus Pseudotabrizicola, strain XJSPT, P. sediminis KCTC 72015T, P. alkalilacus KCTC 62173T, and P. algicola KCTC 72206T, all possessed a complete anoxygenic photosystem II (M00597), namely the L and M subunits of photosynthetic reaction center. The pufML gene sequences of strain XJSPT are given in Supplementary Table S1. Moreover, strain XJSPT had complete phosphatidylcholine (PC) and phosphatidylethanolamine (PE) biosynthesis pathway (M00091 and M00093, respectively) and isoprenoid biosynthesis pathway (M00096 and M00364), which was consistent with P. sediminis KCTC 72015T, P. alkalilacus KCTC 62173T, and P. algicola KCTC 72206T (Figure 4). The potential secondary metabolites synthesized by strain XJSPT were identified using antiSMASH. The results showed that the genome of strain XJSPT encoded eight identified gene clusters about the biosynthesis of secondary metabolites (Supplementary Table S2). One of the eight gene clusters, for terpene, showed 100% similarity to a known biosynthetic gene cluster-encoding carotenoid [37].

3.5. Phenotypic Characteristics

Colonies of strain XJSPT were cream-colored, smooth, and circular after incubating for 3 days at 33 °C on MA medium, and the color of colonies would change to light opaque-pink after a week under a low light condition (10 µmol photons m−2 s−2). Cells of strain XJSPT were Gram-stain-negative, and rod-shaped with widths of 0.3–0.5 μm and lengths of 0.8–2.0 μm (Supplementary Figure S2). Strain XJSPT was unable to undergo autotrophic and heterotrophic growth under light and anaerobic conditions. The activities of esterase (C4), esterase lipase (C8) and leucine arylamidase were positive but ɑ-galactosidase activity was negative, which was consistent with P. sediminis KCTC 72015T and P. alkalilacus KCTC 62173T [38]. However, there were several characteristic differences between strain XJSPT and related species summarized in Table 3, which could distinguish strain XJSPT from related species. Spectral analysis showed that the typical maxima absorptions were at 486, 867, and 895 nm, which indicated the presence of BChl a and carotenoids (Supplementary Figure S3). Strain XJSPT was found to be sensitive to (μg per disc) chloramphenicol (30), rifampicin (5), cefotaxime sodium (30), ceftriaxone (30), acetylspiramycin (30), clarithromycin (15), tobramycin (10), ampicillin (10), norfloxacin (30), and neomycin (30), but resistant to penicillin (10), erythromycin (15), tetracycline (30), vancomycin (30), lincomycin (2), gentamycin (10), nalidixic acid (30), streptomycin (10), and kanamycin (30).

