전체메뉴
검색
Article Search

JMB Journal of Microbiolog and Biotechnology

QR Code QR Code

Research article


References

  1. Adipah S. 2019. Introduction of petroleum hydrocarbons contaminants and its human effects. J. Environ. Sci. Public Health 3: 1-9.
  2. U.S. EPA. 2004. Cleaning Up the Nation’s Waste Sites: Markets and Technology Trends, 2004 edition. 2004th Ed. Office of Solid Waste And Emergency Response, EPAU.S. EPA/National Service Center for Environmental Publications (NSCEP), Cincinnati, OH.
  3. Williams S., Ladd D., Farmer J. 2006. Scientific investigations report 2005-5104 Fate and Transport of petroleum hydrocarbons in soil and ground water at big south fork national river and recreation area, Tennessee and Kentucky, 2002-2003. U.S. Geol. Surv. Sci. Investig. Rep. 2005-5104: 29.
    CrossRef
  4. Somboon K, Doble A, Bulmer D, Baslé A, Khalid S, van den Berg B. 2020. Uptake of monoaromatic hydrocarbons during biodegradation by FadL channel-mediated lateral diffusion. Nat. Commun. 11: 6331.
    Pubmed PMC CrossRef
  5. You J, Du M, Chen H, Zhang X, Zhang S, Chen J, et al. 2018. BTEX degradation by a newly isolated bacterium: performance, kinetics, and mechanism. Int. Biodeterior. Biodegrad. 129: 202-208.
    CrossRef
  6. Jiang B, Zhou Z, Dong Y, Tao W, Wang B, Jiang J, et al. 2015. Biodegradation of benzene, toluene, ethylbenzene, and o-, m-, and p-xylenes by the newly isolated bacterium Comamonas sp. JB. Appl. Biochem. Biotechnol. 176: 1700-1708.
    Pubmed CrossRef
  7. Lee SH, Jin HM, Lee HJ, Kim JM, Jeon CO. 2012. Complete genome sequence of the BTEX-degrading bacterium Pseudoxanthomonas spadix BD-a59. J. Bacteriol. 194: 544.
    Pubmed PMC CrossRef
  8. Lee SK, Lee SB. 2001. Isolation and characterization of a thermotolerant bacterium Ralstonia sp. strain PHS1 that degrades benzene, toluene, ethylbenzene, and o-xylene. Appl. Microbiol. Biotechnol. 56: 270-275.
    Pubmed CrossRef
  9. Meyer-Cifuentes I, Martinez-Lavanchy PM, Marin-Cevada V, Böhnke S, Harms H, Müller JA, et al. 2017. Isolation and characterization of Magnetospirillum sp. strain 15-1 as a representative anaerobic toluene-degrader from a constructed wetland model. PLoS One 12: e0174750.
    Pubmed PMC CrossRef
  10. Surendra SV, Mahalingam BL, Velan M. 2017. Degradation of monoaromatics by Bacillus pumilus MVSV3. Braz. Arch. Biol. Technol. 60: 16160319.
    CrossRef
  11. Wongbunmak A, Khiawjan S, Suphantharika M, Pongtharangkul T. 2017. BTEX- and naphthalene-degrading bacterium Microbacterium esteraromaticum strain SBS1-7 isolated from estuarine sediment. J. Hazard. Mater. 339: 82-90.
    Pubmed CrossRef
  12. Wongbunmak A, Khiawjan S, Suphantharika M, Pongtharangkul T. 2020. BTEX biodegradation by Bacillus amyloliquefaciens subsp. plantarum W1 and its proposed BTEX biodegradation pathways. Sci. Rep. 10: 17408.
    Pubmed PMC CrossRef
  13. Yrjälä K, Paulin L, Romantschuk M. 1997. Novel organization of catechol meta-pathway genes in Sphingomonas sp. HV3 pSKY4 plasmid. FEMS Microbiol. Lett. 154: 403-408.
    Pubmed CrossRef
  14. Zhang S, You J, Kennes C, Cheng Z, Ye J, Chen D, et al. 2018. Current advances of VOCs degradation by bioelectrochemical systems: a review. Chem. Eng. J. 334: 2625-2637.
    CrossRef
  15. Lee Y, Lee Y, Jeon CO. 2019. Biodegradation of naphthalene, BTEX, and aliphatic hydrocarbons by Paraburkholderia aromaticivorans BN5 isolated from petroleum-contaminated soil. Sci. Rep. 9: 860.
    Pubmed PMC CrossRef
  16. El-Naas MH, Acio JA, El Telib AE. 2014. Aerobic biodegradation of BTEX: progresses and prospects. J. Environ. Chem. Eng. 2: 1104-1122.
    CrossRef
  17. Lee K, Gibson DT. 1996. Toluene and ethylbenzene oxidation by purified naphthalene dioxygenase from Pseudomonas sp. strain NCIB 9816-4. Appl. Environ. Microbiol. 62: 3101-3106.
    Pubmed PMC CrossRef
  18. Bühler B, Witholt B, Hauer B, Schmid A. 2002. Characterization and application of xylene monooxygenase for multistep biocatalysis. Appl. Environ. Microbiol. 68: 560-568.
    Pubmed PMC CrossRef
  19. Yu H, Kim BJ, Rittmann BE. 2001. The roles of intermediates in biodegradation of benzene, toluene, and p-xylene by Pseudomonas putida F1. Biodegradation 12: 455-463.
    Pubmed CrossRef
  20. Bickerdike SR, Holt RA, Stephens GM. 1997. Evidence for metabolism of o-xylene by simultaneous ring and methyl group oxidation in a new soil isolate. Microbiology 143: 2321-2329.
    Pubmed CrossRef
  21. Lee SK, Lee SB. 2002. Substrate utilization patterns during BTEX biodegradation by an o-xylene-degrading bacterium Ralstonia sp. PHS1. J. Microbiol. Biotechnol. 12: 909-915.
  22. Jeong E, Hirai M, Shoda M. 2008. Removal of O-xylene using biofilter inoculated with Rhodococcus sp. BTO62. J. Hazard. Mater. 152: 140-147.
    Pubmed CrossRef
  23. Taki H, Syutsubo K, Mattison RG, Harayama S. 2007. Identification and characterization of o-xylene-degrading Rhodococcus spp. which were dominant species in the remediation of o-xylene-contaminated soils. Biodegradation 18: 17-26.
    Pubmed CrossRef
  24. Bertoni G, Bolognese F, Galli E, Barbieri P. 1996. Cloning of the genes for and characterization of the early stages of toluene and oxylene catabolism in Pseudomonas stutzeri OX1. Appl. Environ. Microbiol. 62: 3704-3711.
    Pubmed PMC CrossRef
  25. Boeckmann B, Bairoch A, Apweiler R, Blatter MC, Estreicher A, Gasteiger E, et al. 2003. The SWISS-PROT protein knowledgebase and its supplement TrEMBL in 2003. Nucleic Acids Res. 31: 365-370.
    Pubmed PMC CrossRef
  26. Buchfink B, Reuter K, Drost HG. 2021. Sensitive protein alignments at tree-of-life scale using DIAMOND. Nat. Methods 18: 366-368.
    Pubmed PMC CrossRef
  27. Galperin MY, Wolf YI, Makarova KS, Alvarez R V, Landsman D, Koonin E V. 2021. COG database update: focus on microbial diversity, model organisms, and widespread pathogens. Nucleic Acids Res. 49: D274-D281.
    Pubmed PMC CrossRef
  28. Jeske L, Placzek S, Schomburg I, Chang A, Schomburg D. 2019. BRENDA in 2019: a European ELIXIR core data resource. Nucleic Acids Res. 47: D542-D549.
    Pubmed PMC CrossRef
  29. Kanehisa M, Sato Y, Kawashima M, Furumichi M, Tanabe M. 2016. KEGG as a reference resource for gene and protein annotation. Nucleic Acids Res. 44: D457-D462.
    Pubmed PMC CrossRef
  30. Bertelli C, Laird MR, Williams KP, Lau BY, Hoad G, Winsor GL, et al. 2017. IslandViewer 4: expanded prediction of genomic islands for larger-scale datasets. Nucleic Acids Res. 45: W30-W35.
    Pubmed PMC CrossRef
  31. Tamura K, Stecher G, Kumar S. 2021. MEGA11: molecular evolutionary genetics analysis version 11. Mol. Biol. Evol. 38: 3022-3027.
    Pubmed PMC CrossRef
  32. Kim J, Na SI, Kim D, Chun J. 2021. UBCG2: up-to-date bacterial core genes and pipeline for phylogenomic analysis. J. Microbiol. 59: 609-615.
    Pubmed CrossRef
  33. Lee I, Kim YO, Park SC, Chun J. 2016. OrthoANI: an improved algorithm and software for calculating average nucleotide identity. Int. J. Syst. Evol. Microbiol. 66: 1100-1103.
    Pubmed CrossRef
  34. Meier-Kolthoff JP, Auch AF, Klenk HP, Göker M. 2013. Genome sequence-based species delimitation with confidence intervals and improved distance functions. BMC Bioinformatics 14: 60.
    Pubmed PMC CrossRef
  35. Chun J, Oren A, Ventosa A, Christensen H, Arahal DR, da Costa MS, et al. 2018. Proposed minimal standards for the use of genome data for the taxonomy of prokaryotes. Int. J. Syst. Evol. Microbiol. 68: 461-466.
    Pubmed CrossRef
  36. Wang X, Chen M, Xiao J, Hao L, Crowley DE, Zhang Z, et al. 2015. Genome sequence analysis of the naphthenic acid degrading and metal resistant bacterium Cupriavidus gilardii CR3. PLoS One 10: e0132881.
    Pubmed PMC CrossRef
  37. Ensley BD, Ratzkin BJ, Osslund TD, Simon MJ, Wackett LP, Gibson DT. 1983. Expression of naphthalene oxidation genes in Escherichia coli results in the biosynthesis of indigo. Science 222: 167-169.
    Pubmed CrossRef
  38. Arenghi FLG, Berlanda D, Galli E, Sello G, Barbieri P. 2001. Organization and Regulation of meta cleavage pathway genes for toluene and o-xylene derivative degradation in Pseudomonas stutzeri OX1. Appl. Environ. Microbiol. 67: 3304-3308.
    Pubmed PMC CrossRef
  39. Kaschabek SR, Kasberg T, Müller D, Mars AE, Janssen DB, Reineke W. 1998. Degradation of chloroaromatics: purification and characterization of a novel type of chlorocatechol 2,3-dioxygenase of Pseudomonas putida GJ31. J. Bacteriol. 180: 296-302.
    Pubmed PMC CrossRef
  40. Choi EJ, Jin HM, Lee SH, Math RK, Madsen EL, Jeon CO. 2013. Comparative genomic analysis and benzene, toluene, ethylbenzene, and o-, m-, and p-xylene (BTEX) degradation pathways of Pseudoxanthomonas spadix BD-a59. Appl. Environ. Microbiol 79: 663-671.
    Pubmed PMC CrossRef
  41. Espinoza Tofalos A, Daghio M, Gonzalez M, Papacchini M, Franzetti A, Seeger M. 2018. Toluene degradation by Cupriavidus metallidurans CH34 in nitrate-reducing conditions and in bioelectrochemical systems. FEMS Microbiol. Lett. 365: fny119.
    Pubmed CrossRef
  42. Kumar A, Dewulf J, Luvsanjamba M, Van Langenhove H. 2008. Continuous operation of membrane bioreactor treating toluene vapors by Burkholderia vietnamiensis G4. Chem. Eng. J. 140: 193-200.
    CrossRef
  43. Perez-Pantoja D, De la Iglesia R, Pieper DH, Gonzalez B. 2008. Metabolic reconstruction of aromatic compounds degradation from the genome of the amazing pollutant-degrading bacterium Cupriavidus necator JMP134. FEMS Microbiol. Rev. 32: 736-794.
    Pubmed CrossRef
  44. Shingler V, Powlowski J, Marklund U. 1992. Nucleotide-sequence and functional-analysis of the complete phenol/3,4-dimethylphenol catabolic pathway of Pseudomonas sp. strain-Cf600. J. Bacteriol. 174: 711-724.
    Pubmed PMC CrossRef
  45. Tinberg CE, Song WJ, Izzo V, Lippard SJ. 2011. Multiple roles of component proteins in bacterial multicomponent monooxygenases: phenol hydroxylase and toluene/o-xylene monooxygenase from Pseudomonas sp. OX1. Biochemistry 50: 1788-1798.
    Pubmed PMC CrossRef
  46. Hou YJ, Guo Y, Li DF, Zhou NY. 2021. Structural and biochemical analysis reveals a distinct catalytic site of salicylate 5-monooxygenase NagGH from Rieske dioxygenases. Appl. Environ. Microbiol. 87: e01629-20.
    Pubmed PMC CrossRef
  47. Karlsson A, Parales J V, Parales RE, Gibson DT, Eklund H, Ramaswamy S. 2003. Crystal structure of naphthalene dioxygenase: sideon binding of dioxygen to iron. Science 299: 1039-1042.
    Pubmed CrossRef
  48. Xu JR, Zhang Y. 2010. How significant is a protein structure similarity with TM-score=0.5? Bioinformatics 26: 889-895.
    Pubmed PMC CrossRef
  49. Kim M, Oh HS, Park SC, Chun J. 2014. Towards a taxonomic coherence between average nucleotide identity and 16S rRNA gene sequence similarity for species demarcation of prokaryotes. Int. J. Syst. Evol. Microbiol. 64: 346-351.
    Pubmed CrossRef
  50. Moreno-Hagelsieb G, Latimer K. 2008. Choosing BLAST options for better detection of orthologs as reciprocal best hits. Bioinformatics 24: 319-324.
    Pubmed CrossRef
  51. Burbank LP, Van Horn CR. 2017. Conjugative plasmid transfer in Xylella fastidiosa is dependent on tra and trb operon functions. J. Bacteriol. 199: e00388-17.
    Pubmed PMC CrossRef
  52. Newman LM, Wackett LP. 1995. Purification and characterization of toluene 2-monooxygenase from Burkholderia cepacia G4. Biochemistry 34: 14066-14076.
    Pubmed CrossRef
  53. Cafaro V, Izzo V, Scognamiglio R, Notomista E, Capasso P, Casbarra A, et al. 2004. Phenol hydroxylase and toluene/o-xylene monooxygenase from Pseudomonas stutzeri OX1: interplay between two enzymes. Appl. Environ. Microbiol. 70: 2211-2219.
    Pubmed PMC CrossRef
  54. Arai H, Akahira S, Ohishi T, Maeda M, Kudo T. 1998. Adaptation of Comamonas testosteroni TA441 to utilize phenol: organization and regulation of the genes involved in phenol degradation. Microbiology 144: 2895-2903.
    Pubmed CrossRef
  55. Nordlund I, Powlowski J, Shingler V. 1990. Complete nucleotide sequence and polypeptide analysis of multicomponent phenol hydroxylase from Pseudomonas sp. strain CF600. J. Bacteriol. 172: 6826-6833.
    Pubmed PMC CrossRef
  56. Shields MS, Reagin MJ, Gerger RR, Campbell R, Somerville C. 1995. TOM, a new aromatic degradative plasmid from Burkholderia (Pseudomonas) cepacia G4. Appl. Environ. Microbiol. 61: 1352-1356.
    Pubmed PMC CrossRef
  57. Fox BG, Shanklin J, Somerville C, Munck E. 1993. Stearoyl-acyl carrier protein delta 9 desaturase from Ricinus communis is a diironoxo protein. Proc. Natl. Acad. Sci. USA 90: 2486.
    Pubmed PMC CrossRef
  58. Johnson GR, Olsen RH. 1997. Multiple pathways for toluene degradation in Burkholderia sp. strain JS150. Appl. Environ. Microbiol. 63: 4047.
    Pubmed PMC CrossRef
  59. Harayama S, Polissi A, Rekik M. 1991. Divergent evolution of chloroplast-type ferredoxins. FEBS Lett. 285: 85-88.
    Pubmed CrossRef
  60. Laurie AD, Lloyd-Jones G. 1999. Conserved and hybrid meta-cleavage operons from PAH-degrading Burkholderia RP007. Biochem. Biophys. Res. Commun. 262: 308-314.
    Pubmed CrossRef
  61. Arai H, Ohishi T, Chang MY, Kudo T. 2000. Arrangement and regulation of the genes for meta-pathway enzymes required for degradation of phenol in Comamonas testosteroni TA441. Microbiology 146(Pt 7): 1707-1715.
    Pubmed CrossRef
  62. Pérez-Pantoja D, Donoso R, Agulló L, Córdova M, Seeger M, Pieper DH, et al. 2012. Genomic analysis of the potential for aromatic compounds biodegradation in Burkholderiales. Environ. Microbiol. 14: 1091-1117.
    Pubmed CrossRef
  63. Romine MF, Stillwell LC, Wong KK, Thurston SJ, Sisk EC, Sensen C, et al. 1999. Complete sequence of a 184-kilobase catabolic plasmid from Sphingomonas aromaticivorans F199. J. Bacteriol. 181: 1585.
    Pubmed PMC CrossRef
  64. Erickson BD, Mondello FJ. 1992. Nucleotide sequencing and transcriptional mapping of the genes encoding biphenyl dioxygenase, a multicomponent polychlorinated-biphenyl-degrading enzyme in Pseudomonas strain LB400. J. Bacteriol. 174: 2903.
    Pubmed PMC CrossRef
  65. Froland WA, Andersson KK, Lee SK, Liu Y, Lipscomb JD. 1992. Methane monooxygenase component B and reductase alter the regioselectivity of the hydroxylase component-catalyzed reactions. A novel role for protein-protein interactions in an oxygenase mechanism. J. Biol. Chem. 267: 17588-17597.
    Pubmed CrossRef
  66. Mason JR, Cammack R. 1992. The electron-transport proteins of hydroxylating bacterial dioxygenases. Annu. Rev. Microbiol. 46: 277-305.
    Pubmed CrossRef
  67. Kitayama A, Suzuki E, Kawakami Y, Nagamune T. 1996. Gene organization and low regiospecificity in aromatic-ring hydroxylation of a benzene monooxygenase of Pseudomonas aeruginosa JI104. J. Ferment. Bioeng. 82: 421-425.
    CrossRef
  68. Olsen RH, Kukor JJ, Kaphammer B. 1994. A novel toluene-3-monooxygenase pathway cloned from Pseudomonas pickettii PKO1. J. Bacteriol. 176: 3749-3756.
    Pubmed PMC CrossRef
  69. Mitchell KH, Studts JM, Fox BG. 2002. Combined participation of hydroxylase active site residues and effector protein binding in a para to ortho modulation of toluene 4-monooxygenase regiospecificity. Biochemistry 41: 3176-3188.
    Pubmed CrossRef
  70. Bertoni G, Martino M, Galli E, Barbieri P. 1998. Analysis of the gene cluster encoding toluene/o-xylene monooxygenase from Pseudomonas stutzeri OX1. Appl. Environ. Microbiol. 64: 3626-3632.
    Pubmed PMC CrossRef
  71. Jiang T, Wang W, Han B. 2013. Catalytic hydroxylation of benzene to phenol with hydrogen peroxide using catalysts based on molecular sieves. New J. Chem. 37: 1654-1664.
    CrossRef
  72. Barbieri P, Palladino L, Di Gennaro P, Galli E. 1993. Alternative pathways for o-xylene or m-xylene and p-xylene degradation in a Pseudomonas stutzeri strain. Biodegrad. 4: 71-80.
    CrossRef
  73. Fowler SJ, Gutierrez-Zamora ML, Manefield M, Gieg LM. 2014. Identification of toluene degraders in a methanogenic enrichment culture. FEMS Microbiol. Ecol. 89: 625-636.
    Pubmed CrossRef
  74. Kahng HY, Malinverni JC, Majko MM, Kukor JJ. 2001. Genetic and functional analysis of the tbc operons for catabolism of alkyl- and chloroaromatic compounds in Burkholderia sp. strain JS150. Appl. Environ. Microbiol. 67: 4805-4816.
    Pubmed PMC CrossRef
  75. Siani L, Viggiani A, Notomista E, Pezzella A, Di Donato A. 2006. The role of residue Thr249 in modulating the catalytic efficiency and substrate specificity of catechol-2,3-dioxygenase from Pseudomonas stutzeri OX1. FEBS J. 273: 2963-2976.
    Pubmed CrossRef
  76. Krzywinski M, Schein J, Birol I, Connors J, Gascoyne R, Horsman D, et al. 2009. Circos: an information aesthetic for comparative genomics. Genome Res. 19: 1639-1645.
    Pubmed PMC CrossRef

