<<

bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1 Potential PGPR properties of cellulolytic, -

2 fixing, and phosphate-solubilizing of a

3 rehabilitated tropical forest soil

4 Amelia Tang1¶*, Ahmed Osumanu Haruna123¶, Nik Muhamad Ab. Majid3¶

5 1 Faculty of Agriculture and Food Sciences, Universiti Putra Malaysia Bintulu Campus,

6 Sarawak, Malaysia

7 2 Institute of Tropical Agriculture and Food Security (ITAFoS), Universiti Putra

8 Malaysia, Serdang, Selangor, Malaysia

9 3 Institute of Tropical Forestry and Forest Products (INTROP), Universiti Putra

10 Malaysia, Selangor, Malaysia

11

12 *Corresponding author

13 E-mail: [email protected] (AT)

14 ¶ These authors contributed equally to this work.

15

16

17

18

19

1 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

20 Abstract

21 In the midst of major soil degradation and erosion faced by tropical ecosystems,

22 rehabilitated forests are established to avoid further deterioration of forest land. In this context,

23 cellulolytic, nitrogen-fixing (N-fixing), and phosphate-solubilizing bacteria are very important

24 functional groups in regulating the elemental cycle and plant nutrition, hence replenishing the

25 nutrient content in forest soil. As other potential plant growth-promoting (PGP) rhizobacteria,

26 these functional bacteria could have cross-functional abilities or beneficial traits that are

27 essential for plants and improve their growths. This study was conducted to isolate, identify, and

28 characterize selected PGP properties of these 3 functional groups of bacteria from tropical

29 rehabilitated forest soils at Universiti Putra Malaysia Bintulu Sarawak Campus, Malaysia.

30 Isolated cellulolytic, N-fixing and phosphate-solubilizing bacteria were characterized for

31 respective functional activities, biochemical properties, molecularly identified, and assessed for

32 PGP assays based on seed germination and indole-3-acetic acid (IAA) production. Out of 15

33 identified bacterial isolates exhibiting beneficial phenotypic traits, a third belong to

34 Burkholderia and a fifth to Stenotrophomonas sp. with both genera consisting of members from

35 two different functional groups. Among the tested bacterial strains, isolate Serratia

36 nematodiphila C46d, Burkholderia nodosa NB1, and Burkholderia cepacia PC8 showed

37 outstanding cellulase, N-fixing, and phosphate-solubilizing activities, respectively. The results

38 of the experiments confirmed the multiple PGP traits of selected bacterial isolates based on

39 respective high functional activities, , shoot lengths, and seedling vigour improvements

40 when bacterized on mung bean seeds, as well as presented some significant IAA productions.

41 The results of this study indicated that these functional bacterial strains could potentially be

42 included in future biotechnological screenings to produce beneficial synergistic effects via their

43 versatile properties on improving soil fertility and possible crop growth stimulation.

44 Keywords: rehabilitated tropical forest soil; cellulolytic bacteria; N-fixing bacteria; phosphate-

45 solubilizing bacteria; PGP traits

2 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

46 Introduction

47 Albeit vast in and biodiversity, tropical ecosystems are facing major

48 deforestations, leading to widespread soil erosion which garnered much recent attention

49 [1-3]. As soil erosion aggravates, the soil loss overrides soil establishment, which

50 results in diminishing soil resources and the ecosystem supports they render [4-5], in

51 particularly manifested as low productivity and farming sustainability on the degraded

52 soils [6]. Planted forest or rehabilitation efforts are undertaken to avoid further

53 deterioration of forest land.

54 Bacteria has large biodiversity composition in soils, and thus, is major

55 participant of soil principal processes that regulate whole terrestrial ecosystems

56 operations, by means of completing nutrient, and geochemical cycles, including C,N,S,

57 and P [7-9]. In plant litter mass, 20-30% portion are constituted as cellulose, making the

58 latter as most abundant [10]. Decomposition of cellulose has been

59 established to be primarily accountable by soil , which assimilate the

60 simple sugars resulting from the reduction process of the complex polysaccharides [11-

61 12]. Despite the general ruling capabilities of fungi over bacteria in cellulose

62 decomposition in soils [13-14], these capabilities are also reserved to some, albeit

63 phylogenetically different taxa of bacteria [11]. Previous reports have emphasized

64 bacterial taxa which were recently discovered to possess cellulolytic capabilities, in

65 which they were formerly not known to reduce cellulose [15-16]. To this end, the study

66 of cellulolytic bacterial diversity in tropical soils is warranted.

67 Nitrogen needs by plant are supported by means of organic compound

68 degradation, atmospheric deposition of N, and biological nitrogen-fixation (BNF),

69 whereby BNF supplies 97% of natural N reserve [17] via Bacteria and Archaea [18-19].

70 The N allocation in tropical ecosystems contributed by N-fixing microorganisms

3 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

71 (), prevailing mostly in free-living condition, could range from 12.2 to 36.1

72 kg ha-1 year-1 [20]. Considering the importance of free-living diazotrophic communities

73 to ecosystem processes, there is a need to broaden our knowledge on their diversity in

74 rehabilitated tropical forests. This is essential because there is dearth of information of

75 this kind.

76 In most tropical soils which are usually very acidic, P is one of the most

77 commonly deficient macronutrients, owing to the prevalence of typically high Al and/or

78 Fe level, in which P is fixed as insoluble mineral complexes, causing unavailability of P

79 for root uptake [21]. Apart from chemical treatments, microorganisms have also been

80 applied over the years to improve nutrient availability, especially P in soils, as well as

81 alleviating Al toxicity [22]. Furthermore, tropical acid soils support only limited acid-

82 tolerant plant species and microbes. In soils, phosphate-solubilizers are predominated

83 by bacteria over fungi at 1-50% and 0.1-0.5%, of total population, respectively [23].

84 Bacterial strains consisting of , , and have been cited to

85 be the dominant phosphate-solubilizers [24]. In addition, Burkholderia strains and

86 species which occur in broadly different natural niches have been frequently reported as

87 plant-associated bacteria, albeit their prevalence as free-living in the , as

88 well as epiphytic or endophytic, obligate endosymbionts, or phytopathogens [25]. The

89 phosphate-solubilizing bacteria could release substantial amount of organic acids crucial

90 for fixing Al via a process, thus reducing Al toxicity to plant in the

91 highly weathered soils [26].

92 As other potential PGPR, these functional bacteria could have cross-functional

93 abilities or beneficial traits, apart from their primary functional ability such as the

94 production of PGP phytohormones, polysaccharides, and organic acids that are essential

95 for plant growth and development [22]. In addition, IAA which is known to stimulate

4 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

96 shoot elongation and/or in particularly root structures by those IAA produced by the

97 bacteria [27], could result in more efficient elemental nutrient acquisition by host plants,

98 leading to higher plant growth and development. To be specific, adoption of functional

99 bacterial strains with multiple functional and PGP traits in crop production would not

100 only promote sustainably increased crop productivity, the practice also offers an option

101 for minimizing costs related to chemical fertilizers application, while at the same time

102 reducing the pollution risks from rampant application of chemical fertilizers.

103 However, the isolation and characterization of these particular three types of

104 potential functional bacteria from tropical soils of rehabilitated forest remain scarce.

105 Thus, the aims of this study were to: 1) isolate and identify cellulolytic, N-fixing, and

106 phosphate-solubilizing bacteria from a rehabilitated tropical forest soil at Universiti

107 Putra Malaysia Bintulu Sarawak Campus, Malaysia, and 2) characterize cellulolytic, N-

108 fixing, and phosphate-solubilizing bacteria for their respective functional activities, such

109 as IAA production, early plant growth promotion, some phenotypic, and biochemical

110 properties.

111

112 Materials and Methods

113 Soil sampling

114 Soil samples were collected from a rehabilitated forest site at Universiti Putra

115 Malaysia Bintulu Sarawak Campus through random sampling design using aseptic

116 techniques near rhizospheric zones of planted trees. Soil samples were collected at 0-10

117 cm depth using an auger. Prior to soil sampling, any litter at the surface was manually

118 removed [28]. The plastic bags with soil samples were transported to the laboratory at

119 Universiti Putra Malaysia Bintulu Sarawak Campus for analysis. The soil samples were

5 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

120 homogenized or mixed thoroughly to make composite (consolidated) samples, and

121 surface organic materials, litter, roots, rocks, and macro fauna were removed. Soil

122 samples were kept in tightly sealed plastic bags and stored at 4 °C (for less than 3 days

123 before microbial analysis) to keep the soil moist, and to preserve biological properties.

124

125 Isolation of free-living nitrogen-fixing bacteria

126 Ten-fold serial dilution technique was used to determine the CFU number of the

127 culturable N cycle functional bacteria. Bacterial colonies were enumerated on N-free

128 malate (Nfb) medium [29] with some modifications containing per litre: 0.4 g of

129 KH2PO4, 0.1 g of K2HPO4, 0.2 g of MgSO4.7H2O, 0.1 g of NaCl, 0.02 g of CaCl2, 0.01

130 g of FeCl3, 0.002 g of MoO4Na.2H2O, 5.0 g of malate, 5 mL of bromothymol

131 blue with 0.5% solution and 15.0 mL of agar. The whole media content were

132 adjusted to pH 7 before being sterilized in autoclave (121 oC for 20 minutes).

133 Approximately 20 mL of media was placed in each petri dish before being incubated at

o 134 30 C after triplicate inoculation. The colonies that fix N2 alkalinized the culture

135 medium (blue) due to production. Some colonies exhibiting large and instant

136 blue halo on the medium were isolated, purified on new Nfb media, sub cultured on

137 agar slants, and stored at 4 oC for further studies.

138

139 Isolation of phosphate-solubilizing bacteria

140 Cells and spores were detached from soil surface by shaking at 150 rpm a

141 mixture of 5 g of fresh soil and 100 mL of sterile distilled water for 24 hr. Supernatant

142 was ten-fold serially diluted in sterile distilled water and 0.1 mL aliquots in triplicate

143 were spread over Petri dishes containing differential growth media for phosphate-

144 solubilizing bacteria. This isolation technique is based on the formation of halos around

6 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

145 the colonies of bacteria capable to solubilize a calcium phosphate insoluble compound

146 [30-31]. The inoculated plates were incubated at 30 °C. Culture media employed to

147 differentially isolate phosphate-solubilizing bacterial strains consisted of variations of a

148 chemically defined medium [32-33] with some modifications. All media contained per

-1 -1 149 litre: 50 mL of a salts solution (MgCl2·6H2O, 100 g L ; MgSO4·7H2O, 5 g L ; KCl, 4 g

-1 -1 150 L ; (NH4)2SO4, 2 g L ), 10 g of sucrose as C source, and 2.5 g of tricalcium phosphate,

-1 151 Ca3(PO4)2, as an insoluble source of phosphate. Six mL L of a pH indicator dye

152 (bromophenol blue, 0.4% solution) was added to the media [30]. For

153 solidification, 20 g L-1 of agar was added and pH of medium was adjusted to 7.

154 Phosphate-solubilizing bacterial colonies exhibiting large and instant halos on the media

155 were isolated, further purified on new media, sub cultured on agar slants, and stored at 4

156 oC for further studies.

157

158 Isolation of cellulolytic bacteria

159 This procedure was modified from those as described by Hendricks et al. [34].

160 The cellulolytic bacteria was enumerated by triplicate spread plate inoculation onto

161 cellulose-Congo red agar with 10-fold sterile tap water dilutions of soil, and incubated

o 162 at 30 C. The cellulose-Congo red agar consisted of 0.50 g of K2HPO4, 0.25 g of

163 MgSO4, 1.88 g of cellulose microgranular powder, 0.20 g of Congo red, 5.00 g of agar,

164 2.00 g of gelatin, 100 mL of soil extract, and 900 mL of tap water (used to provide

165 essential trace elements for soil bacteria), with pH adjusted to 7. Following incubation,

166 the colonies exhibiting zones of clearing were counted. Some colonies exhibiting large

167 and instant zones of clearing were isolated, further purified on the same media, sub

168 cultured on agar slants, and stored at 4 oC for further studies.

169

7 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

170 Plate assays for cellulose-hydrolyzing, nitrogen-fixing, and

171 phosphate-solubilizing activities

172 Plate assay for cellulase activity by cellulolytic isolates

173 Based on clear zones formation on cellulose-Congo red agar as modified from

174 Hendricks et al. [34], cellulolytic bacteria were isolated and evaluated for their

175 cellulose-hydrolyzing index values using same medium with following composition:

-1 -1 -1 176 K2HPO4, 0.50 g L ; MgSO4, 0.25 g L ; cellulose powder, 1.88 g L ; Congo red, 0.20 g

177 L-1; agar, 15.00 g L-1; gelatine, 2.00 g L-1; 100 mL of soil extract, and 900 mL of tap

178 water to supply essential trace elements. The medium was maintained at pH 7.

179 Following incubation at 30 °C for over 10 day-period, cellulolytic isolates were

180 assessed for their selective abilities index values based on calculation of ratio of total

181 diameter (colony + halo or clearing zone around colonies) and colony diameter [35].

182

183 Plate assay for nitrogen-fixation activity by nitrogen-fixing isolates

184 Nitrogen-fixing bacteria were cultured on N-free malate (Nfb) medium modified

185 from Döbereiner and Day [29] and incubated at 30 ºC. Bacterial colonies with blue halo

186 formation were isolated and characterized using the same media, after which blue halo

-1 187 index values were calculated. The Nfb medium consisted of: KH2PO4, 0.4 g L ;

-1 -1 -1 -1 188 K2HPO4, 0.1 g L ; MgSO4.7H2O, 0.2 g L ; NaCl, 0.1 g L ; CaCl2, 0.02 g L ; FeCl3,

-1 -1 -1 -1 189 0.01 g L ; Mo.O4Na.2H2O, 0.002 g L ; sodium malate, 5 g L ; agar, 15 g L ;

190 bromothymol blue (0.5% in alcohol solution), 5 mL L-1. The medium was adjusted to

191 pH 7. Following incubation at 30 °C for over 10 day-period, all N-fixing isolates were

192 assessed for their selective abilities index values based on calculation of ratio of total

193 diameter (colony + halo or clearing zone around colonies) and colony diameter [35].