3.6. Chemotaxonomic Properties

The isoprenoid quinone detected in strain XJSPT was Q-10, which was in line with the genus Pseudotabrizicola. The major cellular fatty acids (>10%) were iso-C18:0 and summed feature 8 (comprising C18:1 ω6c and/or C18:1 ω7c) (Supplementary Table S3). The major polar lipids of strain XJSPT were phosphatidylcholine (PC), phophatydilethanolamine (PE), phosphatidylglycerol (PG), one unidentified phospholipid (PL), and one unknown lipid (L). The polar lipids composition of strain XJSPT was similar to that of P. sediminis KCTC 72015T in phosphatidylcholine (PC), phosphatidylglycerol (PG), and phophatydilethanolamine (PE), but the absence of diphosphatidylglycerol (DPG) in strain XJSPT distinguished it from the closest strain P. sediminis KCTC 72015T (Supplementary Figure S4).
Description of Pseudotabrizicola formosa sp. nov.
Pseudotabrizicola formosa (for.mo’sa. L. fem. adj. formosa beautiful, beautifully formed, finely formed).
Cells are Gram-stain-negative and rod-shaped (0.3–0.5 µm wide and 0.8–2.0 µm long). Colonies appear cream-colored or light pink, circular with entire edges, and convex with a diameter of 1.0–1.5 mm. The cell suspension is a light opaque-pink in color. Cells growth occurs at 4–40 °C (optimum 33 °C), with 0–6.0% (w/v) NaCl (optimum 1.0% NaCl) and pH 6.5–9.5 (optimum pH 7.5). Phototrophic growth occurs under aerobic, heterotrophic conditions, and photosynthetic pigments are produced in low light (10 µmol photons m−2 s−2). The activities of catalase, oxidase, and valine arylamidase are positive but the activities of lipase (C14), N-acetyl-β-glucosaminidase, trypsin, α-mannosidase, α-chymotrypsin, acid phosphatase, and β-glucuronidase are negative. Hydrolyses of Tweens 20 and 40 are positive, but negative for hydrolyses of Tween 60, Tween 80, starch, CM-cellulose, alginate, and casein. Acids are produced from d-ribose, d-cellobiose, d-xylose, d-turanose, and d-galactose. In the oxidation test of sole carbon source, positive for d-maltose, d-trehalose, d-cellobiose, gentiobiose, sucrose, d-turanose, β-methyl-d-glucoside, d-salicin, d-mannose, d-fructose, l-fucose, myo-inositol, and l-malic acid. The major cellular fatty acids (>10%) are iso-C18:0 and summed feature 8 (comprising C18:1 ω6c and/or C18:1 ω7c). The main respiratory quinone is Q-10. The predominant polar lipids consist of phosphatidylcholine, phosphatidylglycerol, phophatydilethanolamine, one unidentified phospholipid, and one unknown lipid.
The type strain, XJSPT (= KCTC 52636T = MCCC 1H00184T = SDUM 107003T), was isolated from Sayram Lake water, Xinjiang Province, China. The DNA G+C content of type strain was 63.4 mol%.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/microorganisms10112154/s1, Figure S1: The distribution of core genes, accessory genes and unique genes to different metabolic pathways in the genus Pseudotabrizicola. Figure S2: Scanning electron micrograph of cells of strain XJSPT. Bar, 5 μm. Figure S3: Absorption spectrum of the acetone/methanol (7:2) extract of strain XJSPT. The typical maxima absorptions at 486, 867 and 895 nm, indicating the presence of bacteriochlorophyll α and carotenoids. Figure S4: Two-dimensional TLC plate image of the total polar lipids of strain XJSPT (a) and Pseudotabrizicola sediminis KCTC 72015T (b). PC, phosphatidylcholine; PG, phosphatidylglycerol; DPG, diphosphatidylglycerol; PE, phosphatidylethanolamine; APL, unidentified aminophospholipid; AL, unidentified aminolipid; PL, unidentified phospholipid; L, unidentified lipid. Table S1: pufML gene sequences of strain XJSPT. Table S2: Secondary metabolites of strain XJSPT predicted by antiSMASH. Table S3: Cellular fatty acid composition (%) of strain XJSPT and related species.

Author Contributions

Y.-Q.Y. performed experimental operation, data collection and analysis and finished the manuscript. J.-R.H. isolated the strain XJSPT and performed material preparation, experimental operation. J.-X.Z. helped perform comparative genome analysis. Z.-J.D. and M.-Q.Y. offered experiment guidance and critical revision of manuscripts. All authors contributed to the article and approved the submitted version. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Natural Science Foundation of China (32070002, 32200003), National Science and Technology Fundamental Resources Investigation Program of China (2019FY100700), and Shandong University (Weihai) interdisciplinary cultivation project “Research on resources & environment and high-quality development of marine economy”. J.-X.Z. is a recipient of the 2022 Faculty of Medicine, Nursing and Health Sciences Bridging Fellowship (BPF22-2451802925), Monash University.

Data Availability Statement

The 16S rRNA gene sequence of Pseudotabrizicola formosa XJSPT has been deposited at GenBank database with the accession number KY457223. The GenBank accession number for draft genome sequence of Pseudotabrizicola formosa XJSPT is PJOM00000000.

