Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Harnessing the intracellular triacylglycerols for titer improvement of polyketides in Streptomyces

A Publisher Correction to this article was published on 18 December 2019

This article has been updated

Abstract

Pharmaceutically important polyketides such as avermectin are mainly produced as secondary metabolites during the stationary phase of growth of Streptomyces species in fermenters. The source of intracellular metabolites that are funneled into polyketide biosynthesis has proven elusive. We applied multi-omics to reveal that intracellular triacylglycerols (TAGs), which accumulates in primary metabolism, are degraded during stationary phase. This process could channel carbon flux from both intracellular TAGs and extracellular substrates into polyketide biosynthesis. We devised a strategy named ‘dynamic degradation of TAG’ (ddTAG) to mobilize the TAG pool and increase polyketide biosynthesis. Using ddTAG we increased the titers of actinorhodin, jadomycin B, oxytetracycline and avermectin B1a in Streptomyces coelicolor, Streptomyces venezuelae, Streptomyces rimosus and Streptomyces avermitilis. Application of ddTAG increased the titer of avermectin B1a by 50% to 9.31 g l−1 in a 180-m3 industrial-scale fermentation, which is the highest titer ever reported. Our strategy could improve polyketide titers for pharmaceutical production.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Identification of intracellular metabolites with relationship to polyketides production.
Fig. 2: Cellular TAG pool contributes to Act yield during stationary phase.
Fig. 3: High-yielding mechanism of mobilizing cellular TAG pool.
Fig. 4: Identification of SCO6196 responsible for cellular TAG mobilization.
Fig. 5: The ddTAG strategy for polyketide titer improvement.

Similar content being viewed by others

Data availability

The genome sequence of S. coelicolor strain is available at NCBI (accession number NC_003888). Time-series transcriptome data of S. coelicolor stains are available at NCBI GEO database (accession numbers GSE2983, GSE18489, GSE30569, GSE30570 and GSE53562). Other data supporting the findings of this study are included in the published article and supplementary information. Requests for any additional information can be made to the corresponding authors.

Code availability

The custom Python scripts used for flux sampling analysis are provided in Supplementary Note 4.

Change history

References

  1. Chater, K. F. Streptomyces inside-out: a new perspective on the bacteria that provide us with antibiotics. Phil. Trans. R. Soc. Lond. B. 361, 761–768 (2006).

    Article  CAS  Google Scholar 

  2. Zhang, L. & Demain, A. L. Natural Products: Drug Discovery and Therapeutic Medicine 1st edn (Humana Press, 2005).

  3. Ashforth, E. J. et al. Bioprospecting for antituberculosis leads from microbial metabolites. Nat. Prod. Rep. 27, 1709–1719 (2010).

    Article  CAS  PubMed  Google Scholar 

  4. 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. 32, 255–268 (2014).

    Article  CAS  PubMed  Google Scholar 

  5. Liu, G., Chater, K. F., Chandra, G., Niu, G. & Tan, H. Molecular regulation of antibiotic biosynthesis in Streptomyces. Microbiol. Mol. Biol. Rev. 77, 112–143 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Kang, Q., Bai, L. & Deng, Z. Toward steadfast growth of antibiotic research in China: from natural products to engineered biosynthesis. Biotechnol. Adv. 30, 1228–1241 (2012).

    Article  CAS  PubMed  Google Scholar 

  7. Wang, Q. et al. Abyssomicins from the South China Sea deep-sea sediment Verrucosispora sp.: natural thioether Michael addition adducts as antitubercular prodrugs. Angew. Chem. Int. Ed. Engl. 52, 1231–1234 (2013).

    Article  CAS  PubMed  Google Scholar 

  8. Nieselt, K. et al. The dynamic architecture of the metabolic switch in Streptomyces coelicolor. BMC Genomics 11, 10 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  9. Bibb, M. J. Regulation of secondary metabolism in streptomycetes. Curr. Opin. Microbiol. 8, 208–215 (2005).