Article

Research article

J. Microbiol. Biotechnol. 2023; 33(7): 875-885

Published online July 28, 2023 https://doi.org/10.4014/jmb.2301.01011

Copyright © The Korean Society for Microbiology and Biotechnology.

Comparative Genomic Analysis and BTEX Degradation Pathways of a Thermotolerant Cupriavidus cauae PHS1

Chandran Sathesh-Prabu1†, Jihoon Woo1†, Yuchan Kim1†, Suk Min Kim1 , Sun Bok Lee2 , Che Ok Jeon3*, Donghyuk Kim1*, and Sung Kuk Lee1*

Chandran Sathesh-Prabu1, Jihoon Woo1 and Yuchan Kim1 contributed equally to this work.

1School of Energy and Chemical Engineering, Ulsan National Institute of Science and Technology (UNIST), Ulsan 44919, Republic of Korea
2Department of Chemical Engineering, Pohang University of Science and Technology (POSTECH), Pohang, Gyeongbuk 37673, Republic of Korea
3Department of Life Science, Chung-Ang University, Seoul 06974, Republic of Korea

Correspondence to:Che Ok Jeon,         cojeon@cau.ac.kr
Donghyuk Kim,      dkim@unist.ac.kr
Sung Kuk Lee,       sklee77@gmail.com

Received: January 4, 2023; Revised: March 30, 2023; Accepted: March 31, 2023

Abstract

Volatile organic compounds such as benzene, toluene, ethylbenzene, and isomers of xylenes (BTEX) constitute a group of monoaromatic compounds that are found in petroleum and have been classified as priority pollutants. In this study, based on its newly sequenced genome, we reclassified the previously identified BTEX-degrading thermotolerant strain Ralstonia sp. PHS1 as Cupriavidus cauae PHS1. Also presented are the complete genome sequence of C. cauae PHS1, its annotation, species delineation, and a comparative analysis of the BTEX-degrading gene cluster. Moreover, we cloned and characterized the BTEX-degrading pathway genes in C. cauae PHS1, the BTEX-degrading gene cluster of which consists of two monooxygenases and meta-cleavage genes. A genome-wide investigation of the PHS1 coding sequence and the experimentally confirmed regioselectivity of the toluene monooxygenases and catechol 2,3-dioxygenase allowed us to reconstruct the BTEX degradation pathway. The degradation of BTEX begins with aromatic ring hydroxylation, followed by ring cleavage, and eventually enters the core carbon metabolism. The information provided here on the genome and BTEX-degrading pathway of the thermotolerant strain C. cauae PHS1 could be useful in constructing an efficient production host.

Keywords: BTEX, Cupriavidus cauae, degradation, genome analysis, thermotolerant

Introduction

Soil and water contamination by petroleum hydrocarbons, including monoaromatic and polycyclic aromatic hydrocarbons, is a global environmental and health concern [1, 2]. Volatile organic compounds such as benzene, toluene, ethylbenzene, and three isomers of xylenes (BTEX) are the most common monoaromatic compounds in petroleum and have been classified as priority pollutants by the U.S. Environmental Protection Agency [3]. The high cost and inefficiency of physical and chemical methods for cleaning up contaminated sites have led to increased focus on BTEX removal by hydrocarbon-degrading bacteria through biodegradation and bioremediation, which are cost-effective, eco-friendly, and highly efficient biotechnology-based processes [4, 5]. Several microorganisms, including members of the genera Pseudomonas, Ralstonia, Burkholderia, Sphingomonas, Thauera, Dechloromonas, Rhodococcus, Cladophialophora, Magnetospirillum, Pseudoxanthomonas, Comamonas, Bacillus, Microbacterium, and Acinetobacter, have been reported to degrade all or some components of BTEX [6-14].

BTEX compounds are degraded through various metabolic pathways depending on the microorganism [12, 15, 16]. In general, aerobic degradation of BTEX begins with the incorporation of an oxygen atom and is catalyzed by monooxygenases or dioxygenases. The resulting catechol intermediates are then mineralized through either the ortho- or meta-cleavage pathway and catalyzed by catechol 1,2 dioxygenase or catechol 2,3 dioxygenase (C23O), respectively. This ultimately produces low-molecular mass compounds, such as aldehydes and pyruvate, which are further oxidized and incorporated into the metabolic pathways as sources of carbon and energy [10, 12, 16].

Monoaromatic hydrocarbon-degrading bacteria often have a broad range of substrate capabilities due to the multicomponent mono/di-oxygenases they express. For example, even though no gene encoding a benzene monooxygenase for the conversion of benzene to phenol was identified in Paraburkholderia aromaticivorans BN5, the toluene 4-monooxygenase (tmoABCDE) and phenol 2-hydroxylase (dmpKLMNOP) can catalyze the oxidation of benzene [15]. The naphthalene 1,2-dioxygenase of Pseudomonas sp. strain NCIB 9816-4 can convert all three xylene isomers into benzyl alcohols and aldehyde derivatives [17]. This low substrate specificity allows the bacteria to degrade related molecules through the same catabolic pathway. In contrast, the xylene monooxygenase of P. putida mt-2 or B. amyloliquefaciens subsp. plantarum strain W1 was found to only convert p-xylene and m-xylene, but not o-xylene [12, 18]. Generally, the three isomers of xylene are catabolized through the oxidation of a methyl group of xylene to 2-methylbenzyl alcohol by xylene monooxygenase or the direct oxidation of the aromatic ring by xylene dioxygenase [15, 19, 20].

Previously, we isolated and characterized the first BTEX-degrading thermotolerant bacterium, designated as PHS1, from a hot spring in Pohang, Korea. This bacterium, Ralstonia sp. PHS1, can grow efficiently on a variety of aromatic substrates at 35-45°C (optimally at 42°C) [8]. In mixtures of BTEX compounds, the strain could degrade all six BTEX components, with benzene being degraded the fastest, followed by toluene, o-xylene, ethylbenzene, and finally m- and p-xylene.

The m-xylene and p-xylene were found to be co-metabolized by strain PHS1 in the presence of the growth-supporting BTEX compounds [21]. The unique properties of strain PHS1, such as thermotolerance and the ability to degrade o-xylene, give the strain an advantage in the field of environmental restoration or for the treatment of hot industrial effluents contaminated with BTEX [8]. It is worth noting that among BTEX, o-xylene is the most recalcitrant compound in terms of microbial degradation, and only a limited number of microbial species capable of metabolizing o-xylene have been reported [5, 22, 23]. In addition, to fully grasp the physiological and genomic features of a particular strain, it is important that its complete genomic information be available. Therefore, in this study, we reclassified the PHS1 strain as C. cauae PHS1 and report its complete genome sequence, annotation, and species delineation along with a comparative analysis of the BTEX-degrading gene cluster, and proposed BTEX-degrading pathways.

Materials and Methods

Whole Genome Sequencing

To extract genomic DNA, strain PHS1 was cultured overnight in Luria-Bertani broth (LB; composition per liter: 5 g yeast extract, 10 g peptone, and 10 g NaCl) at 42°C with shaking at 200 rpm. The cells were collected by centrifugation at 12,000 ×g for 15 min at 4°C, and the LaboPass™ Genomic DNA Isolation Kit (Cosmogenetech Co., Ltd., Korea) was used to isolate the genomic DNA of PHS1.

For PacBio long-read genome sequencing library construction, 8 μg of gDNA was used, and the sequencing library was constructed using the PacBio DNA Template Prep Kit 1.0 (Pacific Biosciences, USA) following the manufacturer's instructions. The final SMRTbell library was sequenced using PacBio DNA Sequencing Kit 4.0 (Pacific Biosciences) and 8 SMRT cells with P6-C4 chemistry on the PacBio RS II sequencing platform (Pacific Biosciences). For the Illumina short-read library, 100 ng of gDNA was fragmented using a Covaris instrument (Covaris Inc., USA), and the library was prepared using a TruSeq Nano DNA High-Throughput Library Prep Kit (Illumina Inc., USA) according to the manufacturer's instructions. The library was then sequenced on the HiSeq platform (Illumina) after PCR amplification. The libraries were validated using an Agilent 2100 Bioanalyzer (Agilent Technologies, USA).

De novo assembly was performed using HGAP v3.0 based on PacBio reads, and gap-filling and error correction of the draft assemblies were performed using Pilon v1.21 with the aid of Illumina reads.

Isolation of a BTEX-Degrading Gene Cluster and Plasmid Construction

All plasmids and strains used in this study are listed in Table S1. All DNA manipulations were performed in Escherichia coli TOP10 (USA). A genomic library was constructed by ligating partially Sau3AI-digested genomic DNA fragments of 5 to 15 kb from strain PHS1 into the pBluescript vector (pSK), which was digested with BamHI and treated with alkaline phosphatase. The DNA sequences were analyzed using Sanger DNA di-deoxy dye-terminator sequencing (Bionex Co., Ltd., Korea).

Transformants containing aromatic oxygenases were identified on LB agar plates containing 2 mM indole [24]. After incubating them for 3 days, the colonies were checked for the accumulation of blue pigment, which is a result of indigo formation. Two recombinant cells containing pSB1 or pSB2 were selected for further study. To analyze the initial ring-oxidation metabolites produced by BTEX monooxygenase 1 (Btxm1), pSB1-1 was generated by KpnI-digestion and self-ligation of pSB1 to eliminate beta-cleavage genes and prevent further reactions. pSB2 was used for functional analysis of BTEX monooxygenase 2 (Btxm2). All genes on these plasmids were expressed under the transcriptional control of the lac promoter in the pBluescript vector in E. coli TOP10.

Metabolite Isolation and Identification

E. coli cells containing pSB1-1 or pSB2 grown in LB medium were induced with 100 μM IPTG for 6 h, and then washed twice in 0.1 M potassium phosphate buffer (pH 7.0). The cells were resuspended in the same buffer and supplemented with 1.5 mM toluene, and then shaken at 37°C for 2 h. The metabolite extraction and gas chromatography analysis were performed using a previously described procedure [8]. Metabolites produced from the oxidation of toluene were identified by comparing their retention times with those of authentic samples.

Genomic Analysis

The genomic content of strain PHS1 was annotated based on the COG database (cog-20), UniProt (Swiss-Prot, TrEMBL), BRENDA, and KEGG using DIAMOND v2.0.11 [25-28] and BlastKOALA [29]. The subjects with the highest e-value, as well as identity and query coverage over 50%, were used for analysis based on the BLAST (DIAMOND) results. To predict genomic islands in the PHS1 genome, IslandViewer 4 was used with three methods: IslandPath-DIMOB, SIGI-HMM, and IslandPick [30].

To find gene candidates that may be involved in BTEX degradation, sequences of proteins involved in aromatics degradation were collected from the KEGG orthology and NCBI database. The collected sequences included: xylene monooxygenase, benzoate (toluate) 1,2-dioxygenase, benzylalcohol dehydrogenase, benzaldehyde dehydrogenase, 4-hydroxybenzaldehyde dehydrogenase, 4-cresol dehydrogenase, 4-hydroxybenzoate dehydrogenase, phenylpropionate dehydrogenase, dihydroxycyclohexadiene carboxylate dehydrogenase, naphthalene dioxygenase, ethylbenzene dioxygenase, ethylbenzene hydroxylase, toluene methyl-monooxygenase, aryl-alcohol dehydrogenase, phenol 2-monooxygenase, and benzene/toluene/chlorobenzene dioxygenase. The collected protein sequences were compared with all the coding sequences of the strain using NCBI BLAST+ 2.11.0.

Phylogenetic Analysis

For phylogenetic analysis, the 16S rRNA gene sequences of strain PHS1 and Cupriavidus representative strains were extracted from their corresponding whole genome sequences based on gene annotations. The 16S rRNA gene sequences of the PHS1 strain and closely related type strains were aligned using the fast secondary-structure aware infernal aligner available at the Ribosomal Database Project (https://pyro.cme.msu.edu/aligner/form.spr). A phylogenetic tree based on the 16S rRNA gene sequences was constructed using the neighbour-joining algorithm with bootstrap values (1,000 replications) using MEGA11 software [31]. The 16S rRNA gene sequence similarity values between the PHS1 strain and closely related type strains were calculated using the Nucleotide Similarity Search program on the EzBioCloud server (www.ezbiocloud.net/identify). For the genome-based phylogenetic analysis, the up-to-date bacterial core gene pipeline (UBCG; https://help.ezbiocloud.net/ubcg-gene-set/) (Kim et al., 2021) was used to extract 92 core genes from the genomes of the PHS1 strain and closely related type strains, which were then concatenated and aligned. A maximum-likelihood tree based on the concatenated nucleotide sequences of the 92 core genes was reconstructed using MEGA11 software.

