IJSEM Papers in Press. Published March 12, 2014 as doi:10.1099/ijs.0.059667-0

1

Burkholderia cordobensis sp. nov. from agricultural soils in Argentina

2

Walter O. Draghi12; Charlotte Peeters2; Margo Cnockaert2; Cindy Snauwaert3; Luis G. Wall4;

3

Angeles Zorreguieta1 and Peter Vandamme2*

4 5

1

6

Tecnológicas, Patricias Argentinas 435, C1405BWE Buenos Aires, Argentina.

7

2

8

Ledeganckstraat 35, B-9000 Ghent, Belgium

9

3

Fundación Instituto Leloir and IIBA - Consejo Nacional de Investigaciones Científicas y

Laboratory of Microbiology, Dept. Biochemistry and Microbiology, Ghent University, K. L.

BCCM/LMG Bacteria Collection, Faculty of Sciences, Ghent University, K. L.

10

Ledeganckstraat 35, B-9000 Ghent, Belgium.

11

4

12

Bernal, B1876BXD, Buenos Aires, Argentina

Science and Technology Dept. National University of Quilmes. Roque Sáenz Peña 352,

13 14

*

15

e-mail: [email protected]

16

Keywords: Burkholderia, Burkholderia cordobensis sp. nov., Burkholderia glathei,

17

Burkholderia sp. YI23

Author for correspondence: Peter Vandamme, Tel : +32 9 264 5113, Fax : +32 9 264 5092,

18 19

Running title: Burkholderia cordobensis sp. nov.

20

The Contents Category: New Taxa, subsection Proteobacteria

21 22 23 24 25

26

The GenBank/EMBL/DDBJ accession numbers for the 16S rRNA and gyrB gene sequences

27

of Burkholderia cordobensis sp. nov. LMG 27620T and LMG 27621 are HG324048 and

28

HG324055, and HG324049 and HG324056, respectively. The GenBank/EMBL/DDBJ

29

accession number for the gyrB gene sequence of Burkholderia grimmiae LMG 27580T is

30

HG324054.

31

Abstract

32

Two gram-negative, rod-shaped bacteria were isolated from agricultural soils in the Córdoba

33

province in central Argentina. Their 16S rRNA gene sequences demonstrated that they belong

34

to the genus Burkholderia, with Burkholderia zhejiangensis as most closely related, formally

35

named species which was confirmed through comparative gyrB sequence analysis. Whole cell

36

fatty acid analysis supported their assignment to the genus Burkholderia. Burkholderia sp.

37

strain YI23 for which a whole genome sequence is available represents the same taxon as

38

demonstrated by its highly similar 16S rRNA (100%) and gyrB (99.1-99.7%) gene sequences.

39

The results of DNA-DNA hybridization experiments and the physiological and biochemical

40

characterization further substantiated the genotypic and phenotypic distinctiveness of the

41

Argentinian soil isolates for which the name Burkholderia cordobensis sp. nov. is proposed,

42

with strain LMG 27620T (=CCUG 64368T) as the type strain.

43

44

Burkholderia species belong to the β-proteobacteria and are aerobic, chemoorganotrophic and

45

motile (except for Burkholderia mallei) Gram negative rods, which display a wide

46

physiological versatility allowing them to inhabit extremely different ecological niches

47

(Coenye & Vandamme, 2003; Compant et al., 2008). Several Burkholderia species are human

48

or animal pathogens; these include members of the Burkholderia cepacia complex, a group of

49

closely related species commonly isolated from respiratory specimens of cystic fibrosis

50

patients (Vandamme & Dawyndt, 2011) and species belonging to the Burkholderia

51

pseudomallei group which includes B. pseudomallei, the etiological agent of melioidosis in

52

humans and B. mallei, the causative agent of glanders in animals (Galyov et al., 2010).