    Article  CAS  PubMed  Google Scholar 

  10. Alam, M. T. et al. Metabolic modeling and analysis of the metabolic switch in Streptomyces coelicolor. BMC Genomics 11, 202 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  11. Gao, Q., Tan, G. Y., Xia, X. & Zhang, L. Learn from microbial intelligence for avermectins overproduction. Curr. Opin. Biotechnol. 48, 251–257 (2017).

    Article  CAS  PubMed  Google Scholar 

  12. Rigali, S. et al. Feast or famine: the global regulator DasR links nutrient stress to antibiotic production by Streptomyces. EMBO Rep. 9, 670–675 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Masuma, R., Tanaka, Y., Tanaka, H. & Omura, S. Production of nanaomycin and other antibiotics by phosphate-depressed fermentation using phosphate-trapping agents. J. Antibiot. 39, 1557–1564 (1986).

    Article  CAS  Google Scholar 

  14. Mendes, M. V. et al. The two-component phoR-phoP system of Streptomyces natalensis: Inactivation or deletion of phoP reduces the negative phosphate regulation of pimaricin biosynthesis. Metab. Eng. 9, 217–227 (2007).

    Article  CAS  PubMed  Google Scholar 

  15. Wentzel, A., Sletta, H., Stream, C., Ellingsen, T. E. & Bruheim, P. Intracellular metabolite pool changes in response to nutrient depletion induced metabolic switching in Streptomyces coelicolor. Metabolites 2, 178–194 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Jankevics, A. et al. Metabolomic analysis of a synthetic metabolic switch in Streptomyces coelicolor A3(2). Proteomics 11, 4622–4631 (2011).

    Article  CAS  PubMed  Google Scholar 

  17. Huang, J., Lih, C. J., Pan, K. H. & Cohen, S. N. Global analysis of growth phase responsive gene expression and regulation of antibiotic biosynthetic pathways in Streptomyces coelicolor using DNA microarrays. Genes Dev. 15, 3183–3192 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. D’Huys, P. J. et al. Genome-scale metabolic flux analysis of Streptomyces lividans growing on a complex medium. J. Biotechnol. 161, 1–13 (2012).

    Article  PubMed  CAS  Google Scholar 

  19. Kieser, T., Bibb, M. J., Buttner, M. J., Chater, K. F. & Hopwood, D. A. Practical Streptomyces Genetics (The John Innes Foundation, 2000).

  20. Olukoshi, E. R. & Packter, N. M. Importance of stored triacylglycerols in Streptomyces: possible carbon source for antibiotics. Microbiology 140, 931–943 (1994).

    Article  CAS  PubMed  Google Scholar 

  21. Alvarez, H. M. & Steinbuchel, A. Triacylglycerols in prokaryotic microorganisms. Appl. Microbiol. Biotechnol. 60, 367–376 (2002).

    Article  CAS  PubMed  Google Scholar 

  22. Li, S., Wang, W., Li, X., Fan, K. & Yang, K. Genome-wide identification and characterization of reference genes with different transcript abundances for Streptomyces coelicolor. Sci. Rep. 5, 15840 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Waldvogel, E. et al. The PII protein GlnK is a pleiotropic regulator for morphological differentiation and secondary metabolism in Streptomyces coelicolor. Appl. Microbiol. Biotechnol. 92, 1219–1236 (2011).

