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:

A genetically encoded tool to increase cellular NADH/NAD+ ratio in living cells

Abstract

Impaired redox metabolism is a key contributor to the etiology of many diseases, including primary mitochondrial disorders, cancer, neurodegeneration and aging. However, mechanistic studies of redox imbalance remain challenging due to limited strategies that can perturb redox metabolism in various cellular or organismal backgrounds. Most studies involving impaired redox metabolism have focused on oxidative stress; consequently, less is known about the settings where there is an overabundance of NADH reducing equivalents, termed reductive stress. Here we introduce a soluble transhydrogenase from Escherichia coli (EcSTH) as a novel genetically encoded tool to promote reductive stress in living cells. When expressed in mammalian cells, EcSTH, and a mitochondrially targeted version (mitoEcSTH), robustly elevated the NADH/NAD+ ratio in a compartment-specific manner. Using this tool, we determined that metabolic and transcriptomic signatures of the NADH reductive stress are cellular background specific. Collectively, our novel genetically encoded tool represents an orthogonal strategy to promote reductive stress.

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: Screening of bacterial STHs for their ability to elevate NADH levels in mammalian cells.
Fig. 2: Determination of EcSTH kinetic parameters and visualization of EcSTH activity in live cells using genetically encoded sensors.
Fig. 3: Metabolic features of the NADH reductive stress in HeLa cells.
Fig. 4: RNA-seq of HeLa cells expressing LbNOX, EcSTH and mitoEcSTH.
Fig. 5: The anti-proliferative effect of EcSTH and mitoEcSTH expression is cellular background specific.
Fig. 6: In vivo elevation of the hepatic cytosolic NADH/NAD+ ratio upregulates GDF15 expression.

Similar content being viewed by others

Data availability

RNA-seq data presented in this work are available at the Gene Expression Omnibus database under accession number GEO GSE232115. The H. sapiens GRCh38 reference genome was used for mapping RNA-seq reads. Source data are provided with this paper.

References

  1. Veech, R. L. et al. The ‘great’ controlling nucleotide coenzymes. IUBMB Life 71, 565–579 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Walsh, C. T., Tu, B. P. & Tang, Y. Eight kinetically stable but thermodynamically activated molecules that power cell metabolism. Chem. Rev. 118, 1460–1494 (2018).

    Article  CAS  PubMed  Google Scholar 

  3. Xiao, W., Wang, R. S., Handy, D. E. & Loscalzo, J. NAD(H) and NADP(H) redox couples and cellular energy metabolism. Antioxid. Redox Signal 28, 251–272 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Titov, D. V. et al. Complementation of mitochondrial electron transport chain by manipulation of the NAD+/NADH ratio. Science 352, 231–235 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Cracan, V., Titov, D. V., Shen, H., Grabarek, Z. & Mootha, V. K. A genetically encoded tool for manipulation of NADP+/NADPH in living cells. Nat. Chem. Biol. 13, 1088–1095 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Xiao, W. & Loscalzo, J. Metabolic responses to reductive stress. Antioxid. Redox Signal 32, 1330–1347 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. McReynolds, M. R. et al. NAD+ flux is maintained in aged mice despite lower tissue concentrations. Cell Syst. 12, 1160–1172 e1164 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Svensson, S. et al. Activities of human alcohol dehydrogenases in the metabolic pathways of ethanol and serotonin. Eur. J. Biochem. 262, 324–329 (1999).

    Article  CAS  PubMed  Google Scholar 

  9. Goodman, R. P., Calvo, S. E. & Mootha, V. K. Spatiotemporal compartmentalization of hepatic NADH and NADPH metabolism. J. Biol. Chem. 293, 7508–7516 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Gall, W. E. et al. alpha-hydroxybutyrate is an early biomarker of insulin resistance and glucose intolerance in a nondiabetic population. PLoS ONE 5, e10883 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  11. Thompson Legault, J. et al. A metabolic signature of mitochondrial dysfunction revealed through a monogenic form of Leigh syndrome. Cell Rep. 13, 981–989 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Cobb, J. et al. α-Hydroxybutyric acid is a selective metabolite biomarker of impaired glucose tolerance. Diabetes Care 39, 988–995 (2016).

    Article  CAS  PubMed  Google Scholar 

  13. Edenberg, H. J. The genetics of alcohol metabolism: role of alcohol dehydrogenase and aldehyde dehydrogenase variants. Alcohol Res. Health 30, 5–13 (2007).