53

However, Burkholderia strains including B. cepacia complex bacteria (Parke & Gurian-

54

Sherman, 2001) may have beneficial characteristics as well, which include biological nitrogen

55

fixation, phytohormone synthesis, plant defense induction and xenobiotic activity (Suárez-

56

Moreno et al., 2011; Vial et al., 2011). Also, a growing number of Burkholderia species lives

57

associated with other organisms: these include endophytes from plant tissues including root

58

nodules, and inhabitants of insect guts or fungal mycelia and spores (Gyaneshwar et al., 2011;

59

Kikuchi et al., 2005; Levy et al., 2003; Mahenthiralingam et al., 2008; Verstraete et al.,

60

2013).

61

Burkholderia strains occur commonly in pristine, agricultural or polluted soils in which their

62

population density can reach levels of 103-105 cfu/g soil (Pallud et al., 2001; Salles et al.,

63

2006b) but antagonistic Burkholderia species can be lost when soils are continuously used

64

under arable management schemes (Salles et al., 2006a). As part of the BIOSPAS Consortium

65

(www.biospas.org/en), we analyzed the composition of Burkholderia populations across

66

Argentinian agricultural soils which were subjected to different agricultural managements. A

67

thorough description of site characteristics and treatment definitions was published elsewhere

68

(Figuerola et al., 2012). From these samples, 1 g of soil was suspended into 10 ml saline

69

solution (0.85% w/v NaCl), shaken at 240 rpm for 30 min, vortexed for 1 min, sonicated in a

70

water bath during 30 sec and centrifuged for 2 min at 500 rpm. Two hundred µl of soil

71

suspension was plated onto modified semi-selective medium PCAT (but using citrulline

72

instead of tryptamine as the main nitrogen source; PCAT refers to Pseudomonas cepacia,

73

azelaic acid and tryptamine) composed of, in gl-1, MgSO4, 0.1; citrulline, 0.2; azelaic acid, 2;

74

K2HPO4, 4; KH2PO4, 4; yeast extract, 0.02 and agar, 15. The pH was set at 5.7. The medium

75

was supplemented with cycloheximide (0.2 gl-1) and crystal violet (0.002 gl-1) to inhibit

76

growth of fungi and Gram-positive bacteria, respectively (Burbage et al., 1982). After 5 days

77

of incubation at 28 °C, colonies were randomly selected, subcultured on LB medium

78

(composed of, in gl-1: Tryptone, 10; Yeast Extract, 5; NaCl,

79

15) to check purity, and maintained at -80 °C as a suspension with 20% (v/v) glycerol. Two

80

isolates from Monte Buey soils, in the Córdoba province (32°58′14″S; 62°27′06″W),

81

MMP81T (= LMG 27620T, isolated in 2010) and MAN52 (= LMG 27621, isolated in 2011)

82

were further characterized in the present study. When analyzing our data we found highly

83

similar 16S rRNA and gyrB sequences in the whole genome sequence of strain Burkholderia

84

sp. YI23, which was isolated from a golf course soil in South Korea and which showed the

85

ability to degrade the pesticide fenitrothion (GenBank accession numbers CP003087,

86

CP003088, CP003089, CP003090, CP003091, and CP003092 (Lim et al., 2012).

87

For PCR, genomic DNA was prepared using the method as described by (Niemann et al.,

88

1997). Random amplified polymorphic DNA patterns were generated using the primers

89

TGCGCGCGGG and AGCGGGCCAA, as previously described (Williams et al., 1990) and

90

demonstrated that isolates LMG 27620T and LMG 27621 represent genetically distinct strains

91

(Figure S1).

92

The nearly complete sequences of the 16S rRNA gene of strains LMG 27620T and LMG

93

27621 were obtained as described previously (Peeters et al., 2013). The Mothur software

10 and agar,

94

package (Schloss et al., 2009) was used to align the 16S rRNA gene sequences of strains

95

LMG 27620T, LMG 27621 and YI23 (1484 bp) with those of type strains of phylogenetically

96

related Burkholderia species (1388–1525 bp) against the Silva reference database (www.arb-

97

silva.de). Phylogenetic analysis was conducted in MEGA5 (Tamura et al., 2011) (Figure 1).