    Article  CAS  PubMed  Google Scholar 

  24. Zhuo, Y. et al. Reverse biological engineering of hrdB to enhance the production of avermectins in an industrial strain of Streptomyces avermitilis. Proc. Natl Acad. Sci. USA 107, 11250–11254 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Zhang, Y. et al. Characterization of a pathway-specific activator of milbemycin biosynthesis and improved milbemycin production by its overexpression in Streptomyces bingchenggensis. Microb. Cell Fact. 15, 152 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Yin, S. et al. Identification of a cluster-situated activator of oxytetracycline biosynthesis and manipulation of its expression for improved oxytetracycline production in Streptomyces rimosus. Microb. Cell Fact. 14, 46 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  27. Vemuri, G. N., Eiteman, M. A., McEwen, J. E., Olsson, L. & Nielsen, J. Increasing NADH oxidation reduces overflow metabolism in Saccharomyces cerevisiae. Proc. Natl Acad. Sci. USA 104, 2402–2407 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Kumelj, T., Sulheim, S., Wentzel, A. & Almaas, E. Predicting strain engineering strategies using iKS1317: a genome-scale metabolic model of Streptomyces coelicolor. Biotechnol. J. 14, e1800180 (2019).

    Article  PubMed  CAS  Google Scholar 

  29. Horbal, L., Fedorenko, V. & Luzhetskyy, A. Novel and tightly regulated resorcinol and cumate-inducible expression systems for Streptomyces and other actinobacteria. Appl. Microbiol. Biotechnol. 98, 8641–8655 (2014).

    Article  CAS  PubMed  Google Scholar 

  30. Wang, W. et al. Development of a synthetic oxytetracycline-inducible expression system for streptomycetes using de novo characterized genetic parts. ACS Synth. Biol. 5, 765–773 (2016).

    Article  PubMed  CAS  Google Scholar 

  31. Chae, T. U., Choi, S. Y., Kim, J. W., Ko, Y. S. & Lee, S. Y. Recent advances in systems metabolic engineering tools and strategies. Curr. Opin. Biotechnol. 47, 67–82 (2017).

    Article  CAS  PubMed  Google Scholar 

  32. Bai, C. et al. Exploiting a precise design of universal synthetic modular regulatory elements to unlock the microbial natural products in Streptomyces. Proc. Natl Acad. Sci. USA 112, 12181–12186 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Wang, W. et al. Angucyclines as signals modulate the behaviors of Streptomyces coelicolor. Proc. Natl Acad. Sci. USA 111, 5688–5693 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Winder, C. L. et al. Global metabolic profiling of Escherichia coli cultures: an evaluation of methods for quenching and extraction of intracellular metabolites. Anal. Chem. 80, 2939–2948 (2008).

    Article  CAS  PubMed  Google Scholar 

  35. Perera, M. A., Choi, S. Y., Wurtele, E. S. & Nikolau, B. J. Quantitative analysis of short-chain acyl-coenzymeAs in plant tissues by LC–MS-MS electrospray ionization method. J. Chromatogr. B. 877, 482–488 (2009).

    Article  CAS  Google Scholar 

  36. Rodriguez, E., Navone, L., Casati, P. & Gramajo, H. Impact of malic enzymes on antibiotic and triacylglycerol production in Streptomyces coelicolor. Appl. Environ. Microbiol. 78, 4571–4579 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Arabolaza, A., Rodriguez, E., Altabe, S., Alvarez, H. & Gramajo, H. Multiple pathways for triacylglycerol biosynthesis in Streptomyces coelicolor. Appl. Environ. Microbiol. 74, 2573–2582 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Stein, S. E. An integrated method for spectrum extraction and compound identification from gas chromatography/mass spectrometry data. J. Am. Soc. Mass Spectrom. 10, 770–781 (1999).

    Article  CAS  Google Scholar 

  39. Lommen, A. & Kools, H. MetAlign 3.0: performance enhancement by efficient use of advances in computer hardware. Metabolomics 8, 719–726 (2012).

    Article  CAS  PubMed  Google Scholar 

  40. Xia, J., Sinelnikov, I. V., Han, B. & Wishart, D. S. MetaboAnalyst 3.0–making metabolomics more meaningful. Nucleic Acids Res. 43, W251–W257 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Yuan, J., Bennett, B. D. & Rabinowitz, J. D. Kinetic flux profiling for quantitation of cellular metabolic fluxes. Nat. Protoc. 3, 1328–1340 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Camacho, C. et al. BLAST+: architecture and applications. BMC Bioinforma. 10, 421 (2009).