Acknowledgments

This work was supported by the Physical-Chemical Materials Analytical and Testing Center of Shandong University at Weihai.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AAPBAerobic anoxygenic phototrophic bacteria
ANIAverage nucleotide identity
DDHDNA–DNA hybridization
MCCCMarine Culture Collection of China
SDUMShandong University Collection of Marine Microorganisms
KCTCKorean Collection for Type Cultures
PGAPProkaryotic Genome Annotation Pipeline
BPGABacterial Pan-Genome Analysis Tool
MES2-Morpholinoethanesulfonic acid
PIPES1,4-Piperazinediethanesulfonic acid
HEPESN-(2-Hydroxyethyl) piperazine-N’-2-ethanesulfonic acid
CAPSO3-(Cyclohexylamino)-2-hydroxy-1-propanesulfonic acid
TricineN-[Tris(hydroxymethyl)methyl]glycine

References

  1. Yurkov, V.V.; Beatty, J.T. Aerobic Anoxygenic Phototrophic Bacteria. Microbiol. Mol. Biol. Rev. 1998, 62, 695–724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Jiao, N.; Zhang, Y.; Zeng, Y.; Hong, N.; Liu, R.; Chen, F.; Wang, P. Distinct distribution pattern of abundance and diversity of aerobic anoxygenic phototrophic bacteria in the global ocean. Environ. Microbiol. 2007, 9, 3091–3099. [Google Scholar] [CrossRef]
  3. Kolber, Z.S.; Van Dover, C.L.; Niederman, R.A.; Falkowski, P.G. Bacterial photosynthesis in surface waters of the open ocean. Nature 2000, 407, 177–179. [Google Scholar] [CrossRef] [PubMed]
  4. Sato-Takabe, Y.; Nakao, H.; Kataoka, T.; Yokokawa, T.; Hamasaki, K.; Ohta, K.; Suzuki, S. Abundance of Common Aerobic Anoxygenic Phototrophic Bacteria in a Coastal Aquaculture Area. Front. Microbiol. 2016, 7, 1996. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Lehours, A.-C.; Enault, F.; Boeuf, D.; Jeanthon, C. Biogeographic patterns of aerobic anoxygenic phototrophic bacteria reveal an ecological consistency of phylogenetic clades in different oceanic biomes. Sci. Rep. 2018, 8, 4105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Tarhriz, V.; Thiel, V.; Nematzadeh, G.; Hejazi, M.S.; Imhoff, J.F. Tabrizicola aquatica gen. nov. sp. nov., a novel alphaproteobacterium isolated from Qurugöl Lake nearby Tabriz city, Iran. Antonie Leeuwenhoek 2013, 104, 1205–1215. [Google Scholar] [CrossRef]
  7. Ma, T.; Xue, H.; Piao, C.; Liu, C.; Yang, M.; Bian, D.; Li, Y. Reclassification of 11 Members of the Family Rhodobacteraceae at Genus and Species Levels and Proposal of Pseudogemmobacter hezensis sp. nov. Front. Microbiol. 2022, 13, 849695. [Google Scholar] [CrossRef]
  8. Liu, Z.-X.; Dorji, P.; Liu, H.-C.; Li, A.-H.; Zhou, Y.-G. Tabrizicola sediminis sp. nov., one aerobic anoxygenic photoheterotrophic bacteria from sediment of saline lake. Int. J. Syst. Evol. Microbiol. 2019, 69, 2565–2570. [Google Scholar] [CrossRef]
  9. Tarhriz, V.; Hirose, S.; Fukushima, S.-I.; Hejazi, M.A.; Imhoff, J.F.; Thiel, V. Emended description of the genus Tabrizicola and the species Tabrizicola aquatica as aerobic anoxygenic phototrophic bacteria. Antonie Leeuwenhoek 2019, 112, 1169–1175. [Google Scholar] [CrossRef]
  10. Park, C.-Y.; Chun, S.-J.; Jin, C.; Van Le, V.; Cui, Y.; Kim, S.-Y.; Ahn, C.-Y.; Oh, H.-M. Tabrizicola algicola sp. nov. isolated from culture of microalga Ettlia sp. Int. J. Syst. Evol. Microbiol. 2020, 70, 6133–6141. [Google Scholar] [CrossRef]
