1932

Abstract

Microbial natural products have provided an important source of therapeutic leads and motivated research and innovation in diverse scientific disciplines. In recent years, it has become evident that bacteria harbor a large, hidden reservoir of potential natural products in the form of silent or cryptic biosynthetic gene clusters (BGCs). These can be readily identified in microbial genome sequences but do not give rise to detectable levels of a natural product. Herein, we provide a useful organizational framework for the various methods that have been implemented for interrogating silent BGCs. We divide all available approaches into four categories. The first three are endogenous strategies that utilize the native host in conjunction with classical genetics, chemical genetics, or different culture modalities. The last category comprises expression of the entire BGC in a heterologous host. For each category, we describe the rationale, recent applications, and associated advantages and limitations.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-biochem-081420-102432
2021-06-20
2024-03-29
Loading full text...

Full text loading...

/deliver/fulltext/biochem/90/1/annurev-biochem-081420-102432.html?itemId=/content/journals/10.1146/annurev-biochem-081420-102432&mimeType=html&fmt=ahah

Literature Cited

  1. 1. 
    Demain AL, Sanchez S. 2009. Microbial drug discovery: 80 years of progress. J. Antibiot. 62:5–16
    [Google Scholar]
  2. 2. 
    Newman DJ, Cragg GM. 2020. Natural products as sources of new drugs over the nearly four decades from 01/1981 to 09/2019. J. Nat. Prod. 83:770–803
    [Google Scholar]
  3. 3. 
    Medema MH, Blin K, Cimermancic P, de Jager V, Zakrzewski P et al. 2011. antiSMASH: rapid identification, annotation and analysis of secondary metabolite biosynthesis gene clusters in bacterial and fungal genome sequences. Nucleic Acids Res 39:W339–46
    [Google Scholar]
  4. 4. 
    Skinnider MA, Dejong CA, Rees PN, Johnston CW, Li H et al. 2015. Genomes to natural products PRediction Informatics for Secondary Metabolomes (PRISM). Nucleic Acids Res 43:9645–62
    [Google Scholar]
  5. 5. 
    Nett M, Ikeda H, Moore BS. 2009. Genomic basis for natural product biosynthetic diversity in the actinomycetes. Nat. Prod. Rep. 26:1362–84
    [Google Scholar]
  6. 6. 
    Sugimoto Y, Camacho FR, Wang S, Chankhamjon P, Odabas A et al. 2019. A metagenomic strategy for harnessing the chemical repertoire of the human microbiome. Science 366:eaax9176
    [Google Scholar]
  7. 7. 
    Wilson MR, Zha L, Balskus EP. 2017. Natural product discovery from the human microbiome. J. Biol. Chem. 292:8546–52
    [Google Scholar]
  8. 8. 
    Gerwick WH, Moore BS. 2012. Lessons from the past and charting the future of marine natural products drug discovery and chemical biology. Chem. Biol. 19:85–98
    [Google Scholar]
  9. 9. 
    Morita M, Schmidt EW. 2018. Parallel lives of symbionts and hosts: chemical mutualism in marine animals. Nat. Prod. Rep. 35:357–78
    [Google Scholar]
  10. 10. 
    Covington BC, Spraggins JM, Ynigez-Gutierrez AE, Hylton ZB, Bachmann BO. 2018. Response of secondary metabolism of hypogean actinobacterial genera to chemical and biological stimuli. Appl. Environ. Microbiol. 84:e01125–18
    [Google Scholar]
  11. 11. 
    Derewacz DK, McNees CR, Scalmani G, Covington CL, Shanmugam G et al. 2014. Structure and stereochemical determination of hypogeamicins from a cave-derived actinomycete. J. Nat. Prod. 77:1759–63
    [Google Scholar]
  12. 12. 
    Smith TE, Pond CD, Pierce E, Harmer ZP, Kwan J et al. 2018. Accessing chemical diversity from the uncultivated symbionts of small marine animals. Nat. Chem. Biol. 14:179–85
    [Google Scholar]
  13. 13. 
    Amos GCA, Awakawa T, Tuttle RN, Letzel A-C, Kim MC et al. 2017. Comparative transcriptomics as a guide to natural product discovery and biosynthetic gene cluster functionality. PNAS 114:E11121–30
    [Google Scholar]
  14. 14. 
    Askenazi M, Driggers EM, Holtzman DA, Norman TC, Iverson S et al. 2003. Integrating transcriptional and metabolite profiles to direct the engineering of lovastatin-producing fungal strains. Nat. Biotechnol. 21:150–56
    [Google Scholar]
  15. 15. 
    Xiang S-H, Li J, Yin H, Zheng J-T, Yang X et al. 2009. Application of a double-reporter-guided mutant selection method to improve clavulanic acid production in Streptomyces clavuligerus. Metab. Eng. 11:310–18
    [Google Scholar]
  16. 16. 
    Guo F, Xiang S, Li L, Wang B, Rajasärkkä J et al. 2015. Targeted activation of silent natural product biosynthesis pathways by reporter-guided mutant selection. Metab. Eng. 28:134–42
    [Google Scholar]
  17. 17. 