The average nucleotide identity (ANI) and digital DNA-DNA hybridization (DDH) values between the PHS1 strain and closely related type strains were calculated using the Orthologous Average Nucleotide Identity Tool (OAT) software available on the EzBioCloud webserver (www.ezbiocloud.net/sw/oat) [33] and the server-based Genome-to-Genome Distance Calculator version 2.1 (http://ggdc.dsmz.de/distcalc2.php) [34], respectively.

To construct a single-gene tree of bacterial monooxygenase alpha subunits, pairwise alignments of collected protein sequences from reference articles were carried out using MUSCLE built in MEGA 11 with the default option. The gene tree was constructed using the maximum-likelihood method with the Jones-Taylor-Thornton (JTT) model and 1,000 iterations [31].

Results and Discussion

Reclassification and Species Delineation of the PHS1 Strain

Previously, strain PHS1 was identified as belonging to the genus Ralstonia [8]. However, the 16S rRNA gene sequence of PHS1 was re-examined in the present study using various computational tools and algorithms. The 16S rRNA gene sequence of strain PHS1 and closely related type strains were aligned using the fast secondary-structure aware infernal aligner available at the Ribosomal Database Project (https://pyro.cme.msu.edu/aligner/form.spr). A phylogenetic tree based on 16S rRNA gene sequences, constructed using the neighbour-joining algorithm, showed that strain PHS1 was not a species of Ralstonia (Fig. S1).

For genome-based phylogenetic analysis, a maximum-likelihood tree showing the phylogenetic relationships between strain PHS1 and closely related taxa was constructed (Fig. 1). The highest sequence similarity was observed with species belonging to the genus Cupriavidus.

Figure 1. Phylogenomic tree based on the concatenated 92 core genes showing the phylogenetic relationships between strain PHS1 and closely related taxa. Numbers on nodes correspond to bootstrap values for branches (1,000 replicates), shown only bootstrap values over 70%. R. insidiosa AU2944 was used as the outgroup. Bar, 0.02 substitutions per nucleotide.

Furthermore, the average nucleotide identity (ANI) and digital DNA-DNA hybridization (DDH) values between strain PHS1 and closely related type strains were calculated [33,34]. Strain PHS1 showed very high similarities with C. cauae MKL-01 and C. gilardii FDAARGOS_639, with 99.87% and 99.74% 16S rRNA similarities and 98.33% and 91.33% ANI values, respectively (Table 1). Although the ANI value of C. cauae was much higher than that of C. gilardii, there was no significant difference observed in 16S rRNA similarities. Considering the possibility that strain PHS1 may be closer to other strains of C. gilardii, further comparisons were carried out using ANI and DDH with all strains of C. cauae (1) and C. gilardii (10) deposited in the NCBI assembly database (Table S2). Most comparisons with C. gilardii strains showed low ANI and DDH values below the threshold for delineating the same species (ANI: 95%; DDH: 70%) [35], except for the second highest values for C. gilardii W2-2 with 98.16% and 85.5% values, respectively (Table S2).

Table 1 . 16S rRNA and ANI comparison of Cupriavidus representative strains with strain PHS1..

Strains16S rRNA BLAST identity16S rRNA BLAST coverageANI value*
C. cauae MKL-0199.879998.33
C. gilardii FDAARGOS_63999.749991.33
C. nantongensis X198.769982.72
C. oxalaticus T298.449982.50
C. respiraculi LMG 2151098.769982.31
C. neocaledonicus STM 616098.569981.91
C. taiwanensis LMG 1942498.839981.80
C. alkaliphilus ASC-869NANA81.79
C. necator H1698.379981.70
C. lacunae S23NANA81.68
C. campinensis S14E4C98.3010081.58
C. malaysiensis USMAA102098.7010081.34
C. laharis LMG 2399298.059981.23
C. yeoncheonensis LMG 31506NANA81.03
C. basilensis DSM 1185397.989981.00
C. numazuensis LMG 26411NANA80.82
C. agavae ASC-984298.249980.79
C. pinatubonensis HN-298.2410080.46
C. pauculus MF198.899980.40
C. metallidurans NDB4MOL198.8910080.24
C. pampae LMG 3228998.5010079.92
C. plantarum MA2-19b99.2210079.82

*Average nucleotide identity; All the values are expressed as a percentage (%)..

NA, not available..



The ANI and digital DDH values between strain PHS1 and C. cauae MKL-01, the most closely related type strain, were less than 98.3% and 86.0%, respectively. This suggests that strain PHS1 represents a strain member of C. cauae.

General Characteristics of the PHS1 Genome

The genomic features of strain PHS1 are summarized in Table 2. The genome contains two circular chromosomes, with a total of 5,075 genes (Fig. 2). This is similar to the genome structure of Cupriavidus, which also has two circular chromosomes [36]. The sizes of the chromosomes in strain PHS1 are 3,527,866 bp (3,120 genes) and 2,275,253 bp (1,955 genes), respectively. Additionally, 126 pseudogenes, 4 complete rRNA clusters, and 57 tRNA genes were predicted from annotation by the NCBI PGAP pipeline. Of these, 3 complete rRNA clusters and 51 tRNA genes are located on chromosome 1. The average GC content of the PHS1 genome is 67.73%, with similar GC content for each chromosome (67.77% and 67.68%, respectively). Genomic islands were also predicted using IslandViewer4 and a total of 52 were found in the PHS1 genome.

Table 2 . General characteristics of the PHS1 strain genome..

FeatureChromosome 1Chromosome 2Total
Bioproject no.PRJNA749683
Biosample no.SAMN20397041
GenBank accession no.GCA_026210475.1
Size (bp)3,527,8662,275,2535,803,119
G+C content (%)67.7767.6867.73
Total genes3,1201,9555,075
Protein coding sequences3,0571,9464,877
Pseudogenes7650126
Complete rRNAs314
tRNA genes51657
ncRNA, regulatory, etc.303
Genomic islands312152


Figure 2. Genomic characteristics of strain PHS1. Circular visualization of two chromosomes in strain PHS1. Orange and blue tiles indicate the CDS with forward and reverse direction. Yellow tiles indicate pseudogenes. Purple tiles indicate noncoding RNA including rRNA and tRNA. Black histogram indicates the GC contents of each coding sequence. The genomic island including the btx operon is represented as dotted lines and red histogram. Data were visualized using Circos [76].

Clusters of orthologous genes (COG) analysis was conducted on each genome to investigate the distribution of genes in each chromosome. The predicted CDS of the PHS1 genome were compared with protein sequences in the COG database [27]. Among the 26 categories in the COG database, genes involved in 11 categories (J, L, D, V, M, N, W, U, O, F, and H) were unevenly distributed across the two chromosomes (Fig. S2). In other categories, such as K (transcription) or C (energy production and conversion), the related genes were evenly distributed across the two chromosomes. These distributions and the conservation of the two circular chromosomes suggest that they may play complementary roles in various aspects of cellular functions, not just metabolism. The complete genome information for strain PHS1 has been deposited in NCBI under the GenBank assembly accession number GCA_026210475.1 (https://www.ncbi.nlm.nih.gov/assembly/GCA_026210475.1).

Comparative Analysis of the BTEX-Degrading Gene Cluster

The BTEX-degrading gene cluster (BDGC) identified in the genome of strain PHS1 was located at the position 2,217,391 - 2,240,837 in chromosome 1. We found it consisted of genes encoding two monooxygenases, Btxm1 and Btxm2, as well as enzymes involved in the meta-cleavage pathway (Fig. 3A). The functions of these genes were predicted bioinformatically and are summarized in Table 3. The genomic DNA library of strain PHS1 was screened to clone the genes for BTEX catabolism using the enzymatic oxidation of indole to indigo by aromatic oxygenases [37]. Two types of aromatic oxygenases with different characteristics were isolated from the genomic library. In particular, the accumulation of a ring fission product [38], indicated by a yellow color, was observed in E. coli TOP10 cell cultures harboring the pSB1 plasmid (Fig. 3A) when exposed to toluene. This indicates that the pSB1 insert codes for enzymes are involved in the transformation of toluene into the corresponding meta-cleavage products catalyzed by C23O [39]. The other type of oxygenase was found in the strain TOP10 harboring pSB2 (Fig. 3A). These findings suggest that strain PHS1 has the ability to efficiently degrade BTEX compounds through the meta-cleavage pathway, using the Btxm1 and Btxm2 monooxygenases and the associated enzymes.

Table 3 . Gene contents of BTEX-degrading gene cluster..

Locus tagGeneLocationStrand% G+CLengthProduct
KZ686_09940btxA2217391-2217615+55.5674Phenol hydroxylase subunit (PH)
KZ686_09945btxB2217703-2218698+56.63331Phenol hydroxylase subunit (β)
KZ686_09950btxC2218738-2219007+51.8589Phenol hydroxylase subunit
KZ686_09955btxD2219030-2220589+55.19519Phenol hydroxylase subunit (α)
KZ686_09960btxE2220586-2220942+58.82118Phenol hydroxylase subunit (γ)
KZ686_09965btxF2221021-2222085+57.37354Phenol hydroxylase subunit
KZ686_09970btxG2222088-2222444+59.38118Ferredoxin (Fn)
KZ686_09975btxH2222462-2223406+55.87314Catechol 2,3-dioxygenase (C23O)
KZ686_09980btxI2223428-2223877+61.78149Unknown protein
KZ686_09985IS2223931-2224748-61.37IS element family transposase
KZ686_09990btxJ2224862-2226373+63.295032-Hydroxymuconic semialdehyde dehydrogenase (HMSD)
KZ686_09995btxK2226376-2227158+61.562602-hydroxypent-2,4-dienoate hydratase (OEH)
KZ686_10000btxL2227237-2228148+61.07303Acetaldehyde dehydrogenase (ADA)
KZ686_10005btxM2228168-2229214+62.183484-Hydroxy-2-oxovalerate aldolase (HOA)
KZ686_10010btxN2229211-2229999+62.862624-oxalocrotonate decarboxylase (4OD)
KZ686_10015btxO2230011-2230202+59.38634-Oxalocrotonate tautomerase (4OT)
KZ686_10020btxP2230262-2231764+57.49500Toluene monooxygenase subunit (α) (TMO)
KZ686_10025btxQ2231833-2232099+58.4388Toluene monooxygenase subunit (γ)
KZ686_10030btxR2232155-2232490+54.46111Ferredoxin
KZ686_10035btxS2232531-2232845+56.83104Toluene monooxygenase subunit
KZ686_10040btxT2232900-2233886+56.33328Toluene monooxygenase subunit (β)
KZ686_10045btxU2233986-2235008+55.33340Toluene monooxygenase subunit
KZ686_10050btxV2235074-2235679+53.14201Glutathione S-transferase (GST)
KZ686_10055IS2235750-2236567-60.27IS element family transposase
KZ686_10060adh2236737-2237798+55.08353Alcohol dehydrogenase
KZ686_100652237943-2239271+61.63442Putative transporter
KZ686_10070btxW2240007-2240837+56.922762-Hydroxymuconate-semialdehyde hydrolase (HMSH)


Figure 3. Gene arrangements and comparison of genes in BTEX-degrading gene cluster. (A) Genes encoding two toluene monooxygenase (green) and meta-cleavage pathway (blue) were adjacently located in chromosome 1 of strain PHS1 genome. Screened plasmids, pSB1, pSB1-1, and pSB2, are indicated as black arrow. IS denotes insertion sequence. (B) The integration site of the approximately 75 kbp genetic fragment containing the btx operon. The nucleotide sequences flanking the insertion sequence in PHS1 were almost identical to those in MKL-01, except for 12 bp duplicated sequences. (C) Protein sequences of genes in BTEX-degrading gene cluster are compared with Cupravidus representative strains and several strains reported as degraders of aromatics. (D) Protein sequence tree of known aromatic monooxygenases subunit α were reconstructed based on multiple sequence alignments. Numbers at nodes indicate the percentage of node resampled.