  11. Mu, D.-S.; Liang, Q.-Y.; Wang, X.-M.; Lu, D.-C.; Shi, M.-J.; Chen, G.-J.; Du, Z.-J. Metatranscriptomic and comparative genomic insights into resuscitation mechanisms during enrichment culturing. Microbiome 2018, 6, 230. [Google Scholar] [CrossRef] [PubMed]
  12. Ye, Y.-Q.; Han, Z.-T.; Liu, X.-J.; Ye, M.-Q.; Du, Z.-J. Maribellus maritimus sp. nov., isolated from marine sediment. Arch. Microbiol. 2021, 204, 40. [Google Scholar] [CrossRef]
  13. Edgar, R.C. MUSCLE: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004, 32, 1792–1797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Felsenstein, J. Confidence limits on phylogenies: An approach using the bootstrap. Evolution 1985, 39, 783–791. [Google Scholar] [CrossRef] [PubMed]
  15. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular Evolutionary Genetics Analysis across Computing Platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef] [PubMed]
  16. Zerbino, D.R.; Birney, E. Velvet: Algorithms for de novo short read assembly using de Bruijn graphs. Genome Res. 2008, 18, 821–829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Hyatt, D.; Chen, G.-L.; Locascio, P.F.; Land, M.L.; Larimer, F.W.; Hauser, L.J. Prodigal: Prokaryotic gene recognition and translation initiation site identification. BMC Bioinform. 2010, 11, 119. [Google Scholar] [CrossRef] [Green Version]
  18. Zhao, Y.; Wu, J.; Yang, J.; Sun, S.; Xiao, J.; Yu, J. PGAP: Pan-genomes analysis pipeline. Bioinformatics 2012, 28, 416–418. [Google Scholar] [CrossRef] [Green Version]
  19. Richter, M.; Rosselló-Móra, R.; Oliver Glöckner, F.O.; Peplies, J. JSpeciesWS: A web server for prokaryotic species circumscription based on pairwise genome comparison. Bioinformatics 2016, 32, 929–931. [Google Scholar] [CrossRef]
  20. Meier-Kolthoff, J.P.; Carbasse, J.S.; Peinado-Olarte, R.L.; Göker, M. TYGS and LPSN: A database tandem for fast and reliable genome-based classification and nomenclature of prokaryotes. Nucleic Acids Res. 2022, 50, D801–D807. [Google Scholar] [CrossRef]
  21. Nguyen, L.-T.; Schmidt, H.A.; Von Haeseler, A.; Minh, B.Q. IQ-TREE: A Fast and Effective Stochastic Algorithm for Estimating Maximum-Likelihood Phylogenies. Mol. Biol. Evol. 2015, 32, 268–274. [Google Scholar] [CrossRef] [PubMed]
  22. Chernomor, O.; Von Haeseler, A.; Minh, B.Q. Terrace Aware Data Structure for Phylogenomic Inference from Supermatrices. Syst. Biol. 2016, 65, 997–1008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Chaumeil, P.-A.; Mussig, A.J.; Hugenholtz, P.; Parks, D.H. GTDB-Tk: A toolkit to classify genomes with the Genome Taxonomy Database. Bioinformatics 2019, 36, 1925–1927. [Google Scholar] [CrossRef] [PubMed]
  24. Chaudhari, N.M.; Gupta, V.; Dutta, C. BPGA- an ultra-fast pan-genome analysis pipeline. Sci. Rep. 2016, 6, 24373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Kanehisa, M.; Sato, Y.; Kawashima, M.; Furumichi, M.; Tanabe, M. KEGG as a reference resource for gene and protein annotation. Nucleic Acids Res. 2016, 44, D457–D462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Blin, K.; Shaw, S.; Kloosterman, A.M.; Charlop-Powers, Z.; van Wezel, G.P.; Medema, M.H.; Weber, T. antiSMASH 6.0: Improving cluster detection and comparison capabilities. Nucleic Acids Res. 2021, 49, W29–W35. [Google Scholar] [CrossRef] [PubMed]