    Xu Z, Wang Y, Chater KF, Ou HY, Xu HH et al. 2017. Large-scale transposition mutagenesis of Streptomyces coelicolor identifies hundreds of genes influencing antibiotic biosynthesis. Appl. Environ. Microbiol. 83:e02889–16
    [Google Scholar]
  18. 18. 
    Ahmed Y, Rebets Y, Tokovenko B, Brötz E, Luzhetskyy A. 2017. Identification of butenolide regulatory system controlling secondary metabolism in Streptomyces albus J1074. Sci. Rep. 7:9784
    [Google Scholar]
  19. 19. 
    Park JD, Moon K, Miller C, Rose J, Xu F et al. 2020. Thailandenes, cryptic polyene natural products isolated from Burkholderia thailandensis using phenotype-guided transposon mutagenesis. ACS Chem. Biol. 15:1195–203
    [Google Scholar]
  20. 20. 
    Mao D, Yoshimura A, Wang R, Seyedsayamdost MR. 2020. Reporter-guided transposon mutant selection for activation of silent gene clusters in Burkholderia thailandensis. ChemBioChem 21:1826–31
    [Google Scholar]
  21. 21. 
    Yoshimura A, Covington B, Gallant E, Zhang C, Li A, Seyedsayamdost M. 2020. Unlocking cryptic metabolites using mass spectrometry–guided transposon mutant selection. ACS Chem. Biol. 15:2766–74
    [Google Scholar]
  22. 22. 
    Hermenau R, Mehl JL, Ishida K, Dose B, Pidot SJ et al. 2019. Genomics-driven discovery of NO-donating diazeniumdiolate siderophores in diverse plant-associated bacteria. Angew. Chem. Int. Ed. 58:13024–29
    [Google Scholar]
  23. 23. 
    Zhang B, Tian W, Wang S, Yan X, Jia X et al. 2017. Activation of natural products biosynthetic pathways via a protein modification level regulation. ACS Chem. Biol. 12:1732–36
    [Google Scholar]
  24. 24. 
    Horinouchi S, Beppu T. 1993. A-factor and streptomycin biosynthesis in Streptomyces griseus. Anton. Leeuw. Int. J. G 64:177–86
    [Google Scholar]
  25. 25. 
    Fuqua C, Greenberg EP. 2002. Listening in on bacteria: acyl-homoserine lactone signalling. Nat. Rev. Mol. Cell. Biol. 3:685–95
    [Google Scholar]
  26. 26. 
    Maddocks SE, Oyston PCF. 2008. Structure and function of the LysR-type transcriptional regulator (LTTR) family proteins. Microbiology 154:3609–23
    [Google Scholar]
  27. 27. 
    Cai X, Teta R, Kohlhaas C, Crüsemann M, Ueoka R et al. 2013. Manipulation of regulatory genes reveals complexity and fidelity in hormaomycin biosynthesis. Chem. Biol. 20:839–46
    [Google Scholar]
  28. 28. 
    Pan Y, Wang L, He X, Tian Y, Liu G, Tan H. 2011. SabR enhances nikkomycin production via regulating the transcriptional level of sanG, a pathway-specific regulatory gene in Streptomyces ansochromogenes. BMC Microbiol 11:164
    [Google Scholar]
  29. 29. 
    Krause J, Handayani I, Blin K, Kulik A, Mast Y. 2020. Disclosing the potential of the SARP-type regulator PapR2 for the activation of antibiotic gene clusters in streptomycetes. Front. Microbiol. 11:225
    [Google Scholar]
  30. 30. 
    Daniel-Ivad M, Hameed N, Tan S, Dhanjal R, Socko D et al. 2017. An engineered allele of afsQ1 facilitates the discovery and investigation of cryptic natural products. ACS Chem. Biol. 12:628–34
    [Google Scholar]
  31. 31. 
    Xie C, Deng J-J, Wang H-X. 2015. Identification of AstG1, a LAL family regulator that positively controls ansatrienins production in Streptomyces sp. XZQH13. Curr. Microbiol. 70:859–64
    [Google Scholar]
  32. 32. 
    Laureti L, Song L, Huang S, Corre C, Leblond P et al. 2011. Identification of a bioactive 51-membered macrolide complex by activation of a silent polyketide synthase in Streptomyces ambofaciens. PNAS 108:6258–63
    [Google Scholar]
  33. 33. 
    Thanapipatsiri A, Gomez-Escribano JP, Song L, Bibb MJ, Al-Bassam M et al. 2016. Discovery of unusual biaryl polyketides by activation of a silent Streptomyces venezuelae biosynthetic gene cluster. ChemBioChem 17:2189–98
    [Google Scholar]
  34. 34. 
    Zhou Z, Xu Q, Bu Q, Guo Y, Liu S et al. 2015. Genome mining-directed activation of a silent angucycline biosynthetic gene cluster in Streptomyces chattanoogensis. ChemBioChem 16:496–502
    [Google Scholar]
  35. 35. 
    Biggins JB, Gleber CD, Brady SF. 2011. Acyldepsipeptide HDAC inhibitor production induced in Burkholderia thailandensis. Org. Lett. 13:1536–39
    [Google Scholar]
  36. 36. 