After comparing the nucleotide sequences of the BDGC with representative strains of Cupriavidus, we found that C. metallidurans CH34 showed full conservation of the BDGC (Fig. 3C). Additionally, analysis of the BDGC with other genera known to be involved in aromatic degradation revealed that the Btxm1 and meta-cleavage pathway of B. vietnamiensis G4 and Pseudomonas sp. CF600 are closely related and homologous to the BDGC. The BDGC (comprising Btxm1, Btxm2, and the meta-cleavage pathway) was also conserved in P. aromaticivorans BN5, with nearly 100% similarity at the nucleotide level (Fig. 3C) [15, 40-45]. Interestingly, no genes in the BDGC were conserved in closely related strains, such as C. cauae and C. gilardii, but were rather conserved in phylogenetically distant strains. Additionally, the GC content of the BDGC (57.5%) was significantly lower than that of the genome (67.73%). It was also predicted that the BDGC is a genomic island using IslandPath-DIMOB, SIGI-HMM, and IslandPick algorithms in the IslandsViewer 4 software (Fig. 2) [30]. These findings suggest that the BDGC in strain PHS1 was likely acquired horizontally rather than being conserved across genera or species.

To determine whether any genes other than the BDGC may be involved in the biodegradation of BTEX compounds, the entire PHS1 genome was annotated using various databases, including UniProt, KEGG, and BRENDA (Table S3). Then, all hydroxylases and oxygenases that might be involved in the early stages of BTEX degradation were investigated. As a result, none of the genes were found to be potentially involved in BTEX degradation, except for KZ18760-KZ18775, which were predicted to be dioxygenases for aromatic compounds [46-48].

In addition to analyzing the BDGC, the protein orthologous sequences known to be involved in the early stages (before cleavage of the aromatic ring) of BTEX degradation were collected and compared to the protein sequences of the PHS1 strain genome using BLAST [49]. For example, a total of 281 orthologous sequences of p-cresol dehydrogenase (pchCF), which catalyzes the methyl group of p-cresol (a derivative of toluene) to yield p-hydroxybenzylaldehyde, were collected. The BLAST results showed that the highest match of pchF orthologs had only 27.1% sequence similarity with the PHS1 genomic content, and none of the protein sequences in strain PHS1 genome were aligned with pchC orthologs. This suggests that the PHS1 strain does not encode p-cresol dehydrogenase in its genome. Following an examination of the gene candidates that were listed in the Materials and Methods section as potential contributors to BTEX degradation, we found that strain PHS1 lacks any known enzymes that support the early stages of BTEX degradation, except for those found in the btx operon (as shown in Table S4). As such, we have concluded that the PHS1 strain does not harbor any additional gene candidates involved in BTEX degradation.

To further investigate the differences between the C. cauae PHS1 and C. cauae MKL-01 genomes, the coding sequences from both strains were compared based on an ortholog search. A total of 4,369 orthologous genes were identified between the two genomes using the Reciprocal Best Hit method with a Smith-Waterman algorithm, with a 50% identity and 80% coverage threshold in both directions [50]. A total of 508 and 511 non-orthologous genes were respectively identified in PHS1 and MKL-01 (Fig. S3). The BTEX-degrading operon (btx operon) was not present in MKL-01. Most of the genes within the btx operon were classified as COG category Q (secondary metabolites biosynthesis, transport, and catabolism). The number of COG category X genes present in PHS1, which are indicative of the mobilome involved in horizontal gene transfer, was 25, a number greater than the 13 found in MKL-01. Additionally, we identified 12 COG category U genes in PHS1 that are associated with intracellular trafficking, secretion, and vesicular transport. Of these 12 genes, 11 were located adjacent to the btx operon. Notably, 10 of these genes upstream of the btx operon appeared to form the trb operon and traG gene, which are known to play a crucial role in facilitating the conjugative transfer of plasmids (KZ686_09850, KZ686_09855, KZ686_09860, KZ686_09865, KZ686_09875, KZ686_09880, KZ686_09885, KZ686_09890, KZ686_09895, and KZ686_09905) [51]. These findings suggest that horizontal gene transfer may occur in PHS1, as shown in Figs. S2 and 3B. The integration site of the btx operon in MKL-01 was analyzed by comparing the conserved nucleotide sequences of the flanking region with those of the strain PHS1. The genomic island containing the btx operon was 74,621 bp and may have been integrated into the 217,057 to 217,068 region of contig 32 of the MKL-01 genome (an incompletely assembled genome). The flanking genes of the MKL-01 strain were F1599_RS10900 and F1599_RS10905 (Fig. S3B).

BTEX-Monooxygenase 1 (Btxm1)

Analysis of the nucleotide sequences of the Btxm1 genes (btxABCDEF) showed that they were closely related in gene order and homologous to a group of T2MOs from Burkholderia sp. JS150 and B. cepacia G4, and phenol hydroxylases (PH) from C. testosteroni TA441, P. stutzeri OX1, and P. putida CF600 [52-55] (Fig. 3D). Specifically, BtxABCDEF showed high overall sequence identity (> 98%) to TomABCDEF from the approximately 108-kb degradative plasmid TOM of B. cepacia G4, which is known to metabolize o-xylene as a carbon source [56].

BtxD, the largest polypeptide of Btxm1, was homologous to the large oxygenase subunit of other bacterial multicomponent monooxygenases (BMM) containing two dinuclear iron-binding domains with the amino acid sequence Asp-Glu-X-Arg-His, which is found in several enzymes that catalyze reactions involving activated oxygen [57]. BtxB shares homology with the small oxygenase subunits of PHs and TMOs. BtxF is homologous to a number of other bacterial iron-sulfur flavoproteins that serve as oxidoreductases for several enzyme systems, including monooxygenases, aromatic dioxygenases, and reductases involved in the biosynthesis of deoxy-sugars [58]. Database comparisons using the deduced amino acid sequences of btxA, btxC, and btxE as query sequences showed that each of these three gene products has homologous counterparts in PHs and T2MOs.

Meta-Cleavage Pathway

The meta-pathway genes (btxGHIJKLMNO) are closely related to the corresponding phenol-degrading aph and dmp genes of the strains C. testosteroni TA441 and Pseudomonas sp. CF600, respectively (Fig. 3). In particular, the meta-pathway gene products of Burkholderia sp. RP007 that degrade polycyclic aromatic hydrocarbons (PAHs), PhnT2/E2/X, showed 95% overall sequence identity to BtxG/H/I, respectively. BtxG, which is homologous to ferredoxins, contains the amino acid motif Cys-XXXX-Cys-XX-Cys that is characteristic of chloroplast-type ferredoxins [59]. In addition to PhnE2 of Burkholderia sp. RP007, the PHS1 C23O (btxH) has 97% identity with TomB from o-xylene-degrading B. cepacia G4. A fragment encoding an ORF of 149 amino acids was found downstream of C23O, which is preceded by a Shine-Dalgarno-like sequence, AGGAGA. The predicted amino acid sequence of this ORF shares sequence identity with an unknown protein and has no assigned function that is associated with other catabolic gene clusters, including meta-cleavage genes in Burkholderia sp. RP007 [60], C. testosteroni TA441 [61], and Sphingomonas sp. HV3 [13]. A putative 2-hydroxymuconic semialdehyde dehydrogenase (BtxJ) gene is located downstream of the insertion sequence (IS), starting 23 bp downstream of the end of the IS. The btxJ is followed by the genes btxKLM, which encode 2-hydroxypent-2,4-dienoate hydratase, acetaldehyde dehydrogenase, and 4-hydroxy-2-oxovalerate aldolase, respectively. These enzymes catalyze the final three steps of the meta-pathway that are common to other meta-degradation pathways for various aromatic compounds. The genes for the 4-oxalocrotonate branch enzymes, 4-oxalocrotonate decarboxylase (BtxN) and 4-oxalocrotonate tautomerase (BtxO), follow BtxM [62].

Most meta-cleavage pathways can be classified into two types [60]: P-type (gene order: Fn, C23O, HMSD, HMSH) from Pseudomonas spp. [44] and S-type (gene order: GST, HMSH, C23O, U, HMSD) from Sphingomonas spp. [13,63]. However, the meta-pathway of strain PHS1 (gene order: Fn, C23O, U, HMSD) is a mixture of both types, similar to the PAH-degrading pathway of Burkholderia RP007 [60]. Homologous polypeptides were also found in other enzymes, particularly in the phenol-degrading aph and dmp systems of C. testosteroni TA441 and Pseudomonas sp. strain CF600, respectively, and the PCB-degrading bph system of Pseudomonas sp. strain LB400 [64]. The hydrolytic branch involving the enzyme HMSH (BtxW), which corresponds to DmpD in strain CF600, is required for the degradation of 2- and/or 3-methylated phenols [44]. However, the gene for HMSH was not found in the aph and btx gene clusters of strain TA441 and PHS1, respectively [61]. Unlike strain TA441, strain PHS1 is able to utilize various methylphenols such as o-, m-, p-cresol, and 2,3-dimethylphenol [8]. This suggests that the gene for HMSH may exist downstream of the GST gene, as in the S-type meta-pathway (GST-HMSH-C23O-U-HMSD) [60]. In strain PHS1, the HMSH gene is immediately downstream of the btx cluster and designated as KZ686_10070, with over 70% similarity to the HMSH gene of strain CF600. In C metallidurans CH34 (GenBank accession no. NC 007973), the HMSH gene is located downstream of the GST gene.

BTEX-Monooxygenase 2 (Btxm2)

The Btxm2 enzyme is composed of BtxPQRSTU and is similar to toluene/o-xylene monooxygenase (ToMO). BtxP and BtxT are similar to the large and small oxygenase subunits, respectively. Like BtxD in Btxm1, BtxP contains a pair of dinuclear iron-binding ligands with the amino acid sequence Asp-Glu-X-Arg-His between amino acids 133 and 137, and 230 and 234. BtxU is similar to the NADH-ferredoxin oxidoreductase components of several mono- and dioxygenase systems. BtxS is a coupling/effector protein that affects the activity and specificity of other bacterial oxygenase systems [65]. BtxR has a region between amino acids 40 and 70 that can be aligned with a domain found in many proteins classified as ferredoxin types. In addition, two conserved Cys residues followed by His residues (Cys45, His47, and Cys64, His67 in BtxR), which are believed to be involved in coordinating a Rieske-type iron-sulfur cluster, are present [66].