  27. Tank, M.; Thiel, V.; Imhoff, J.F. Phylogenetic relationship of phototrophic purple sulfur bacteria according to pufL and pufM genes. Int. Microbiol. 2009, 12, 175–185. [Google Scholar]
  28. Dong, X.Z.; Cai, M.Y. Determination of biochemical characteristics. In Manual for the Systematic Identification of General Bacteria, 14th ed.; Dong, X.Z., Cai, M.Y., Eds.; Science Press: Beijing, China, 2001; pp. 370–398. [Google Scholar]
  29. Jorgensen, J.H.; Turnidge, J.D. Susceptibility Test Methods: Dilution and Disk Diffusion Methods. In Manual of Clinical Microbiology, 11th ed.; Jorgensen, J.H., Carroll, K.C., Funke, G., Pfaller, M.A., Landry, M.L., Richter, S.S., Warnock, D.W., Eds.; Wiley: New York, NY, USA, 2015; pp. 1253–1273. ISBN 9781683672807. [Google Scholar]
  30. Biebl, H.; Allgaier, M.; Tindall, B.J.; Koblizek, M.; Lünsdorf, H.; Pukall, R.; Wagner-Döbler, I. Dinoroseobacter shibae gen. nov., sp. nov., a new aerobic phototrophic bacterium isolated from dinoflagellates. Int. J. Syst. Evol. Microbiol. 2005, 55, 1089–1096. [Google Scholar] [CrossRef] [Green Version]
  31. Komagata, K.; Suzuki, K.-I. 4 Lipid and Cell-Wall Analysis in Bacterial Systematics. Methods Microbiol. 1987, 19, 161–207. [Google Scholar] [CrossRef]
  32. Liang, Q.-Y.; Xu, Z.-X.; Zhang, J.; Chen, G.-J.; Du, Z.-J. Salegentibacter sediminis sp. nov., a marine bacterium of the family Flavobacteriaceae isolated from coastal sediment. Int. J. Syst. Evol. Microbiol. 2018, 68, 2375–2380. [Google Scholar] [CrossRef]
  33. Ye, Y.-Q.; Hao, Z.-P.; Yue, Y.-Y.; Ma, L.; Ye, M.-Q.; Du, Z.-J. Characterization of Kordiimonas marina sp. nov. and Kordiimonas laminariae sp. nov. and Comparative Genomic Analysis of the Genus Kordiimonas, A Marine-Adapted Taxon. Front. Mar. Sci. 2022, 9, 985. [Google Scholar] [CrossRef]
  34. Minnikin, D.; O’Donnell, A.; Goodfellow, M.; Alderson, G.; Athalye, M.; Schaal, A.; Parlett, J. An integrated procedure for the extraction of bacterial isoprenoid quinones and polar lipids. J. Microbiol. Methods 1984, 2, 233–241. [Google Scholar] [CrossRef]
  35. Goris, J.; Konstantinidis, K.T.; Klappenbach, J.A.; Coenye, T.; Vandamme, P.; Tiedje, J.M. DNA–DNA hybridization values and their relationship to whole-genome sequence similarities. Int. J. Syst. Evol. Microbiol. 2007, 57, 81–91. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Konstantinidis, K.T.; Tiedje, J.M. Genomic insights that advance the species definition for prokaryotes. Proc. Natl. Acad. Sci. USA 2005, 102, 2567–2572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Lang, H.P.; Cogdell, R.J.; Gardiner, A.T.; Hunter, C.N. Early steps in carotenoid biosynthesis: Sequences and transcriptional analysis of the crtI and crtB genes of Rhodobacter sphaeroides and overexpression and reactivation of crtI in Escherichia coli and R. sphaeroides. J. Bacteriol. 1994, 176, 3859–3869. [Google Scholar] [CrossRef] [Green Version]