8 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

194 Plate assay for phosphate-solubilization activity by phosphate-

195 solubilizing isolates

196 Phosphate-solubilizing bacterial isolation technique is based on the formation of

197 clear halos around the colonies capable to solubilize calcium phosphate [30-31]. Media

198 containing insoluble phosphate [32-33] with some modifications were used to isolate

199 and assess the phosphate-solubilizing bacterial activities. Its composition was: 50 mL L-

1 -1 -1 -1 200 of a salts solution (MgCl2·6H2O, 100 g L ; MgSO4·7H2O, 5 g L ; KCl, 4 g L ;

-1 -1 -1 -1 201 (NH4)2SO4, 2 g L ); agar, 15 g L ; 10 g L of sucrose as C source, and 2.5 g L of

202 tricalcium phosphate, Ca3(PO4)2, as an insoluble source of phosphate, and content was

203 adjusted to pH 7 before being autoclaved. Six mL L-1 of a pH indicator dye

204 (bromophenol blue, 0.4% ethanol solution) was also added to the media [30]. Following

205 incubation at 30 °C for over 10 day-period, phosphate-solubilizing bacterial isolates

206 were assessed for their selective abilities index values based on calculation of ratio of

207 total diameter (colony + halo or clearing zone around colonies) and colony diameter

208 [35].

209

210 Quantitative assays for functional activities

211 Quantitative assay for cellulase activity

212 For quantitative determination of cellulase activity, cellulolytic isolates were

213 previously inoculated in nutrient broth, shaken at 120 rpm at 30 °C for 16 hr and size of

6 7 214 the culture as preinoculum for cellulase assay was fixed to OD600=0.5 (10 to 10 cfu

215 mL-1). One mL of each isolate preinoculum was seeded in 40 mL cellulose broth as

216 basal medium containing cellulose as sole C source. The composition of cellulose broth

-1 -1 -1 217 per litre was: 5.0 g L cellulose microgranular, 2.5 g L NaNO3, 1.0 g L KH2PO4, 0.6

9 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

-1 -1 -1 -1 -1 218 g L MgSO4·7H2O, 0.1 g L NaCl, 0.1 g L CaCl2·6H2O, 0.01 g L FeCl3, 2.0 g L

219 gelatin, and 0.1 g L-1 yeast extract with pH 7 adjusted using diluted NaOH as modified

220 from Lu et al. [36]. Inoculated broth was incubated at 30 °C under agitation at 180 rpm

221 and CMCase activity was analyzed for up to 5 days. For CMCase activity, 0.5 mL

222 culture filtrate from broth as crude enzyme source, or supernatant cellulase, also

223 specified as CMCase, was incubated with 1 mL of 1% carboxymethylcellulose in 0.1 M

224 citrate buffer, pH 5 at 40 °C in agitation for 30 min. The amount of reducing sugars

225 (glucose) was measured using Somogyi-Nelson method [37-38] at 520 nm. One unit of

226 enzyme activity (IU mL-1) is equivalent to the amount of enzyme per mL culture filtrate

227 that liberates 1 µg of glucose per minute [39].

228

229 Quantitative assay for in vitro biological nitrogen-fixation activity

230 The procedure for quantitative estimation of BNF activity was modified from

231 those described by Soares et al. [40]. Each N-fixing isolate was grown as a preinoculum

232 in nutrient broth under agitation at 120 rpm at 30 oC for 16 hr. Optical density was used

6 7 233 to adjust the inoculum size [OD600= 0.5 is equivalent to 10 to 10 colony forming units

234 (cfu) mL-1]. Inoculum was transferred (100 µL) to new media, NFb broth (containing

235 similar composition as Nfb medium except without agar). The inoculated media were

o 236 incubated at 30 C under constant agitation for 72 hr for N2 fixation quantification. A

237 control medium was also included by using 100 µL nutrient broth as inoculant. The

238 amount of N2 fixation was measured by sulfur digestion and distillation with NaOH 10

239 mol L-1, as described by Bremner and Keeney [41] and done in triplicates. The samples

240 were digested with 2 mL concentrated H2SO4 (d = 1.84), 1 mL H2O2 30% and 0.7 g

o 241 digestion mixture (100 g Na2SO4 + 10 g CuSO4.5H2O + 1 g selenium) at 350–375 C.

242 The products of digestion were distillated with 10 mL NaOH 10 mol L-1, and the

10 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

243 ammonia trapped in 10 mL of 2% boric acid was measured by titration with H2SO4

244 0.0025 mol L-1.

245

246 Quantitative assay for inorganic phosphate-solubilization activity

247 Quantitative estimation of inorganic phosphate-solubilization was done with

248 some modifications from Delvasto et al. [42]. Preinoculum cultures were previously

249 prepared by inoculating phosphate-solubilizing isolate in nutrient broth under agitation

6 7 250 at 120 rpm at 30 °C for 16 hr. Inoculum size was adjusted to OD600=0.5 (10 to 10 cfu

251 mL-1). Prior to the assay, preinoculum cultures were washed 3 times and resuspended

252 using saline solution 0.85% NaCl, and one mL isolate suspension was inoculated into

253 100 mL liquid medium containing insoluble calcium phosphate as in solid medium,

254 except without agar. Inoculated broth was incubated at 30 °C under agitation at 180 rpm

255 for up to 5 days. Saline solution without inoculant in the broth was left as control.

256 Soluble P content in culture filtrates was determined by the blue

257 method [43]. Appropriate amount of aliquots from samples containing 1-50 µg mL-1 P

258 were pipetted into 50 mL volumetric flasks containing approximately 15 mL distilled

259 water and 8 mL ascorbic acid antimony sulfomolybdo working reagent (as color

260 developing solution to form blue color complex in the present of P from samples and/or

261 standard solutions), swirled to mix, and filled to the mark with distilled water. A series

262 of standard P solutions (KH2PO4) aliquots were pipetted into a set of volumetric flasks

263 containing the same amount of color developing solutions to make up the range of

264 expected P concentrations in the culture filtrates, swirled to mix, and made up to the

265 mark with distilled water. A blank solution equal to the aliquot size of sample was also

266 included to each flask of standard solutions. A standard curve was plotted using UV-vis

267 spectrophotometer (Unico) at 882 nm. The absorbance of samples were measured by

11 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

268 means of the standard curve using the same wavelength and converted into P

269 concentrations. Values of soluble P are expressed as µg mL-1. The final pH of liquid

270 medium was measured using a pH meter. All values of pH and P concentrations in

271 samples were measured in triplicates.

272

273 Morphological and physiological characterization of selected

274 isolates

275 Selected isolates were each studied for morphological properties, which

276 included color, size, and colony characteristics such as form, margin, and elevation on

277 nutrient agar. The spore staining procedure, done according to Schaeffer-Fulton method

278 and standard Gram staining procedure were conducted as described by Cappuccino and

279 Sherman [44], both with some modification on fixation, whereby bacterial smears were

280 fixed using methanol, without heat fixation. For each procedure, the smears were finally

281 examined under oil immersion at total magnification power of 1000 times. The

282 morphology of each bacterial colony such as the color, shape, and arrangement of the

283 stained cells were recorded. Any presence of endospores for each studied bacterial

284 isolate was also noted.

285 The biochemical tests consisting of several enzymatic, physiological activities,

286 or differential growth tests were carried out. Citrate utilization was determined on

287 Simmons citrate agar; hydrogen sulfide production, along with lactose and glucose

288 fermentation were determined on Kligler Iron Agar; Gram negative strains and lactose

289 fermentation were detected using Mac Conkey agar; gelatinase production, starch

290 , casein proteolysis, lipase production, and indole production were each

291 determined using nutrient gelatine, starch agar, skim milk agar, Tween 80 agar and

292 BD’s Indole Reagent Droppers (modified Kovacs’ reagent) via standard tube method,

12 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

293 respectively. Catalase test was done using 3% hydrogen peroxide solution. test

294 was conducted on Biolife’s Motility Medium whereas nitrate reduction test was done

295 using the test kit by Fluka. The aforementioned physiological and biochemical test

296 results were processed and compared with the results of systematic bacteriology

297 identification according to Bergey’s Manual of Determinative Bacteriology and

298 Bergey’s Manual of Systematic Bacteriology [45-46].

299

300 Molecular method for bacterial identification

301 Deoxyribonucleic acid extraction

302 Bacterial genomic DNA was extracted using sodium dodecyl sulfate (SDS)

303 method. Bacterial inoculum from streaked culture plate overnight was recultured

304 overnight in 1.5 mL nutrient broth in capped tubes with one control (broth without

305 inoculum) in the water bath shaker at 30 oC. The 1.5 mL suspension was

306 centrifuged at 13040 g for 3 min at 4 oC. After removing the supernatant, the cells were

307 resuspended or washed with 180 µL TE buffer (10 mM Tris, 1mM EDTA, pH 8.0) by

308 vortex-mixing and incubated on ice for 5 min. Seventy five µL of 2% SDS solution was

309 added, followed by invert-mixing step to lyse the bacterial cells and incubation on ice

310 for 5-10 min. Then 250 µL of 3 M potassium acetate was added, followed by gentle

311 invert-mixing, and incubation on ice for 5 min. The sample was subsequently

312 centrifuged at 13040 g for 5 min at 4 oC, with special attention on the cells pellet

313 formation at the bottom of the tube. The supernatant containing bacterial DNA was

314 carefully transferred to a clean 1.5 mL tube on ice. Absolute ethanol was then added at

315 approximately two times the volume of the supernatant on the same tube, invert-mixed,

316 and followed by centrifugation at 12000 g for 10 min at 4 oC.

13 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

317 Later, supernatant was discarded and 1 mL of 75% ethanol was added to the

318 DNA pellet for washing step. DNA pellet was resuspended by pipette-mixing the

319 suspension in repeated gentle motion, followed by centrifugation at 12000 g for 10 min

320 at 4 oC. The supernatant was discarded with extra care taken at this stage so as not to

321 disrupt the DNA pellet while at the same time ensuring the entire removal of ethanol

322 residues from the pellet and within the tube itself. The pellet was air-dried for about 10

323 to 20 min, till no ethanol residue is visible. Finally, 20 µL of TE buffer (10 mM Tris,

324 1mM EDTA, pH 8.0) was added, pipette-mixed gently, and stored at -20 oC.

325

326 Polymerase chain reaction amplification of extracted DNA

327 Amplification of DNA extraction products was performed in a final volume of

328 25 µL each. For each reaction, 1.25 µL 0.5 µM for each of both primers (16S-F and

329 16S-R1), 2.5 µL 25mM MgCl2, 2.5 µL 10X reaction buffer [(NH4)2SO4], 0.5 µL

330 working stock dNTPs or dNTP mix combined to give a final concentration of 0.2 mM

331 each, 0.25 µL of Taq DNA Polymerase (5 U/µL), 15.75 µL of sterilized distilled water

332 and 1 µL of template DNA. Primers 16S-F and 16S-R1 had the sequences of 5’-

333 AAGAGTTTGATCATGGCTCA-3’ and 5’-

334 TAAGGAGGTGATCCAACCGCAGGTTC-3’, respectively. The PCR was performed

335 in a thermal cycler (Bioer XP Cycler) using cycling conditions consisting of an initial

336 denaturation at 95 oC for 3 min, followed by 35 cycles consisting of denaturation at 94

337 oC for 30 s, annealing at 50 oC for 30 s, and extension at 72 oC for 1 min. A final

338 extension was performed at 72 oC for 5 min. A blank that contained all the components

339 of the reaction mixture without the DNA template was included as a control. The PCR

340 products were analyzed by running on 1% agarose gel electrophoresis.

341

14 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

342 Agarose gel electrophoresis

343 A 1% (w/v) agarose gel was prepared by mixing 0.25 g agarose and 25 mL of

344 1X TAE buffer. The solution was heated gently in a microwave oven until the agarose

345 was fully dissolved. The gel was allowed to cool to the touch before being poured to the

346 gel cask with the well comb in place. The gel was allowed to solidify for 20-30 minutes

347 before removing the comb and mould barriers, placed onto electrophoresis tank, and 1X

348 TAE buffer solution was poured evenly into the electrophoresis tank until it just

349 submerged the surface of the gel. The samples were each prepared by combining 5 µL

350 of PCR products with 3 µL of 6x loading dye and 1 µL of EZ Vision dye before being

351 loaded into the performed wells. One µL of DNA marker or ladder bearing 1.5 kb size

352 was mixed well with 2 µL of 6x loading dye, 1 µL of EZ Vision dye, and 5 µL of 1x

353 TAE buffer solution before being loaded into assigned well. The gel was then run at 90

354 V for about an hour. The EZ Vision dye-stained fragments were visualized and

355 photographed with UV light in a gel imager (AlphaInnotech Fluorchem). Any PCR

356 bands obtained for the specific isolates were recorded and the gel image was captured.

357 Once PCR bands were sighted on the gel image, preparations of the samples were done

358 which included PCR scale-up to 50 µL per reaction, and purification of PCR products

359 prior to sending them out for outsourced sequencing service.

360

361 Purification of PCR products for sequencing

362 Direct precipitation is a general method used for purification of PCR products.

363 Fifty µL of PCR product was transferred to a 1.5 mL tube. Sodium acetate (pH 8) at 0.1

364 volume or equivalent to 4 µL was added into the tube, followed by absolute ethanol

365 (99%) at 2.5 volume or equivalent to 100 µL. The tube was inverted a few times for

366 homogeneous mixing and kept in ice for 30 min. The tube was centrifuged at 13,040 g

15 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

367 for 15 min at 4 oC. Supernatant was decanted and 200 µL of 75% ethanol was added

368 into the tube, followed by vortex-mixing for 1 min. The tube was centrifuged at 13,040

369 g for 5 min at 4 oC. The pellet was air-dried until no traces of ethanol could be seen.