Based on multiple sequence alignments of the 8 alpha subunits of known TMOs, an evolutionary tree of protein sequences was constructed (Fig. 3D). Btxm2 is similar to the benzene monooxygenase from P. aeruginosa JI104, ToMO from P. stutzeri OX1, toluene 3-monooxygenase (T3MO) from R. pickettii PKO1, and toluene 4-monooxygenase (T4MO) from P. mendocina KR1 [67-70].

Proposed BTEX Biodegradation Pathways in PHS1

Previously, we reported that the main products of benzene and o-xylene degradation by strain PHS1 were phenol and 2,3-dimethylphenol, respectively, based on mass spectrometric analysis [8]. These observations suggest that the enzyme involved in the first step of individual BTEX degradation directly mono-oxidizes the aromatic nucleus [71, 72].

Toluene has been studied as a representative model compound of monoaromatic hydrocarbons [73]. Btxm1 only produced o-cresol from toluene (Fig. 4A), indicating that its activity is similar to that of the multicomponent T2MO from B. cepacia G4 [52] and B. cepacia JS150 [74]. In contrast, Btxm2 produced o-, m-, and p-cresol compounds (Fig. 4A), indicating that Btxm2 has low regioselectivity, similar to the benzene monooxygenase of P. aeruginosa JI104 [67] and ToMo of Pseudomonas sp. OX1 [24], which can produce three cresol compounds from toluene. The three isomers of cresols may then be further converted into 3-methylcatechol or 4-methylcatechol by Btxm1 or Btxm2 activity. The phenol hydroxylase (PH) from P. stutzeri OX1 is known to have catalytic activity for converting these three isomers of cresols into 3-methlycatechol (from o-cresol and m-cresol) and 4-methylcatechol (from p-cresol) [38]. Next, 3- or 4-methylcatechol may be cleaved by catechol 2,3-dioxygenase (C23O) to yield ring cleavage products, 2-Hydroxy-6-oxo-hept-2,4-dienoate or (2E,4Z)-2-Hydroxy-5-methyl-6-oxohexa-2,4-dienoate, depending on the substrate specificity of C23O of P. stutzeri OX1 [75]. Finally, the intermediates may be degraded by BtxIJKLMNO to yield pyruvate and acetyl-CoA, which are universally utilizable metabolites that enter the central carbon metabolism (Fig. 4B).

Figure 4. Proposed BTEX- biodegradation pathways in strain PHS1. (A) Gas chromatographs of metabolites produced from toluene by cells of TOP10. (a) Btxm1, (b) Btxm2, (c) standards (in order of o-, p-, m-cresol). (B) Reactions in green outline indicate the catalytic reactions of two toluene monooxygenase, while those in blue outline indicate the reactions catalyzed by meta-cleavage enzymes.

In the case of ethylbenzene degradation, it is likely to be initiated by Btxm2, considering the ethylbenzene selectivity of BMO from P. aeruginosa JI104, ToMO from P. stutzeri OX1, and T3MO from R. pickettii PKO1 [24, 67, 68]. The main products reported from ethylbenzene degradation are 3-ethylphenol (produced by BMO, ToMO, and T3MO) and 4-ethylphenol (produced by T3MO). Therefore, considering the phylogenetic relationship of toluene monooxygenases, it is likely that the degradation of ethylbenzene in strain PHS1 begins with the aromatic ring hydroxylation rather than the hydroxylation of the ethyl group by Btxm2.

The two toluene monooxygenases (Btxm1 and Btxm2) in the BDGC of strain PHS1 may catalyze the hydroxylation of aromatics, but they are expected to have different specificities for non-hydroxylated and monohydroxylated compounds, as well as regioselectivity related to the location of the methyl group. In P. stutzeri OX1, the Kcat/Km value for benzene with ToMO is much larger than that of phenol, while PH is much more effective at converting phenol than benzene [53]. In addition, PH can act on various monohydroxylated aromatics, such as cresol and dimethylphenol [38]. It appears that Btxm1 and Btxm2 also catalyze different steps in the two hydroxylations of aromatics in a complementary manner. PH mainly catalyzes the second hydroxylation of toluene and xylene derivatives, as well as benzene, rather than the first one [38,53]. Therefore, it is speculated that Btxm2 mainly catalyzes the first hydroxylation of various aromatic compounds with low regioselectivity, and Btxm1 mainly catalyzes the second hydroxylation at the ortho-position of phenolic compounds, as well as the first hydroxylation at the ortho-position of toluene.

In conclusion, our results suggest that new catabolic pathways may have arisen through enzyme structural evolution to allow for a wider range of substrates or the acquisition of catabolic functions from different bacteria, in order to promote rapid adaptation to diverse environmental conditions. The BDGC of PHS1, comprising the two TMOs and the meta-pathway, is thought to have an efficient architecture with complementary activities of the two TMOs and cooperating meta-cleavage enzymes, as well as the advantage of a clustered structure of gene arrangements. Further analysis revealed that the BDGC is conserved in certain related strains but not in others, suggesting that it may have been acquired horizontally rather than being conserved across genera or species.

Supplemental Materials

Acknowledgments

This work was supported by the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT, and Future Planning (MSIT) (2020R1A4A1018332 and 2015M3D3A1A01064919). This research was also supported by the Innovative Science Project of the Circle Foundation in 2020.

Conflict of Interest

The authors have no financial conflicts of interest to declare.

Fig 1.

Figure 1.Phylogenomic tree based on the concatenated 92 core genes showing the phylogenetic relationships between strain PHS1 and closely related taxa. Numbers on nodes correspond to bootstrap values for branches (1,000 replicates), shown only bootstrap values over 70%. R. insidiosa AU2944 was used as the outgroup. Bar, 0.02 substitutions per nucleotide.
Journal of Microbiology and Biotechnology 2023; 33: 875-885https://doi.org/10.4014/jmb.2301.01011

Fig 2.

Figure 2.Genomic characteristics of strain PHS1. Circular visualization of two chromosomes in strain PHS1. Orange and blue tiles indicate the CDS with forward and reverse direction. Yellow tiles indicate pseudogenes. Purple tiles indicate noncoding RNA including rRNA and tRNA. Black histogram indicates the GC contents of each coding sequence. The genomic island including the btx operon is represented as dotted lines and red histogram. Data were visualized using Circos [76].
Journal of Microbiology and Biotechnology 2023; 33: 875-885https://doi.org/10.4014/jmb.2301.01011

Fig 3.

Figure 3.Gene arrangements and comparison of genes in BTEX-degrading gene cluster. (A) Genes encoding two toluene monooxygenase (green) and meta-cleavage pathway (blue) were adjacently located in chromosome 1 of strain PHS1 genome. Screened plasmids, pSB1, pSB1-1, and pSB2, are indicated as black arrow. IS denotes insertion sequence. (B) The integration site of the approximately 75 kbp genetic fragment containing the btx operon. The nucleotide sequences flanking the insertion sequence in PHS1 were almost identical to those in MKL-01, except for 12 bp duplicated sequences. (C) Protein sequences of genes in BTEX-degrading gene cluster are compared with Cupravidus representative strains and several strains reported as degraders of aromatics. (D) Protein sequence tree of known aromatic monooxygenases subunit α were reconstructed based on multiple sequence alignments. Numbers at nodes indicate the percentage of node resampled.
Journal of Microbiology and Biotechnology 2023; 33: 875-885https://doi.org/10.4014/jmb.2301.01011

Fig 4.

Figure 4.Proposed BTEX- biodegradation pathways in strain PHS1. (A) Gas chromatographs of metabolites produced from toluene by cells of TOP10. (a) Btxm1, (b) Btxm2, (c) standards (in order of o-, p-, m-cresol). (B) Reactions in green outline indicate the catalytic reactions of two toluene monooxygenase, while those in blue outline indicate the reactions catalyzed by meta-cleavage enzymes.
Journal of Microbiology and Biotechnology 2023; 33: 875-885https://doi.org/10.4014/jmb.2301.01011

Table 1 . 16S rRNA and ANI comparison of Cupriavidus representative strains with strain PHS1..

Strains16S rRNA BLAST identity16S rRNA BLAST coverageANI value*
C. cauae MKL-0199.879998.33
C. gilardii FDAARGOS_63999.749991.33
C. nantongensis X198.769982.72
C. oxalaticus T298.449982.50
C. respiraculi LMG 2151098.769982.31
C. neocaledonicus STM 616098.569981.91
C. taiwanensis LMG 1942498.839981.80
C. alkaliphilus ASC-869NANA81.79
C. necator H1698.379981.70
C. lacunae S23NANA81.68
C. campinensis S14E4C98.3010081.58
C. malaysiensis USMAA102098.7010081.34
C. laharis LMG 2399298.059981.23
C. yeoncheonensis LMG 31506NANA81.03
C. basilensis DSM 1185397.989981.00
C. numazuensis LMG 26411NANA80.82
C. agavae ASC-984298.249980.79
C. pinatubonensis HN-298.2410080.46
C. pauculus MF198.899980.40
C. metallidurans NDB4MOL198.8910080.24
C. pampae LMG 3228998.5010079.92
C. plantarum MA2-19b99.2210079.82

*Average nucleotide identity; All the values are expressed as a percentage (%)..

NA, not available..


Table 2 . General characteristics of the PHS1 strain genome..

FeatureChromosome 1Chromosome 2Total
Bioproject no.PRJNA749683
Biosample no.SAMN20397041
GenBank accession no.GCA_026210475.1
Size (bp)3,527,8662,275,2535,803,119
G+C content (%)67.7767.6867.73
Total genes3,1201,9555,075
Protein coding sequences3,0571,9464,877
Pseudogenes7650126
Complete rRNAs314
tRNA genes51657
ncRNA, regulatory, etc.303
Genomic islands312152

Table 3 . Gene contents of BTEX-degrading gene cluster..