    Article  CAS  Google Scholar 

  43. Rice, P., Longden, I. & Bleasby, A. EMBOSS: the European molecular biology open software suite. Trends Genet. 16, 276–277 (2000).

    Article  CAS  PubMed  Google Scholar 

  44. Xia, J., Xu, Z., Xu, H., Feng, X. & Bo, F. The regulatory effect of citric acid on the co-production of poly(epsilon-lysine) and poly(L-diaminopropionic acid) in Streptomyces albulus PD-1. Bioprocess Biosyst. Eng. 37, 2095–2103 (2014).

    Article  CAS  PubMed  Google Scholar 

  45. Weitzman, P. D. J. Regulation of α-ketoglutarate dehydrogenase activity in Acinetobacter. FEBS Lett. 22, 323–326 (1972).

    Article  CAS  PubMed  Google Scholar 

  46. Kasuya, F. et al. Determination of acyl-CoA esters and acyl-CoA synthetase activity in mouse brain areas by liquid chromatography-electrospray ionization-tandem mass spectrometry. J. Chromatogr. B. 929, 45–50 (2013).

    Article  CAS  Google Scholar 

  47. Coze, F., Gilard, F., Tcherkez, G., Virolle, M. J. & Guyonvarch, A. Carbon-flux distribution within Streptomyces coelicolor metabolism: a comparison between the actinorhodin-producing strain M145 and its non-producing derivative M1146. PLoS ONE 8, e84151 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Ebrahim, A., Lerman, J. A., Palsson, B. O. & Hyduke, D. R. COBRApy: constraints-based reconstruction and analysis for Python. BMC Syst. Biol. 7, 74 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Schuetz, R., Kuepfer, L. & Sauer, U. Systematic evaluation of objective functions for predicting intracellular fluxes in Escherichia coli. Mol. Syst. Biol. 3, 119 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. Megchelenbrink, W., Huynen, M. & Marchiori, E. optGpSampler: an improved tool for uniformly sampling the solution-space of genome-scale metabolic networks. PLoS ONE 9, e86587 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  51. Zhao, Y., Xiang, S., Dai, X. & Yang, K. A simplified diphenylamine colorimetric method for growth quantification. Appl. Microbiol. Biotechnol. 97, 5069–5077 (2013).

    Article  CAS  PubMed  Google Scholar 

  52. Wang, W. et al. An engineered strong promoter for streptomycetes. Appl. Environ. Microbiol. 79, 4484–4492 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We dedicate this paper to the memory of our beloved mentor, friend and colleague, K. Yang. We thank H. Tan, Y. Tao, Y. Chen, M. Li, S. Chen, S. Gao and Q. Gao for helpful discussions. We acknowledge the financial support from the National Natural Science Foundation of China (grants 31430002 and 31720103901 to L.Z.; 31922002 and 31570031 to W.W.; 31772242 to S.L. and 31672092 to W.X.). This work was also supported by the Open Project Funding of the State Key Laboratory of Bioreactor Engineering, the 111 Project (B18022), the Major Basic Program of the Natural Science Foundation of Shandong Province (ZR2017ZB0206), the Shanghai Science and Technology Commission (18JC1411900) and the Shandong Taishan Scholar Program of China to L.Z. and the International Partnership Program of Chinese Academy of Sciences (153211KYSB20170014) and the Young Scientists Innovation Promotion Association of CAS (2016087) to W.W.

Author information

Authors and Affiliations

Authors

Contributions

L.Z., W.W. and S.L. conceived and supervised the project. W.W., S.L., Z.L. and W.X. designed and performed the main experiments. J.Z. performed the large-scale fermentations. G.A. performed the GC–MS analyses. G.S. and S.M.L. collected LC–MS/MS data for isotope tracing experiments. K.F., G.T., Z.Y., H.L., P. J., Y.L., X.C., X.X., X.L., H.K.D., C.Y., Y.Y., G.A. and S.Z. participated in the experiments. W.W., S.L. and L.Z. wrote the manuscript.