    Ahn SK, Cuthbertson L, Nodwell JR. 2012. Genome context as a predictive tool for identifying regulatory targets of the TetR family transcriptional regulators. PLOS ONE 7:e50562
    [Google Scholar]
  37. 37. 
    Cuthbertson L, Nodwell JR. 2013. The TetR family of regulators. Microbiol. Mol. Biol. Rev. 77:440–75
    [Google Scholar]
  38. 38. 
    Sidda JD, Song L, Poon V, Al-Bassam M, Lazos O et al. 2014. Discovery of a family of γ-aminobutyrate ureas via rational derepression of a silent bacterial gene cluster. Chem. Sci. 5:86–89
    [Google Scholar]
  39. 39. 
    Zhang Y, Zou Z, Niu G, Tan H. 2013. jadR* and jadR2 act synergistically to repress jadomycin biosynthesis. Sci. China Life Sci. 56:584–90
    [Google Scholar]
  40. 40. 
    Bunet R, Song L, Mendes MV, Corre C, Hotel L et al. 2011. Characterization and manipulation of the pathway-specific late regulator AlpW reveals Streptomyces ambofaciens as a new producer of kinamycins. J. Bacteriol. 193:1142–53
    [Google Scholar]
  41. 41. 
    Wang B, Guo F, Dong SH, Zhao H. 2019. Activation of silent biosynthetic gene clusters using transcription factor decoys. Nat. Chem. Biol. 15:111–14
    [Google Scholar]
  42. 42. 
    Rigali S, Titgemeyer F, Barends S, Mulder S, Thomae AW et al. 2008. Feast or famine: The global regulator DasR links nutrient stress to antibiotic production by Streptomyces. EMBO Rep 9:670–75
    [Google Scholar]
  43. 43. 
    Higo A, Hara H, Horinouchi S, Ohnishi Y. 2012. Genome-wide distribution of AdpA, a global regulator for secondary metabolism and morphological differentiation in Streptomyces, revealed the extent and complexity of the AdpA regulatory network. DNA Res 19:259–73
    [Google Scholar]
  44. 44. 
    McKenzie NL, Thaker M, Koteva K, Hughes DW, Wright GD, Nodwell JR. 2010. Induction of antimicrobial activities in heterologous streptomycetes using alleles of the Streptomyces coelicolor gene absA1. J. Antibiot. 63:177–82
    [Google Scholar]
  45. 45. 
    Gehrke EJ, Zhang X, Pimentel-Elardo SM, Johnson AR, Rees CA et al. 2019. Silencing cryptic specialized metabolism in Streptomyces by the nucleoid-associated protein Lsr2. eLife 8:e47691
    [Google Scholar]
  46. 46. 
    Mao D, Bushin LB, Moon K, Wu Y, Seyedsayamdost MR 2017. Discovery of scmR as a global regulator of secondary metabolism and virulence in Burkholderia thailandensis E264. PNAS 114:E2920–28
    [Google Scholar]
  47. 47. 
    Gupta A, Bedre R, Thapa SS, Sabrin A, Wang G et al. 2017. Global awakening of cryptic biosynthetic gene clusters in Burkholderia thailandensis. ACS Chem. Biol. 12:3012–21
    [Google Scholar]
  48. 48. 
    Xu J, Zhang J, Zhuo J, Li Y, Tian Y, Tan H. 2017. Activation and mechanism of a cryptic oviedomycin gene cluster via the disruption of a global regulatory gene, adpA, in Streptomyces ansochromogenes. J. Biol. Chem. 292:19708–20
    [Google Scholar]
  49. 49. 
    Nazari B, Kobayashi M, Saito A, Hassaninasab A, Miyashita K, Fujii T. 2013. Chitin-induced gene expression in secondary metabolic pathways of Streptomyces coelicolor A3(2) grown in soil. Appl. Environ. Microbiol. 79:707–13
    [Google Scholar]
  50. 50. 
    Liao CH, Xu Y, Rigali S, Ye BC. 2015. DasR is a pleiotropic regulator required for antibiotic production, pigment biosynthesis, and morphological development in Saccharopolyspora erythraea. Appl. Microbiol. Biotechnol. 99:10215–24
    [Google Scholar]
  51. 51. 
    Mukherji R, Zhang S, Chowdhury S, Stallforth P. 2020. Chimeric LuxR transcription factors rewire natural product regulation. Angew. Chem. Int. Ed. 59:6192–95
    [Google Scholar]
  52. 52. 
    Olano C, García I, González A, Rodriguez M, Rozas D et al. 2014. Activation and identification of five clusters for secondary metabolites in Streptomyces albus J1074. Microb. Biotechnol. 7:242–56
    [Google Scholar]
  53. 53. 
    Trottmann F, Franke J, Richter I, Ishida K, Cyrulies M et al. 2019. Cyclopropanol warhead in malleicyprol confers virulence of human- and animal-pathogenic Burkholderia species. Angew. Chem. Int. Ed. 58:14129–33
    [Google Scholar]
  54. 54. 
    Biggins JB, Kang HS, Ternei MA, DeShazer D, Brady SF. 2014. The chemical arsenal of Burkholderia pseudomallei is essential for pathogenicity. J. Am. Chem. Soc. 136:9484–90
    [Google Scholar]
  55. 55. 