98

Uncorrected pairwise distances were calculated using MEGA5 (Tamura et al., 2011) and all

99

ambiguous positions were removed for each sequence pair. Strains LMG 27620T, LMG 27621

100

and YI23 had identical 16S rRNA gene sequences, which were highly similar to those of

101

Burkholderia zhejiangensis OP-1T (99.7%) and Burkholderia grimmiae R27T (99.3%) (Figure

102

1).

103

Partial sequences of the gyrB gene of strains LMG 27620T and LMG 27621 and their nearest

104

neighbors as determined by 16S rRNA sequence proximity were obtained using the method

105

described by (Spilker et al., 2009) as modified by (De Meyer et al., 2013). Sequence

106

assembly was performed using BioNumerics v5.10 (Applied Maths). Sequences (589-704 bp)

107

were aligned based on amino acid sequences using Muscle (Edgar, 2004) in MEGA5 (Tamura

108

et al., 2011). Phylogenetic analysis of nucleotide sequences was conducted in MEGA5

109

(Tamura et al., 2011) (Figure 2). Uncorrected pairwise distances were calculated using

110

MEGA5 (Tamura et al., 2011). Strains LMG 27620T, LMG 27621 and YI23 formed a

111

homogeneous cluster (sequences were >99% similar), which was supported by a 75%

112

bootstrap value (Figure 2) and grouped with B. zhejiangensis LMG 27258T (95.3-95.9%) and

113

B. grimmiae LMG 27580T (90.8-91.6%) as nearest neighbors.

114

For DNA-DNA hybridization and for the determination of the DNA G+C content, high-

115

molecular weight DNA was prepared as described by (Pitcher et al., 1989). DNA-DNA

116

hybridizations were performed using a microplate method and biotinylated probe DNA (Ezaki

117

et al., 1989). The hybridization temperature was 45°C ± 1°C. Reciprocal reactions (A x B and

118

B x A) were performed and the variation was within the limits of this method (Goris et al.,

119

1998). DNA-DNA hybridization experiments were performed between strain LMG 27620T

120

and B. zhejiangensis LMG 27258T, its nearest named phylogenetic neighbor (Figure 1 and 2).

121

The DNA-DNA hybridization value between both strains was 44%. The G+C content of DNA

122

was determined by HPLC according to the method of (Mesbah & Whitman, 1989) using a

123

Waters Breeze HPLC system and XBridge Shield RP18 column thermostabilised at 37°C. The

124

solvent was 0.02 M NH4H2PO4 (pH 4.0) with 1.5% (v/v) acetonitrile. Non-methylated lambda

125

phage (Sigma) and E. coli DNA were used as calibration reference and control, respectively.

126

The DNA G+C content of strain LMG 27620T was 63.6 mol%, which is within the range

127

reported for Burkholderia genus (59-69.9 mol%) (Gillis et al., 1995; Yabuuchi et al., 1992)

128

and which corresponds to the G+C content of 63.5%, reported for the Burkholderia sp. strain

129

YI23 whole genome sequence (Lim et al., 2012).

130

Phenotypic analysis of strains LMG 27620T and LMG 27621 and of B. zhejiangensis LMG

131

27258T was performed on Tryptone Soya Agar (TSA) at 28°C unless indicated otherwise

132

(LMG medium 14; http://bccm.belspo.be/db/media_search_form.php). Cell morphology and

133

motility were observed by phase contrast microscopy. Oxidase activity was detected by

134

immersion of cells in 1% N,N,N′,N′-tetramethyl-p-phenylenediamine solution and catalase

135

activity was determined by bubbling through flooding a colony with 10% H2O2. Lipase

136

activity was determined according to the method described by (Sierra, 1957). Growth on

137

MacConkey medium was observed after 48h of incubation at 28 °C. Starch hydrolysis was

138

observed after 48h of incubation at 28 °C in LMG14 medium amended with 2% of starch.