Corresponding authors

Correspondence to Weishan Wang, Wensheng Xiang or Lixin Zhang.

Ethics declarations

Competing interests

The authors have filed a provisional patent for this work to the China National Intellectual Property Administration (CNIPA). L.Z., W.W., S.L., Z.L., G.T., K,F. and J.Z. are inventors on the provisional patent application (CN201910411123.7, filed 17 May 2019).

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Integrated supplementary information

Supplementary Figure 1 Trends of the metabolic pathways in M145 and HY01.

Black bold lines indicate the overall trends of the metabolites involved in the five investigated pathways, and light gray lines show the trends of individual metabolites. Pink background indicates the late stationary phase. (a and g) Overall trends of metabolites in different metabolic pathways in M145 and HY01, respectively. Trend of each metabolite investigated in different pathways were as follows: (b and h) glucose-6-phosphate, dihydroxyacetone phosphate, glyceraldehyde 3-phosphate, D-glycerate 3-phosphate, and pyruvate for glycolysis (EMP) in M145 and HY01, respectively; (c and i) D-altro-heptulose 7-phosphate, D-erythrose 4-phosphate, 6-phospho-D-gluconate, D-ribose, and D-ribulose 5-phosphate for pentose phosphate pathway (PPP) in M145 and HY01, respectively; (d and j) succinic acid, isocitric acid, fumaric acid, cirtrate, DL-malic acid, and α-ketoglutarate for tricarboxylic acid cycle (TCA) in M145 and HY01, respectively; (e and k) serine, glycine, L-valine, L-threonine, L-pyroglutamic acid, L-proline, L-leucine, L-isoleucine, L-aspartic acid, L-alanine, DL-ornithine, 5-oxo-L-proline, 3-phosphoserine, N-acetyl-L-glutamic acid, and α-ketoadipic acid for amino acid metabolism (AAM) in M145 and HY01, respectively; (f and l) myristic acid, pentadecanoic acid, hexadecanoic acid, margaric acid, octadecanoic acid, MG(16:0/0:0/0:0), and MG(18:0/0:0/0:0) for lipid metabolism (LPM) in M145 and HY01, respectively. Levels of metabolites determined at 20 h were set as one. Data were the average and s.d. of five independent experimental replicates.

Supplementary Figure 2 Analysis of cellular TAGs by GC-MS.

(a) Fatty acid moieties of cellular TAG pool identified by GC-MS. Fatty acid abbreviations used were as follows: iC12:0, isolauric acid; iC13:0, isotridecanoic acid; aiC13:0, anteisotridecanoic acid; iC14:0, isomyristic acid; C14:0, myristic acid; iC15:0, isopentadecanoic acid ; aiC15:0, anteisopentadecanoic acid ; C15:0, pentadeclic acid; iC16:0, isopalmitic acid; C16:1-Δ9, 9-hexadecenoic acid; C16:0, palmitic acid; iC17:0, isoheptadecanoic acid; aiC17:0, anteisoheptadecanoic acid; C17:0, heptadecanoic acid; C18:1-Δ9, 9-octadecenoic acid; and C18:0, stearic acid. (b) Trends of fatty acid moieties of cellular TAG pool in M145. The cellular TAGs from the time-course samples were firstly hydrolyzed to release fatty acid moieties, and then the fatty acids were quantified using GC-MS. Peak areas of the mass chromatograms were used to indicate the relative levels of the corresponding fatty acid moieties. Results shown are average and s.d. of five independent experimental replicates.

Supplementary Figure 3 Change of cellular TAGs and their derived intermediate metabolite during stationary phase in M145 and HY01.