370 Fifteen µL of sterile distilled water was added to resuspend the pellet by tapping. The

371 concentrated PCR product was short-spun and ready to be sent for sequencing. The

372 quantification or estimation of the concentration of the concentrated PCR products was

373 done by using the gel imager software (FluorChem 5500) which was within the range of

374 5 to 100 ng per µL.

375

376 Sequencing of 16S rDNA

377 Determination of the 16S rDNA sequences was done by transferring

378 the purified PCR products containing the concentrated amplified DNA, along with

379 primer R2 to First BASE Laboratories Sdn Bhd, a life sciences-based research company

380 located in Peninsular Malaysia.

381

382 Identification of bacterial sequences

383 The sequences obtained for selected bacterial isolates were scanned manually

384 for any errors on the bases interpretation provided by the base sequencing service which

385 consisted primarily of adenine (A), (G), (C), and thymine (T) bases.

386 The sequence bases were analyzed using Sequence Scanner Software v1.0 which was

387 registered under Applied Biosystems®. Each of the corrected sequences was searched

388 on BLAST website which compared nucleotide or sequences to sequence

389 databases and calculated statistical significance of matches. The BLAST website was

390 available at http://blast.ncbi.nlm.nih.gov/Blast.cgi. The closest matches of top-scoring

391 sequences from the BLAST databases were used to presume the identity or closest

16 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

392 identities of the particular bacteria. In the case of multiple identities produced based on

393 sequence matches alone, biochemical properties offered phenotypic information

394 necessary for further comparison in order to finalize each isolate to only one most

395 presumed identity.

396

397 Plant growth-promoting assays

398 Germination assays of maize and green gram seeds

399 Phytotoxic assay for 15 selected bacterial inoculants on green gram seeds were

400 done via germination check of both root and shoot of the germinating seeds prior to pot

401 experiment. Seeds of green gram (Vigna radiata) were obtained from a local market.

402 Prior to the assay, the seeds were sorted for uniformity of size, and all visibly damaged

403 seeds were discarded. The remaining seeds were washed with tap water and surface

404 sterilized in 0.1% sodium hypochlorite solution for 5 min before rinsing at least three

405 times with sterile distilled water to thoroughly remove sterilants. The seeds were then

406 soaked in broth containing 16 hr cultured bacterial isolates according to treatments for 2

407 hr at 30 °C. The seed treatments consisted of 15 selected bacterial isolates. Two control

408 treatments were included whereby first control was designated as seeds treated with

409 sterile distilled water, and another control treatment with sterile nutrient broth.

410 The treated green gram seeds were placed on petri dishes containing sterilized

411 moist sheets of Whatman filter paper per treatment. Each treatment was based on 100

412 seeds of each plant . The seeds were incubated for up to 7 days at about 30 °C and

413 moisture content of growth support was maintained with an equal volume of autoclaved

414 distilled water on daily basis. The number of germinated green gram seeds was firstly

415 counted on third and fourth day of incubation, respectively, and followed by second

416 counting and measurement of shoots and roots or radicle length for both plants on

17 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

417 seventh day. Percentage of germinated seeds and vigor index were calculated. Vigor

418 index was calculated as the sum of mean shoot and root length multiplied by percentage

419 of germination [47]. All experimental sets and apparatus used were sterilized where

420 necessary.

421

422 Indole-3-acetic acid production assay

423 Some selected bacterial isolates were subjected to in vitro colorimetric

424 determination of IAA production using Salkowski’s reagent (2% of 0.5 M FeCl3 in 35%

425 perchloric acid) as modified from Asghar et al. [48]. Based on seed germination test

426 results, 7 bacterial isolates were selected for IAA production assay. Each selected

427 isolate was grown to exponential growth phase (OD ≥ 1.0) within 16 hr to be used as

6 7 428 pre-inoculum for the bioassay. The inoculum was adjusted to OD600=0.5 (10 to 10 cfu

429 mL-1) before inoculation at 500 µL into 50 mL of nutrient broth without addition of

430 tryptophan and incubated at 30 °C under agitation at 100 rpm for up to 48 hr. Non

431 inoculated broth was included as control. Each culture was centrifuged at 10,000 rpm

432 for 10 min at 4 °C.

433 One millilitre of culture supernatant was mixed with 2 mL of Salkowski’s

434 reagent in a test tube and kept in the dark for 20-25 min before measuring absorbance

435 using a UV-Vis spectrophotometer at 535 nm. The assay for each isolate was conducted

436 in triplicates. Prior to samples measurement, a series of IAA standards prepared within

437 the range of 0 to 100 µg mL-1 reacted with Salkowski’s reagent were measured to

438 construct a standard linear graph as a reference to the samples’ IAA concentration

439 values. Any appearance of reddish to pinkish color in the solution indicated IAA

440 presence in the cultured medium and depending on the color intensity of the test

18 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

441 solutions, coloring responses were classified into two; pink type (pale pink to deep

442 pink), and red type (dark red to red).

443

444 Data analysis of bioassays 445 Analysis of Variance (ANOVA) was used to test for significance differences of

446 measurements of each bioassay whereas Tukey’s Test was employed for the

447 significance difference among different treatments at p ≥ 0.05. The Statistical Analysis

448 System (SAS Ver. 9.2) was used for the analysis.

449

450 Results and Discussion

451 Functional activity assays of respective isolates

452 In this study, 14 out of 62 cellulolytic, 12 out of 39 N-fixing, and 8 out of 36

453 phosphate-solubilizing bacterial isolates obtained from of both planted and

454 indigenous tree species at the rehabilitated forest were screened based on their

455 functional activities on their respective selective media and distinct morphology for

456 further characterization. Among these 3 functional groups, phosphate-solubilizing group

457 had the least number of characterized isolates due to the fact that phosphate-solubilizing

458 bacterial isolates tend to lose their ability after successive transfers or subcultures on

459 agar medium as reported previously [49]. Diazotrophic isolates also encountered similar

460 phenomenon whereby upon successive cultivation, some isolates failed to develop blue

461 halo or unable to grow on N-free media. da Silva et al. [50] has reported that a

462 diazotrophic isolate lost its characteristic pellicle growth upon successive cultures

463 which made its activity assessment not possible. However, none of the

464 cellulolytic isolates lost their functional ability.

19 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

465 Cellulolytic microorganisms play major role in the synthesis of cellulases crucial

466 for the breakdown of complex polysaccharides into simple sugars. There was no

467 specific correlation between cellulose-hydrolyzing index and cellulase (CMCase)

468 activity for the 14 tested cellulolytic strains (Table 1).

469 Table 1. Cellulose-hydrolyzing index and cellulase (CMCase) activities (U mL-1 470 cell-free culture broth after 5 days at 30 °C, 180 rpm, and medium pH 7) of 471 cellulolytic isolates.

Activity Index Cellulase Activity (U mL-1) Isolates Mean value ± S.E. C14 5.25bc ± 1.50 0.016f ± 0.0015 C39d 5.10bc ± 0.44 0.015f ± 0.0017 C41d 5.29bc ± 0.88 0.078b ± 0.0012 C42d 8.52ab ± 1.93 0.077b ± 0.0007 C46d 10.41a ± 0.43 0.089a ± 0.0009 C22b 10.31a ± 0.62 0.035e ± 0.0006 C34c 4.57c ± 3.24 0.047d ± 0.0010 C45d 5.93bc ± 1.23 0.059c ± 0.0012 C4 3.59c ± 0.14 0.079b ± 0.0015 C25b 3.70c ± 0.79 0.078b ± 0.0012 C29b 5.57bc ± 0.70 0.080b ± 0.0007 C40d 4.63c ± 0.61 0.077b ± 0.0007 C43d 4.60c± 2.26 0.078b ± 0.0009 CB15 5.54bc ± 0.36 0.060c ± 0.0010 472 S.E. is standard error of the mean. 473 Means with same letter are not significantly different at p ≥ 0.05 within columns using 474 Tukey’s test. 475

476 Among the 14 cellulolytic isolates assessed, isolates C46d and C42d consistently

477 showed higher cellulolytic activity based on both cellulose-hydrolyzing index (10.41

478 and 8.52, respectively), and cellulase activity (0.089 U mL-1 and 0.077 U mL-1,

479 respectively). The lowest cellulose-hydrolyzing index and cellulose activity were in the

480 range of 1.33-3.73 U mL-1 and 0.015-0.016 U mL-1, respectively. The cellulase

481 (CMCase) activities of the cellulolytic isolates in this present study were comparable to

482 those of bacterial strains isolated from soil at different culture conditions as reported by

20 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

483 Sethi et al. [51]. The cellulase (CMCase) activity demonstrated by C46d was especially

484 higher than those reported for Bacillus amyloliquefaciens SS35 [52] and Bacillus

485 pumilus EB3 [53] at 0.079 U mL-1.

486 For nitrogenase activity assessment, there was positive relationship (r=0.63, p ≤

487 0.05) between N-fixing index values and BNF amount. Among the 12 diazotrophic

488 isolates evaluated, NB1 exhibited the most active N-fixing activity as demonstrated by

489 the consistent highest means of both activity index and nitrogenase activity

490 quantification at 10.82 and 2.21 µg N mL-1 day-1, respectively (Table 2).

491 Table 2. Nitrogen-fixation index and biological nitrogen-fixation quantification of 492 diazotrophic isolates.

Activity Index Quantitative activity (µg N mL-1 day-1) Isolate Mean value ± S.E.

NB1 10.82a ± 0.26 2.21a ± 0.34 NB4 1.84e ± 0.05 1.11cd ± 0.27 NB5 2.03e ± 0.06 0.65cd ± 0.00 NB7 1.97e ± 0.05 1.23bc ± 0.24 NB8 1.80e ± 0.05 0.47d ± 0.00 NC9 4.63c ± 0.12 0.84cd ± 0.00 NB10 2.35e ± 0.32 1.01cd ± 0.19 NC11 1.95e ± 0.04 0.44d ± 0.10 NB11 1.88e ± 0.07 1.87ab ± 0.34 N20c 5.92b ± 0.46 1.24bc ± 0.30 N27d 3.61d ± 0.21 0.73cd ± 0.00 N29d 1.94e ± 0.08 0.60cd ± 0.00 493 S.E. is standard error of the mean. 494 Means with same letter are not significantly different at p ≥ 0.05 within columns using 495 Tukey’s test. 496

497 NB11 also presented comparably high N-fixing activity at 1.87 µg N mL-1 day-1

498 although producing fairly small halo on plate assay, yielding mean index value of 1.88.

499 Strain N20c demonstrated second highest mean N-fixing index (5.92) but at fairly

500 moderate amount of N-fixing activity (1.24 µg N mL-1 day-1). The lowest N-fixing

501 activity was recorded in the range values of 0.34-1.38 µg N mL-1 day-1 by the rest of

21 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

502 assessed isolates. Plate assays revealed lowest N-fixing index values in the range of

503 1.75-2.67. Although there was lack of reports on N-fixing activity (index) evaluation

504 based on plate assay, the quantitative values were within range of other reported

505 diazotrophic strains [40]. The ability of diazotrophic bacteria to convert atmospheric N2

506 into ammonia, the available form of N to be used by living organisms for is

507 undeniably of vital importance to soil system especially through plants either through

508 direct influence on plant growth, and indirectly through production of phytohormones

509 that boost plant growth [50, 54].

510 Being one of the most limited but important plant growth nutrient, P mainly

511 occurs in insoluble forms, whereby phosphate-solubilizing rhizobacteria plays a crucial

512 role in converting it to available forms [55]. Among the 8 phosphate-solubilizing

513 strains, PB3 and PB5 showed superior phosphate-solubilization index values in plate

514 assay (4.21 and 4.10, respectively) whereas in enzymatic assay, the highest activity was

515 demonstrated by isolate PC8 (307.7 µg mL-1) (Table 3).

516 Table 3. Phosphate-solubilization index and quantitative phosphate-solubilization 517 estimation based on soluble P by phosphate-solubilizing isolates.

Quantitative Phosphate-solubilizing phosphate- Final index solubilization (Soluble mean pH Isolate P in µg mL-1) in broth Mean value ± S.E. culture

PB5 4.21a ± 0.22 96.00d ± 2.65 4.27 PC1 2.38cd ± 0.15 9.14f ± 0.75 4.93 PB7 3.18b ± 0.23 159.00c ± 1.53 4.15 PB3 4.10a ± 0.23 209.17b ± 8.97 4.11 PC6 1.85ef ± 0.05 31.53e ± 0.39 4.63 PC8 2.74bc ± 0.22 307.67a ± 8.41 4.04 PB11 1.66f ± 0.08 3.25f ± 0.52 4.57 PB6 2.19de ± 0.24 12.67f ± 0.55 4.20 518 S.E. is standard error of the mean. 519 Means with same letter are not significantly different at p ≥ 0.05 within columns using 520 Tukey’s test. 521

22 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

522 The amount of phosphate-solubilization index and activities with final broth

523 culture pH reported in this present study is comparable to the range reported by Pandey

524 et al. [56], Gulati et al. [57], Song et al. [58]. Isolate PC8 showed higher phosphate-

525 solubilization activity (307.67 µg mL-1) as compared with Pseudomonas putida (B0) at

526 maximum activity of 247 µg mL-1 [56]. There was no significant relationship observed

527 between phosphate-solubilizing index and soluble phosphate. Similar observations had

528 been previously reported by Nautiyal [31] and Chang and Yang [59]. Nautiyal [31]

529 stated the inability of phosphate-solubilizing Pseudomonas aerogenes and

530 Pseudomonas sp.1 to produce halo on solid media but could solubilize substantial

531 amount of P in different type of broth cultures. Nonetheless, there was a significant

532 inverse correlation (r=-0.74, p ≤ 0.05) between soluble phosphate and pH of culture

533 broth, and this observation corroborates previous findings [60-61]. The pH of broth

534 declined to as low 4.04 as observed in PC8 activity from initial pH 7. This observation

535 is consistent with earlier finding where there was a drop in pH due to bacterial activity

536 is accompanied with mineral phosphate-solubilization and increment of its activity

537 efficiency [56]. In their study on various organic acids secretion by 10 bacteria and 3

538 fungal strains, Rashid et al. [62] reported lowering of pH as part of the mechanism of

539 microbial phosphate-solubilization.