139

Deoxyribonuclease activity (DNase) was observed after 48h of incubation at 28 °C on BD

140

Difco™ DNase Test Agar, according to the method of (Jeffries et al., 1957). Casein hydrolysis

141

was observed after 48 h of incubation at 28 °C on TSA plates amended with 1.3% of skim

142

milk, through the observation of clear halos around colonies. Other biochemical tests were

143

performed by inoculating API 20NE and API ZYM strips (BioMérieux) according to the

144

manufacturer’s instructions and incubating for 48h at 28°C, or for 4 h at 28 °C, respectively.

145

Growth was tested at 28°C on nutrient broth (NB, from BD Difco™) at pH4 - pH9 using

146

appropriate biological buffers (acetate, citrate-Na2HPO4, phosphate buffer and Tris-

147

hydrochloride). Growth on LMG 14 medium was tested at 4 °C, 15 °C, 28 °C, 30 °C, 37 °C,

148

40 °C and 45°C on LMG medium14. The results of the phenotypic and biochemical tests are

149

given in the species description below and in Table 1. They allow a clear differentiation of the

150

taxon represented by strains LMG 27620T and LMG 27621 and its nearest phylogenetic

151

neighbor, B. zhejiangensis: the former exhibits oxidase but not arginine dihydrolase or urease

152

activity, reduces nitrate, hydrolyses Tween 80, and does not assimilate caprate; opposite test

153

results were obtained for the latter. In addition, on the basis of phenotypic data published by

154

(Tian et al., 2013) using the same test gallery, strains LMG 27620T and LMG 27621 can be

155

differentiated from B. grimmiae R27T: the former exhibits no arginine dihydrolase or urease

156

activity, and assimilates phenyl acetate. In addition, it does not hydrolyse starch and grows on

157

MacConkey agar; opposite test results were obtained for B. grimmiae R27T. Differential

158

characteristics towards other Burkholderia species in this same lineage are presented in Table

159

1.

160

161

Whole-cell fatty acid methyl esters were extracted according to the MIDI protocol

162

(http://www.microbialid.com/PDF/TechNote_101.pdf).

163

temperature, medium, and physiological age (overlap area of the second and third quadrant

164

from a quadrant streak for late log phase growing cells) were as in the MIDI protocol. The

165

profiles were generated using an Agilent Technologies 6890N gas chromatograph (Santa

166

Clara, CA USA), identified and clustered using the Microbial Identification System software

167

and MIDI TSBA database version 5.0. The most abundant fatty acids of strains LMG 27620T

All

characteristics

such

as

168

and LMG 27621 were: C18:1 ω7c, summed feature 3 (most likely C16:1 ω7c) and C16:0. Moderate

169

to low amounts of summed feature 2 (most likely C14:0 3OH), cyclo-C17:0, C16:0 3-OH, C14:0,

170

cyclo-C19:0 ω8c, C16:1 2-OH and C16:0 2-OH have been detected. The presence of C16:0 3-OH

171

supports the placement of these strains in the genus Burkholderia (Yabuuchi et al., 1992) but

172

the overall profile is very similar to that of its nearest neighbors (Table 2).

173

The present phenotypic, chemotaxonomic and genotypic data demonstrate that strains LMG

174

27620T, LMG 27621 and YI23 represent a novel Burkholderia species that can be

175

distinguished from its nearest phylogenetic neighbors both phenotypically as well as

176

genotypically. We therefore propose to classify these bacteria as a new species for which the

177

name Burkholderia cordobensis sp. nov. is proposed, with strain LMG 27620T

178

(=CCUG64368T) as the type strain.

179

180

Description of Burkholderia cordobensis sp. nov.

181

Burkholderia cordobensis (cor.do.ben´sis. N. L. fem. adj. cordobensis pertaining to the

182

Argentinian province Córdoba.