(a) Levels of TAG pools in M145 and HY01 at transition phase. Columns show the consumed amount of fatty acid moieties from TAGs at 48 h. The level of each fatty acid moiety in M145 and HY01 shows no significant difference (Student t-test, p < 0.05). (b-e) Levels of the metabolites derived from TAG degradation. (b) Schematic of the metabolic pathway related to TAG degradation. (c-e) Levels of metabolites related to TAG degradation. Metabolite levels are determined by LC-MS/MS. G3P, glycerol-3-phosphate; 3PG, 3-phosphoglycerate; DHAP, dihydroxyacetone phosphate. Data shown are the average and s.d. of three independent experiments. The levels of significance are *** p < 0.001, ** p < 0.01, * p < 0.05.

Supplementary Figure 4 Contributions of metabolites on Act production during stationary phase.

(a) Contributions of fatty acid moieties of cellular TAG pool on Act production during stationary phase. (b) Contributions of intercellular amino acids on Act production during stationary phase. VIP score is used to identify important features from the examined samples (Nat. Protoc. 6, 743-760, 2011). Here we aimed to identify important metabolites for Act production using VIP score. VIP score was obtained by comparing the data matrix of M145 and HY01 from different period of stationary phase (48-60 h, 60-72 h, and 72-96 h) using MetaboAnalyst Software.

Supplementary Figure 5 Dynamic profile of genes and metabolites related to TAG metabolism.

(a-d) Transcript profiles of the genes in fatty acid metabolism under diverse culture conditions. The selected genes are the same as those shown in Fig. 2e, f. Time-series transcriptome data were obtained from NCBI GEO database: (a) GSE2983, M145 cultivated in modified R5 medium (Genes Dev. 15, 3183-3192, 2001); (b) GSE18489, M145 cultivated in a fermentation medium (BMC Genomics 11, 10, 2010); (c) GSE30569, the glnK mutant of M145 cultivated in SSBM-E medium (Appl. Microbiol. Biotechnol. 92, 1219-1236, 2011); (d) GSE30570, M145 cultivated in SSBM-E medium (Appl. Microbiol. Biotechnol. 92, 1219-1236, 2011). (e-g) Profiles of cellular TAG pools in industrial Streptomyces avermitilis A56 (Proc. Natl. Acad. Sci. U. S. A. 107, 11250-11254, 2010; Synth. Syst. Biotechnol. 1, 7-16, 2016), Streptomyces bingchenggesis BC04 (Microb. Cell Fact. 15, 152, 2016), and Streptomyces rimosus M2R (M4018::2SFOtcR) (Microb. Cell Fact. 14, 46, 2015), respectively. The amount of cellular TAG pool was analyzed by TLC.

Supplementary Figure 6 Isotopic feeding experiments in M145 and HY01.

(a) Feeding experiments with [U-13C] oleate. 2.5 nM [U-13C] oleate was added in cultures of M145 and HY01 at 72 h. The samples were harvested after 24 h. (b) Feeding experiments with [U-13C] glucose. 8.6 nM [U-13C] glucose was added in cultures of M145 and HY01 at 72 h. The samples were harvested after 48 h labeling. Data shown are the average and s.d. of three independent experiments.

Supplementary Figure 7 Levels of intermediates.

The level is indicated by the peak area of metabolites. (a) Levels of acetyl-CoA obtained from [U-13C] oleate feeding experiments. (b) Levels of acetyl-CoA obtained from [U-13C] glucose feeding experiments. (c–f) Levels of the TCA intermediates obtained from [U-13C] oleate feeding experiments. (g–j) Levels of the TCA intermediates obtained from [U-13C] glucose feeding experiments. Data of oleate and glucose labeling experiments were collected at 96 h and 120 h, respectively (Supplementary Fig. 6). Data shown are the average and s.d. of three independent experiments.

Supplementary Figure 8 Comparison of labeling ratio of TCA intermediates in M145 and HY01.