    Zhang MM, Wong FT, Wang Y, Luo S, Lim YH et al. 2017. CRISPR–Cas9 strategy for activation of silent Streptomyces biosynthetic gene clusters. Nat. Chem. Biol. 13:607–9
    [Google Scholar]
  56. 56. 
    Craney A, Ozimok C, Pimentel-Elardo SM, Capretta A, Nodwell JR. 2012. Chemical perturbation of secondary metabolism demonstrates important links to primary metabolism. Chem. Biol. 19:1020–27
    [Google Scholar]
  57. 57. 
    Seyedsayamdost MR. 2014. High-throughput platform for the discovery of elicitors of silent bacterial gene clusters. PNAS 111:7266–71
    [Google Scholar]
  58. 58. 
    Davies J, Spiegelman GB, Yim G. 2006. The world of subinhibitory antibiotic concentrations. Curr. Opin. Microbiol. 9:445–53
    [Google Scholar]
  59. 59. 
    Okada BK, Wu Y, Mao D, Bushin LB, Seyedsayamdost MR. 2016. Mapping the trimethoprim-induced secondary metabolome of Burkholderia thailandensis. ACS Chem. Biol. 11:2124–30
    [Google Scholar]
  60. 60. 
    Li A, Mao D, Yoshimura A, Rosen PC, Martin WL et al. 2020. Multi-omic analyses provide links between low-dose antibiotic treatment and induction of secondary metabolism in Burkholderia thailandensis. mBio 11:e03210–19
    [Google Scholar]
  61. 61. 
    Klaus JR, Deay J, Neuenswander B, Hursh W, Gao Z et al. 2018. Malleilactone is a Burkholderia pseudomallei virulence factor regulated by antibiotics and quorum sensing. J. Bacteriol. 200:e00008–18
    [Google Scholar]
  62. 62. 
    Xu F, Nazari B, Moon K, Bushin LB, Seyedsayamdost MR. 2017. Discovery of a cryptic antifungal compound from Streptomyces albus J1074 using high-throughput elicitor screens. J. Am. Chem. Soc. 139:9203–12
    [Google Scholar]
  63. 63. 
    Xu F, Wu Y, Zhang C, Davis KM, Moon K et al. 2019. A genetics-free method for high-throughput discovery of cryptic microbial metabolites. Nat. Chem. Biol. 15:161–68
    [Google Scholar]
  64. 64. 
    Zhang C, Seyedsayamdost MR. 2020. Discovery of a cryptic depsipeptide from Streptomyces ghanaensis via MALDI-MS-guided high-throughput elicitor screening. Angew. Chem. Int. Ed. 59:23005–9
    [Google Scholar]
  65. 65. 
    Moon K, Xu F, Zhang C, Seyedsayamdost MR. 2019. Bioactivity-HiTES unveils cryptic antibiotics encoded in actinomycete bacteria. ACS Chem. Biol. 14:767–74
    [Google Scholar]
  66. 66. 
    Moon K, Xu F, Seyedsayamdost MR. 2019. Cebulantin, a cryptic lanthipeptide antibiotic uncovered using bioactivity-coupled HiTES. Angew. Chem. Int. Ed. 58:5973–77
    [Google Scholar]
  67. 67. 
    Panthee S, Takahashi S, Hayashi T, Shimizu T, Osada H. 2019. β-carboline biomediators induce reveromycin production in Streptomyces sp. SN-593. Sci. Rep. 9:5802
    [Google Scholar]
  68. 68. 
    Truong TT, Seyedsayamdost M, Greenberg EP, Chandler JR. 2015. A Burkholderia thailandensis acyl-homoserine lactone-independent orphan LuxR homolog that activates production of the cytotoxin malleilactone. J. Bacteriol. 197:3456–62
    [Google Scholar]
  69. 69. 
    Shima J, Hesketh A, Okamoto S, Kawamoto S, Ochi K. 1996. Induction of actinorhodin production by rpsL (encoding ribosomal protein S12) mutations that confer streptomycin resistance in Streptomyces lividans and Streptomyces coelicolor A3(2). J. Bacteriol. 178:7276–84
    [Google Scholar]
  70. 70. 
    Tamehiro N, Hosaka T, Xu J, Hu H, Otake N, Ochi K. 2003. Innovative approach for improvement of an antibiotic-overproducing industrial strain of Streptomyces albus. Appl. Environ. Microbiol. 69:6412–17
    [Google Scholar]
  71. 71. 
    Ochi K, Hosaka T. 2013. New strategies for drug discovery: activation of silent or weakly expressed microbial gene clusters. Appl. Microbiol. Biotechnol. 97:87–98
    [Google Scholar]
  72. 72. 
    Imai Y, Sato S, Tanaka Y, Ochi K, Hosaka T. 2015. Lincomycin at subinhibitory concentrations potentiates secondary metabolite production by Streptomyces spp. Appl. Environ. Microbiol. 81:3869–79
    [Google Scholar]
  73. 73. 
    Hosaka T, Ohnishi-Kameyama M, Muramatsu H, Murakami K, Tsurumi Y et al. 2009. Antibacterial discovery in actinomycetes strains with mutations in RNA polymerase or ribosomal protein S12. Nat. Biotechnol. 27:462–64
    [Google Scholar]
  74. 74. 