183

Gram negative, aerobic, motile, non-spore forming rods about 0.4-0.7 in width and 1.2-1.8 in

184

length. Colonies are round, with entire margins, a convex elevation, a white-creamy color, and

185

a moist appearance and are 1-2 mm in diameter after 48 h of growth on TSA medium at 28

186

°C. Growth on MacConkey agar. Growth occurs between 28°C and 37 °C, and from pH 6 to

187

pH 8 in liquid nutrient broth at 28 °C. Catalase and oxidase activity is present, but no arginine

188

dihydrolase, urease, β-glucosidase, gelatinase or β-galactosidase activity. When tested by

189

using API ZYM strips, activities of the following enzymes are present: alkaline and acid

190

phospatase, leucyl arylamidase, and phosphoamidase; enzymatic activities were absent for C4-

191

lipase, C8-lipase, C14-lipase, valine and cystine arylamidase, trypsin, chymotrypsin, α-

192

galactosidase, β-galactosidase, β-glucuronidase, α-glucosidase, β-glucosidase, N-acetyl-β-

193

glucosaminidase, α- mannosidase and α-fucosidase. Tween 20, Tween 40, Tween 60 and

194

Tween 80 are degraded, but casein and starch were not hydrolysed. When tested by using API

195

20NE strips nitrate was reduced; glucose, arabinose (except for strain LMG 27621), mannose,

196

mannitol, N-acetyl-glucosamine, gluconate, malate and phenyl-acetate were assimilated, but

197

not maltose, caprate, adipate or citrate. No fermentation of glucose, tryptophanase, arginine

198

dihydrolase, urease, or PNPG beta-galactosidase activity; no hydrolysis of aesculin or gelatin

199

liquefaction. The most abundant fatty acids were C18:1 ω7c, summed feature 3 (probably C16:1

200

ω7c) and C16:0.

201

The type strain, LMG 27620T(=CCUG 64368T), was isolated from agricultural soils at

202

Córdoba province, Argentina. The DNA G+C content is 63.6%. The whole genome sequence

203

of strain YI23 is 8.89 Mb and consists of three chromosomes and three plasmids (Lim et al.,

204

2012).

205

206

Acknowledgements

207

This work was partially supported by grants PAE 36976-PID 52 and PICT 2010/2691

208

(Agencia Nacional de Promoción Científica y Tecnológica, Argentina). W.O.D., A.Z. and

209

L.G.W. are members of Research Career at CONICET, Argentina.

210

Bibliography

211

Burbage, D. A., Sasser, M. & Lumsden, R. D. (1982). A medium selective for Pseudomonas

212 213 214

cepacia. Phytopathology 72. Coenye, T. & Vandamme, P. (2003). Diversity and significance of Burkholderia species occupying diverse ecological niches. Environ Microbiol 5, 719-729.

215

Compant, S., Nowak, J., Coenye, T., Clement, C. & Ait Barka, E. (2008). Diversity and

216

occurrence of Burkholderia spp. in the natural environment. FEMS Microbiol Rev 32,

217

607-626.

218

De Meyer, S. E., Cnockaert, M., Ardley, J. K., Trengove, R. D., Garau, G., Howieson, J.

219

G. & Vandamme, P. (2013). Burkholderia rhynchosiae sp. nov. isolated from

220

Rhynchosia ferulifolia root nodules from South Africa. Int J Syst Evol Microbiol 63,

221

3944-3949.

222 223

Edgar, R. C. (2004). MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res 32, 1792-1797.

224

Ezaki, T., Hashimoto, Y. & Yabuuchi, E. (1989). Fluorometric deoxyribonucleic acid-

225

deoxyribonucleic acid hybridization in microdilution wells as an alternative to

226

membrane filter hybridization in which radioisotopes are used to determine genetic

227

relatedness among bacterial strains. Int J Syst Bacteriol 39, 224-229.

228

Figuerola, E. L., Guerrero, L. D., Rosa, S. M., Simonetti, L., Duval, M. E., Galantini, J.

229

A., Bedano, J. C., Wall, L. G. & Erijman, L. (2012). Bacterial indicator of

230

agricultural management for soil under no-till crop production. PLoS One 7, e51075.

231

Galyov, E. E., Brett, P. J. & DeShazer, D. (2010). Molecular insights into Burkholderia

232

pseudomallei and Burkholderia mallei pathogenesis. Annu Rev Microbiol 64, 495-517.