(a) Labeling ratio obtained from [U-13C] oleate feeding experiments. (b) Labeling ratio obtained from [U-13C] glucose feeding experiments. The labeling ratio of TCA intermediates in HY01 were normalized with that of the direct input acetyl-CoA using the following equation: normalized labeling ratio = labeling_ratio × (labeling ratio of acetyl-CoAHY01 / labeling ratio of acetyl-CoAM145). Differences were analyzed by Student t-test, and p < 0.05 was considered statistically significant. The levels of significance are *** p < 0.001, ** p < 0.01, * p < 0.05. Data shown are the average and s.d. of three independent experiments.

Supplementary Figure 9 Cellular levels of reducing equivalents and ATP.

(a–c) Cellular levels of NADH, ATP and NADPH, respectively, in M145 and HY01 during stationary phase (72, and 96 h). Differences were analyzed by Student t-test, and p < 0.05 was considered statistically significant. The levels of significance are *** p < 0.001, ** p < 0.01, * p < 0.05. Data shown are the average and s.d. of three independent experiments.

Supplementary Figure 10 Carbon flux distribution of S. coelicolor by metabolic flux analysis.

(a) Simplified metabolic network of S. coelicolor. Flux for biomass biosynthesis from the G6P node (v2) including fluxes from five different precursors (glucose-6-phosphate, fructose-6-phosphate, ribose-5-phosphate, erythrose-4-phosphate, glyceraldehyde-3-phosphate). 15 fatty acid moieties from cellular TAGs are detected in S. coelicolor, here flux v19 represents the sum of specific production rate of AcCoA from degradation of these fatty acids. The biochemical reactions, stoichiometric equations and experimentally determined fluxes are shown in Supplementary Table 68. (b) Actual fluxes of M145, HY01 and M145-DT in fermentation with (+) or without (-) glucose feeding. Data for metabolic flux analysis were collected during stationary phase (72–120 h). Data shown here were the average of three independent experiments. Abbreviations: Glc, glucose; G6P, glucose-6-phosphate; 3PG, 3-phosphate-glycerate; PEP, phosphoenolpyruvate; PYR, pyruvate; AcCoA, acetyl coenzyme A; AKG, alpha-ketoglutaric acid; OAA, oxaloacetic acid; TAG, triacylglycerol; Act, actinorhodin.

Supplementary Figure 11 Sampling analysis of flux distribution in M145 and HY01.

The genome-scale metabolic model (GEM) of S. coelicolor (iKS1317) (Biotechnol. J 14, e1800180, 2019) was constrained by the experimentally measured rate (Supplementary Table 8) with the minimization of glucose uptake rate as the objective function. During the simulation, the sampling algorithm optGpSampler (PLoS One 9, e86587, 2014) was adopted. For each condition, 1000 samples were used for valid sampling distributions. Here we showed two reactions from TCA cycle as examples. (a) Flux changes of reactions catalyzed by citrate synthase (CS) in batch fermentation (without glucose feeding). (b) Flux changes of reactions catalyzed by CS in fed-batch fermentation (with glucose feeding). (c) Flux changes of reactions catalyzed by aconitase (ACONTa) in batch fermentation (without glucose feeding). (d) Flux changes of reactions catalyzed by ACONTa in fed-batch fermentation (with glucose feeding).

Supplementary Figure 12 Construction of sco6196 deletion and overexpression strains.

6196DM, sco6196 deletion strain; 6196OE, sco6196 overexpression strain. (a) Confirmation of 6196DM.The primer pair 96vF/96vR for PCR amplification is listed in Supplementary table 14. (b) Effect of SCO6196 on cell growth. Data shown here were the average of three independent experiments. (c) Effect of SCO6196 on Act production. The picture shown here was the bottom of the solid plates. (d) Act production of M145, 6196OE and 6196DM in liquid SMM medium. (e) TLC assay of the cellular TAG pools in M145 and 6196DM at transition phase (48 h).

Supplementary Figure 13 Biochemical analysis of substrate specificity of SCO6196 in vitro.