Locus tagGeneLocationStrand% G+CLengthProduct
KZ686_09940btxA2217391-2217615+55.5674Phenol hydroxylase subunit (PH)
KZ686_09945btxB2217703-2218698+56.63331Phenol hydroxylase subunit (β)
KZ686_09950btxC2218738-2219007+51.8589Phenol hydroxylase subunit
KZ686_09955btxD2219030-2220589+55.19519Phenol hydroxylase subunit (α)
KZ686_09960btxE2220586-2220942+58.82118Phenol hydroxylase subunit (γ)
KZ686_09965btxF2221021-2222085+57.37354Phenol hydroxylase subunit
KZ686_09970btxG2222088-2222444+59.38118Ferredoxin (Fn)
KZ686_09975btxH2222462-2223406+55.87314Catechol 2,3-dioxygenase (C23O)
KZ686_09980btxI2223428-2223877+61.78149Unknown protein
KZ686_09985IS2223931-2224748-61.37IS element family transposase
KZ686_09990btxJ2224862-2226373+63.295032-Hydroxymuconic semialdehyde dehydrogenase (HMSD)
KZ686_09995btxK2226376-2227158+61.562602-hydroxypent-2,4-dienoate hydratase (OEH)
KZ686_10000btxL2227237-2228148+61.07303Acetaldehyde dehydrogenase (ADA)
KZ686_10005btxM2228168-2229214+62.183484-Hydroxy-2-oxovalerate aldolase (HOA)
KZ686_10010btxN2229211-2229999+62.862624-oxalocrotonate decarboxylase (4OD)
KZ686_10015btxO2230011-2230202+59.38634-Oxalocrotonate tautomerase (4OT)
KZ686_10020btxP2230262-2231764+57.49500Toluene monooxygenase subunit (α) (TMO)
KZ686_10025btxQ2231833-2232099+58.4388Toluene monooxygenase subunit (γ)
KZ686_10030btxR2232155-2232490+54.46111Ferredoxin
KZ686_10035btxS2232531-2232845+56.83104Toluene monooxygenase subunit
KZ686_10040btxT2232900-2233886+56.33328Toluene monooxygenase subunit (β)
KZ686_10045btxU2233986-2235008+55.33340Toluene monooxygenase subunit
KZ686_10050btxV2235074-2235679+53.14201Glutathione S-transferase (GST)
KZ686_10055IS2235750-2236567-60.27IS element family transposase
KZ686_10060adh2236737-2237798+55.08353Alcohol dehydrogenase
KZ686_100652237943-2239271+61.63442Putative transporter
KZ686_10070btxW2240007-2240837+56.922762-Hydroxymuconate-semialdehyde hydrolase (HMSH)