233

Gillis, M., Van Van, T., Bardin, R., Goor, M., Hebbar, P., Willems, A., Segers, P.,

234

Kersters, K., Heulin, T. & other authors (1995). Polyphasic taxonomy in the genus

235

Burkholderia leading to an emended description of the genus and proposition of

236

Burkholderia vietnamiensis sp. nov. for N2-fixing isolates from rice in Vietnam. Int J

237

Syst Bacteriol 45, 274-289.

238

Goris, J., Suzuki, K., De Vos, P., Nakase, T. & Kersters, K. (1998). Evaluation of a

239

microplate DNA - DNA hybridization method compared with the initial renaturation

240

method. Can J Microbiol 44, 1148-1153.

241

Gyaneshwar, P., Hirsch, A. M., Moulin, L., Chen, W. M., Elliott, G. N., Bontemps, C.,

242

Estrada-de Los Santos, P., Gross, E., Dos Reis, F. B. & other authors (2011).

243

Legume-nodulating betaproteobacteria: diversity, host range, and future prospects.

244

Mol Plant Microbe Interact 24, 1276-1288.

245 246 247 248

Hasegawa, M., Kishino, H. & Yano, T.-a. (1985). Dating of the human-ape splitting by a molecular clock of mitochondrial DNA. J Mol Evol 22, 160-174. Jeffries, C. D., Holtman, D. F. & Guse, D. G. (1957). Rapid method for determining the activity of microorganisms on nucleic acids. J Bacteriol 73, 590-591.

249

Kikuchi, Y., Meng, X. Y. & Fukatsu, T. (2005). Gut symbiotic bacteria of the genus

250

Burkholderia in the broad-headed bugs Riptortus clavatus and Leptocorisa chinensis

251

(Heteroptera: Alydidae). Appl Environ Microbiol 71, 4035-4043.

252

Levy, A., Chang, B. J., Abbott, L. K., Kuo, J., Harnett, G. & Inglis, T. J. (2003). Invasion

253

of spores of the arbuscular mycorrhizal fungus Gigaspora decipiens by Burkholderia

254

spp. Appl Environ Microbiol 69, 6250-6256.

255

Lim, J. S., Choi, B. S., Choi, A. Y., Kim, K. D., Kim, D. I., Choi, I. Y. & Ka, J. O. (2012).

256

Complete genome sequence of the fenitrothion-degrading Burkholderia sp. strain

257

YI23. J Bacteriol 194, 896-896.

258

Mahenthiralingam, E., Baldwin, A. & Dowson, C. G. (2008). Burkholderia cepacia

259

complex bacteria: opportunistic pathogens with important natural biology. J Appl

260

Microbiol 104, 1539-1551.

261

Mesbah, M. & Whitman, W. B. (1989). Measurement of deoxyguanosine/thymidine ratios in

262

complex mixtures by high-performance liquid chromatography for determination of

263

the mole percentage guanine + cytosine of DNA. J Chromatogr 479, 297-306.

264

Niemann, S., Pühler, A., Tichy, H. V., Simon, R. & Selbitschka, W. (1997). Evaluation of

265

the resolving power of three different DNA fingerprinting methods to discriminate

266

among isolates of a natural Rhizobium meliloti population. J Appl Microbiol 82, 477-

267

484.

268

Pallud, C., Viallard, V., Balandreau, J., Normand, P. & Grundmann, G. (2001).

269

Combined use of a specific probe and PCAT medium to study Burkholderia in soil. J

270

Microbiol Methods 47, 25-34.

271

Parke, J. L. & Gurian-Sherman, D. (2001). Diversity of the Burkholderia cepacia complex

272

and implications for risk assessment of biological control strains. Annu Rev

273

Phytopathol 39, 225-258.

274

Peeters, C., Zlosnik, J. E. A., Spilker, T., Hird, T. J., LiPuma, J. J. & Vandamme, P.