    Wang G, Hosaka T, Ochi K. 2008. Dramatic activation of antibiotic production in Streptomyces coelicolor by cumulative drug resistance mutations. Appl. Environ. Microbiol. 74:2834–40
    [Google Scholar]
  75. 75. 
    Goodwin CR, Covington BC, Derewacz DK, McNees CR, Wikswo JP et al. 2015. Structuring microbial metabolic responses to multiplexed stimuli via self-organizing metabolomics maps. Chem. Biol. 22:661–70
    [Google Scholar]
  76. 76. 
    Derewacz DK, Goodwin CR, McNees CR, McLean JA, Bachmann BO 2013. Antimicrobial drug resistance affects broad changes in metabolomic phenotype in addition to secondary metabolism. PNAS 110:2336–41
    [Google Scholar]
  77. 77. 
    Wang L, Zhao Y, Liu Q, Huang Y, Hu C, Liao G. 2012. Improvement of A21978C production in Streptomyces roseosporus by reporter-guided rpsL mutation selection. J. Appl. Microbiol. 112:1095–101
    [Google Scholar]
  78. 78. 
    Webersinke F, Klein H, Flieher M, Urban A, Jäger H, Forster C. 2018. Control of fermentation by-products and aroma features of beer produced with Scottish ale yeast by variation of fermentation temperature and wort aeration rate. J. Am. Soc. Brew. Chem. 76:147–55
    [Google Scholar]
  79. 79. 
    Liao Y, Wei ZH, Bai L, Deng Z, Zhong JJ. 2009. Effect of fermentation temperature on validamycin A production by Streptomyces hygroscopicus 5008. J. Biotechnol. 142:271–74
    [Google Scholar]
  80. 80. 
    Görke B, Stülke J. 2008. Carbon catabolite repression in bacteria: many ways to make the most out of nutrients. Nat. Rev. Microbiol. 6:613–24
    [Google Scholar]
  81. 81. 
    Soltero FV, Johnson MJ. 1953. The effect of the carbohydrate nutrition on penicillin production by Penicillium chrysogenum Q-176. Appl. Microbiol. 1:52–57
    [Google Scholar]
  82. 82. 
    Kim ES, Hong HJ, Choi CY, Cohen SN. 2001. Modulation of actinorhodin biosynthesis in Streptomyces lividans by glucose repression of afsR2 gene transcription. J. Bacteriol. 183:2198–203
    [Google Scholar]
  83. 83. 
    Rokem JS, Lantz AE, Nielsen J. 2007. Systems biology of antibiotic production by microorganisms. Nat. Prod. Rep. 24:1262–87
    [Google Scholar]
  84. 84. 
    Bode HB, Bethe B, Höfs R, Zeeck A. 2002. Big effects from small changes: possible ways to explore nature's chemical diversity. ChemBioChem 3:619–27
    [Google Scholar]
  85. 85. 
    Sproule A, Correa H, Decken A, Haltli B, Berrué F et al. 2019. Terrosamycins A and B, bioactive polyether ionophores from Streptomyces sp. RKND004 from Prince Edward Island sediment. Mar. Drugs 17:347
    [Google Scholar]
  86. 86. 
    Romano S, Jackson SA, Patry S, Dobson ADW. 2018. Extending the “one strain many compounds” (OSMAC) principle to marine microorganisms. Mar. Drugs 16:244
    [Google Scholar]
  87. 87. 
    Covington BC, McLean JA, Bachmann BO. 2017. Comparative mass spectrometry–based metabolomics strategies for the investigation of microbial secondary metabolites. Nat. Prod. Rep. 34:6–24
    [Google Scholar]
  88. 88. 
    Lincke T, Behnken S, Ishida K, Roth M, Hertweck C 2010. Closthioamide: an unprecedented polythioamide antibiotic from the strictly anaerobic bacterium Clostridium cellulolyticum. Angew. Chem. Int. Ed. 49:2011–13
    [Google Scholar]
  89. 89. 
    Klapper M, Götze S, Barnett R, Willing K, Stallforth P. 2016. Bacterial alkaloids prevent amoebal predation. Angew. Chem. Int. Ed. 55:8944–47
    [Google Scholar]
  90. 90. 
    Arp J, Götze S, Mukherji R, Mattern DJ, García-Altares M et al. 2018. Synergistic activity of cosecreted natural products from amoebae-associated bacteria. PNAS 115:3758–63
    [Google Scholar]
  91. 91. 
    Beemelmanns C, Guo H, Rischer M, Poulsen M. 2016. Natural products from microbes associated with insects. Beilstein J. Org. Chem. 12:314–27
    [Google Scholar]
  92. 92. 
    Tobias NJ, Wolff H, Djahanschiri B, Grundmann F, Kronenwerth M et al. 2017. Natural product diversity associated with the nematode symbionts Photorhabdus and Xenorhabdus. Nat. Microbiol. 2:1676–85
    [Google Scholar]
  93. 93. 
    Tianero MD, Balaich JN, Donia MS. 2019. Localized production of defence chemicals by intracellular symbionts of Haliclona sponges. Nat. Microbiol. 4:1149–59
    [Google Scholar]
  94. 94. 