References

  1. Adipah S. 2019. Introduction of petroleum hydrocarbons contaminants and its human effects. J. Environ. Sci. Public Health 3: 1-9.
  2. U.S. EPA. 2004. Cleaning Up the Nation’s Waste Sites: Markets and Technology Trends, 2004 edition. 2004th Ed. Office of Solid Waste And Emergency Response, EPAU.S. EPA/National Service Center for Environmental Publications (NSCEP), Cincinnati, OH.
  3. Williams S., Ladd D., Farmer J. 2006. Scientific investigations report 2005-5104 Fate and Transport of petroleum hydrocarbons in soil and ground water at big south fork national river and recreation area, Tennessee and Kentucky, 2002-2003. U.S. Geol. Surv. Sci. Investig. Rep. 2005-5104: 29.
    CrossRef
  4. Somboon K, Doble A, Bulmer D, Baslé A, Khalid S, van den Berg B. 2020. Uptake of monoaromatic hydrocarbons during biodegradation by FadL channel-mediated lateral diffusion. Nat. Commun. 11: 6331.
    Pubmed KoreaMed CrossRef
  5. You J, Du M, Chen H, Zhang X, Zhang S, Chen J, et al. 2018. BTEX degradation by a newly isolated bacterium: performance, kinetics, and mechanism. Int. Biodeterior. Biodegrad. 129: 202-208.
    CrossRef
  6. Jiang B, Zhou Z, Dong Y, Tao W, Wang B, Jiang J, et al. 2015. Biodegradation of benzene, toluene, ethylbenzene, and o-, m-, and p-xylenes by the newly isolated bacterium Comamonas sp. JB. Appl. Biochem. Biotechnol. 176: 1700-1708.
    Pubmed CrossRef
  7. Lee SH, Jin HM, Lee HJ, Kim JM, Jeon CO. 2012. Complete genome sequence of the BTEX-degrading bacterium Pseudoxanthomonas spadix BD-a59. J. Bacteriol. 194: 544.
    Pubmed KoreaMed CrossRef
  8. Lee SK, Lee SB. 2001. Isolation and characterization of a thermotolerant bacterium Ralstonia sp. strain PHS1 that degrades benzene, toluene, ethylbenzene, and o-xylene. Appl. Microbiol. Biotechnol. 56: 270-275.
    Pubmed CrossRef
  9. Meyer-Cifuentes I, Martinez-Lavanchy PM, Marin-Cevada V, Böhnke S, Harms H, Müller JA, et al. 2017. Isolation and characterization of Magnetospirillum sp. strain 15-1 as a representative anaerobic toluene-degrader from a constructed wetland model. PLoS One 12: e0174750.
    Pubmed KoreaMed CrossRef
  10. Surendra SV, Mahalingam BL, Velan M. 2017. Degradation of monoaromatics by Bacillus pumilus MVSV3. Braz. Arch. Biol. Technol. 60: 16160319.
    CrossRef
  11. Wongbunmak A, Khiawjan S, Suphantharika M, Pongtharangkul T. 2017. BTEX- and naphthalene-degrading bacterium Microbacterium esteraromaticum strain SBS1-7 isolated from estuarine sediment. J. Hazard. Mater. 339: 82-90.
    Pubmed CrossRef
  12. Wongbunmak A, Khiawjan S, Suphantharika M, Pongtharangkul T. 2020. BTEX biodegradation by Bacillus amyloliquefaciens subsp. plantarum W1 and its proposed BTEX biodegradation pathways. Sci. Rep. 10: 17408.
    Pubmed KoreaMed CrossRef
  13. Yrjälä K, Paulin L, Romantschuk M. 1997. Novel organization of catechol meta-pathway genes in Sphingomonas sp. HV3 pSKY4 plasmid. FEMS Microbiol. Lett. 154: 403-408.
    Pubmed CrossRef
  14. Zhang S, You J, Kennes C, Cheng Z, Ye J, Chen D, et al. 2018. Current advances of VOCs degradation by bioelectrochemical systems: a review. Chem. Eng. J. 334: 2625-2637.
    CrossRef
  15. Lee Y, Lee Y, Jeon CO. 2019. Biodegradation of naphthalene, BTEX, and aliphatic hydrocarbons by Paraburkholderia aromaticivorans BN5 isolated from petroleum-contaminated soil. Sci. Rep. 9: 860.
    Pubmed KoreaMed CrossRef
  16. El-Naas MH, Acio JA, El Telib AE. 2014. Aerobic biodegradation of BTEX: progresses and prospects. J. Environ. Chem. Eng. 2: 1104-1122.
    CrossRef
  17. Lee K, Gibson DT. 1996. Toluene and ethylbenzene oxidation by purified naphthalene dioxygenase from Pseudomonas sp. strain NCIB 9816-4. Appl. Environ. Microbiol. 62: 3101-3106.
    Pubmed KoreaMed CrossRef
  18. Bühler B, Witholt B, Hauer B, Schmid A. 2002. Characterization and application of xylene monooxygenase for multistep biocatalysis. Appl. Environ. Microbiol. 68: 560-568.
    Pubmed KoreaMed CrossRef
  19. Yu H, Kim BJ, Rittmann BE. 2001. The roles of intermediates in biodegradation of benzene, toluene, and p-xylene by Pseudomonas putida F1. Biodegradation 12: 455-463.
    Pubmed CrossRef
  20. Bickerdike SR, Holt RA, Stephens GM. 1997. Evidence for metabolism of o-xylene by simultaneous ring and methyl group oxidation in a new soil isolate. Microbiology 143: 2321-2329.
    Pubmed CrossRef
  21. Lee SK, Lee SB. 2002. Substrate utilization patterns during BTEX biodegradation by an o-xylene-degrading bacterium Ralstonia sp. PHS1. J. Microbiol. Biotechnol. 12: 909-915.
  22. Jeong E, Hirai M, Shoda M. 2008. Removal of O-xylene using biofilter inoculated with Rhodococcus sp. BTO62. J. Hazard. Mater. 152: 140-147.
    Pubmed CrossRef
  23. Taki H, Syutsubo K, Mattison RG, Harayama S. 2007. Identification and characterization of o-xylene-degrading Rhodococcus spp. which were dominant species in the remediation of o-xylene-contaminated soils. Biodegradation 18: 17-26.
    Pubmed CrossRef
  24. Bertoni G, Bolognese F, Galli E, Barbieri P. 1996. Cloning of the genes for and characterization of the early stages of toluene and oxylene catabolism in Pseudomonas stutzeri OX1. Appl. Environ. Microbiol. 62: 3704-3711.
    Pubmed KoreaMed CrossRef
  25. Boeckmann B, Bairoch A, Apweiler R, Blatter MC, Estreicher A, Gasteiger E, et al. 2003. The SWISS-PROT protein knowledgebase and its supplement TrEMBL in 2003. Nucleic Acids Res. 31: 365-370.
    Pubmed KoreaMed CrossRef
  26. Buchfink B, Reuter K, Drost HG. 2021. Sensitive protein alignments at tree-of-life scale using DIAMOND. Nat. Methods 18: 366-368.
    Pubmed KoreaMed CrossRef
  27. Galperin MY, Wolf YI, Makarova KS, Alvarez R V, Landsman D, Koonin E V. 2021. COG database update: focus on microbial diversity, model organisms, and widespread pathogens. Nucleic Acids Res. 49: D274-D281.
    Pubmed KoreaMed CrossRef
  28. Jeske L, Placzek S, Schomburg I, Chang A, Schomburg D. 2019. BRENDA in 2019: a European ELIXIR core data resource. Nucleic Acids Res. 47: D542-D549.
    Pubmed KoreaMed CrossRef
  29. Kanehisa M, Sato Y, Kawashima M, Furumichi M, Tanabe M. 2016. KEGG as a reference resource for gene and protein annotation. Nucleic Acids Res. 44: D457-D462.
    Pubmed KoreaMed CrossRef
  30. Bertelli C, Laird MR, Williams KP, Lau BY, Hoad G, Winsor GL, et al. 2017. IslandViewer 4: expanded prediction of genomic islands for larger-scale datasets. Nucleic Acids Res. 45: W30-W35.
    Pubmed KoreaMed CrossRef
  31. Tamura K, Stecher G, Kumar S. 2021. MEGA11: molecular evolutionary genetics analysis version 11. Mol. Biol. Evol. 38: 3022-3027.
    Pubmed KoreaMed CrossRef
  32. Kim J, Na SI, Kim D, Chun J. 2021. UBCG2: up-to-date bacterial core genes and pipeline for phylogenomic analysis. J. Microbiol. 59: 609-615.
    Pubmed CrossRef
  33. Lee I, Kim YO, Park SC, Chun J. 2016. OrthoANI: an improved algorithm and software for calculating average nucleotide identity. Int. J. Syst. Evol. Microbiol. 66: 1100-1103.
    Pubmed CrossRef
  34. Meier-Kolthoff JP, Auch AF, Klenk HP, Göker M. 2013. Genome sequence-based species delimitation with confidence intervals and improved distance functions. BMC Bioinformatics 14: 60.
    Pubmed KoreaMed CrossRef
  35. Chun J, Oren A, Ventosa A, Christensen H, Arahal DR, da Costa MS, et al. 2018. Proposed minimal standards for the use of genome data for the taxonomy of prokaryotes. Int. J. Syst. Evol. Microbiol. 68: 461-466.
    Pubmed CrossRef
  36. Wang X, Chen M, Xiao J, Hao L, Crowley DE, Zhang Z, et al. 2015. Genome sequence analysis of the naphthenic acid degrading and metal resistant bacterium Cupriavidus gilardii CR3. PLoS One 10: e0132881.
    Pubmed KoreaMed CrossRef
  37. Ensley BD, Ratzkin BJ, Osslund TD, Simon MJ, Wackett LP, Gibson DT. 1983. Expression of naphthalene oxidation genes in Escherichia coli results in the biosynthesis of indigo. Science 222: 167-169.
    Pubmed CrossRef
  38. Arenghi FLG, Berlanda D, Galli E, Sello G, Barbieri P. 2001. Organization and Regulation of meta cleavage pathway genes for toluene and o-xylene derivative degradation in Pseudomonas stutzeri OX1. Appl. Environ. Microbiol. 67: 3304-3308.
    Pubmed KoreaMed CrossRef
  39. Kaschabek SR, Kasberg T, Müller D, Mars AE, Janssen DB, Reineke W. 1998. Degradation of chloroaromatics: purification and characterization of a novel type of chlorocatechol 2,3-dioxygenase of Pseudomonas putida GJ31. J. Bacteriol. 180: 296-302.