  38. Phurbu, D.; Wang, H.; Tang, Q.; Lu, H.; Zhu, H.; Jiang, S.; Xing, P.; Wu, Q.L. Tabrizicola alkalilacus sp. nov., isolated from alkaline Lake Dajiaco on the Tibetan Plateau. Int. J. Syst. Evol. Microbiol. 2019, 69, 3420–3425. [Google Scholar] [CrossRef]
Figure 1. Neighbor-joining phylogenetic tree based on 16S rRNA gene sequences of strain XJSPT and other closely related species. Filled circles indicate branches that were recovered with all three methods (neighbor-joining, maximum-likelihood, and minimum-evolution). Percentages bootstrap values above 50% (1000 replicates) are shown at branch nodes. Rhodospirillum rubrum ATCC 11170T was used as the out-group. Bar, 0.02 substitutions per nucleotide position.
Figure 1. Neighbor-joining phylogenetic tree based on 16S rRNA gene sequences of strain XJSPT and other closely related species. Filled circles indicate branches that were recovered with all three methods (neighbor-joining, maximum-likelihood, and minimum-evolution). Percentages bootstrap values above 50% (1000 replicates) are shown at branch nodes. Rhodospirillum rubrum ATCC 11170T was used as the out-group. Bar, 0.02 substitutions per nucleotide position.
Microorganisms 10 02154 g001
Figure 2. The IQ-TREE based on 120 ubiquitous single-copy proteins. Percentages bootstrap values (1000 replicates) are shown at branch nodes. Rhodospirillum rubrum ATCC 11170T was used as the out-group. Bar, 0.20 substitutions per nucleotide position.
Figure 2. The IQ-TREE based on 120 ubiquitous single-copy proteins. Percentages bootstrap values (1000 replicates) are shown at branch nodes. Rhodospirillum rubrum ATCC 11170T was used as the out-group. Bar, 0.20 substitutions per nucleotide position.
Microorganisms 10 02154 g002
Figure 3. Comparisons of Pseudotabrizicola orthologous protein groups in four Pseudotabrizicola genomes. (A) Venn diagram displaying the numbers of core gene families and unique genes for each of the four Pseudotabrizicola strains. (B) Percentage of core, accessory, and unique genes in each of the four genomes.
Figure 3. Comparisons of Pseudotabrizicola orthologous protein groups in four Pseudotabrizicola genomes. (A) Venn diagram displaying the numbers of core gene families and unique genes for each of the four Pseudotabrizicola strains. (B) Percentage of core, accessory, and unique genes in each of the four genomes.
Microorganisms 10 02154 g003
Figure 4. Heat maps of complete and incomplete metabolic pathways in the genomes of strain XJSPT, P. sediminis KCTC 72015T, P. alkalilacus KCTC 62173T, and P. algicola KCTC 72206T.
Figure 4. Heat maps of complete and incomplete metabolic pathways in the genomes of strain XJSPT, P. sediminis KCTC 72015T, P. alkalilacus KCTC 62173T, and P. algicola KCTC 72206T.
Microorganisms 10 02154 g004
Table 1. Comparisons of the 16S rRNA gene sequence similarity, average nucleotide identity (ANI), and DNA–DNA hybridization (DDH) values between strain XJSPT and members of the genus Pseudotabrizicola.
Table 1. Comparisons of the 16S rRNA gene sequence similarity, average nucleotide identity (ANI), and DNA–DNA hybridization (DDH) values between strain XJSPT and members of the genus Pseudotabrizicola.
StrainsXJSPT
16S rRNA Gene Similarity (%)ANI (%)DDH (%)
P. sediminis KCTC 72015T99.587.332.5
P. alkalilacus KCTC 62173T99.183.426.3