    Seyedsayamdost MR, Wang R, Kolter R, Clardy J. 2014. Hybrid biosynthesis of roseobacticides from algal and bacterial precursor molecules. J. Am. Chem. Soc. 136:15150–53
    [Google Scholar]
  95. 95. 
    Xiao Y, Wei X, Ebright R, Wall D. 2011. Antibiotic production by myxobacteria plays a role in predation. J. Bacteriol. 193:4626–33
    [Google Scholar]
  96. 96. 
    Okada BK, Seyedsayamdost MR. 2017. Antibiotic dialogues: induction of silent biosynthetic gene clusters by exogenous small molecules. FEMS Microbiol. Rev. 41:19–33
    [Google Scholar]
  97. 97. 
    Pettit RK. 2009. Mixed fermentation for natural product drug discovery. Appl. Microbiol. Biotechnol. 83:19–25
    [Google Scholar]
  98. 98. 
    Ueda K, Kawai S, Ogawa H, Kiyama A, Kubota T et al. 2000. Wide distribution of interspecific stimulatory events on antibiotic production and sporulation among Streptomyces species. J. Antibiot. 53:979–82
    [Google Scholar]
  99. 99. 
    Yamanaka K, Oikawa H, Ogawa HO, Hosono K, Shinmachi F et al. 2005. Desferrioxamine E produced by Streptomyces griseus stimulates growth and development of Streptomyces tanashiensis. Microbiology 151:2899–905
    [Google Scholar]
  100. 100. 
    Traxler MF, Seyedsayamdost MR, Clardy J, Kolter R. 2012. Interspecies modulation of bacterial development through iron competition and siderophore piracy. Mol. Microbiol. 86:628–44
    [Google Scholar]
  101. 101. 
    Seyedsayamdost MR, Traxler MF, Zheng S-L, Kolter R, Clardy J. 2011. Structure and biosynthesis of amychelin, an unusual mixed-ligand siderophore from Amycolatopsis sp. AA4. J. Am. Chem. Soc. 133:11434–37
    [Google Scholar]
  102. 102. 
    Lee N, Kim W, Chung J, Lee Y, Cho S et al. 2020. Iron competition triggers antibiotic biosynthesis in Streptomyces coelicolor during coculture with Myxococcus xanthus. ISME J 14:1111–24
    [Google Scholar]
  103. 103. 
    Coisne S, Béchet M, Blondeau R. 1999. Actinorhodin production by Streptomyces coelicolor A3(2) in iron-restricted media. Lett. Appl. Microbiol. 28:199–202
    [Google Scholar]
  104. 104. 
    Xiao Y, Wei X, Ebright R, Wall D. 2011. Antibiotic production by myxobacteria plays a role in predation. J. Bacteriol. 193:4626–33
    [Google Scholar]
  105. 105. 
    Goldman BS, Nierman WC, Kaiser D, Slater SC, Durkin AS et al. 2006. Evolution of sensory complexity recorded in a myxobacterial genome. PNAS 103:15200–5
    [Google Scholar]
  106. 106. 
    Lloyd DG, Whitworth DE. 2017. The myxobacterium Myxococcus xanthus can sense and respond to the quorum signals secreted by potential prey organisms. Front. Microbiol. 8:439
    [Google Scholar]
  107. 107. 
    Ellis BM, Fischer CN, Martin LB, Bachmann BO, McLean JA. 2019. Spatiochemically profiling microbial interactions with membrane scaffolded desorption electrospray ionization–ion mobility–imaging mass spectrometry and unsupervised segmentation. Anal. Chem. 91:13703–11
    [Google Scholar]
  108. 108. 
    Vos M, Velicer GJ. 2009. Social conflict in centimeter-and global-scale populations of the bacterium Myxococcus xanthus. Curr. Biol. 19:1763–67
    [Google Scholar]
  109. 109. 
    Onaka H, Tabata H, Igarashi Y, Sato Y, Furumai T. 2001. Goadsporin, a chemical substance which promotes secondary metabolism and morphogenesis in streptomycetes. I. Purification and characterization. J. Antibiot. 54:1036–44
    [Google Scholar]
  110. 110. 
    Amano S-I, Morota T, Kano Y-K, Narita H, Hashidzume T et al. 2010. Promomycin, a polyether promoting antibiotic production in Streptomyces spp. J. Antibiot. 63:486–91
    [Google Scholar]
  111. 111. 
    Khalil ZG, Cruz-Morales P, Licona-Cassani C, Marcellin E, Capon RJ 2019. Inter-kingdom beach warfare: Microbial chemical communication activates natural chemical defences. ISME J 13:147–58
    [Google Scholar]
  112. 112. 
    Onaka H, Mori Y, Igarashi Y, Furumai T. 2011. Mycolic acid-containing bacteria induce natural-product biosynthesis in Streptomyces species. Appl. Environ. Microbiol. 77:400–6
    [Google Scholar]
  113. 113. 
    Hoshino S, Wakimoto T, Onaka H, Abe I 2015. Chojalactones A–C, cytotoxic butanolides isolated from Streptomyces sp. cultivated with mycolic acid containing bacterium. Org. Lett. 17:1501–4
    [Google Scholar]
  114. 114. 