    Pubmed KoreaMed CrossRef
  40. Choi EJ, Jin HM, Lee SH, Math RK, Madsen EL, Jeon CO. 2013. Comparative genomic analysis and benzene, toluene, ethylbenzene, and o-, m-, and p-xylene (BTEX) degradation pathways of Pseudoxanthomonas spadix BD-a59. Appl. Environ. Microbiol 79: 663-671.
    Pubmed KoreaMed CrossRef
  41. Espinoza Tofalos A, Daghio M, Gonzalez M, Papacchini M, Franzetti A, Seeger M. 2018. Toluene degradation by Cupriavidus metallidurans CH34 in nitrate-reducing conditions and in bioelectrochemical systems. FEMS Microbiol. Lett. 365: fny119.
    Pubmed CrossRef
  42. Kumar A, Dewulf J, Luvsanjamba M, Van Langenhove H. 2008. Continuous operation of membrane bioreactor treating toluene vapors by Burkholderia vietnamiensis G4. Chem. Eng. J. 140: 193-200.
    CrossRef
  43. Perez-Pantoja D, De la Iglesia R, Pieper DH, Gonzalez B. 2008. Metabolic reconstruction of aromatic compounds degradation from the genome of the amazing pollutant-degrading bacterium Cupriavidus necator JMP134. FEMS Microbiol. Rev. 32: 736-794.
    Pubmed CrossRef
  44. Shingler V, Powlowski J, Marklund U. 1992. Nucleotide-sequence and functional-analysis of the complete phenol/3,4-dimethylphenol catabolic pathway of Pseudomonas sp. strain-Cf600. J. Bacteriol. 174: 711-724.
    Pubmed KoreaMed CrossRef
  45. Tinberg CE, Song WJ, Izzo V, Lippard SJ. 2011. Multiple roles of component proteins in bacterial multicomponent monooxygenases: phenol hydroxylase and toluene/o-xylene monooxygenase from Pseudomonas sp. OX1. Biochemistry 50: 1788-1798.
    Pubmed KoreaMed CrossRef
  46. Hou YJ, Guo Y, Li DF, Zhou NY. 2021. Structural and biochemical analysis reveals a distinct catalytic site of salicylate 5-monooxygenase NagGH from Rieske dioxygenases. Appl. Environ. Microbiol. 87: e01629-20.
    Pubmed KoreaMed CrossRef
  47. Karlsson A, Parales J V, Parales RE, Gibson DT, Eklund H, Ramaswamy S. 2003. Crystal structure of naphthalene dioxygenase: sideon binding of dioxygen to iron. Science 299: 1039-1042.
    Pubmed CrossRef
  48. Xu JR, Zhang Y. 2010. How significant is a protein structure similarity with TM-score=0.5? Bioinformatics 26: 889-895.
    Pubmed KoreaMed CrossRef
  49. Kim M, Oh HS, Park SC, Chun J. 2014. Towards a taxonomic coherence between average nucleotide identity and 16S rRNA gene sequence similarity for species demarcation of prokaryotes. Int. J. Syst. Evol. Microbiol. 64: 346-351.
    Pubmed CrossRef
  50. Moreno-Hagelsieb G, Latimer K. 2008. Choosing BLAST options for better detection of orthologs as reciprocal best hits. Bioinformatics 24: 319-324.
    Pubmed CrossRef
  51. Burbank LP, Van Horn CR. 2017. Conjugative plasmid transfer in Xylella fastidiosa is dependent on tra and trb operon functions. J. Bacteriol. 199: e00388-17.
    Pubmed KoreaMed CrossRef
  52. Newman LM, Wackett LP. 1995. Purification and characterization of toluene 2-monooxygenase from Burkholderia cepacia G4. Biochemistry 34: 14066-14076.
    Pubmed CrossRef
  53. Cafaro V, Izzo V, Scognamiglio R, Notomista E, Capasso P, Casbarra A, et al. 2004. Phenol hydroxylase and toluene/o-xylene monooxygenase from Pseudomonas stutzeri OX1: interplay between two enzymes. Appl. Environ. Microbiol. 70: 2211-2219.
    Pubmed KoreaMed CrossRef
  54. Arai H, Akahira S, Ohishi T, Maeda M, Kudo T. 1998. Adaptation of Comamonas testosteroni TA441 to utilize phenol: organization and regulation of the genes involved in phenol degradation. Microbiology 144: 2895-2903.
    Pubmed CrossRef
  55. Nordlund I, Powlowski J, Shingler V. 1990. Complete nucleotide sequence and polypeptide analysis of multicomponent phenol hydroxylase from Pseudomonas sp. strain CF600. J. Bacteriol. 172: 6826-6833.
    Pubmed KoreaMed CrossRef
  56. Shields MS, Reagin MJ, Gerger RR, Campbell R, Somerville C. 1995. TOM, a new aromatic degradative plasmid from Burkholderia (Pseudomonas) cepacia G4. Appl. Environ. Microbiol. 61: 1352-1356.
    Pubmed KoreaMed CrossRef
  57. Fox BG, Shanklin J, Somerville C, Munck E. 1993. Stearoyl-acyl carrier protein delta 9 desaturase from Ricinus communis is a diironoxo protein. Proc. Natl. Acad. Sci. USA 90: 2486.
    Pubmed KoreaMed CrossRef
  58. Johnson GR, Olsen RH. 1997. Multiple pathways for toluene degradation in Burkholderia sp. strain JS150. Appl. Environ. Microbiol. 63: 4047.
    Pubmed KoreaMed CrossRef
  59. Harayama S, Polissi A, Rekik M. 1991. Divergent evolution of chloroplast-type ferredoxins. FEBS Lett. 285: 85-88.
    Pubmed CrossRef
  60. Laurie AD, Lloyd-Jones G. 1999. Conserved and hybrid meta-cleavage operons from PAH-degrading Burkholderia RP007. Biochem. Biophys. Res. Commun. 262: 308-314.
    Pubmed CrossRef
  61. Arai H, Ohishi T, Chang MY, Kudo T. 2000. Arrangement and regulation of the genes for meta-pathway enzymes required for degradation of phenol in Comamonas testosteroni TA441. Microbiology 146(Pt 7): 1707-1715.
    Pubmed CrossRef
  62. Pérez-Pantoja D, Donoso R, Agulló L, Córdova M, Seeger M, Pieper DH, et al. 2012. Genomic analysis of the potential for aromatic compounds biodegradation in Burkholderiales. Environ. Microbiol. 14: 1091-1117.
    Pubmed CrossRef
  63. Romine MF, Stillwell LC, Wong KK, Thurston SJ, Sisk EC, Sensen C, et al. 1999. Complete sequence of a 184-kilobase catabolic plasmid from Sphingomonas aromaticivorans F199. J. Bacteriol. 181: 1585.
    Pubmed KoreaMed CrossRef
  64. Erickson BD, Mondello FJ. 1992. Nucleotide sequencing and transcriptional mapping of the genes encoding biphenyl dioxygenase, a multicomponent polychlorinated-biphenyl-degrading enzyme in Pseudomonas strain LB400. J. Bacteriol. 174: 2903.
    Pubmed KoreaMed CrossRef
  65. Froland WA, Andersson KK, Lee SK, Liu Y, Lipscomb JD. 1992. Methane monooxygenase component B and reductase alter the regioselectivity of the hydroxylase component-catalyzed reactions. A novel role for protein-protein interactions in an oxygenase mechanism. J. Biol. Chem. 267: 17588-17597.
    Pubmed CrossRef
  66. Mason JR, Cammack R. 1992. The electron-transport proteins of hydroxylating bacterial dioxygenases. Annu. Rev. Microbiol. 46: 277-305.
    Pubmed CrossRef
  67. Kitayama A, Suzuki E, Kawakami Y, Nagamune T. 1996. Gene organization and low regiospecificity in aromatic-ring hydroxylation of a benzene monooxygenase of Pseudomonas aeruginosa JI104. J. Ferment. Bioeng. 82: 421-425.
    CrossRef
  68. Olsen RH, Kukor JJ, Kaphammer B. 1994. A novel toluene-3-monooxygenase pathway cloned from Pseudomonas pickettii PKO1. J. Bacteriol. 176: 3749-3756.
    Pubmed KoreaMed CrossRef
  69. Mitchell KH, Studts JM, Fox BG. 2002. Combined participation of hydroxylase active site residues and effector protein binding in a para to ortho modulation of toluene 4-monooxygenase regiospecificity. Biochemistry 41: 3176-3188.
    Pubmed CrossRef
  70. Bertoni G, Martino M, Galli E, Barbieri P. 1998. Analysis of the gene cluster encoding toluene/o-xylene monooxygenase from Pseudomonas stutzeri OX1. Appl. Environ. Microbiol. 64: 3626-3632.
    Pubmed KoreaMed CrossRef
  71. Jiang T, Wang W, Han B. 2013. Catalytic hydroxylation of benzene to phenol with hydrogen peroxide using catalysts based on molecular sieves. New J. Chem. 37: 1654-1664.
    CrossRef
  72. Barbieri P, Palladino L, Di Gennaro P, Galli E. 1993. Alternative pathways for o-xylene or m-xylene and p-xylene degradation in a Pseudomonas stutzeri strain. Biodegrad. 4: 71-80.
    CrossRef
  73. Fowler SJ, Gutierrez-Zamora ML, Manefield M, Gieg LM. 2014. Identification of toluene degraders in a methanogenic enrichment culture. FEMS Microbiol. Ecol. 89: 625-636.
    Pubmed CrossRef
  74. Kahng HY, Malinverni JC, Majko MM, Kukor JJ. 2001. Genetic and functional analysis of the tbc operons for catabolism of alkyl- and chloroaromatic compounds in Burkholderia sp. strain JS150. Appl. Environ. Microbiol. 67: 4805-4816.
    Pubmed KoreaMed CrossRef
  75. Siani L, Viggiani A, Notomista E, Pezzella A, Di Donato A. 2006. The role of residue Thr249 in modulating the catalytic efficiency and substrate specificity of catechol-2,3-dioxygenase from Pseudomonas stutzeri OX1. FEBS J. 273: 2963-2976.
    Pubmed CrossRef
  76. Krzywinski M, Schein J, Birol I, Connors J, Gascoyne R, Horsman D, et al. 2009. Circos: an information aesthetic for comparative genomics. Genome Res. 19: 1639-1645.
    Pubmed KoreaMed CrossRef