    PubMed  PubMed Central  Google Scholar 

  14. Partipilo, M. et al. Minimal pathway for the regeneration of redox cofactors. JACS Au 1, 2280–2293 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Sanchez, A. M., Rews, J., Hussein, I., Bennett, G. N. & San, K. Y. Effect of overexpression of a soluble pyridine nucleotide transhydrogenase (UdhA) on the production of poly(3-hydroxybutyrate) in Escherichia coli. Biotechnol. Prog. 22, 420–425 (2006).

    Article  CAS  PubMed  Google Scholar 

  16. Hou, J., Lages, N. F., Oldiges, M. & Vemuri, G. N. Metabolic impact of redox cofactor perturbations in Saccharomyces cerevisiae. Metab. Eng. 11, 253–261 (2009).

    Article  CAS  PubMed  Google Scholar 

  17. Yu, T. et al. Metabolic reconfiguration enables synthetic reductive metabolism in yeast. Nat. Metab. 4, 1551–1559 1 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Sharma, R. et al. Circulating markers of NADH-reductive stress correlate with mitochondrial disease severity. J. Clin. Invest. 131, e136055 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Balsa, E. et al. Defective NADPH production in mitochondrial disease complex I causes inflammation and cell death. Nat. Commun. 11, 2714 1 (2020).

    Article  Google Scholar 

  20. Hosios, A. M. & Vander Heiden, M. G. The redox requirements of proliferating mammalian cells. J. Biol. Chem. 293, 7490–7498 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Luengo, A. et al. Increased demand for NAD+ relative to ATP drives aerobic glycolysis. Mol. Cell. 81, 691–707 e696 (2021).

    Article  CAS  PubMed  Google Scholar 

  22. King, M. P. & Attardi, G. Human cells lacking mtDNA: repopulation with exogenous mitochondria by complementation. Science 246, 500–503 (1989).

    Article  CAS  PubMed  Google Scholar 

  23. Zhao, Y. et al. SoNar, a highly responsive NAD+/NADH sensor, allows high-throughput metabolic screening of anti-tumor agents. Cell Metab. 21, 777–789 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Tao, R. et al. Genetically encoded fluorescent sensors reveal dynamic regulation of NADPH metabolism. Nat. Methods 14, 720–728 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Niu, X. et al. Cytosolic and mitochondrial NADPH fluxes are independently regulated. Nat. Chem. Biol. https://doi.org/10.1038/s41589-023-01283-9 (2023).

    Article  PubMed  Google Scholar 

  26. Schafer, F. Q. & Buettner, G. R. Redox environment of the cell as viewed through the redox state of the glutathione disulfide/glutathione couple. Free Radic. Biol. Med 30, 1191–1212 (2001).

    Article  CAS  PubMed  Google Scholar 

  27. Liu, S. et al. Glycerol-3-phosphate biosynthesis regenerates cytosolic NAD+ to alleviate mitochondrial disease. Cell Metab. 33, 1974–1987 e1979 (2021).

    Article  CAS  PubMed  Google Scholar 

  28. Huang, L. et al. A family of metal-dependent phosphatases implicated in metabolite damage-control. Nat. Chem. Biol. 12, 621–627 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Zhang, Q. & Bartels, D. Octulose: a forgotten metabolite? J. Exp. Bot. 68, 5689–5694 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Martinez-Reyes, I. & Chandel, N. S. Mitochondrial TCA cycle metabolites control physiology and disease. Nat. Commun. 11, 102 3 (2020).

    Article  Google Scholar 

  31. Goodman, R. P. et al. Hepatic NADH reductive stress underlies common variation in metabolic traits. Nature 583, 122–126 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Williamson, D. H., Lund, P. & Krebs, H. A. The redox state of free nicotinamide-adenine dinucleotide in the cytoplasm and mitochondria of rat liver. Biochem. J. 103, 514–527 (1967).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Kim, W. et al. Polyunsaturated fatty acid desaturation is a mechanism for glycolytic NAD+ recycling. Cell Metab. 29, 856–870 e857 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Sullivan, L. B. et al. Supporting aspartate biosynthesis is an essential function of respiration in proliferating cells. Cell 162, 552–563 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Yang, L. et al. Serine catabolism feeds NADH when respiration is impaired. Cell Metab. 31, 809–821 e806 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Unterlass, J. E. et al. Structural insights into the enzymatic activity and potential substrate promiscuity of human 3-phosphoglycerate dehydrogenase (PHGDH). Oncotarget 8, 104478–104491 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Reid, M. A. et al. Serine synthesis through PHGDH coordinates nucleotide levels by maintaining central carbon metabolism. Nat. Commun. 9, 5442 6 (2018).