540

541 Cultural, biochemical, and molecular identification of selected

542 functional bacterial isolates

543 The isolation of bacteria near the root zones of vegetations thriving in the

544 rehabilitated forest generally to the recovery of putative beneficial rhizospheric or

545 even free-living functional bacteria. In this present study, growth in the selective media

546 proved to be a successful strategy to select bacteria based on their phenotypic activities

23 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

547 on respective media. From the isolates assessed for their functional activities, 5 isolates

548 of each functional group were selected to be further characterized on the basis of

549 distinct colony morphology and high functional activities. After being selected based on

550 these combined criteria, the isolates were grouped by morphological similarities and

551 phenotypic characteristics.

552 On the basis of morphological characteristics of the 15 selected isolates depicted

553 in Table 4, they exhibited varying properties. It was observed that the majority or 13

554 isolates exhibited round colonies in configuration or form; 10 colonies were convex

555 whereas the rest shown raised elevations; 14 colonies were opaque in density; colonies

556 of all of the isolates were almost creamy or yellowish in colour; 13 showed Gram

557 negative reactions; all isolates were rods or short rods in single or in pairs arrangement;

558 13 without endospore production; and they were mostly motile as exhibited by 11

559 isolates. The following morphological properties presented more diverse characteristics:

560 colony margin was represented by 8 entires, 3 undulates and serrates, respectively, and

561 1 lobate; colony size was represented by 6 moderates, 5 smalls, 3 larges and 1 pinpoint;

562 and the shapes of endospores produced by 2 isolates were of rods and ,

563 respectively.

564

24 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

565 Table 4. Morphological characteristics of selected functional isolates.

Colony Isolate morphology CB15 C22d C29b C41d C46d NB1 NB10 NB11 N27d NC9 PB3 PB5 PB6 PB7 PC8

Configuration Irregular Round Round Irregula Round Round Round Round Round Round Round Round Round Round Round / Form r Undulat Entire Entire Serrate Entire Entire Undulat Undulate Entire Entire Entire Serrate Serrate Entire Lobate Margin e e

Raised Convex Convex Raised Conve Convex Raised Convex Convex Convex Convex Raised Raised Convex Convex Elevation x Opaque Opaque Opaque Opaque Opaqu Opaque Opaque Opaque Opaque Opaque Opaque Opaque Transparen Opaque Opaque Density e t Yellowis Yellowis Pigments Cream Cream Yellow Cream Cream Cream Cream Cream h cream Cream h cream Cream Cream Cream Cream Gram’s + ------+ - - reaction Rods Rods Rods Short Short Short Short Rods Short Short rods Rods Rods Short rods Rods Short rods Shape rods rods rods rods rods Large Moderat Small Large Small Pinpoin Large Moderat Small Moderate Moderate Moderate Moderate Small Small Size e t e Single Single Single/pair Single Single Single Single Single Single Single/pair Single Single/pair Single/pair Single/pair Single/pair Arrangement s s s s s s Endospore + ------+ ------production Endospore Rods ------Coccus ------shape Motility + + + - + - + - - + + + + + + 566 +, positive result; -, negative result.

25 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

567 The isolates were primarily identified based on molecular method, and in cases

568 of multiple identities produced based on BLAST generated database, morphological

569 (Table 4) and biochemical (Table 5) properties facilitated in the downsizing of

570 identification task.

571 Table 5. Biochemical properties of selected functional bacterial isolates.

Isolate Catalase MCA KIA SA SCA SMA T80A NG Indole NRT CB15 + ˗ Glu ferm +, Lacferm - ˗ ˗ + + + ˗ + C22d + + Glu ferm +, Lacferm - ˗ + + + + ˗ + C29b + + Glu ferm -, Lacferm - ˗ + + + + ˗ + C41d + ˗ Glu ferm +, Lacferm - ˗ ˗ + ˗ + ˗ + C46d + + Glu ferm +, Lacferm - ˗ + + + + ˗ + NB1 + + Glu ferm -, Lacferm - ˗ + ˗ + ˗ ˗ + NB10 + ˗ Glu ferm +, Lacferm - + + + + + ˗ ˗ NB11 + ˗ Glu ferm -, Lacferm + + ˗ ˗ + ˗ ˗ + N27d + + Glu ferm -, Lacferm - ˗ + + ˗ + ˗ + NC9 + + Glu ferm +, Lacferm - ˗ + + + + ˗ + PB3 + + Glu ferm -, Lacferm - ˗ + + + + ˗ + PB5 + ˗ Glu ferm -, Lacferm - + + + ˗ + ˗ + PB6 + ˗ Glu ferm -, Lacferm - + + + ˗ + ˗ + PB7 + + Glu ferm -, Lacferm - ˗ + + + + ˗ + PC8 + + Glu ferm -, Lacferm - ˗ + + + + ˗ + 572 +, positive result; -, negative result; MCA, Mac Conkey agar assay; KIA, Kligler Iron 573 Agar assay; Glu ferm, Glucose fermentation; Lacferm, Lactose fermentation; SA, 574 Starch agar assay; SCA, Simmons citrate agar; SMA, Skim milk agar; T80A, Tween 80 575 Agar assay; NG, Nutrient gelatine assay; Indole, Indole production assay; NRT, Nitrate 576 reduction test assay. 577

578 In Table 6, the selected cellulolytic isolates in this present study were identified

579 as Serratia nematodiphila strain SP6 (C46d), Serratia marcescens subspecies sakuensis

580 isolate PSB23 (C22b), Stenotrophomonas maltophilia strain KNUC605 (C29b),

581 Bacillus thuringiensis (Bt) (CB15), and Stenotrophomonas sp. Ellin162 (C41d). The N-

582 fixing isolates were consisted of Burkholderia nodosa strain Br3470 (NB1), Delftia

583 lacustris strain 332 (NC9), Stenotrophomonas sp. strain SeaH-As3w (N27d), Ralstonia

584 sp. strain MCT1 (NB10), and basilensis strain CCUG 49340T (NB11). The

26 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

585 phosphate-solubilizing isolates were represented by Burkholderia cepacia strain 5.5B

586 (PB5), Burkholderia cepacia strain ATCC 35254 (PB3), Burkholderia uboniae (PB7),

587 Burkholderia cepacia strain 2EJ5 (PC8), and Gluconacetobacter diazotrophicus PAl 5

588 (PB6).

589 Table 6. Identities of cellulolytic, N-fixing, and phosphate-solubilizing bacterial 590 isolates based on 16S rDNA analysis and BLAST databases.

Similarity Accession Phenotype Isolate Closest Identity (%) No. Serratia nematodiphila strain 98 C46d JN315887.1 SP6 (324/329) Serratia marcescens subspecies 97 C22b HQ242736.1 sakuensis isolate PSB23 (578/594) Stenotrophomonas maltophilia 99 Cellulolytic C29b HM047520.1 strain KNUC605 (945/949) 99 CB15 Bacillus thuringiensis FR846529.1 (1018/1021) 100 C41d Stenotrophomonas sp. Ellin162 AF409004.1 (989/989) Burkholderia nodosa strain 98 NB1 AM284972.1 Br3470 (943/967) 99 NC9 Delftia lacustris strain 332 EU888308.1 (1009/1014) Stenotrophomonas sp. SeaH- 100 N-fixing N27d FJ607350.1 As3w (1029/1029) 99 NB10 Ralstonia sp. MCT1 DQ232889.1 (999/1007) Cupriavidus basilensis strain 99 NB11 FN597608.1 CCUG 49340T (961/969) 99 PB5 Burkholderia cepacia strain 5.5B FJ652617.1 (874/885) Burkholderia cepacia strain 99 PB3 AY741346.1 ATCC 35254 (990/993) Phosphate- 99 PB7 Burkholderia uboniae AB030584.1 solubilizing (993/997) 99 PC8 Burkholderia cepacia strain 2EJ5 GQ383907.1 (1008/1016) Gluconacetobacter 97 PB6 CP001189.1 diazotrophicus PAl 5 (778/801) 591

592 These identities were matched with BLAST databases with the range of 97 to

593 100% base sequence similarities (Table 6). It was observed that identification of

27 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

594 bacterial isolates over the three phenotypic groups revealed the predominance of Gram-

595 negative over Gram-positive bacteria in the rehabilitated forest soils. The Gram-positive

596 bacteria belong to only one taxonomic grouping, Firmicutes, represented by the genus

597 Bacillus, whereas Gram-negative bacteria included Alpha- with genus

598 Gluconacetobacter; Beta-Proteobacteria represented by genera Burkholderia, Delftia,

599 Ralstonia, and/or Cupriavidus; and Gamma-Proteobacteria by genera Serratia and

600 Stenotrophomonas.

601 From the overall identification results of each functional group of isolates,

602 bacteria from genus Burkholderia seemed to predominate in these tropical forest soils at

603 one third of total identified isolates with distinguishing phosphate-solubilizing and N-

604 fixing capabilities over other genera. Stenotrophomonas genus ranked second in

605 abundance to Burkholderia. Burkholderia genus is a well-known classification of

606 bacteria wherein its members possess vast versatility in niches and nutrition regulation

607 such as N-fixing and phosphate-solubilization, as well as possessing numerous PGP

608 attributes. Burkholderia species occur in the natural environment, even though some

609 strains are also found in clinical samples, especially the B. cepacia complex species

610 from patients with cystic fibrosis [63].

611 Based on our 16S rDNA sequences analysis results, the top 4 isolates with

612 highest phosphate-solubilizing and N-fixing activities belong to genus Burkholderia, in

613 which 3 of them were most closely related to B. cepacia (99% similarity) and B. nodosa

614 (98% similarity), respectively. Serratia marcescens (of close similarity by C22b at

615 97%) or Serratia sp. (with C46d closely matched to Serratia nematodiphila by 98%) are

616 known to have both cellulolytic and phosphate-solubilizing abilities [64-65]. Similar to

617 the results of this present study, Stenotrophomonas spp. have been reported previously

618 to possess both N-fixing [66] and cellulose-hydrolyzing abilities [67].

28 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

619 Stenotrophomonas species commonly prevail in nature, whereby S. maltophilia and S.

620 rhizophila are reported to occur as rhizospheric and endophytic bacteria accountable for

621 beneficial plant interactions [67]. The N-fixing isolates in this present study displayed

622 higher diversity in terms of the species, as well as activities.

623 Diazotrophic isolate N27d revealed 100% matching identity with

624 Stenotrophomonas sp. (strain SeaH-As3w) and it is believed to have cellolulytic

625 potential similar to the C41d and C29b isolates. Previously, many studies have reported

626 bacterial belonging to genus Burkholderia especially B. cepacia strains as having

627 biocontrol ability against many phyto-pathogenic fungi such as Fusarium sp., Pythium

628 aphanidermatum, Pythium ultimum, Phytophthora capsici, Aspergillus flavus, Botrytis

629 cinerea and Rhizoctonia solani [68-72] as part of its main contributing PGP factors.

630 The phosphate-solubilizing bacteria showed the lowest diversity that only three

631 species identities from two genera were recorded among the five identified isolates.

632 Isolates PB3, PB5, PB7, and PC8 were closely matched (99% similarity) to

633 Burkholderia sp. These isolates together with the N-fixing NB1 might potentially carry

634 both the N-fixing and phosphate-solubilizing activities [73-74]. Diazotrophic isolate

635 Burkholderia nodosa NB1 along with Burkholderia mimosarum, have been the latest

636 classified novel species with N-fixing capacity and legume-nodulating bacteria that

637 establish effective symbioses with legumes of Mimosa species [75-76]. All

638 Burkholderia strains in this present study demonstrated variability in cultural and

639 biochemical properties, which is typical for members of this genus.

640 The ability of Burkholderia strains to solubilize inorganic phosphates by various

641 studies [77-80] have been well documented and nevertheless, a beneficial property in

642 plant- growth promotion. Among the Burkholderia strains isolated in this study, this

643 trait seemed most prevalent. Few Burkholderia strains are known for other PGP ability,

29 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

644 which is to fix N2, either as free-living bacteria [81], or even nodulating ones that also

645 fix N2 on media including Burkholderia nodosa strain as reported by da Silva et al. [82].

646 The genus Delftia (of close similarity by NC9 to Delftia lacustris at 99%)

647 belongs to a newly grouped genus closely linked to Comamonas, in which Delftia

648 terephthalate has been reported as a diazotrophic PGPR capable of degrading

649 terephthalate, as well as antagonizing rice blast and blight causing Xanthomonas oryzae,

650 Rhizoctonia solani, and Pyricularia oryzae [83-84]. One of the diazotrophic bacterium

651 IHB B 4037 showing closely matched identity with Delftia lacustris isolated from

652 rhizosphere of tea plant in India showed the highest nitrogenase activity in a study

653 conducted by Gulati et al. [85]. Delftia tsuruhatensis was also revealed to be a

654 with biological control ability against several plant pathogens especially the

655 three main rice pathogens in the likes of Xanthomonas oryzae pv. Oryzae, Rhizoctonia

656 solani, and Pyricularia oryzae Cavara [86]. In summary, the variability among the

657 bacterial strains of the same genera pointed to the fact that bacterial phenotypic

658 properties are influenced by genetic, as well as environmental factors [87].

659

660 Seed germination, shoot, and root elongation assays

661 Seed germination, shoot, and root elongation assays were done on 15 selected

662 functional isolates on Green Gram seeds (Tables 7 - 9) to represent an attribute of early

663 plant growth promotion activity by the isolates. After a week of growth, Green Gram

664 roots originating from seeds treated with all selected bacterial strains exhibited different

665 effects on root lengths, shoot lengths, and seedling vigor indexes compared with two

666 controls, treated with distilled water and nutrient broth, respectively.