    Derewacz DK, Covington BC, McLean JA, Bachmann BO. 2015. Mapping microbial response metabolomes for induced natural product discovery. ACS Chem. Biol. 10:1998–2006
    [Google Scholar]
  115. 115. 
    Hoshino S, Okada M, Wakimoto T, Zhang H, Hayashi F et al. 2015. Niizalactams A–C, multicyclic macrolactams isolated from combined culture of Streptomyces with mycolic acid-containing bacterium. J. Nat. Prod. 78:3011–17
    [Google Scholar]
  116. 116. 
    Hoshino S, Wong CP, Ozeki M, Zhang H, Hayashi F et al. 2018. Umezawamides, new bioactive polycyclic tetramate macrolactams isolated from a combined-culture of Umezawaea sp. and mycolic acid-containing bacterium. J. Antibiot. 71:653–57
    [Google Scholar]
  117. 117. 
    Cueto M, Jensen PR, Kauffman C, Fenical W, Lobkovsky E, Clardy J. 2001. Pestalone, a new antibiotic produced by a marine fungus in response to bacterial challenge. J. Nat. Prod. 64:1444–46
    [Google Scholar]
  118. 118. 
    Oh DC, Jensen PR, Kauffman CA, Fenical W. 2005. Libertellenones A–D: induction of cytotoxic diterpenoid biosynthesis by marine microbial competition. Bioorg. Med. Chem. 13:5267–73
    [Google Scholar]
  119. 119. 
    Schroeckh V, Scherlach K, Nützmann HW, Shelest E, Schmidt-Heck W et al. 2009. Intimate bacterial-fungal interaction triggers biosynthesis of archetypal polyketides in Aspergillus nidulans. PNAS 106:14558–63
    [Google Scholar]
  120. 120. 
    Nützmann HW, Reyes-Dominguez Y, Scherlach K, Schroeckh V, Horn F et al. 2011. Bacteria-induced natural product formation in the fungus Aspergillus nidulans requires Saga/Ada-mediated histone acetylation. PNAS 108:14282–87
    [Google Scholar]
  121. 121. 
    Nichols D, Cahoon N, Trakhtenberg EM, Pham L, Mehta A et al. 2010. Use of ichip for high-throughput in situ cultivation of “uncultivable” microbial species. Appl. Environ. Microbiol. 76:2445–50
    [Google Scholar]
  122. 122. 
    Ling LL, Schneider T, Peoples AJ, Spoering AL, Engels I et al. 2015. A new antibiotic kills pathogens without detectable resistance. Nature 517:455–59
    [Google Scholar]
  123. 123. 
    Hao Y-Q, Zhao X-F, Zhang D-Y. 2016. Field experimental evidence that stochastic processes predominate in the initial assembly of bacterial communities. Environ. Microbiol. 18:1730–39
    [Google Scholar]
  124. 124. 
    Pishchany G, Mevers E, Ndousse-Fetter S, Horvath DJ Jr., Paludo CR et al. 2018. Amycomicin is a potent and specific antibiotic discovered with a targeted interaction screen. PNAS 115:10124–29
    [Google Scholar]
  125. 125. 
    Schloss PD, Handelsman J. 2005. Metagenomics for studying unculturable microorganisms: cutting the Gordian knot. Genome Biol 6:229
    [Google Scholar]
  126. 126. 
    Garcie C, Tronnet S, Garénaux A, McCarthy AJ, Brachmann AO et al. 2016. The bacterial stress-responsive Hsp90 chaperone (HtpG) is required for the production of the genotoxin colibactin and the siderophore yersiniabactin in Escherichia coli. J. Infect. Dis. 214:916–24
    [Google Scholar]
  127. 127. 
    Washio K, Lim SP, Roongsawang N, Morikawa M. 2010. Identification and characterization of the genes responsible for the production of the cyclic lipopeptide arthrofactin by Pseudomonas sp. MIS38. Biosci. Biotech. Biochem 74:992–99
    [Google Scholar]
  128. 128. 
    Vivien E, Megessier S, Pieretti I, Cociancich S, Frutos R et al. 2005. Xanthomonas albilineans HtpG is required for biosynthesis of the antibiotic and phytotoxin albicidin. FEMS Microbiol. Lett. 251:81–89
    [Google Scholar]
  129. 129. 
    Hong ST, Carney JR, Gould SJ. 1997. Cloning and heterologous expression of the entire gene clusters for PD 116740 from Streptomyces strain WP 4669 and tetrangulol and tetrangomycin from Streptomyces rimosus NRRL 3016. J. Bacteriol. 179:470–76
    [Google Scholar]
  130. 130. 
    Brady SF, Chao CJ, Handelsman J, Clardy J. 2001. Cloning and heterologous expression of a natural product biosynthetic gene cluster from eDNA. Org. Lett. 3:1981–84
    [Google Scholar]
  131. 131. 
    Blodgett JAV, Zhang JK, Metcalf WW. 2005. Molecular cloning, sequence analysis, and heterologous expression of the phosphinothricin tripeptide biosynthetic gene cluster from Streptomyces viridochromogenes DSM 40736. Antimicrob. Agents Chemother. 49:230–40
    [Google Scholar]
  132. 132. 