275

(2013). Burkholderia pseudomultivorans sp. nov., a novel Burkholderia cepacia

276

complex species from human respiratory samples and the rhizosphere. Syst Appl

277

Microbiol 36, 483-489.

278 279

Pitcher, D. G., Saunders, N. A. & Owen, R. J. (1989). Rapid extraction of bacterial genomic DNA with guanidium thiocyanate. Lett Appl Microbiol 8, 151-156.

280

Salles, J. F., van Elsas, J. D. & van Veen, J. A. (2006a). Effect of agricultural management

281

regime on Burkholderia community structure in soil. Microb Ecol 52, 267-279.

282

Salles, J. F., Samyn, E., Vandamme, P., van Veen, J. A. & van Elsas, J. D. (2006b).

283

Changes in agricultural management drive the diversity of Burkholderia species

284

isolated from soil on PCAT medium. Soil Biol Biochem 38, 661-673.

285

Schloss, P. D., Westcott, S. L., Ryabin, T., Hall, J. R., Hartmann, M., Hollister, E. B.,

286

Lesniewski, R. A., Oakley, B. B., Parks, D. H. & other authors (2009). Introducing

287

mothur: open-source, platform-independent, community-supported software for

288

describing and comparing microbial communities. Appl Environ Microbiol 75, 7537-

289

7541.

290

Sierra, G. (1957). A simple method for the detection of lipolytic activity of micro-organisms

291

and some observations on the influence of the contact between cells and fatty

292

substrates. Antonie van Leeuwenhoek 23, 15-22.

293

Spilker, T., Baldwin, A., Bumford, A., Dowson, C. G., Mahenthiralingam, E. & LiPuma,

294

J. J. (2009). Expanded multilocus sequence typing for Burkholderia species. J Clin

295

Microbiol 47, 2607-2610.

296

Suárez-Moreno, Z., Caballero-Mellado, J., Coutinho, B., Mendonça-Previato, L., James,

297

E. & Venturi, V. (2011). Common features of environmental and potentially

298

beneficial plant-associated Burkholderia. Microb Ecol, 1-18.

299 300

Tamura, K. (1992). Estimation of the number of nucleotide substitutions when there are strong transition-transversion and G+C-content biases. Mol Biol Evol 9, 678-687.

301

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

302

MEGA5: molecular evolutionary genetics analysis using maximum likelihood,

303

evolutionary distance, and maximum parsimony methods. Mol Biol Evol 28, 2731-

304

2739.

305

Tian, Y., Kong, B. H., Liu, S. L., Li, C. L., Yu, R., Liu, L. & Li, Y. H. (2013). Burkholderia

306

grimmiae sp. nov., isolated from a xerophilous moss (Grimmia montana). Int J Syst

307

Evol Microbiol 63, 2108-2113.

308 309

Vandamme, P. & Dawyndt, P. (2011). Classification and identification of the Burkholderia cepacia complex: past, present and future. Syst Appl Microbiol 34, 87-95.

310

Vandamme, P., De Brandt, E., Houf, K., Salles, J. F., Dirk van Elsas, J., Spilker, T. &

311

LiPuma, J. J. (2013). Burkholderia humi sp. nov., Burkholderia choica sp. nov.,

312

Burkholderia telluris sp. nov., Burkholderia terrestris sp. nov. and Burkholderia udeis

313

sp. nov.: Burkholderia glathei-like bacteria from soil and rhizosphere soil. Int J Syst

314

Evol Microbiol 63, 4707-4718.

315 316

Verstraete, B., Janssens, S., Smets, E. & Dessein, S. (2013). Symbiotic β-proteobacteria beyond legumes: Burkholderia in Rubiaceae. PLoS One 8, e55260.

317

Vial, L., Chapalain, A., Groleau, M. C. & Deziel, E. (2011). The various lifestyles of the

318

Burkholderia cepacia complex species: a tribute to adaptation. Environ Microbiol 13,

319

1-12.

320

Williams, J. G. K., Kubelik, A. R., Livak, K. J., Rafalski, J. A. & Tingey, S. V. (1990).