30 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

667 Table 7. Effects of selected cellulolytic bacterial inoculation on early growth stage of green gram.

Mean number of Mean number of Mean length of Mean length of % germinated Vigor index Treatment seed germinated seed germinated shoots (cm) radicle (cm) seeds (Mean ± S.E.) (3rd day) (7th day) (Mean ± S.E.) (Mean ± S.E.)

C22b 50.0 50.0 12.84a ± 0.06 5.16ab ± 0.04 100 1799.50a ± 2.50 CB15 50.0 50.0 12.38a ± 0.45 5.47ab ± 0.75 100 1785.00ab ± 30.00 C41d 50.0 50.0 11.75a ± 0.03 5.67a ± 0.13 100 1741.50ab ± 10.50 C46d 50.0 50.0 11.73a ± 0.23 5.28ab ± 0.53 100 1700.00ab ± 30.00 C29b 50.0 50.0 11.81a ± 0.35 5.04ab ± 0.40 100 1684.50b ± 5.50 Control (Nutrient 50.0 50.0 9.26b ± 0.18 3.36bc ± 0.02 100 1261.00c ± 19.00 broth) Control 49.5 49.5 7.91b ± 0.18 2.87c ± 0.10 99 1067.00d ± 17.00 (dH2O) 668 Data are presented as mean ± standard error (S.E.) of duplicates of 50 seedlings for each replicate per treatment for root elongation assay and 669 seedling vigor index. The roots emerging from seeds failing to germinate after 2 days were previously marked and were not measured. The 670 seedling vigor index (SVI) was calculated using formula: SVI = % of germination x seedling length (root length + shoot length). In the same 671 column, significant differences according to Tukey’s test at p ≤ 0.05 level are indicated by different letters. 672

31 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

673 Table 8. Effects of selected N-fixing bacterial inoculation on early growth stage of green gram. Mean number of Mean number of Mean length of Mean length of % Vigor index Treatment seed germinated seed germinated shoots (cm) radicle (cm) germinated (Mean ± S.E.) (3rd day) (7th day) (Mean ± S.E.) (Mean ± S.E.) seeds NB1 50.0 50.0 11.27a ± 0.06 3.84ab ± 0.02 100 1510.00a ± 4.00 NC9 50.0 50.0 11.00a ± 0.32 4.10a ± 0.34 100 1510.00a ± 66.00 NB10 50.0 50.0 10.81ab ± 0.15 3.89ab ± 0.24 100 1469.00a ± 9.00 NB11 50.0 50.0 10.43ab ± 0.27 4.09a ± 0.26 100 1451.50ab ± 52.50 N27d 50.0 50.0 10.31ab ± 0.61 3.68ab ± 0.19 100 1399.00ab ± 42.00 Control (Nutrient 50.0 50.0 9.26bc ± 0.18 3.36ab ± 0.02 100 1261.00bc ± 19.00 broth) Control 49.5 49.5 7.91c ± 0.18 2.87b ± 0.10 99 1067.00c ± 17.00 (dH2O) 674 Data are presented as mean ± standard error (S.E.) of duplicates of 50 seedlings for each replicate per treatment for root elongation assay and 675 seedling vigor index. The roots emerging from seeds failing to germinate after 2 days were previously marked and were not measured. The 676 seedling vigor index (SVI) was calculated using formula: SVI = % of germination x seedling length (root length + shoot length). In the same 677 column, significant differences according to Tukey’s Test at p ≤ 0.05 level are indicated by different letters. 678 679 680 681 682 683

32 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

684 Table 9. Effects of selected phosphate-solubilizing bacterial inoculation on early growth stage of green gram.

Mean number Mean number Mean length of Mean length of of seed of seed % germinated Vigor index Treatment shoots (cm) radicle (cm) germinated (3rd germinated (7th seeds (Mean ± S.E.) (Mean ± S.E.) (Mean ± S.E.) day) day) PB3 50.0 50.0 12.25a ± 0.08 5.14a ± 0.02 100 1738.50a ± 9.50 PB7 49.5 50.0 12.08ab ± 0.40 4.89ab ± 0.33 100 1696.00ab ± 7.00 PB5 50.0 50.0 11.78abc ± 0.21 4.61ab ± 0.35 100 1638.50b ± 14.50 PB6 50.0 50.0 10.97bc ± 0.06 4.18abc ± 0.31 100 1514.50c ± 25.50 PC8 50.0 50.0 10.71c ± 0.18 3.73bcd ± 0.04 100 1443.00c ± 14.00 Control (Nutrient 50.0 50.0 9.26d ± 0.18 3.36cd ± 0.02 100 1261.00d ± 19.00 broth) Control 49.5 49.5 7.91e ± 0.18 2.87d ± 0.10 99 1067.00e ± 17.00 (dH2O) 685 Data are presented as mean ± standard error (S.E.) of duplicates of 50 seedlings for each replicate per treatment for root elongation assay and 686 seedling vigor index. The roots emerging from seeds failing to germinate after 2 days were previously marked and were not measured. The 687 seedling vigor index (SVI) was calculated using formula: SVI = % of germination x seedling length (root length + shoot length). In the same 688 column, significant differences according to Tukey’s test at p ≤ 0.05 level are indicated by different letters.

33 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

689 As shown in Table 7, inoculation with cellulolytic isolate C41d

690 (Stenotrophomonas sp.) resulted in significantly longer root lengths compared to both

691 control treatments, but was not significantly different from the rest of cellulolytic

692 isolates. On the other hand, the treatments with all selected cellulolytic strains resulted

693 in significantly higher shoot lengths compared to both controls but not different with

694 one another, which ranged from 11.46 to 12.9 cm. In terms of SVI, all cellulolytic

695 strains exhibited significantly higher values than both controls, with isolate C22b

696 (Serratia marcescens subsp. sakuensis) presented outstanding values but was not

697 different with isolates CB15, C41d, and C46d (Bt, Stenotrophomonas sp., and Serratia

698 nematodiphila) which ranged from 1670 to 1815.

699 In Table 8, the treatments with N-fixing isolates NC9 (Delftia lacustris) and

700 NB11 (Cupriavidus basilensis) showed higher root lengths and shoot lengths occurring

701 in the range of 3.76 cm to 4.44 cm and 10.68 cm to 11.33 cm, respectively, whereas SVI

702 values for isolates NB1, NC9, and NB10 (Burkholderia nodosa, Delftia lacustris, and

703 Ralstonia sp.) revealed significantly higher SVI values than both controls, ranging from

704 1444 to 1576, but were not significantly different from the rest of N-fixing strains.

705 Isolates PB3, PB5, and PB7 (Burkholderia cepacia, Burkholderia cepacia, and

706 Burkholderia uboniae) resulted in the longest root lengths among the phosphate-

707 solubilizing strains in the range of 4.26 cm to 5.16 cm and SVI in the range of 1624 to

708 1748 using these three isolates (Table 9). Significantly higher shoot lengths however,

709 were achieved with phosphate-solubilizing isolates than both controls, ranging from

710 10.53 cm to 12.33 cm. Isolate PB3 had the most outstanding effect on shoot length, but

711 not different from those of isolates PB5 and PB7, among other phosphate-solubilizing

712 bacterial strains.

34 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

713 A previous study by Satya et al. [88] confirmed similar observation on

714 Burkholderia sp. as their application using Burkholderia sp. strain TNAU-1

715 demonstrated improvement in root and shoot lengths, as well as increased seedling

716 vigour on mung bean. According to Egamberdieva et al. [89], inoculation of one of the

717 salt tolerant rhizobacterial strains, Stenotrophomonas rhizophila ep-17 resulted in

718 improved dry weight of cucumber by up to 68% in various salt concentrations as high as

719 10 dSm-1 in comparison to non-inoculated plants exposed to salt stress.

720 Hameeda et al. [90] demonstrated that one of phosphate-solubilizing bacterial

721 strains in the study, Serratia marcescens designated as EB 67 tested on maize growth by

722 paper towel method was able to produce higher root and plumule lengths, plant weight

723 (43%), and seed vigor index, whereas another strain also included in the study, Serratia

724 sp. designated as EB75 resulted in increased root length, plant weight by 23%, and seed

725 vigor index over uninoculated control.

726 Kang et al. [91] reported that inoculation of a novel cold stress resistant

727 rhizobacteria Serratia nematodiphila PEJ1011 on pepper plants resulted in the latter’s

728 significantly improved growth by means of both increased root and shoot lengths, and

729 fresh weight of the peppers on both normal and cold temperature as compared with

730 controls under normal and cold temperature, respectively. Most previous studies on Bt

731 reported it as a safe and effective biological control agent for the suppression of a broad

732 type of insect pests [92-95] due to its ability to produce parasporal crystals consisting of

733 protein molecules called delta endotoxins which are toxic to insect pests [96]. Besides

734 its apparent biological control attribute, current study by Khan et al. [97] also reported

735 plant growth promotion via increase in plant growth parameters on mung bean seeds

736 coated with various Bt strains used in the study. In their study, all the tested strains

737 improved shoot weight, with strain BT-64 in particular, was reported to have resulted in

35 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

738 97%, 42%, 46%, and 71% increment in shoot weight, shoot length, root length, and root

739 weight, respectively over control.

740 Apart from having high nitrogenase activity strain from rhizospheric soil of tea

741 plants [85], Jin et al. [98] found that Delftia lacustris isolated from rhizospheric soils of

742 tobacco to be one of strongly antagonistic bacteria against pathogen Phytophthora

743 nicotianae which could have potentially helped promoting growth of tobacco crop.

744 According to a study by Yrjȁlȁ et al. [99], Cupriavidus basilensis of

745 family, and especially strains of the genus Cupriavidus have been shown to be isolated

746 from hybrid aspen root in mostly pristine soil as compared to rhizospheric soils

747 contaminated with polyaromatic hydrocarbon (PAH) constituents.

748 On the other hand, the addition of PAH mixture (anthracene, phenanthrene, and

749 pyrene) resulted in the apparent shift in the diversity of uncultured Burkholderia in

750 rhizosphere of hybrid aspen and lowest fraction (five out of 38 sequences) of this genus

751 was found in pristine rhizosphere. Their study indicated the main difference in terms of

752 soil diversity that could influence the abundance between both genera. Cupriavidus

753 basilensis designated HMF14 has also been recently identified as an important furan-

754 degrading bacterium, and was isolated by employing an elegant assay based on the

755 germination of lettuce seeds [100]. Furanic compounds as toxic fermentation inhibitors

756 are derived from lignocellulose hydrolysates in which occurrences of the former in high

757 amount could pose carcinogenicity effects to humans [100].

758 Cupriavidus (synonym Ralstonia) is closely linked to Burkholderia, but

759 compared to the latter genus it consists of very limited N-fixing species [101]. Both

760 genera, in particular consisting of Burkholderia spp. and Cupriavidus taiwanensis have

761 been reported to be N-fixing symbionts of legume Mimosa with the ability to nodulate

762 the latter [102-104]. Some strains of Burkholderia have been known as Beta-rhizobia

36 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

763 which could exclusively nodulate large legume of genus Mimosa (Mimosoideae) [101].

764 These legume-nodulating burkholderias are not closely related to phytopathogenic or

765 pathogenic group of burkholderias including the Burkholderia cepacia complex, but

766 instead are closely linked to PGPR, mostly diazotrophic, endophytic environmental

767 Burkholderia species and strains [105].

768 Burkholderia nodosa designated as LMG 23741 has also been isolated from

769 acidic soils of Amazon which was a nodulating N-fixing bacterium [82], as compared to

770 the strain in this present study which did not nodulate, but was a free-living or

771 rhizospheric N fixer. Both strains are known to be absent in ability to solubilize

772 inorganic phosphates. The strain reported by da Silva et al. [82] also did not possess

773 anti-fungal characteristic. Besides the N-fixing isolate Ralstonia sp. characterized in

774 this present study, Ralstonia taiwanensis is the first N-fixing beta-proteobacterium

775 strain having the ability to nodulate the roots of Mimosa from which it was isolated

776 [106]. However, some Ralstonia spp. are environmental variant microorganisms and

777 there are also phytopathogenic strains such as Ralstonia solanacearum that cause

778 bacterial wilt occurrence on a broad range of crops [107].

779 Burkholderia cepacia designated as LMG 1222 by da Silva et al. [82] was also

780 reported as a non N-fixing bacteria isolated from acidic Amazon soils which could

781 solubilize inorganic phosphate, as well as possessing anti-fungal property. However,

782 there is lack of literature on beneficial properties of Burkholderia uboniae and hence,

783 the results of this present study could contribute primary evaluation on the potential of

784 this particular strain on early plant growth promotion. Many studies reported bacterial

785 belonging to genus Burkholderia especially B. cepacia strains as having biocontrol

786 ability against many phyto-pathogenic fungi such as Fusarium sp., Pythium

787 aphanidermatum, Pythium ultimum, Phytophthora capsici, Aspergillus flavus, Botrytis

37 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

788 cinerea, and Rhizoctonia solani [68-72] as part of its main contributing PGP factors.

789 Gluconacetobacter diazotrophicus is one of diazotrophic species reported to form

790 anomalous symbiotic relationship with plants by thriving on the root system surface as

791 rhizobacteria, whereas a number of strains are known to infect plant tissues as

792 endophytic bacteria and fix N2, resulting in enhanced plant growth [108-111].