(a) SDS-PAGE of the purified recombinant protein SCO6196. (b) Relative activities of SCO6196 when adding different fatty acids. When using C18:1-Δ9 as a substrate, the activity of SCO6196 was set to one. Data shown here were the average of three independent experiments. (c) Specific daughter ions used for the quantification of the corresponding acyl-CoAs from C18:1-Δ9, iC14:0, iC15:0, aiC15:0, iC16:0, aiC17:0, respectively. These daughter ions were selected based on m/z 425.9 and 408 in common and specific ions of [M-H-80] - and [M-H-347]-.

Supplementary Figure 14 Construction and application of the ddTAG module.

(a) Map of the plasmid used to construct the ddTAG module. The ddTAG module is controlled by a cumate inducible system (Appl. Microbiol. Biotechnol. 98, 8641-8655, 2014). (b) Induction range of sco6196 transcripts quantified by RT-qPCR. (c) Comparison of the Act titers of M145, HY01 and M145-DT. (d) Comparison of the amount of consumed TAGs between M145 and M145-DT. Columns show the consumed amount of fatty acid moieties from cellular TAG pool between 48 h and 96 h. (e) Comparison of the redox status (NADH/NAD+) in M145-DT and M145. (f) Effects of cumate on oxytetracycline (Otc) titer of M4018 and M-DT. (f) Effects of cumate on B1a titer of A56 and A56-DT. Data shown in (b-g) are the average and s.d. of three independent experiments. Significant differences were analyzed by one-way ANOVA, and p < 0.05 was considered statistically significant. The levels of significance are *** p < 0.001, ** p < 0.01, * p < 0.05.

Supplementary Figure 15 Carbon flux distribution of S. avermitilis by metabolic flux analysis.

(a) Simplified metabolic network of S. avermitilis. Flux for biomass biosynthesis from the G6P node (v12) including fluxes from five different precursors (glucose-6-phosphate, fructose-6-phosphate, ribose-5-phosphate, erythrose-4-phosphate, glyceraldehyde-3-phosphate). 15 detected fatty acid moieties from cellular TAGs in S. avermitilis were detected, and flux v21 represents the sum of specific production rate of AcCoA from degradation of these 15 fatty acids. Similarly, v22 and v23 represent the sum of specific production rate of PCoA from degradation of two fatty acids (C15:0 and C17:0) and the sum of specific production rate of MBCoA from degradation of three fatty acids (aiC13:0, aiC15:0, and aiC17:0), respectively. The biochemical reactions, stoichiometric equations and experimentally determined fluxes are shown in Supplementary Table 1113. (b) Actual fluxes of A56 and A56-DT in fed-batch fermentation. The inset showed the mass balance of the AcCoA node, and fluxes were expressed in unit of the acetyl unit. Data for metabolic flux analysis were collected during stationary phase (120–196 h). Data shown here were the average of three independent experiments. Abbreviations: PCoA, propionyl coenzyme A; MCoA, malonyl coenzyme A; MMCoA, methylmalonyl coenzyme A; MBCoA, 2-methylbutyryl coenzyme A; SucCoA, succinyl coenzyme A; dTDP-Ole, deoxythymidine diphosphate-oleandrose; L-Asp, L-aspartate; L-Thr, L-threonine; Ile, L-isoleucine; B1a, avermectin B1a; other abbreviations were shown in Supplementary Fig. 10.

Supplementary information

Supplementary Materials

Supplementary Figs. 1–15, Tables 1, 3–9 and 11–14 and Notes 1–4.

Reporting Summary

Supplementary Table 2

The identified metabolites in this work.

Supplementary Table 10

Details of 888 ACSs.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, W., Li, S., Li, Z. et al. Harnessing the intracellular triacylglycerols for titer improvement of polyketides in Streptomyces. Nat Biotechnol 38, 76–83 (2020). https://doi.org/10.1038/s41587-019-0335-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41587-019-0335-4

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research