321

DNA polymorphisms amplified by arbitrary primers are useful as genetic markers.

322

Nucleic Acids Res 18, 6531-6535.

323

Yabuuchi, E., Kosako, Y., Oyaizu, H., Yano, I., Hotta, H., Hashimoto, Y., Ezaki, T. &

324

Arakawa, M. (1992). Proposal of Burkholderia gen. nov. and transfer of seven

325

species of the genus Pseudomonas homology group II to the new genus, with the type

326

species Burkholderia cepacia (Palleroni and Holmes 1981) comb. nov. Microbiol

327

Immunol 36, 1251-1275.

328 329

330

Table 1. Phenotypic characteristics distinguishing B. cordobensis sp. nov. from its nearest

331

phylogenetic neighbor species.

332

Species: 1. B. cordobensis sp. nov. (LMG 26720T, LMG 27621); 2: B. zhejiangensis (LMG

333

27258T, LMG 26180, LMG 26181); 3, B. grimmiae (R27T); 4, B. choica (LMG 22940T); 5, B.

334

glathei (LMG 14190T); 6, B. humi (LMG 22934T);7, B. sordidicola (LMG 22029T); 8, B.

335

telluris (LMG 22936T); 9, B. terrestris (LMG 22937T); 10, B. udeis (LMG 27134T). Results

336

from B. cordobensis sp. nov. and its nearest neighbor B. zhejiangensis are from the present

337

study. Test results of the type strain are given first, followed by remaining strains in order as

338

presented in the title. Data from B. grimmiae are extracted from (Tian et al., 2013). Data from

339

the remaining species were taken from (Vandamme et al., 2013). +, Present; 2, absent; W,

340

weak reaction; ND, Not determined.

341

1 Growth at: 37 °C 40 °C pH 8 Hydrolysis of Tween 60 Nitrate reduction (API -20NE) β-Galactosidase activity (API -20NE) Assimilation of (API -20NE): Arabinose Mannose Mannitol N-Acetylglucosamine Gluconate Caprate Malate Citrate Phenylacetate Enzyme activity (API ZYM) C4-Lipase C8-Lipase Valine arylamidase Cystine arylamidase β-Galactosidase 342

+-ww ++ w+

2

3

4 5 6 7 8 9 10

+++ + w +++ + +++ + +++ ND + +++ + -

+ -

+ -

w +

w + +

-

+ +

--

---

-

- - - + - - w

+++ ++ ++ ++ -++ -++

+++ +++ --+++ +++ +++ +++ --++w

+ + + + + + -

w w w w w -

w + w + + + + + +

+ + + + w + + +

+ + + + + v -

+ + + + + + + + +

+ + + + w + + +

+ + + + + + -

------

-----------

+ + -

w + w + -

+ w w -

+ + -

+ + +

+ w -

+ w w -

+ + + +

343

Table 2. Average fatty acid composition of B. cordobensis sp. nov. and of its nearest

344

phylogenetic neighbor species. Species: 1, B. cordobensis sp. nov. (LMG 27620T and LMG

345

27621); 2, B. zhejiangensis (LMG 27258T, LMG 26180, LMG 26181); 3, B. grimmiae (LMG

346

27258T); 4, B. choica; 5, B. glathei; 6, B. humi; 7, B. sordidicola; 8, B. telluris; 9, B. terrestris;

347

10, B. udeis. Data for B. cordobensis sp. nov. and B. grimmiae are from the present study. The

348

remaining data were extracted from (Vandamme et al., 2013) and were generated in the same

349

cultivation and analysis conditions. Those fatty acids for which the mean amount for all taxa

350

was < 1% are not included, therefore, the percentages may not add up to 100 %. TR, Trace

351

amount (

Burkholderia cordobensis sp. nov., from agricultural soils.

Two Gram-negative, rod-shaped bacteria were isolated from agricultural soils in Córdoba province in central Argentina. Their 16S rRNA gene sequences d...
336KB Sizes 5 Downloads 3 Views