    Eliot AC, Griffin BM, Thomas PM, Johannes TW, Kelleher NL et al. 2008. Cloning, expression, and biochemical characterization of Streptomyces rubellomurinus genes required for biosynthesis of antimalarial compound FR900098. Chem. Biol. 15:765–70
    [Google Scholar]
  133. 133. 
    Bonet B, Teufel R, Crüsemann M, Ziemert N, Moore BS. 2015. Direct capture and heterologous expression of Salinispora natural product genes for the biosynthesis of enterocin. J. Nat. Prod. 78:539–42
    [Google Scholar]
  134. 134. 
    Ross AC, Gulland LE, Dorrestein PC, Moore BS. 2015. Targeted capture and heterologous expression of the Pseudoalteromonas alterochromide gene cluster in Escherichia coli represents a promising natural product exploratory platform. ACS Synth. Biol. 4:414–20
    [Google Scholar]
  135. 135. 
    Tang X, Li J, Millán-Aguiñaga N, Zhang JJ, O'Neill EC et al. 2015. Identification of thiotetronic acid antibiotic biosynthetic pathways by target-directed genome mining. ACS Chem. Biol. 10:2841–49
    [Google Scholar]
  136. 136. 
    Yamanaka K, Reynolds KA, Kersten RD, Ryan KS, Gonzalez DJ et al. 2014. Direct cloning and refactoring of a silent lipopeptide biosynthetic gene cluster yields the antibiotic taromycin A. PNAS 111:1957–62
    [Google Scholar]
  137. 137. 
    Kang H-S, Brady SF. 2014. Arixanthomycins A–C: phylogeny-guided discovery of biologically active eDNA-derived pentangular polyphenols. ACS Chem. Biol. 9:1267–72
    [Google Scholar]
  138. 138. 
    Kang H-S, Brady SF. 2014. Mining soil metagenomes to better understand the evolution of natural product structural diversity: pentangular polyphenols as a case study. J. Am. Chem. Soc. 136:18111–19
    [Google Scholar]
  139. 139. 
    Hover BM, Kim SH, Katz M, Charlop-Powers Z, Owen JG et al. 2018. Culture-independent discovery of the malacidins as calcium-dependent antibiotics with activity against multidrug-resistant Gram-positive pathogens. Nat. Microbiol. 3:415–22
    [Google Scholar]
  140. 140. 
    Jiang W, Zhao X, Gabrieli T, Lou C, Ebenstein Y, Zhu TF. 2015. Cas9-assisted targeting of chromosome segments CATCH enables one-step targeted cloning of large gene clusters. Nat. Commun. 6:8101
    [Google Scholar]
  141. 141. 
    Tu Q, Herrmann J, Hu S, Raju R, Bian X et al. 2016. Genetic engineering and heterologous expression of the disorazol biosynthetic gene cluster via Red/ET recombineering. Sci. Rep. 6:21066
    [Google Scholar]
  142. 142. 
    Horbal L, Marques F, Nadmid S, Mendes MV, Luzhetskyy A. 2018. Secondary metabolites overproduction through transcriptional gene cluster refactoring. Metab. Eng. 49:299–315
    [Google Scholar]
  143. 143. 
    Bian X, Plaza A, Yan F, Zhang Y, Müller R. 2015. Rational and efficient site-directed mutagenesis of adenylation domain alters relative yields of luminmide derivatives in vivo. Biotechnol. Bioeng. 112:1343–53
    [Google Scholar]
  144. 144. 
    Li S, Wang J, Li X, Yin S, Wang W, Yang K 2015. Genome-wide identification and evaluation of constitutive promoters in streptomycetes. Microb. Cell. Fact. 14:172
    [Google Scholar]
  145. 145. 
    Luo Y, Zhang L, Barton KW, Zhao H. 2015. Systematic identification of a panel of strong constitutive promoters from Streptomyces albus. ACS Synth. Biol. 4:1001–10
    [Google Scholar]
  146. 146. 
    Smanski MJ, Bhatia S, Zhao D, Park Y, Woodruff LBA et al. 2014. Functional optimization of gene clusters by combinatorial design and assembly. Nat. Biotechnol. 32:1241–49
    [Google Scholar]
  147. 147. 
    Panter F, Krug D, Baumann S, Müller R. 2018. Self-resistance guided genome mining uncovers new topoisomerase inhibitors from myxobacteria. Chem. Sci. 9:4898–908
    [Google Scholar]
  148. 148. 
    Billingsley JM, DeNicola AB, Tang Y. 2016. Technology development for natural product biosynthesis in Saccharomyces cerevisiae. Curr. Opin. Biotechnol. 42:74–83
    [Google Scholar]
  149. 149. 
    Krug D, Müller R. 2014. Secondary metabolomics: the impact of mass spectrometry-based approaches on the discovery and characterization of microbial natural products. Nat. Product. Rep. 31:768–83
    [Google Scholar]
  150. 150. 
    Bachmann BO, Van Lanen SG, Baltz RH 2014. Microbial genome mining for accelerated natural products discovery: Is a renaissance in the making?. J. Ind. Microbiol. Biotechnol. 41:175–84
    [Google Scholar]
/content/journals/10.1146/annurev-biochem-081420-102432
Loading
/content/journals/10.1146/annurev-biochem-081420-102432
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error