P. algicola KCTC 72206T97.781.423.9
Table 2. Genome statistics of strain XJSPT and members of the genus Pseudotabrizicola.
Table 2. Genome statistics of strain XJSPT and members of the genus Pseudotabrizicola.
1234
Genome size (bp)3,702,7584,040,6974,610,0614,491,281
Contigs148611029
N50 length (bp)812,613296,200284,500694,770
G+C content (mol %)63.463.062.964.4
Genes3552389744764391
Protein-coding genes3470375543374253
tRNA genes46434946
rRNA genes3373
ncRNA genes3333
GenBank IDPJOM
00000000
NZ_RPEM
00000000
NZ_QWEY
00000000
NZ_JAAIKE
000000000
Strains: 1, XJSPT; 2, P. sediminis KCTC 72015T; 3, P. alkalilacus KCTC 62173T; 4, P. algicola KCTC 72206T.
Table 3. Differential characteristics between strain XJSPT and related species.
Table 3. Differential characteristics between strain XJSPT and related species.
Characteristic123 a
Colony colorcream or light pinkopaque-pinkcream
Temperature range (°C)4–404–3515–37
NaCl range (%, w/v)0–6.01.0–2.00–3.0
pH range6.5–9.57.0–9.06.0–10.0
Voges–Proskauer reaction++
Enzyme activity:
Arginine dihydrolase+
Urease+
Alkaline phosphatase++
Cystine arylamidase+
ɑ-glucosidase++
β-glucosidase+
β-galactosidase+
Hydrolysis of:
Tween 20+NA
Tween 40++
Acid production from:
Glycerol++
d-tagatosew
d-arabitol+NA
Oxidation of:
ɑ-d-glucose+w+
Inosine++
d-glucose-6-PO4++
Major fatty acids (>10%)iso-C18:0, Summed feature 8iso-C18:0, Summed feature 8Summed feature 8
Polar lipidsPC, PG, PE, PL, LPC, PG, DPG, PE, APL, AL, PL, LPG, DPG, PE, PL, L
DNA G+C content (mol %)63.463.062.9
Strains: 1, XJSPT; 2, P. sediminis KCTC 72015T; 3, P. alkalilacus KCTC 62173T. All data were from this study unless indicated otherwise. +, positive; −, negative; w, weakly positive; NA, no data available. Summed features are groups of two or three fatty acids that cannot be separated by GLC using the MIDI system. Summed feature 8 comprised C18:1 ω7c and/or C18:1 ω6c. PC, phosphatidylcholine; PG, phosphatidylglycerol; DPG, diphosphatidylglycerol; PE, phosphatidylethanolamine; APL, unidentified aminophospholipid; AL, unidentified aminolipid; PL, unidentified phospholipid; L, unidentified lipid. a Data from Phurbu et al. (2019) [38].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ye, Y.-Q.; Han, J.-R.; Zhao, J.-X.; Ye, M.-Q.; Du, Z.-J. Genomic Analysis and Characterization of Pseudotabrizicola formosa sp. nov., a Novel Aerobic Anoxygenic Phototrophic Bacterium, Isolated from Sayram Lake Water. Microorganisms 2022, 10, 2154. https://doi.org/10.3390/microorganisms10112154

AMA Style

Ye Y-Q, Han J-R, Zhao J-X, Ye M-Q, Du Z-J. Genomic Analysis and Characterization of Pseudotabrizicola formosa sp. nov., a Novel Aerobic Anoxygenic Phototrophic Bacterium, Isolated from Sayram Lake Water. Microorganisms. 2022; 10(11):2154. https://doi.org/10.3390/microorganisms10112154

Chicago/Turabian Style

Ye, Yu-Qi, Ji-Ru Han, Jin-Xin Zhao, Meng-Qi Ye, and Zong-Jun Du. 2022. "Genomic Analysis and Characterization of Pseudotabrizicola formosa sp. nov., a Novel Aerobic Anoxygenic Phototrophic Bacterium, Isolated from Sayram Lake Water" Microorganisms 10, no. 11: 2154. https://doi.org/10.3390/microorganisms10112154

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop