Next Article in Journal / Special Issue
Comparative Genomics among Closely Related Streptomyces Strains Revealed Specialized Metabolite Biosynthetic Gene Cluster Diversity
Previous Article in Journal
Relationship between Maternal and First Year of Life Dispensations of Antibiotics and Antiasthmatics
Previous Article in Special Issue
Natural Products from Actinobacteria Associated with Fungus-Growing Termites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mining Actinomycetes for Novel Antibiotics in the Omics Era: Are We Ready to Exploit This New Paradigm?

Fundación MEDINA, Avda Conocimiento 34, 18016 Granada, Spain
Antibiotics 2018, 7(4), 85; https://doi.org/10.3390/antibiotics7040085
Submission received: 6 August 2018 / Revised: 20 September 2018 / Accepted: 21 September 2018 / Published: 25 September 2018
(This article belongs to the Special Issue Actinomycetes: The Antibiotics Producers)

Abstract

:
The current spread of multi-drug resistance in a number of key pathogens and the lack of therapeutic solutions in development to address most of the emerging infections in the clinic that are difficult to treat have become major concerns. Microbial natural products represent one of the most important sources for the discovery of potential new antibiotics and actinomycetes have been one of the most relevant groups that are prolific producers of these bioactive compounds. Advances in genome sequencing and bioinformatic tools have collected a wealth of knowledge on the biosynthesis of these molecules. This has revealed the broad untapped biosynthetic diversity of actinomycetes, with large genomes and the capacity to produce more molecules than previously estimated, opening new opportunities to identify the novel classes of compounds that are awaiting to be discovered. Comparative genomics, metabolomics and proteomics and the development of new analysis and genetic engineering tools provide access to the integration of new knowledge and better understanding of the physiology of actinomycetes and their tight regulation of the production of natural products antibiotics. This new paradigm is fostering the development of new genomic-driven and culture-based strategies, which aims to deliver new chemical classes of antibiotics to be developed to the clinic and replenish the exhausted pipeline of drugs for fighting the progression of infection diseases in the near future.

1. Introduction

Actinomycetes are accepted as one of the most relevant bacterial groups as prolific producers of secondary metabolites (SM). For decades, they were intensively exploited by industrial discovery programs that deliveredmost of the natural products, which were subsequently used as scaffolds to derive a large number of the antibiotics that are currently in the clinic [1,2]. Despite this past success, the current spread of multi-drug resistance in a number of key pathogens (Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa and Enterobacter spp.) and the lack of therapeutic solutions in development to address most of the emerging infections in the clinic that are difficult to treat have become major health concerns and require a new and sustained research effort to respond to the need for new antibiotics [3,4,5]. The progressive abandonment of the field by large pharmaceutical companies, moving away from the high costs of development and poor incentives as part of a broken economic model, has left academic groups and small biotech companies as the only players in the discovery field. These research teams are confronted with the major challenge of identifying novel antibiotic classes that they may not have the capacity to develop alone to reach the market [6,7]. The international initiatives launched in the last few years are supporting the development of new antibiotics and are revitalizing new preclinical and clinical developments. In contrast, basic research and early discovery efforts still remain poorly addressed by these programs and the existing research gaps are not enabling the establishment of a sustainable early discovery pipeline for future new compounds from the innovative approaches that have emerged in the field in last few decades [8,9]. This limited number of clinical development programs is in direct contrast to the dynamic and prolific activity that has developed in the field of microbial natural products during the last decade [10,11]. Research in natural products has evolved to represent one of the most important sources for the discovery of potential new antibiotics, compared to the lack of success of synthetic molecules. The poor suitability of synthetic libraries that were used to select molecules from target-based in vitro screens, or to develop rational drug designs based on ligand binding devoid of suitable physiochemical properties for penetrating bacterial membranes has been extensively discussed in previous reports [12,13,14,15]. Despite the rediscovery problem of known compounds, which has been represented as one of the major burdens for the continued investment in traditional NP discovery programs in the past, the advances in genome sequencing and bioinformatic tools have permitted us to collect a wealth of knowledge on the biosynthesis of natural products, which has revealed the broad untapped biosynthetic diversity of microorganisms, especially actinomycetes [16]. These talented filamenting bacteria have been shown to encode a previously unexpected diversity of biosynthetic gene clusters, opening new opportunities to identify novel classes of compounds that are awaiting to be discovered [16,17,18]. New advances in cultivation techniques from unexploited microbial niches have provided new insights into the diversity of actinobacteria and the possibility to access a broader chemical space of bioactive compounds. Comparative genomics, metabolomics and proteomics and the development of new analysis tools are generating a wealth of integrated information, which is fostering the emergence of new strategies that are aimed at obtaining a better understanding of the physiology of actinomycetes and their production of natural products. The new methods developed to mine genomes and predict chemical structures, to activate silent clusters and enable pathway engineering and to directly express metagenome gene clusters in heterologous hosts are providing the necessary tools to set up the basis for a new antibiotic discovery paradigm. Many recent excellent reviews cover each of the individual aspects of these new developments in the field and their impact on the identification of new antibiotics [19,20,21,22]. This short review is mostly focused on revisiting the most recent contributions in the field of actinomycetes in terms of their natural products, which have an impact on the discovery of novel scaffolds. These new approaches could bring new opportunities for the production of novel classes of compounds to address the current therapeutic needs in the multi-resistance era.

2. Exploiting Diversity of Cultured Actinomycetes

The continued interest in exploring the environment to untap novel sources of microbial diversity has resulted in extensive prospective studies on a broad variety of sources, which ranges from specific terrestrial extreme environments to unique microbial assemblages in plant–host associations and marine ecosystems. The distribution of some microbial species presents biogeographic patterns that are mostly determined by micro-environmental conditions and the most recent efforts have been focused at exploiting these still poorly explored habitats to discover new chemical diversity. The exploration of the extreme conditions of desert and arid habitats, such as the Atacama desert with high salinity and high levels of UV radiation, has permitted the isolation of new species of actinomycetes that are well adapted to surviving under these conditions [23,24]. These actinomycetes communities have shown to produce a wide diversity of novel compounds from different natural product classes, such as the new ansamycins chaxamycins or the β-diketone polyketides asenjonamides A–C. This only names a few compounds with antibacterial activities that have been described recently [25,26] (Table 1) (Figure 1). Cave microbiomes are another pristine eco-system that have frequently been studied in the search for novel producing strains and many reports have highlighted the isolation of novel species of actinomycetes that produce bioactive compounds [27]. The actinobacterial diversity in calcite moonmilk deposits are one of the most recent examples showing the potential of these new isolates to create products with antibacterial activities [28]. Alternatively, the use of diffusion chamber methods to grow previously uncultured soil bacteria has permitted to isolate new antibiotics, such as lassomycin [29].
The marine environment has been traditionally another source of new actinomycetes [42,43]. The broad diversity of marine ecosystems, ranging from mangroves, shallow waters, deep sea sediments and associated invertebrates, have continued to attract the interest of microbiologists. Their isolation programs have ensured a continued discovery of new strains that produce new compounds or new analogs with biological activity (Table 1 and Figure 1) [30,31,32,33,34,43,44,45,46,47,48]. Despite this prolific description of marine-derived strains and new antibiotic producers, this environment remains poorly studied in terms of microbial diversity and functional diversity. Marine sediments have been the focus of recent studies, which has shown that actinobacteria are in fact only a minor component of this microbial community. More interestingly, these findings have suggested that the production of many of the secondary metabolites have a deep impact on the microbial community composition [49]. There is no doubt that the parallel advances in the metagenomic assessment of microbial diversity have allowed us to explore the dynamics of the microbial populations of interest in the communities currently being studied. These technologies are opening new avenues to investigate the potential roles of these members and the effects of the antibacterial compounds on the microbiome composition. The results derived from these studies will guide the isolation and selection of the most promising members of these microbial communities and their screening for the production of novel bioactive compounds.

3. Genomics Driven Discovery

In parallel, the exponential increase in the number of partial and complete genome sequencing projects on the actinomycetes species available in the public databases not only have confirmed their broad biosynthetic diversity across the different lineages, but also have enabled intensive genome mining approaches to untap new natural products scaffolds. The identification of new biosynthetic pathways of potentially interesting new molecules has fostered the search for new biosynthetic clusters (BGCs) on draft genomes, which has taken advantage of improved bioinformatic tools for cluster identification and gene annotation, such as AntiSMASH [50,51,52]. Furthermore, the increasing numbers of almost complete genomes has permitted the extensive comparative genomic analyses of members of this bacterial group, and the identification of the genomic components and the evolutionary history of many different species [53,54,55]. These comparative studies are revealing the existence of a core genome within some members of actinomycetes as well as a divergence of BGCs among the different lineages. These results are providing a basis for understanding the functional evolution of species as shown for Streptomyces [56,57]. Another relevant aspect of the impact of the increasing number of BGCs sequence information on antibiotic discovery is the possibility of developing specific targeted genome mining searches in genomic libraries based on specific genomic signatures related to the biosynthesis of privileged scaffolds or functionalizations that could drive the discovery of novel compounds and chemical spaces (Table 1 and Figure 1) [35,36]. The integration of genomics with transcriptomics, proteomics and metabolomics is providing unique information for assessing the functional evolution of actinomycetes species. These data are encouraging the development of new approaches for the expression and engineering of these biosynthetic pathways and the identification of new bioactive compounds from cultured actinomycetes not only from underexplored habitats but also from large microbial collections that are still the untapped treasures of biosynthetic diversity (Figure 2).
One of the major challenges that still remains in this context is the efficient cloning and expression of BGCs that are originally silent or poorly expressed in their natural host after the use of refactoring by the replacement of the regulatory elements and further detection of the synthetized compounds [37,58,59,60]. Thus far, many different in vitro DNA assembly or direct capture methods have been described to clone BGCs in heterologous hosts [38,39,61,62,63]. Many BGCs cannot be detected by the rule-based bioinformatic tools due to the absence of signature genes, but the application of prediction tools based on the frequencies of Pfam domains occurring in BGCs have improved the identification of additional clusters [51]. New genomic bacterial artificial chromosome (BAC) libraries built from large 100-Kb fragments of Streptomyces spp. genomic DNA are also being used in the high throughput functional screening approaches to identify non-predicted BGCs by heterologous expression [64].

4. Eliciting Production from Silent Pathways

The activation of silent BGCs in the well-characterized producing strains by direct gene manipulation is an alternative approach when the strains are amenable to genetic engineering. Many recent successful examples have been reported, which have described the use of the newest genetic engineering tools in modifying metabolic pathways, altering metabolic fluxes that are blocking unrelated metabolic pathways, inactivating transcriptional repressors, over-expressing pathway-specific activator genes or even multiplying the BGC copies in the original producing strains to increase titers [63,65,66,67]. The recent development of optimal ribosomal binding sequences and strong terminators to be applied in the control of the metabolic pathways of actinomycetes has opened new possibilities for gene expression modulation and represent promising tools for metabolic engineering (Table 2) [68].
Despite this success, a large proportion of wild-type and industrial strains remain recalcitrant to being manipulated and many silent BGCs cannot be directly activated using genetic tools. Culture-based approaches based on multiple nutritional conditions have been one of the most common methods employed to explore the media components required for the production of new compounds. The concept of using small molecules as elicitors by perturbing biological systems and signaling pathways dates back several decades and has been extensively used to activate silent or poorly expressed pathways with a broad range of small molecules, including the sub-inhibitory concentrations of antibiotics [69,70]. The first large scale elicitor screening reported was performed with more than 30,000 compounds on S. coelicolor and it identified a small number of compounds that are able to stimulate the production of some of the secondary metabolites by several times [89]. Many examples of hormesis have been described with sub-inhibitory concentrations of antibiotics and other well-known bioactive natural products, which has elicited a response that is associated with the major activation of secondary metabolism, induction of cryptic gene clusters and production of novel compounds [70,90]. The activation effect cannot be predicted from the antibiotic mode of actions and the lack of universal effectors to awaken all silent BGCs has added another level of complexity in the identification of new effector molecules [70,71]. The same effect is pursued by co-cultivation, which is an approach that has been used extensively with many different types of cultivation formats and strain combinations. The approach takes advantage of the effect that small concentrations of signaling or antibiotic molecules from one of the strains can have on another strain. The difficulty to scale-up this approach as a general method is related to the impossibility of predicting which combinations will result in an effective response, which normally does not account for more than 15–20% of the cases studied [71]. Mycolic acids have been shown to play a role in the physical interaction and the activation of some silent pathways. New antibacterial compounds, such as alchivemycin A and B, arcyriaflavin E or ciromicins, were described after co-culturing different Rhodoccoci with the species of Streptomyces, Tsukamurella and Nocardiopsis (Table 1) (Figure 1) [40,41]. In other situations, the induction does not require cell-to-cell interaction and is only mediated by diffusible small effector molecules. One of the most recent examples is the production of the cryptic natural product keyicin from the co-cultivation of the producer Micromonospora strain with Rhodococcus [91] (Table 2).
Most recent studies on the differential metabolomic analysis of metabolites in response to cultivation conditions have shown that the chemical potential of actinomycetes is far from being fully characterized [92]. From a methodological perspective, the modern LC-MS and NMR analytical tools and differential metabolomic analysis have been the determinant for detecting the production of novel compounds in complex mixtures and mapping the response to external chemical signals. Comparative metabolomics have been efficiently used to identify the induced expression of secondary metabolites from S. coelicolor cryptic genes resulting from the exposure to multiplexed perturbations and to identify the subsets of primary and secondary metabolites that respond similarly across a large variety of stimuli [72]. The major challenge for these methods is related to the identification of novel bioactive molecules within the complexity of the metabolomic profiles. New dereplication and identification approaches are continuously being developed, which are based on the similarity of MS/MS patterns in natural product databases and NMR-based metabolomics [73,74,75]. Proteomining is another method developed to support this identification. The analysis links natural products to biosynthetic enzymes as it correlates protein expression profiles of biosynthetic enzymes to the metabolome of the producing strains based on the statistical analysis of strains cultured under different conditions [76,77].

5. Harnessing Regulation of Primary and Secondary Metabolisms

The production of secondary metabolites in actinomycetes is tightly regulated and responds to external stimuli from the environment. This regulation is the result of different classes of pathway-specific regulatory elements involving two-component systems, extra-cytoplasmic sigma factors or pathway specific regulators, such as the most recently described LmbU family [78,79,80]. Triggering the expression of BGCs frequently involves an additional transcriptional response through the master regulators involved in global regulatory metabolic networks that are not always pathway specific [81]. Understanding the right combinations of regulatory elements and transcription factors that regulate a BGC has been proposed as the “cracking the code” approach to be followed in identifying the key regulatory signals and the eliciting signals needed to set up culture conditions to activate a specific BGC [83]. The regulon predictor program PREDetector was developed to identify signaling cascades deduced from the in silico searches of regulatory elements, which revealed that primary metabolism transcription factors were also involved in controlling pathway-specific regulators [84]. One of the best examples are the pleiotropic regulators, DasR and CebR, that control the uptake of chitin and cellulose, for which responsive elements were identified upstream of many pathway-specific regulators [85,86,87]. The increasing number of available new genomes are enabling the new in silico approaches derived from comparative genomics to detect specific binding motifs of transcription factor orthologues [82]. Regulon prediction has been proposed as a successful strategy to identify the regulatory networks involved in the control of BGC expression and the most promising strains to be explored for the induction of silent pathways. This finely tuned participation of regulatory requirements for the specialized metabolite production also requires the precise provision of chemical precursors from primary metabolism. Recent reports highlight the role of primary metabolism gene expansions in secondary metabolite producing strains with impact on metabolic adaptation and strain fitness and how they may represent another target for future genetic engineering interventions to improve production and activate silent pathways [88] (Table 2).

6. Conclusions and Future Prospects

Microbial prospection studies have continued to reveal that there is a huge and still poorly explored diversity of actinomycetes in the environment that is waiting to be mined for the isolation of new bioactive compounds. In addition, thousands of selected strains that are preserved in microbial collections and distributed across laboratories worldwide also should be revisited as they represent a unique reservoir of silent biosynthetic diversity that traditional approaches did not manage to unlock in its full extension. The continued increase in new genome sequences and the development of new genome-mining and genome-directed engineering tools to trigger the production of new natural products are paralleling the development of integrated analytical tools and open access databases. The application of comparative genomics, metabolomics and proteomics to culture-based studies is providing a wealth of knowledge on the physiology and the regulation systems, opening new avenues to approach the activation of silent or poorly expressed pathways. There is no doubt that all these advances are setting the new foundations for a new paradigm in natural product discovery, especially from actinomycetes. A sustained and integrated multidisciplinary research effort should respond to the major challenge of discovering new chemical classes of antibiotics that will be required to replenish the preclinical development pipeline in the near future.

Funding

This research received no external funding.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Berdy, J.J. Thoughts and facts about antibiotics: Where we are now and where we are heading. J. Antibiot. 2012, 65, 385–395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Landwehr, W.; Wolf, C.; Wink, J. Actinobacteria and myxobacteria-two of the most important bacterial resources for novel antibiotics. Curr. Top. Microbiol. Immunol. 2016, 398, 273–302. [Google Scholar] [PubMed]
  3. Fischbach, M.A.; Walsh, C.T. Antibiotics for emerging pathogens. Science 2009, 325, 1089–1093. [Google Scholar] [CrossRef] [PubMed]
  4. Silver, L.L. Challenges of antibacterial discovery. Clin. Microbiol. Rev. 2011, 24, 71–109. [Google Scholar] [CrossRef] [PubMed]
  5. Walsh, C.T.; Wencewicz, T.A. Prospects for new antibiotics: A molecule-centered perspective. J. Antibiot. 2014, 67, 7–22. [Google Scholar] [CrossRef] [PubMed]
  6. Rex, J.H.; Outterson, K. Antibiotic reimbursement in a sales-delinked model: Context and a benchmark-based global approach. Lancet Infect. Dis. 2016, 16, 500–505. [Google Scholar] [CrossRef]
  7. Editorial. Wanted: A reward for antibiotic development. Nat. Biotechnol. 2018, 36, 555. [Google Scholar] [CrossRef] [PubMed]
  8. Rex, J.H.; Goldberger, M.; Eisenstein, B.; Harney, C. The evolution of the regulatory framework for antibacterial agents. Ann. N. Y. Acad. Sci. 2014, 1323, 11–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Simpkin, V.L.; Renwick, M.J.; Kelly, R.; Mossialos, E. Incentivising innovation in antibiotic drug discovery and development: Progress, challenges and next steps. J. Antibiot. 2017, 70, 1087–1096. [Google Scholar] [CrossRef] [PubMed]
  10. Antibiotics Currently in Global Clinical Development. Available online: http://www.pewtrusts.org/-/media/assets/2018/03/antibiotics_clinical_dev_table_february2018.pdf (accessed on 1 February 2018).
  11. Brown, E.D.; Wright, G.D. Antibacterial drug discovery in the resistance era. Nature 2016, 529, 336–343. [Google Scholar] [CrossRef] [PubMed]
  12. Payne, D.J.; Gwynn, M.N.; Holmes, D.J.; Pompliano, D.L. Drugs for bad bugs: Confronting the challenges of antibacterial discovery. Nat. Rev. Drug Discov. 2007, 6, 29–40. [Google Scholar] [CrossRef] [PubMed]
  13. Silver, L.L. Are natural products still the best source for antibacterial discovery? The bacterial entry factor. Expert Opin. Drug Discov. 2008, 3, 487–500. [Google Scholar] [CrossRef] [PubMed]
  14. Wright, G.D. Something old, something new: Revisiting natural products in antibiotic drug discovery. Can. J. Microbiol. 2014, 60, 147–154. [Google Scholar] [CrossRef] [PubMed]
  15. Lewis, K. New approaches to antimicrobial discovery. Biochem. Pharmacol. 2017, 134, 87–98. [Google Scholar] [CrossRef] [PubMed]
  16. Baltz, R.H. Gifted microbes for genome mining and natural product discovery. J. Ind. Microbiol. Biotechnol. 2017, 44, 573–588. [Google Scholar] [CrossRef] [PubMed]
  17. Nett, M.; Ikeda, H.; Moore, B.S. Genomic basis for natural product biosynthetic diversity in the actinomycetes. Nat. Prod. Rep. 2009, 26, 1362–1384. [Google Scholar] [CrossRef] [PubMed]
  18. Rutledge, P.J.; Challis, G.L. Discovery of microbial natural products by activation of silent biosynthetic gene clusters. Nat. Rev. Microbiol. 2015, 13, 509–523. [Google Scholar] [CrossRef] [PubMed]
  19. Genilloud, O. Actinomycetes: Still a source of novel antibiotics. Nat. Prod. Rep. 2017, 34, 1203–1232. [Google Scholar] [CrossRef] [PubMed]
  20. Katz, L.; Baltz, R.H. Natural product discovery: Past, present, and future. J. Ind. Microbiol. Biotechnol. 2016, 43, 155–176. [Google Scholar] [CrossRef] [PubMed]
  21. Kealey, C.; Creaven, C.A.; Murphy, C.D.; Brady, C.B. New approaches to antibiotic discovery. Biotechnol. Lett. 2017, 39, 805–817. [Google Scholar] [CrossRef] [PubMed]
  22. Wencewicz, T.A. New antibiotics from nature’s chemical inventory. Bioorg. Med. Chem. 2016, 24, 6227–6252. [Google Scholar] [CrossRef] [PubMed]
  23. Bull, A.T.; Asenjo, J.A.; Goodfellow, M.; Gomez-Silva, B. The atacama desert: Technical resources and the growing importance of novel microbial diversity. Ann. Rev. Microbiol. 2016, 70, 215–234. [Google Scholar] [CrossRef] [PubMed]
  24. Mohammadipanah, F.; Wink, J. Actinobacteria from arid and desert habitats: Diversity and biological activity. Front. Microbiol. 2016, 6, 1541. [Google Scholar] [CrossRef] [PubMed]
  25. Rateb, M.E.; Ebel, R.; Jaspars, M. Natural product diversity of actinobacteria in the atacama desert. Antonie Van Leeuwenhoek 2018, 111, 1467–1477. [Google Scholar] [CrossRef] [PubMed]
  26. Abdelkader, M.S.A.; Philippon, T.; Asenjo, J.A.; Bull, A.T.; Goodfellow, M.; Ebel, R.; Jaspars, M.; Rateb, M.E. Asenjonamides A-C, antibacterial metabolites isolated from Streptomyces asenjonii strain KNN 42.f from an extreme-hyper arid Atacama Desert soil. J. Antibiot. 2018, 71, 425–431. [Google Scholar] [CrossRef] [PubMed]
  27. Ghosh, S.; Kuisiene, N.; Cheeptham, N. The cave microbiome as a source for drug discovery: Reality or pipe dream? Biochem. Pharmacol. 2017, 134, 18–34. [Google Scholar] [CrossRef] [PubMed]
  28. Adam, D.; Maciejewska, M.; Naômé, A.; Martinet, L.; Coppieters, W.; Karim, L.; Baurain, D.; Rigali, S. Isolation, characterization, and antibacterial activity of hard-to-culture actinobacteria from cave moonmilk deposits. Antibiotics 2018, 7, 28–48. [Google Scholar] [CrossRef] [PubMed]
  29. Gavrish, E.; Sit, C.S.; Cao, S.; Kandror, O.; Spoering, A.; Peoples, A.; Ling, L.A.; Hughes, D.; Bissell, A.; Torrey, H.; et al. Lassomycin, a ribosomally synthesized cyclic peptide, kills mycobacterium tuberculosis by targeting the ATP-dependent protease ClpC1P1P2. Chem. Biol. 2014, 21, 509–518. [Google Scholar] [CrossRef] [PubMed]
  30. Jang, K.H.; Nam, S.-J.; Locke, J.B.; Kauffman, C.A.; Beatty, D.S.; Paul, L.A.; Fenical, W. Anthracimycin, a potent anthrax antibiotic from a marine-derived actinomycete. Chem. Int. Ed. 2013, 52, 7822–7824. [Google Scholar] [CrossRef] [PubMed]
  31. Hassan, H.M.; Degen, D.; Jang, K.H.; Ebright, R.H.; Fenical, W.; Salinamide, F. New depsipeptide antibiotic and inhibitor of bacterial RNA polymerase from a marine-derived Streptomyces sp. J. Antibiot. 2015, 68, 206–209. [Google Scholar] [CrossRef] [PubMed]
  32. Palomo, S.; González, I.; Martín, J.; de la Cruz, M.; Vicente, F.; Reyes, F.; Tormo, J.R.; Anderson, M.; Hill, R.T.O.; Genilloud, O. Sponge-derived Kocuria and Micrococcus spp. as sources of the new thiazolyl peptide antibiotic kocurin. Mar. Drugs 2013, 11, 1071–1086. [Google Scholar] [CrossRef] [PubMed]
  33. Zhang, Y.; Adnani, N.; Braun, D.R.; Ellis, G.A.; Barns, K.J.; Parker-Nance, S.; Guzei, I.A.; Bugni, T.S. Micromonohalimanes A and B: Antibacterial halimane-type diterpenoids from a marine Micromonospora species. J. Nat. Prod. 2016, 79, 2968–2972. [Google Scholar] [CrossRef] [PubMed]
  34. Pérez-Bonilla, M.; Oves-Costales, D.; de la Cruz, M.; Kokkini, M.; Martín, J.; Vicente, F.; Genilloud, O.; Reyes, F. Phocoenamicins B and C, New antibacterial spirotetronates isolated from a marine Micromonospora sp. Mar. Drugs 2018, 16. [Google Scholar] [CrossRef] [PubMed]
  35. Ju, K.-S.; Gao, J.; Doroghazi, J.R.; Wang, K.-K.A.; Thibodeaux, C.J.; Lia, S.; Metzger, E.; Fudala, J.; Su, J.; Zhang, J.K.; et al. Discovery of phosphonic acid natural products by mining the genomes of 10,000 actinomycetes. Proc. Natl. Acad. Sci. USA 2015, 112, 12175–12180. [Google Scholar] [CrossRef] [PubMed]
  36. Tang, X.; Li, J.; Millán-Aguiñaga, N.; Zhang, J.J.; O’Neill, E.C.; Ugalde, J.A.; Jensen, P.R.; Mantovani, S.M.; Moore, B.S. Identification of thiotetronic acid antibiotic biosynthetic pathways by target-directed genome mining. ACS Chem. Biol. 2015, 10, 2841–2849. [Google Scholar] [CrossRef] [PubMed]
  37. Yamanaka, K.; Reynolds, K.A.; Kersten, R.D.; Ryan, K.S.; Gonzalez, D.J.; Nizet, V.; Dorrestein, P.C.; Moore, B.S. Direct cloning and refactoring of a silent lipopeptide biosynthetic gene cluster yields the antibiotic taromycin A. Proc. Natl. Acad. Sci. USA 2014, 111, 1957–1962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Bonet, B.; Teufel, R.; Cruüsemann, M.; Ziemert, N.; Moore, B.S. Direct capture and heterologous expression of Salinispora natural product genes for the biosynthesis of enterocin. J. Nat. Prod. 2014, 78, 539–542. [Google Scholar] [CrossRef] [PubMed]
  39. Yang, C.; Huang, C.; Zhang, W.; Zhu, Y.; Zhang, C. Heterologous expression of fluostatin gene cluster leads to a bioactive heterodimer. Org. Lett. 2015, 17, 5324–5327. [Google Scholar] [CrossRef] [PubMed]
  40. Onaka, H. Novel antibiotic screening methods to awaken silent or cryptic secondary metabolic pathways in actinomycetes. J. Antibiot. 2017, 70, 865–870. [Google Scholar] [CrossRef] [PubMed]
  41. Derewacz, D.K.; Covington, B.C.; McLean, J.A.; Bachmann, B.O. Mapping microbial response metabolomes for induced natural product discovery. ACS Chem. Biol. 2015, 10, 1998–2006. [Google Scholar] [CrossRef] [PubMed]
  42. Fenical, W.; Jensen, P.R. Developing a new resource for drug discovery: Marine actinomycete bacteria. Nat. Chem. Biol. 2006, 2, 666–673. [Google Scholar] [CrossRef] [PubMed]
  43. Bull, A.T.; Stach, J.E.M. Marine actinobacteria: New opportunities for natural product search and discovery. Trends Microbiol. 2007, 15, 491–499. [Google Scholar] [CrossRef] [PubMed]
  44. Choudhary, A.; Naughton, L.M.; Montánchez, I.; Dobson, A.D.W.; Rai, D.K. Current status and future prospects of marine natural products (MNPs) as antimicrobials. Mar. Drugs 2017, 15, 272. [Google Scholar] [CrossRef] [PubMed]
  45. Dhakal, D.; Pokhrel, A.R.; Shrestha, B.; Sohng, J.K. Marine rare actinobacteria: Isolation, characterization, and strategies for harnessing bioactive compounds. Front. Microbiol. 2017, 8, 1106. [Google Scholar] [CrossRef] [PubMed]
  46. Xu, D.; Han, L.; Li, C.; Cao, Q.; Zhu, D.; Barrett, N.H.; Harmody, D.; Chen, J.; Zhu, H.; McCarthy, P.J.; et al. Bioprospecting deep-sea actinobacteria for novel anti-infective natural products. Front. Microbiol. 2018, 9, 787. [Google Scholar] [CrossRef] [PubMed]
  47. Jose, P.A.; Jha, B. Intertidal marine sediment harbours actinobacteria with promising bioactive and biosynthetic potential. Sci. Rep. 2017, 7, 10041. [Google Scholar] [CrossRef] [PubMed]
  48. Sarmiento-Vizcaíno, A.; González, V.; Braña, A.F.; Palacios, J.J.; Otero, L.; Fernández, J.; Molina, A.; Kulik, A.; Vázquez, F.; Acuña, J.L.; et al. Pharmacological potential of phylogenetically diverse actinobacteria isolated from deep-sea coral ecosystems of the submarine Avilés canyon in the cantabrian sea. Microb. Ecol. 2017, 73, 338–352. [Google Scholar] [CrossRef] [PubMed]
  49. Patin, N.V.; Schorn, M.; Aguinaldo, K.; Lincecum, T.; Moore, B.S.; Jensen, P.R. Effects of actinomycete secondary metabolites on sediment microbial communities. Appl. Environ. Microbiol. 2017, 83, e02676-16. [Google Scholar] [CrossRef] [PubMed]
  50. Medema, M.H.; Fischbach, M.A. Computational approaches to natural product discovery. Nat. Chem. Biol. 2015, 11, 639–648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Weber, T.; Kim, H.U. The secondary metabolite bioinformatics portal: Computational tools to facilitate synthetic biology of secondary metabolite production. Synth. Syst. Biotechnol. 2016, 1, 69–79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Blin, K.; Medema, M.H.; Kottmann, R.; Lee, S.Y.; Weber, T. The antiSMASH database, a comprehensive database of microbial secondary metabolite biosynthetic gene clusters. Nucleic Acids Res. 2017, 45, D555–D559. [Google Scholar] [CrossRef] [PubMed]
  53. Sánchez-Hidalgo, M.; González, I.; Díaz-Muñoz, C.; Martínez, G.; Genilloud, O. Comparative Genomics and Biosynthetic Potential Analysis of Two Lichen-Isolated Amycolatopsis Strains. Front. Microbiol. 2018, 9, 369. [Google Scholar] [CrossRef] [PubMed]
  54. Adamek, M.; Alanjary, M.; Sales-Ortells, H.; Goodfellow, M.; Bull, A.T.; Winkler, A.; Wibberg, D.; Kalinowski, J.; Ziemert, N. Comparative genomics reveals phylogenetic distribution patterns of secondary metabolites in Amycolatopsis species. BMC Genomics 2018, 19, 426. [Google Scholar] [CrossRef] [PubMed]
  55. Zhou, Z.; Gu, J.; Li, Y.-Q.; Wang, Y. Genome plasticity and systems evolution in Streptomyces. BMC Bioinform. 2012, 13, S8. [Google Scholar] [CrossRef] [PubMed]
  56. Kim, J.-N.; Kim, Y.; Jeong, Y.; Roe, J.-H.; Kim, B.-G.; Cho, B.-K. Comparative Genomics Reveals the Core and Accessory Genomes of Streptomyces Species. J. Microbiol. Biotechnol. 2015, 25, 1599–1605. [Google Scholar] [CrossRef] [PubMed]
  57. Choudoir, M.J.; Pepe-Ranney, C.; Buckley, D.H. Diversification of Secondary Metabolite Biosynthetic Gene Clusters Coincides with Lineage Divergence in Streptomyces. Antibiotics 2018, 7, 12. [Google Scholar] [CrossRef] [PubMed]
  58. Zhang, M.M.; Wang, Y.; Ang, E.L.; Zhao, H. Engineering microbial hosts for production of bacterial natural products. Nat. Prod. Rep. 2016, 33, 963–987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Myronovskyia, M.; Luzhetskyy, A. Native and engineered promoters in natural product discovery. Nat. Prod. Rep. 2016, 33, 1006–1020. [Google Scholar] [CrossRef] [PubMed]
  60. Weber, T.; Charusanti, P.; Musiol-Kroll, E.M.; Jiang, X.; Tong, Y.; Hyun Uk Kim, H.U.; Lee, S.Y. Metabolic engineering of antibiotic factories: New tools for antibiotic production in actinomycetes. Trends Biotechnol. 2015, 33, 15–26. [Google Scholar] [CrossRef] [PubMed]
  61. Ren, H.; Wang, B.; Zhao, H. Breaking the silence: New strategies for discovering novel natural products. Curr. Opin. Biotechnol. 2017, 48, 21–27. [Google Scholar] [CrossRef] [PubMed]
  62. Li, L.; Zhenga, G.; Chena, J.; Gec, M.; Jianga, W.; Lu, Y. New strategies and approaches for engineering biosynthetic gene clusters of microbial natural products. Metab. Eng. 2017, 40, 80–92. [Google Scholar] [CrossRef] [PubMed]
  63. Tong, Y.; Robertsen, H.L.; Blin, K.; Weber, T.; Lee, S.Y. CRISPR-Cas9 Toolkit for Actinomycete Genome Editing. In Synthetic Metabolic Pathways; Humana Press: New York, NY, USA, 2018; pp. 163–184. [Google Scholar]
  64. Xu, M.; Wang, Y.; Zhao, Z.; Gao, G.; Huang, S.-X.; Kang, Q.; He, X.; Lin, S.; Pang, X.; Deng, Z.; et al. Functional genome mining for metabolites encoded by large gene clusters through heterologous expression of a whole-genome bacterial artificial chromosome library in Streptomyces spp. Appl. Environ. Microbiol. 2016, 82, 5795–5805. [Google Scholar] [CrossRef] [PubMed]
  65. Arakawa, K. Manipulation of metabolic pathways controlled by signaling molecules, inducers of antibiotic production, for genome mining in Streptomyces spp. Antonie van Leeuwenhoek 2018, 111, 743–751. [Google Scholar] [CrossRef] [PubMed]
  66. Baltz, R.H. Genetic manipulation of secondary metabolite biosynthesis for improved production in Streptomyces and other actinomycetes. J. Ind. Microbiol. Biotechnol. 2016, 43, 343–370. [Google Scholar] [CrossRef] [PubMed]
  67. Hwang, K.-S.; Kim, H.U.; Charusanti, P.; Palsson, B.O.; Lee, S.Y. Systems biology and biotechnology of Streptomyces species for the production of secondary metabolites. Biotechnol. Adv. 2014, 32, 255–268. [Google Scholar] [CrossRef] [PubMed]
  68. Horbal, L.; Siegl, T.; Luzhetskyy, A. A set of synthetic versatile genetic control elements for the efficient expression of genes in Actinobacteria. Sci. Rep. 2018, 8, 491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Goh, E.B.; Yim, G.; Tsui, W.; McClure, J.; Surette, M.G.; Davies, J. Transcriptional modulation of bacterial gene expression by subinhibitory concentrations of antibiotics Proc. Natl. Acad. Sci. USA 2002, 99, 17025–17030. [Google Scholar] [CrossRef] [PubMed]
  70. Rosen, P.C.; Seyedsayamdost, M.R. Though much is taken, much abides: Finding new antibiotics using old ones. Biochemistry 2017, 56, 4925–4926. [Google Scholar] [CrossRef] [PubMed]
  71. Okada, B.K.; Seyedsayamdost, M.R. Antibiotic dialogues: Induction of silent biosynthetic gene clusters by exogenous small molecules. FEMS Microbiol. Rev. 2017, 41, 19–33. [Google Scholar] [CrossRef] [PubMed]
  72. Goodwin, C.R.; Covington, B.C.; Derewacz, D.K.; McNees, C.R.; Wikswo, J.P.; McLean, J.A.; Bachmann, B.O. Structuring microbial metabolic responses to multiplexed stimuli via self-organizing metabolomics maps. Chem. Biol. 2015, 22, 661–670. [Google Scholar] [CrossRef] [PubMed]
  73. Wang, M.; Carver, J.J.; Phelan, V.V.; Sanchez, L.M.; Garg, N.; Peng, Y.; Nguyen, D.D.; Watrous, J.; Kapono, C.A.; Luzzatto-Knaan, T.; et al. Sharing and community curation of mass spectrometry data with GNPS. Nat. Biotechnol. 2016, 34, 828–837. [Google Scholar] [CrossRef] [PubMed]
  74. Pérez-Victoria, I.; Martín, J.; Reyes, F. Combined LC/UV/MS and NMR strategies for the dereplication of marine natural products. Planta Med. 2016, 82, 857–871. [Google Scholar] [CrossRef] [PubMed]
  75. Wu, C.; Du, C.; Ichinose, K.; Choi, Y.H.; van Wezel, G.P. Discovery of C-Glycosylpyranonaphthoquinones in Streptomyces sp. MBT76 by a Combined NMR-Based Metabolomics and Bioinformatics Workflow. J. Nat. Prod. 2017, 80, 269. [Google Scholar] [CrossRef] [PubMed]
  76. Gubbens, J.; Zhu, H.; Girard, G.; Song, L.; Florea, B.I.; Aston, P.; Ichinose, K.; Filippov, D.V.; Choi, Y.H.; Overkleeft, H.S.; et al. Natural product proteomining, a quantitative proteomics platform, allows rapid discovery of biosynthetic gene clusters for different classes of natural products. Chem. Biol. 2014, 21, 707. [Google Scholar] [CrossRef] [PubMed]
  77. Chao Du, C.; van Wezel, G.P. Mining for microbial gems: Integrating proteomics in the postgenomic natural product discovery pipeline. Proteomics 2018. [Google Scholar] [CrossRef]
  78. Hou, B.; Lin, Y.; Wu, H.; Guo, M.; Petkovic, H.; Tao, L.; Zhu, X.; Ye, J.; Zhang, H. The novel transcriptional regulator LmbU promotes lincomycin biosynthesis through regulating expression of its target genes in Streptomyces lincolnensis. J. Bacteriol. 2018, 200, e00447-17. [Google Scholar] [CrossRef] [PubMed]
  79. Liu, G.; Chater, K.F.; Chandra, G.; Niu, G.; Tan, H. Molecular regulation of antibiotic biosynthesis in Streptomyces. Microbiol. Mol. Biol. Rev. 2013, 77, 112–143. [Google Scholar] [CrossRef] [PubMed]
  80. Van Wezel, G.P.; Mcdowall, K.J. The regulation of the secondary metabolism of Streptomyces: New links and experimental advances. Nat. Prod. Rep. 2011, 28, 1311–1333. [Google Scholar] [CrossRef] [PubMed]
  81. Hoskisson, P.A.; Fernández-Martínez, L.T. Regulation of specialised metabolites in actinobacteria–expanding the paradigms. Environ. Microbiol. Rep. 2018, 10, 231–238. [Google Scholar] [CrossRef] [PubMed]
  82. Bailey, T.L.J.; Grant, C.E.; Noble, W.S. The MEME suite. Nucleic Acids Res. 2015, 43, W39–W49. [Google Scholar] [CrossRef] [PubMed]
  83. Rigali, S.; Anderssen, S.; Naômé, A.; van Wezel, G.P. Cracking the regulatory code of biosynthetic gene clusters as a strategy for natural product discovery. Biochem. Pharmacol. 2018, 153, 24–34. [Google Scholar] [CrossRef] [PubMed]
  84. Hiard, S.; Marée, R.; Colson, S.; Hoskisson, P.A.; Titgemeyer, F.; van Wezel, G.P.B.; Wehenkel, J.L.; Rigali, S. PREDetector: A new tool to identify regulatory elements in bacterial genomes. Biochem. Biophys. Res. Commun. 2007, 357, 861–864. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Rigali, S.; Titgemeyer, F.; Barends, S.; Mulder, S.; Thomae, A.W.; Hopwood, D.A.; van Wezel, G.P. Feast or famine: The global regulator DasR links nutrient stress to antibiotic production by Streptomyces. EMBO Rep. 2008, 9, 670–675. [Google Scholar] [CrossRef] [PubMed]
  86. Craig, M.; Lambert, S.; Jourdan, S.; Tenconi, E.; Colson, S.; Maciejewska, M.; Ongena, M.; Martin, J.F.; van Wezel, G.; Rigali, S. Unsuspected control of siderophore production by N-acetylglucosamine in streptomycetes. Environ. Microbiol. Rep. 2012, 4, 512–521. [Google Scholar] [CrossRef] [PubMed]
  87. Jourdan, S.; Francis, I.M.; Kim, M.J.; Salazar, J.J.C.; Planckaert, S.; Frère, J.-M.; Matagne, A.; Kerff, F.; Devreese, B.; Loria, R.; et al. The CebE/MsiK transporter is a doorway to the cello-oligosaccharide-mediated Induction of Streptomyces scabies pathogenicity. Sci. Rep. 2016, 6, 27144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Schniete, J.K.; Cruz-Morales, P.; Selem-Mojica, N.; Fernández-Martínez, L.T.; Hunter, I.S.; Barona-Gómez, F.; Hoskisson, P.A. Expanding primary metabolism helps generatethe metabolic robustness to facilitate antibioticbiosynthesis in Streptomyces. mBio 2018, 9, e02283-17. [Google Scholar] [CrossRef] [PubMed]
  89. Craney, A.; Ozimok, C.; Pimentel-Elardo, S.M.; Capretta, A.; Nodwell, J.R. Chemical perturbation of secondary metabolism demonstrates important links to primary metabolism. Chem. Biol. 2012, 19, 1020–1027. [Google Scholar] [CrossRef] [PubMed]
  90. Seyedsayamdost, M.R. High-throughput platform for the discovery of elicitors of silent bacterial gene clusters. Proc. Natl. Acad. Sci. USA 2014, 111, 7266–7271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Adnani, N.; Chevrette, M.G.; Adibhatla, S.N.; Zhang, F.; Yu, Q.; Braun, D.R.; Nelson, J.; Simpkins, S.W.; McDonald, B.R.; Myers, C.L.; et al. Coculture of marine invertebrate-associated bacteria and interdisciplinary technologies enable biosynthesis and discovery of a new antibiotic, keyicin. ACS Chem. Biol. 2017, 12, 3093–3102. [Google Scholar] [CrossRef] [PubMed]
  92. Sengesa, C.H.R.; Al-Dilaimib, A.; Marchbankc, D.H.; Wibbergb, D.; Winklerb, A.; Haltlic, B.; Nowrousiand, M.; Kalinowskib, J.; Kerrc, R.G.; Bandow, J.E. The secreted metabolome of Streptomyces chartreusis and implications for bacterial chemistry. Proc. Natl. Acad. Sci. USA 2018, 115, 2490–2495. [Google Scholar] [CrossRef] [PubMed]
Figure 1. New antibiotics and analogs discovered from Actinomycetes.
Figure 1. New antibiotics and analogs discovered from Actinomycetes.
Antibiotics 07 00085 g001aAntibiotics 07 00085 g001b
Figure 2. Omics-driven discovery.
Figure 2. Omics-driven discovery.
Antibiotics 07 00085 g002
Table 1. New antibiotics and analogs described from Actinomycetes since 2013, following different mining approaches.
Table 1. New antibiotics and analogs described from Actinomycetes since 2013, following different mining approaches.
AntibioticStructural ClassProducing SpeciesAntibiotic SpectrumDiscovery ApproachReference
Asenjonamides A–Cdi-ketone polyketidesS. asenjoniiGram positive/
negative
Extreme environment[26]
Lassomycincyclic peptideLentzea kentukyensisM. tuberculosisDiffusion chambers[29]
AnthracimycintricarboxilicStreptomyces sp.Gram positiveMarine source[30]
Salinamide FdepsipeptideStreptomyces sp.Gram positive/
negative
Marine source[31]
KocurinthiazolylpeptideKocuria lacustris /Micrococcus sp.Gram positiveMarine source[32]
Micromonohalimanes A and BditerpenoidsMicromonospora sp.Gram positiveMarine source[33]
Phocoenamicins B and CspirotetronatesMicromonospora sp.Gram positiveMarine source[34]
Argolaphos A/BphosphonopeptideStreptomyces monomyciniGram positive/negativeGenome-driven[35]
Thiolactomycin and analogsthiotetronic acidsSalinispora/S. afghaniensisGram positiveGenome-driven[36]
taromycin AlipopetideSaccharomonospora sp.Gram positiveGenome-driven[37]
EnterocinpolyketideSalinispora pacificaGram positiveGenome-driven[38]
Difluostatin AangucyclineMicromonospora rosariaGram positiveGenome-driven[39]
Alchivemycin A and BheterocyclicS. endus + Tsukamurella pulmonisGram positiveCo-cultivation[40]
CiromicinspolyketideNocardiopsis sp. + Rhodococcus wratislaviensisNot determinedCo-cultivation[41]
Table 2. Strategies to elicit antibiotic production in actinomycetes.
Table 2. Strategies to elicit antibiotic production in actinomycetes.
Eliciting Production ApproachesMethods and TargetsReferences
Genetic engineering:Genome expression modulation:
Transcriptional repressors inactivation; Transcriptional activators overexpression
Optimized ribosomal binding sequences
Strong terminators
Increase BGCs copies
Alter metabolic fluxes
[63,65,66,67,68]
Culture-based approaches:Small molecule signaling:
Sub-inhibitory small molecule elicitors
Co-cultivation: cell–cell signaling
[69,70] [71]
Analytical mining:Comparative metabolomics:
LC-MS-based metabolomics
NMR-based metabolomics
Proteomining
[72,73,74,75] [75] [76,77]
Regulation of primary and secondary metabolismPathway specific regulatory elements:
Two-component systems
Sigma factors
Pathways specific regulators
[78,79,80]
Global regulatory metabolic networks
Master regulators
[81]
Comparative genomics:
Identification of Transcription factor orthologs
[82]
Regulon prediction: identification of regulatory networks
Primary metabolism transcription factors
Signaling cascades
Pleiotropic regulators
[83,84] [85,86,87]
Primary metabolism gene expansion[88]

Share and Cite

MDPI and ACS Style

Genilloud, O. Mining Actinomycetes for Novel Antibiotics in the Omics Era: Are We Ready to Exploit This New Paradigm? Antibiotics 2018, 7, 85. https://doi.org/10.3390/antibiotics7040085

AMA Style

Genilloud O. Mining Actinomycetes for Novel Antibiotics in the Omics Era: Are We Ready to Exploit This New Paradigm? Antibiotics. 2018; 7(4):85. https://doi.org/10.3390/antibiotics7040085

Chicago/Turabian Style

Genilloud, Olga. 2018. "Mining Actinomycetes for Novel Antibiotics in the Omics Era: Are We Ready to Exploit This New Paradigm?" Antibiotics 7, no. 4: 85. https://doi.org/10.3390/antibiotics7040085

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

Article Metrics

Back to TopTop