793

794 Indole-3-acetic acid assay

795 On the basis of respective phenotypic functional activities, Green Gram seed

796 germination, and root elongation assay results, 7 bacterial isolates were selected for

797 IAA production assay (Table 10).

798 Table 10. Indole-3-acetic acid production by selected functional bacterial isolates.

IAA production Phenotype Isolate (µg mL-1) Mean ± S.E. Stenotrophomonas maltophilia C29b 12.38a ± 1.48 Cellulolytic Serratia nematodiphila C46d 11.39ab ± 1.68 Bacillus thuringiensis CB15 0.98c ± 0.16 Burkholderia cepacia PC8 8.45ab ± 0.28 Phosphate- Burkholderia cepacia PB3 7.81b ± 0.11 solubilizing Gluconacetobacter diazotrophicus 7.51b ± 0.17 PB6 Nitrogen-fixing Burkholderia nodosa NB1 0.43c ± 0.45 799 S.E. is the standard error of the mean value. 800 Significant differences according to Tukey’s test at p ≤ 0.05 level are indicated by 801 different letters. 802

803 All the 7 selected strains produced IAA without supplementation of L-

804 tryptophan, from N-fixing isolate Burkholderia nodosa NB1 to most outstanding IAA

805 production by cellulolytic isolates Stenotrophomonas maltophilia C29b and Serratia

806 nematodiphila C46d, and phosphate-solubilizing isolate Burkholderia cepacia PC8

38 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

807 ranging from 8.17 to 13.86 µg mL-1. The IAA produced by the cellulolytic, non N-

808 fixing Stenotrophomonas maltophilia strain in this present study, C29b was at 12.38 µg

809 mL-1 with absence of tryptophan. Stenotrophomonas maltophilia has been previously

810 reported as diazotrophic with IAA production of 2.6 mg mL-1 [112], 112.8 µg mL-1 for

811 S. maltophilia PM-1, and 139.2 µg mL-1 in S. maltophilia PM-26 [113] with L-

812 tryptophan as supplementation.

813 Serratia nematodiphila C46d strain in this present study was found to synthesize

814 IAA at 11.39 µg mL-1. Previously, Serratia nematodiphila designated NII-0928 isolated

815 from Western ghat forest soil was found to possess multiple PGP attributes including

816 phosphate-solubilization and IAA production at higher concentration, at 76.6 µg mL-1

817 and 58.9 µg mL-1, respectively [114]. As stated previously, the variability among the

818 bacterial strains of the same genera points to the fact that bacterial phenotypic properties

819 are influenced by genetic and environmental factors [87]. Serratia nematodiphila

820 LRE07 has also been reported to be one of the heavy metals-resistant endophytic

821 bacteria isolated from Solanum nigrum L. planted in metal-polluted soil which

822 improved plant growth and Cd extraction from soil as phytoremediation activity [115].

823 Bacillus thuringiensis (Bt) strain CB15 in this present study yielded IAA amount within

824 the stated range as reported by Vidal-Quist et al. [116], whereby all of 44 Bt strains

825 screened in their studies were reported to produce IAA in a range of 0.12 to 6.59 µg

826 mL-1. The IAA production range is comparable to the findings of Raddadi et al. [117].

827 Besides being well known for having biocontrol activity against insects [118],

828 phytopathogens including fungi [119-120], and bacteria [121], Bt has the potential to

829 promote plant growth via production of phytohormones, such as , or by

830 regulating levels of plant ethylene [117, 122], stimulation of nodulation in legumes by

831 some strains and hence, Rhizobium-based N2 fixation [123-124], improvement by other

39 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

832 strains of mineral nutrients (such as Fe and P), recovery by the plant [117], and promote

833 protection of plants from abiotic stresses [125].

834 Burkholderia cepacia strains PC8 and PB3 evaluated in this present study

835 without addition of tryptophan produced IAA at 8.45 and 7.81 µg mL-1, respectively.

836 Bevivino et al. [126] reported that rhizospheric strains of Burkholderia cepacia PHP7

837 and TVV75 were found to produce IAA at approximately 0.6 and 0.4 µg mL-1,

838 respectively, both of which were far lacking by at least 14-fold, even with addition of L-

839 tryptophan, in comparison to strains in this present study. This phenomenon could

840 indicate that the B. cepacia strains were actively involved in IAA synthesis in pure

841 culture [127].

842 Phosphate-solubilizer and non N-fixing Gluconacetobacter diazotrophicus strain

843 PB6 produced IAA at 7.51 µg mL-1 without precursor tryptophan, which was higher

844 than those reported previously and improved shoot length and vigor index of green

845 gram than uninoculated controls. Anitha and Thangaraju [128] reported an endophytic

846 diazotroph Gluconacetobacter diazotrophicus strain in their study to have exhibited

847 nitrogenase activity, phosphate-solubilization, and IAA production with and without

848 tryptophan at 4.74 µg mL-1 and 1.4 µg mL-1, respectively. The strain in their study could

849 also colonize roots of paddy in the growth solution under aseptic hydroponic condition.

850 The researchers further stated that inoculation of rice seedlings with the strain also

851 resulted in improved shoot length, root length, and plant biomass than uninoculated

852 control.

853 Despite the least amount of IAA produced by Burkholderia nodosa strain NB1

854 at 0.43 µg mL-1, this is the first results on IAA production ever reported for non-

855 nodulating N-fixing Burkholderia nodosa strain. In fact, there is dearth of reports

856 concerning the beneficial PGP traits of Burkholderia nodosa with exception of it being

40 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

857 reported as N-fixing symbiont of nodulating legumes by being isolated from N-fixing

858 root nodules of legumes Mimosa bimucronata and Mimosa scabrella [129], and its

859 apparent ability as biocontrol agent against bacterial wilt of tomato, as well as Fusarium

860 wilt of tomato and spinach with recorded suppression of disease index by 33-79%

861 [130].

862 The presence of an appropriate precursor such as L-tryptophan may mediate the

863 production of auxins by some microorganisms [131]. The influence of auxins on plant

864 seedlings relies upon their concentrations. For instance, high concentration may

865 suppress growth, whereas low concentration may produce stimulative effects on growth

866 [132]. The absence of tryptophan precursor in IAA production has also been previously

867 observed in other bacterium, Azotobacter brasilense in a mechanism known as a

868 tryptophan-independent pathway which was suggested during experiments using

869 labelled tryptophan [133].

870 Despite this fact, there is still absence of clear genetic or biochemical proof to be

871 established to consolidate this pathway [134]. According to Spaepen and Vanderleyden

872 [134], occurrence of some IAA production despite inactivation of a single pathway on

873 some knockout studies revealed redundancy of IAA biosynthetic pathways in some

874 studied microorganisms, indicating that several pathways do exist and operating in a

875 single . Nevertheless, the ability to secrete IAA without absolute

876 dependence on tryptophan as the main precursor proved to be an added advantage over

877 tryptophan dependent microbes in plant rhizospheres to produce beneficial associative

878 effects with plant roots to enhance plant growth.

879

880

881

41 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

882 Conclusions

883 Out of the 15 identified, bacterial isolates possess beneficial phenotypic traits, a

884 third belong to genus Burkholderia and a fifth to Stenotrophomonas sp. with both

885 genera consisting of members from two different functional groups. Among the tested

886 bacterial strains, isolate C46d, NB1, and PC8 showed outstanding cellulase, N-fixing,

887 and phosphate-solubilizing activities, respectively. The results of the experiments

888 confirmed the multiple PGP traits of the selected bacterial isolates based on their

889 respective high functional activities, root, shoot lengths, and seedling vigour

890 improvements when bacterized on mung bean seeds, as well as presented some

891 significant IAA productions. All these functional bacterial strains could potentially

892 produce beneficial synergistic effects via their versatile properties on improving soil

893 fertility and possible plant growth stimulation in beneficial interactions. Further

894 characterizations of these potent strains will be helpful in elucidating their potential uses

895 in biotechnological applications.

896

897 Acknowledgements

898 Many thanks are extended to the colleagues and staff of Universiti Putra

899 Malaysia Bintulu Sarawak Campus and Department of Agriculture, Kuching, Sarawak

900 for their valuable cooperation and support. Contribution of the late Dr. Wong Sing King

901 was also greatly appreciated for his advice on microbiological aspects of the study.

902

903

904

905

42 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

906 References

907 1. Strassburg BBN, Kelly A, Balmford A, Davies RG, Gibbs HK, Lovett A, et al.

908 Global congruence of carbon storage and biodiversity in terrestrial ecosystems.

909 Conserv Lett. 2010;3(2): 98-105.

910 2. Saatchi SS, Harris NL, Brown S, Lefsky M, Mitchard ETA, Salas W, et al.

911 Benchmark map of forest carbon stocks in tropical regions across three

912 continents. Proceedings of the National Academy of Sciences of the United

913 States of America. 2011;108(24): 9899-9904.

914 3. Tropek R, Sedláček O, Beck J, Keil P, Musilová Z, Šímová I, et al. Forest

915 surveys: comment on high-resolution global maps of 21st-century forest cover

916 change [technical comment- research]. USA (Washington): American

917 Association for the Advancement of Science. Science. 2014 May

918 30;344(6187): 981.

919 4. Lal R. Soil erosion and the global carbon budget. Environ Int. 2003;29(4): 437-

920 450.

921 5. Morgan RPC. Soil erosion and conservation. UK (Oxford): Wiley-Blackwell;

922 2005.

923 6. Grainger A. Controlling tropical deforestation. Earthscan. 1st ed. London;

924 1993. pp. 310. ISBN-10: 1853831425.

925 7. Roesch LF, Fulthorpe RR, Riva A, Casella G, Hadwin AKM, Kent AD, et al.

926 Pyrosequencing enumerates and contrasts soil microbial diversity. ISME J.

927 2007;1: 283–290.

928 8. Bardgett RD, Freeman C, Ostle NJ. Microbial contributions to climate change

929 through carbon cycle feedbacks. The ISME Journal. 2008;2: 805–814.

43 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

930 9. Fulthorpe RR, Roesch LFW, Riva A, Triplett EW. Distantly sampled soils

931 carry few species in common. ISME J. 2008;2: 901–910.

932 10. Berg B, Laskowski R. Litter decomposition: a guide to carbon and nutrient

933 turnover. Amsterdam: Academic Press; 2006.

934 11. Lynd LR, Weimer PJ, van Zyl WH, Pretorious IS. Microbial cellulose

935 utilization: fundamentals and biotechnology. Microbiol Mol Biol Rev.

936 2002;66: 506–577.

937 12. Baldrian P, Valášková V. Degradation of cellulose by basidiomycetous fungi.

938 FEMS Microbiol Rev. 2008;32(3): 501–521.

939 13. Kjoller A, Struwe S. Fungal communities, succession, enzymes, and

940 decomposition. In: Burns RG, Dick RP, editors. Enzymes in the environment:

941 activity, ecology and applications. New York: Marcel Dekker; 2002. pp. 267–

942 284.

943 14. de Boer W, Folman LB, Summerbell RC, Boddy L. Living in a fungal world:

944 impact of fungi on soil bacterial niche development. FEMS Microbiol Rev.

945 2005;29: 795–811.

946 15. Ward NL, Challacombe JF, Janssen PH, Henrissat B, Coutinho PM, Wu M, et

947 al. Three genomes from the phylum Acidobacteria provide insight into the

948 lifestyles of these microorganisms in soils. Appl Environ Microbiol.

949 2009;75(7): 2046–2056.

950 16. Pankratov TA, Ivanova AO, Dedysh SN, Liesack W. Bacterial populations and

951 environmental factors controlling cellulose degradation in an acidic sphagnum

952 peat. Environ Microbiol. 2011;13: 1800–1814.

44 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

953 17. Galloway JN, Townsend AR, Erisman JW, Bekunda M, Cai Z, Freney JR, et

954 al. Transformation of the : recent trends, questions, and potential

955 solutions. Science. 2008;320: 889–892.

956 18. Eady RR. The dinitrogen-fixing bacteria. The prokaryotes. Vol. 1. New York:

957 Springer-Verlag; 1992. pp. 534–553.

958 19. Young JPW. Phylogenetic classification of nitrogen fixing organisms. In:

959 Stacey G, Burris RH, Evans HJ, editors. Biological . New

960 York: Chapman and Hall; 1992. pp. 43–86.

961 20. Cleveland CC, Townsend AR, Schimel DS, Fisher H, Howarth RW, Hedin

962 LO, et al. Global patterns of terrestrial biological nitrogen (N2) fixation in

963 natural ecosystems. Global Biogeochem Cycles. 1999;13(2): 623-645.

964 21. Rengel Z, Marschner P. Nutrient availability and management in the

965 rhizosphere: exploiting genotypic differences. New Phytol. 2005;168: 305-312.

966 22. Panhwar QA, Naher UA, Jusop S, Othman R, Latif MA, Ismail MR.

967 Biochemical and molecular characterization of potential phosphate-solubilizing

968 bacteria in acid sulfate soils and their beneficial effects on rice growth. PLoS

969 ONE. 2014;9(10): e97241.

970 23. Sharma SB, Sayyed RZ, Trivedi MH, Gobi TA. Phosphate solubilizing

971 microbes: sustainable approach for managing phosphorus deficiency in

972 agricultural soils. Springer Plus. 2013;2: 587.

973 24. Rodriguez H, Fraga R. Phosphate solubilizing bacteria and their role in plant

974 growth promotion. Biotechnol Adv. 1999;17(4-5): 319–339. PMID:14538133.

975 25. Janssen PH. Identifying the dominant soil bacteria taxa in libraries of 16S

976 rRNA and 16S rRNA genes. Appl Environ Microbiol. 2006;72: 1719–1728.

45 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

977 26. Panhwar QA, Radziah O, Zaharah AR, Sariah M, Mohd Razi I. Isolation and

978 characterization of phosphorus solubilizing bacteria from aerobic rice. Afr J

979 Biotechnol. 2012;11(11): 2711–2719.

980 27. Finnie JF, Van Staden J. Effect of seed weed concentrate and applied

981 hormones on in vitro cultured tomato roots. J Plant Physiol. 1985;120: 215-

982 222.

983 28. Lotter D. New Tests Of Soil Quality: Evaluating The Health Of Your Soil's

984 Microbial Community. Farmer To Farmer Magazine. January-February 1997.

985 Available from: https://donlotter.net/f2farm.html Cited 14 June 2010.

986 29. Döbereiner J, Day JM. Associative symbioses in tropical grasses:

987 characterization of microorganisms and dinitrogen-fixing sites. In: Newton

988 WE, Nyman CJ, editors. Proceedings of the First International Symposium on

989 N2-fixation. Vol. 2. Washington: Washington University Press; 1976. pp. 518–

990 538.

991 30. Gupta R, Singal R, Shankar A, Kuhad RC, Saxena RK. A modified plate assay

992 for secreening phosphate solubilizing microorganisms. J Gen Appl Microbiol.

993 1994;40(3): 255–260.

994 31. Nautiyal CS. An efficient microbiological growth medium for screening of

995 phosphate solubilizing microorganisms. FEMS Microbiol Lett. 1999;170(1):

996 265–270.

997 32. Kokal H. The origin of phosphorus in iron-making raw materials and methods

998 of removal- a review. Sixty third annual meeting Minnesota section AIME;

999 1990. pp. 225-257.

46 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1000 33. Buis PF. techniques for the removal of phosphorus from iron

1001 ore. Ph.D. Dissertation, Mining engineering. USA: Michigan Technological

1002 University. 1995.

1003 34. Hendricks CW, Doyle JD, Hugley B. A new solid medium for enumerating

1004 cellulose-utilizing bacteria in soil. Appl Environ Microbiol. 1995;61(5): 2016-

1005 2019.

1006 35. Edi-Premono M, Moawad AM, Vlek PLG. Effect of phosphate solubilizing

1007 Pseudomonas putida on the growth of maize and its survival in the

1008 rhizosphere. Indonesian Journal of Crop Science. 1996;11: 13–23.

1009 36. Lu WJ, Wang HT, Yang SJ, Wang ZC, Yong FN. Isolation and

1010 characterization of mesophilic cellulose-degrading bacteria from flower stalks-

1011 vegetable waste co-composting system. J Gen Appl Microbiol. 2005;51: 353-

1012 360.

1013 37. Somogyi M. Notes on sugar determination. J Biol Chem. 1952;195(1): 19-23.

1014 38. Wood TM, Bhat KM. Methods for measuring cellulase activities. Methods

1015 Enzymol. 1988;160: 87-112.

1016 39. Saha S, Roy RN, Sen SK, Ray AK. Characterization of cellulase-producing

1017 bacteria from the digestive tract of tilapia, Oreochromis mossambica (Peters)

1018 and grass carp, Ctenopharyngodon idella (Valenciennes). Aquacult Res.

1019 2006;37(4): 380-388.

1020 40. Soares RA, Roesch LFW, Zanatta G, de Oliveira Camargo FA, Passaglia LMP.

1021 Occurrence and distribution of nitrogen fixing bacterial community associated

1022 with oat (Avena sativa) assessed by molecular and microbiological techniques.

1023 Appl Soil Ecol. 2006;33(3): 221–234.

47 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1024 41. Bremner JM, Keeney DR. A chemical index of soil nitrogen availability.

1025 Nature. 1966;20: 892–893.

1026 42. Delvasto P, Ballester A, García C, Igual JM, Muñoz J, González F, et al.

1027 Mineral-phosphate solubilization activity of iron ore associated microflora. In:

1028 Mendez-Vilas A, editor. Modern multidisciplinary applied microbiology:

1029 exploiting microbes and their interactions. Germany: Wiley; 2006. pp. 241-

1030 245.

1031 43. Murphy J, Riley JP. A modified single solution method for the determination

1032 of phosphate in natural waters. Anal Chim Acta. 1962;27: 31-36.

1033 44. Cappuccino JG, Sherman N. Microbiology: a laboratory manual. 7th ed. San

1034 Francisco (CA): Pearson Education; 2005.

1035 45. Holt JG, Krieg NR, Sneath PHA, Staley JT, Williams ST. Bergey’s manual of

1036 determinative bacteriology. 9th ed. USA: Lippincott Williams & Wilkins;

1037 1994. pp. 787.

1038 46. Garrity GM. Bergey’s manual of systematic bacteriology. The Proteobacteria.

1039 Vol. 2. USA: Springer-Verlag; 2005.

1040 47. Abdul-Baki AA, Anderson JD. Vigor determination in soybean seed by

1041 multiple criteria. Crop Sci. 1973;13(6): 630–633.

1042 48. Asghar HN, Zahir ZA, Khaliq A, Arshad M. Assessment of production

1043 from rhizobacteria isolated from different varieties of rapeseed. Pak J Agri Sci.

1044 2000;37(3-4): 101-104.

1045 49. Kucey RMN. Phosphate solubilizing bacteria and fungi in various cultivated

1046 and virgin Alberta soils. Can J Soil Sci. 1983;63: 671-678.

1047 50. da Silva K, Nóbrega RSA, Lima AS, Barberi A, de Souza Moreira FM.

1048 Density and diversity of diazotrophic bacteria isolated from Amazonian soils

48 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1049 using N-free semi-solid media. Sci Agric. (Piracicaba, Brazil). 2011;68(5):

1050 518-525.

1051 51. Sethi S, Datta A, Lal Gupta B, Gupta S. Optimization of cellulase production

1052 from bacteria isolated from soil. ISRN Biotechnol. 2013: 985685.

1053 52. Singh S, Moholkar VS, Goyal A. Isolation, identification, and characterization

1054 of a cellulolytic Bacillus amyloliquefaciens strain SS35 from rhinoceros dung.

1055 ISRN Microbiol. 2013: 728134.

1056 53. Ariffin H, Abdullah N, Umi Kalsom MS, Shirai Y, Hassan MA. Production

1057 and characterisation of cellulase by Bacillus pumilus EB3. International

1058 Journal of Engineering and Technology (IJET). 2006;3(1): 47-53.

1059 54. Sato A, Watanabe T, Unno Y, Purnomo E, Osaki M, Shinano T. Analysis of

1060 diversity of diazotrophic bacteria associated with the rhizosphere of a tropical

1061 arbor, Melastoma malabathricum L. Microbes Environ. 2009;24: 81-87.

1062 55. Supraja Y, Reddy RS, Reddy SS, Durga Rani CHV. Plant growth promotion

1063 and biocontrol properties of local isolates of fluorescent Pseudomonads. J Res

1064 ANGRAU. 2011;39(3): 1-5.

1065 56. Pandey A, Trivedi P, Kumar B, Palni LM. Characterization of a phosphate

1066 solubilizing and antagonistic strain of Pseudomonas putida (B0) isolated from

1067 a subalpine location in the Indian Central Himalaya. Curr Microbiol.

1068 2006;53(2): 102-107.

1069 57. Gulati A, Rahi P, Vyas P. Characterization of phosphate-solubilizing

1070 fluorescent pseudomonads from the rhizosphere of seabuckthorn growing in

1071 the cold deserts of Himalayas. Curr Microbiol. 2008;56(1): 73–79.

49 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1072 58. Song OR, Lee SJ, Lee YS, Lee SC, Kim KK, Choi YL. Solubilization of

1073 insoluble inorganic phosphate by Burkholderia cepacia DA23 isolated from

1074 cultivated soil. Braz J Microbiol. 2008;39: 151–156. PMID: 24031195.

1075 59. Chang CH, Yang SS. Thermo-tolerant phosphate-solubilizing microbes for

1076 multi-functional preparation. Bioresour Technol. 2009;100(4):

1077 1648-1658.

1078 60. Hwangbo H, Park RD, Kim YW, Rim YS, Park KH, Kim TH, et al. 2-

1079 Ketogluconic production and phosphate solubilization by Enterobacter

1080 intermedium. Curr Microbiol. 2003;47: 87–92.

1081 61. Chen YP, Rekha PD, Arun AB, Shen FT, Lai W-A, Young CC. Phosphate

1082 solubilizing bacteria from subtropical soil and their tricalcium phosphate

1083 solubilizing abilities. Appl Soil Ecol. 2006;34(1): 33–41.

1084 62. Rashid M, Khalil S, Ayub N, Alam S, Latif F. Organic acids production and

1085 phosphate solubilization by phosphate solubilizing microorganisms (PSM)

1086 under in vitro conditions. Pak J Biol Sci. 2004;7(2): 187–196.

1087 63. Compant S, Nowak J, Coenye T, Clement C, Ait Barka E. Diversity and

1088 occurrence of Burkholderia spp. in the natural environment. FEMS Microbiol

1089 Rev. 2008;32: 607–626.

1090 64. Thayer DW. Carboxymethylcellulase produced by facultative bacteria from the

1091 hind-gut of the termite Reticulitermes hesperus. J Gen Microbiol. 1978;106(1):

1092 13-18.

1093 65. Hameeda B, Rupela O, Reddy G, Satyavani K. Application of plant growth-

1094 promoting bacteria associated with composts and macrofauna for growth

1095 promotion of Pearl millet (Pennisetum glaucum L.). Biol Fert Soils.

1096 2006;43(2): 221-227.

50 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1097 66. Reinhardt EL, Ramos PL, Manfio GP, Barbosa HR, Pavan C, Moreira-Filho

1098 CA. Molecular characterization of nitrogen-fixing bacteria isolated from

1099 Brazilian agricultural plants at São Paulo state. Braz J Microbiol. 2008;39(3):

1100 414-422.

1101 67. Dantur K, Enrique R, Welin B, Castagnaro AP. Isolation of cellulolytic

1102 bacteria from the intestine of Diatraea saccharalis larvae and evaluation of

1103 their capacity to degrade sugarcane biomass. AMB Express. 2015;5: 15.

1104 PMCID: PMC4385043.

1105 68. Hebbar KP, Martel MH, Heulin T. Suppression of pre and post-emergence

1106 damping-off in corn by Burkholderia cepacia. Eur J Plant Pathol. 1998;104:

1107 29–36.

1108 69. Heydari A, Misaghi IJ. Biocontrol activity of Burkholderia cepacia against

1109 Rhizoctonia solani in -treated soils. Plant Soil. 1998;202: 109–116.

1110 70. Cain CC, Henry AT, Waldo RH, Casida LJ, Falkinham JO. Identification and

1111 characteristics of a novel Burkholderia strain with broad-spectrum

1112 antimicrobial activity. Appl Environ Microbiol. 2000;66: 4139–4141.

1113 71. Li W, Roberts DP, Dery PD, Meyer SLF, Lohrke S, Lumsden RD, et al. Broad

1114 spectrum antibiotic activity and disease suppression by the potential biocontrol

1115 agent Burkholderia ambifaria BC-F. Crop Prot. 2002;21: 129–135.

1116 72. Vijayasamundeeswari A, Vijayanandraj S, Paranidharan V, Samiyappan R,

1117 Velazhahan R. Integrated management of aflatoxin B1 contamination of

1118 groundnut (Arachis hypogaea L.) with Burkholderia sp. and zimmu (Allium

1119 sativum L. x Allium cepa L.) intercropping. J Plant Interact. 2010;5: 59–68.

51 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1120 73. Estrada P, Mavingui P, Cournoyer B, Fontaine F, Balandreau J, Caballero-

1121 Mellado J. A N2-fixing endophytic Burkholderia sp. associated with maize

1122 plants cultivated in Mexico. Can J Microbiol. 2002;48(4): 285-294.

1123 74. Lin TF, Huang HI, Shen FT, Young CC. The protons of gluconic acid are the

1124 major factor responsible for the dissolution of tricalcium phosphate by

1125 Burkholderia cepacia CC-A174. Bioresour Technol. 2006;97: 957–960.

1126 75. Chen YP, Rekha PD, Arun AB, Shen FT, Lai W-A, Young CC. Phosphate

1127 solubilizing bacteria from subtropical soil and their tricalcium phosphate

1128 solubilizing abilities. Appl Soil Ecol. 2006;34(1): 33–41.

1129 76. Chen WM, de Faria SM, James EK, Elliott GN, Lin KY, Chou JH, et al.

1130 Burkholderia nodosa sp. nov., isolated from root nodules of the woody

1131 Brazilian legumes Mimosa bimucronata and Mimosa scabrella. Int J Syst Evol

1132 Microbiol. 2007;57: 1055–1059.

1133 77. Babu-Khan S, Yeo TC, Martin WL, Rogers RD, Goldstein AH. Cloning of a

1134 mineral phosphate-solubilizing gene from Pseudomonas cepacia. Appl

1135 Environ Microbiol. 1995;61(3): 972–978.

1136 78. Peix A, Mateos PF, Rodriguez-Barrueco C, Martinez-Molina E, Velazquez E.

1137 Growth promotion of common bean (Phaseolus vulgaris L.) by a strain of

1138 Burkholderia cepacia under growth chamber conditions. Soil Biol Biochem.

1139 2001;33: 1927–1935.

1140 79. Caballero-Mellado J, Onofre-Lemus J, Estrada-De Los Santos P, Martinez-

1141 Aguilar L. The tomato rhizosphere, an environment rich in nitrogen-fixing

1142 Burkholderia species with capabilities of interest for agriculture and

1143 bioremediation. Appl Environ Microbiol. 2007;73(16): 5308–5319.

52 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1144 80. Marra LM, Oliveira SM, Soares CRFS, Moreira FMS. Solubilisation of

1145 inorganic phosphates by inoculant strains from tropical legumes. Sci Agric.

1146 2011;68(5): 603–609.

1147 81. Estrada-De Los Santos P, Bustillos-Cristales R, Caballero-Mellado J.

1148 Burkholderia, a genus rich in plant-associated nitrogen fixers with wide

1149 environmental and geographic distribution. Appl Environ Microbiol.

1150 2001;67(6): 2790–2798.

1151 82. da Silva K, Cassetari Ade S, Lima AS, De Brandt E, Pinnock E, Vandamme P,

1152 et al. Diazotrophic Burkholderia species isolated from the Amazon region

1153 exhibit phenotypical, functional and genetic diversity. Syst Appl Microbiol.

1154 2012;35(4): 253-262.

1155 83. Shigematsu T, Yumihara K, Ueda Y, Numaguchi M, Morimura S, Kida K.

1156 Delftia tsuruhatensis sp. nov., a terephthalate-assimilating bacterium isolated

1157 from activated sludge. Int J Syst Evol Microbiol. 2003;53(5): 1479–1483.

1158 84. Validov S, Kamilova F, Qi S, Stephen D, Wang JJ, Makarova N, et al.

1159 Selection of bacteria able to control Fusarium oxysporum f. sp. Radicus-

1160 lycopersici in stone substrate. J Appl Microbiol. 2007;102: 461–471.

1161 85. Gulati A, Sood S, Rahi P, Thakur R, Chauhan S, nee Chadha IC. Diversity

1162 analysis of diazotrophic bacteria associated with the roots of tea [Camellia

1163 sinensis (L.) O. Kuntze]. J Microbiol Biotechnol. 2011;21(6): 545-555.

1164 86. Han J, Sun L, Dong X, Cai Z, Sun X, Yang H, et al. Characterization of a

1165 novel plant growth promoting bacteria strain Delftia tsuruhatensis HR4 both as

1166 a diazotroph and a potential biocontrol agent against various plant pathogens.

1167 Syst Appl Microbiol. 2005;28: 66–76.

53 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1168 87. O’Donnell AG, Seasman M, Macrae A, Waite I, Davies JT. Plants and

1169 fertilisers as drivers of change in microbial community structure and function

1170 in soils. Plant Soil. 2001;232: 135-145.

1171 88. Satya VK, Vijayasamundeeswari A, Paranidharan V, Velazhahan R.

1172 Burkholderia sp. strain TNAU-1 for biological control of root rot in mung bean

1173 (Vigna radiata L.) caused by Macrophomina phaseolina. J Plant Prot Res.

1174 2011;51(3): 273-278.

1175 89. Egamberdieva D, Berg G, Lindstrom K, Rasanen L. Chapter 7: Alleviation of

1176 salt stress in cucumber and tomato by root colonising bacteria. In: Miransari

1177 M, editor. Soil microbiology and biotechnology. USA: Studium Press, LLC;

1178 2013. pp. 113-126. ISBN: 1-626990-14-X.

1179 90. Hameeda B, Harini G, Rupela OP, Wani SP, Reddy G. Growth promotion of

1180 maize by phosphate-solubilizing bacteria isolated from composts and

1181 macrofauna. Microbiol Res. 2008;163(2): 234-242.

1182 91. Kang SM, Khan AL, Waqas M, Young HY, Hamayun M, Joo GJ, et al.

1183 Gibberellin-producing Serratia nematodiphila PEJ1011 ameliorates low

1184 temperature stress in Capsicum anuum L. Eur J Soil Biol. 2015;68(May-June):

1185 85-93.

1186 92. Tailor R, Tippett J, Gibb G, Pells S, Pike D, Jordan L, et al. Identification and

1187 characterization of a novel Bacillus thuringiensis delta-endotoxin entomocidal

1188 to coleopteran and lepidopteran larvae. Mol Microbiol. 1992;6(9): 1211-1217.

1189 93. Borgonie G, Van Driessche R, Leyns F, Arnaut G, De Waele D, Coomans A.

1190 Germination of Bacillus thuringiensis spores in bacteriophagous nematodes

1191 (Nematoda: Rhabditida). J Invertebr Pathol. 1995;65(1): 61-67.

54 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1192 94. Nester EW, Thomashow LS, Metz M, Girdon M. 100 years of Bacillus

1193 thuringiensis. A critical scientific assessment. USA(Washington): ASM; 2002.

1194 95. Wei JZ, Hale K, Carta L, Platzer E, Wong C, Fang SC, et al. Bacillus

1195 thuringiensis Crystal that target nematodes. PNAS. 2003;100: 2760-

1196 2765.

1197 96. Höfte H, Whiteley HR. Insecticidal crystal proteins of Bacillus thuringiensis.

1198 Microbiol Rev. 1989;53(2): 242-255.

1199 97. Khan MQ, Abbasi MW, Zaki MJ, Khan SA. Evaluation of Bacillus

1200 thuringiensis isolates against root-knot nematodes following seed application

1201 in okra and mungbean. Pak J Bot. 2010;42(4): 2903-2910.

1202 98. Jin F, Ding Y, Ding W, Reddy MS, Fernando WGD, Du B. Genetic diversity

1203 and phylogeny of antagonistic bacteria against Phytophthora nicotianae

1204 isolated from tobacco rhizosphere. Int J Mol Sci. 2011;12(5): 3055-3071.

1205 99. Yrjȁlȁ K, Mancano G, Fortelius C, Akerman ML, Sipila TP. The incidence of

1206 Burkholderia in epiphytic and endophytic bacterial cenoses in hybrid aspen

1207 grown on sandy peat. Boreal Environ Res. 2010;15(1): 81-96.

1208 100. Wierckx N, Koopman F, Ruijssenaars HJ, de Winde JH. Microbial degradation

1209 of furanic compounds: biochemistry, genetics, and impact. Appl Microbiol

1210 Biotechnol. 2011;92(6): 1095-1105.

1211 101. James EK, Moulin L, Klonowska A, de Faria SM, Simon MF, Chen WM.

1212 Beta-rhizobial symbioses with legumes. In: Gupta VVSR, Unkovich M, Kaiser

1213 BN, editors. ASNF. Adelaide: The 17th Australian Nitrogen Fixation

1214 Proceedings; 2014. pp. 49-53.

55 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1215 102. Parker MA, Wurtz A, Paynter Q. Nodule of invasive Mimosa pigra

1216 in Australia and in ancestral habitats: a comparative analysis. Biol Invasions.

1217 2007;9: 127-138.

1218 103. Gyaneshwar P, Hirsch AM, Moulin L, Chen WM, Elliott GN, Bontemps C, et

1219 al. Legume-nodulating : diversity, host range, and future

1220 prospects. Mol Plant Microbe Interact. 2011;24(11): 1276-1288.

1221 104. Andrus AD, Andam C, Parker MA. American origin of Cupriavidus bacteria

1222 associated with invasive Mimosa legumes in the Philippines. FEMS Microbiol

1223 Ecol. 2012;80: 747-750.

1224 105. Suárez-Moreno ZR, Caballero-Mellado J, Coutinho BG, Mendonca-Previato L,

1225 James EK, Venturi V. Common features of environmental and potentially

1226 beneficial plant-associated Burkholderia. Microb Ecol. 2012;63(2): 249-266.

1227 106. Chen WM, James EK, Prescott AR, Kierans M, Sprent JI. Nodulation of

1228 Mimosa spp. by the beta-proteobacterium Ralstonia taiwanensis. Mol Plant

1229 Microbe Interact. 2003;16(12): 1051–1061.

1230 107. Palleroni NJ, Doudoroff M. Phylogenetic characterization and

1231 deoxyribonucleic acid homologies of Pseudomonas solanacearum. J Bacteriol.

1232 1971;107: 690-696.

1233 108. Rediers H, Vanderleyden J, De-Mot R. : a Pseudomonas

1234 in disguise? Microbiology. 2004;6: 1117-1119.

1235 109. Pedraza RO. Recent advances in nitrogen-fixing acetic acid bacteria. Int J Food

1236 Microbiol. 2008;6: 25-35.

1237 110. Liu Y, Wang H, Sun X, Yang H, Wang Y, Song W. Study on mechanisms of

1238 colonization of nitrogen-fixing PGPB, Klebsiella pneumonia NG14 on the root

56 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1239 surface of rice and the formation of biofilm. Curr Microbiol. 2011;6: 1113-

1240 1122.

1241 111. Zehr JP. Nitrogen fixation by marine cyanobacteria. Trends Microbiol. 2011;6:

1242 162-173.

1243 112. Othman R, Naher UA, Yusoff SZ. Effect of urea-N on growth and indoleacetic

1244 acid production of Stenotrophomonas maltophilia (Sb16) isolated from rice

1245 growing soils in Malaysia. Chil J Agr Res. 2013;73(2): 187-192.

1246 113. Park M, Kim C, Yang J, Lee H, Shin W, Kim S, et al. Isolation and

1247 characterization of diazotrophic growth promoting bacteria from rhizosphere of

1248 agricultural crops of Korea. Microbiol Res. 2005;160(2): 127-133.

1249 114. Dastager SG, Deepa CK, Pandey A. Potential plant growth-promoting activity

1250 of Serratia nematodiphila NII-0928 on black pepper (Piper nigrum L.). World

1251 J Microbiol Biotechnol. 2011;27(2): 259-265.

1252 115. Chen L, Luo S, Xiao X, Guo H, Chen J, Wan Y, et al. Application of plant

1253 growth-promoting endophytes (PGPE) isolated from Solanum nigrum L. for

1254 phytoextraction of Cd-polluted soils. Appl Soil Ecol. 2010;46: 383-389.

1255 116. Vidal-Quist JC, Rogers HJ, Mahenthiralingam E, Berry C. Bacillus

1256 thuringiensis colonises plant roots in a phylogeny-dependent manner. FEMS

1257 Microbiol Ecol. 2013;86(3): 474-489.

1258 117. Raddadi N, Cherif A, Boudabous A, Daffonchio D. Screening of plant growth

1259 promoting traits of Bacillus thuringiensis. Ann Microbiol. 2008;58: 47-52.

1260 118. Schnepf E, Crickmore N, Van Rie J, Lereclus D, Baum J, Feitelson J, et al.

1261 Bacillus thuringiensis and its pesticidal crystal proteins. Microbiol Mol Biol

1262 Rev. 1998;62(3): 775-806.

57 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1263 119. Sadfi N, Chérif M, Fliss I, Boudabbous A, Antoun H. Evaluation of bacterial

1264 isolates from salty soils and Bacillus thuringiensis strains for the biocontrol of

1265 Fusarium dry rot of potato tubers (Solanum tuberosum L.). J Plant Pathol.

1266 2001;83: 101–108.

1267 120. Gomaa EZ. Chitinase production by Bacillus thuringiensis and Bacillus

1268 licheniformis: their potential in antifungal biocontrol. J Microbiol. 2012;50:

1269 103-111.

1270 121. Dong YH, Zhang XF, Xu JL, Zhang LH. Insecticidal Bacillus thuringiensis

1271 silences Erwinia carotovora virulence by a new form of microbial antagonism,

1272 signal interference. Appl Environ Microbiol. 2004;70: 954-960.

1273 122. de Goes KCGP, de Castro Fisher ML, Cattelan AJ, Nogueira MA, Portela de

1274 Carvalho CG, Martinez de Oliveira AL. Biochemical and molecular

1275 characterization of high population density bacteria isolated from sunflower. J

1276 Microbiol Biotechnol. 2012;22(4): 437-447.

1277 123. Lee KD, Gray EJ, Mabood F, Jung WJ, Charles T, Clark SRD, et al. The class

1278 IId bacteriocin thuricin-17 increases plant growth. Planta. 2009;229(4): 747-

1279 755.

1280 124. Mishra PK, Mishra S, Selvakumar G, Bisht JK, Kundu S, Gupta HS.

1281 Coinoculation of Bacillus thuringiensis-KR1 with Rhizobium leguminosarum

1282 enhances plant growth and nodulation of pea (Pisum sativum L.) and lentil

1283 (Lens culinaris L.). World J Microbiol Biotechnol. 2009;25(5): 753-761.

1284 125. Vardharajula S, Ali SZ, Grover M, Reddy G, Bandi V. Drought-tolerant plant

1285 growth promoting Bacillus spp.: effect on growth, osmolytes and antioxidant

1286 status of maize under drought stress. J Plant Interact. 2011;6: 1-14.

58 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1287 126. Bevivino A, Tabacchioni S, Chiarini L, Carusi MV, Del Gallo M, Visca P.

1288 Phenotypic comparison between rhizosphere and clinical isolates of

1289 Burkholderia cepacia. Microbiology. 1994;140: 1069-1077.

1290 127. Yasmin F, Othman R, Sijam K, Saad MS. Characterization of beneficial

1291 properties of plant growth-promoting rhizobacteria isolated from sweet potato

1292 rhizosphere. Afr J Microbiol Res. 2009;3(11): 815-821.

1293 128. Anitha KG, Thangaraju M. Growth promotion of rice seedling by

1294 Gluconacetobacter diazotrophicus under in vivo conditions. J Cell Plant Sci.

1295 2010;1: 6-12.

1296 129. Chen WM, de Faria SM, James EK, Elliott GN, Lin KY, Chou JH, et al.

1297 Burkholderia nodosa sp. nov., isolated from root nodules of the woody

1298 Brazilian legumes Mimosa bimucronata and Mimosa scabrella. Int J Syst Evol

1299 Microbiol. 2007;57: 1055–1059.

1300 130. Nion YA, Toyota K. Suppression of bacterial wilt and Fusarium wilt by a

1301 Burkholderia nodosa strain isolated from Kalimantan soils, Indonesia.

1302 Microbes Environ. 2008;23(2): 134-141.

1303 131. Ambawade MS, Pathade GR. Production of indole acetic acid (IAA) by

1304 Stenotrophomonas maltophilia BE25 isolated from roots of banana (Musa

1305 spp.). International Journal of Science and Research (IJSR). 2013; 2319-7064.

1306 132. Arshad M, Frankenberger WT Jr. Microbial production of plant hormones.

1307 Plant and Soil. 1991;133(1): 1-8.

1308 133. Prinsen E, Costacurta A, Michiels K, Vanderleyden J, Van Onckelen H.

1309 Azospirillum brasilense indole-3-acetic acid biosynthesis: evidence for a non-

1310 tryptophan dependent pathway. Mol Plant Microbe Interact. 1993;6: 609–615.

59 bioRxiv preprint doi: https://doi.org/10.1101/351916; this version posted June 20, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY 4.0 International license.

1311 134. Spaepen S, Vanderleyden J. Auxin and plant-microbe interactions. In: Estelle

1312 M, Weijers D, Leyser O, Ljung K, editors. Perspectives in biology. Cold

1313 Spring Harbor Laboratory Press; 2010. doi: 10.1101/cshperspect.a001438.

1314

1315

1316

1317

60