    Article  Google Scholar 

  38. Yang, L. et al. Ketogenic diet and chemotherapy combine to disrupt pancreatic cancer metabolism and growth. Med 3, 119–136 (2022).

    Article  CAS  PubMed  Google Scholar 

  39. Jackson, B. et al. Update on the aldehyde dehydrogenase gene (ALDH) superfamily. Hum. Genomics 5, 283–303 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Vasiliou, V., Pappa, A. & Estey, T. Role of human aldehyde dehydrogenases in endobiotic and xenobiotic metabolism. Drug Metab. Rev. 36, 279–299 (2004).

    Article  CAS  PubMed  Google Scholar 

  41. Davis, I., Yang, Y., Wherritt, D. & Liu, A. Reassignment of the human aldehyde dehydrogenase ALDH8A1 (ALDH12) to the kynurenine pathway in tryptophan catabolism. J. Biol. Chem. 293, 9594–9603 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Wang, X. et al. MicroRNA hsa-miR-1301-3p regulates human ADH6, ALDH5A1 and ALDH8A1 in the ethanol–acetaldehyde–acetate metabolic pathway. Mol. Pharmacol. 98, 120–129 (2020).

    Article  PubMed  Google Scholar 

  43. Yatsuga, S. et al. Growth differentiation factor 15 as a useful biomarker for mitochondrial disorders. Ann. Neurol. 78, 814–823 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Fujita, Y. et al. GDF15 is a novel biomarker to evaluate efficacy of pyruvate therapy for mitochondrial diseases. Mitochondrion 20, 34–42 (2015).

    Article  CAS  PubMed  Google Scholar 

  45. Mick, E. et al. Distinct mitochondrial defects trigger the integrated stress response depending on the metabolic state of the cell. eLife 9, e49178 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Kang, S. G. et al. Differential roles of GDF15 and FGF21 in systemic metabolic adaptation to the mitochondrial integrated stress response. iScience 24, 102181 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Argyrou, A. & Blanchard, J. S. Flavoprotein disulfide reductases: advances in chemistry and function. Prog. Nucleic Acid Res. Mol. Biol. 78, 89–142 (2004).

    Article  CAS  PubMed  Google Scholar 

  48. Ichinose, H., Kamiya, N. & Goto, M. Enzymatic redox cofactor regeneration in organic media: functionalization and application of glycerol dehydrogenase and soluble transhydrogenase in reverse micelles. Biotechnol. Prog. 21, 1192–1197 (2005).

    Article  CAS  PubMed  Google Scholar 

  49. Mouri, T., Shimizu, T., Kamiya, N., Goto, M. & Ichinose, H. Design of a cytochrome P450BM3 reaction system linked by two-step cofactor regeneration catalyzed by a soluble transhydrogenase and glycerol dehydrogenase. Biotechnol. Prog. 25, 1372–1378 (2009).

    Article  CAS  PubMed  Google Scholar 

  50. Nissen, T. L., Erlund, M., Nielsen, J., Villadsen, J. & Kielland-Brandt, M. C. Expression of a cytoplasmic transhydrogenase in Saccharomyces cerevisiae results in formation of 2-oxoglutarate due to depletion of the NADPH pool. Yeast 18, 19–32 (2001).

    Article  CAS  PubMed  Google Scholar 

  51. Zhao, H. et al. The soluble transhydrogenase UdhA affecting the glutamate-dependent acid resistance system of Escherichia coli under acetate stress. Biol. Open 7, bio031856 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  52. Sauer, U., Canonaco, F., Heri, S., Perrenoud, A. & Fischer, E. The soluble and membrane-bound transhydrogenases UdhA and PntAB have divergent functions in NADPH metabolism of Escherichia coli. J. Biol. Chem. 279, 6613–6619 (2004).

    Article  CAS  PubMed  Google Scholar 

  53. Wu, Z., et al. Electron transport chain inhibition increases cellular dependence on purine transport and salvage. Preprint at bioRxiv https://doi.org/10.1101/2023.05.11.540429 (2023).

  54. Yang, R. et al. Identification of purine biosynthesis as an NADH-sensing pathway to mediate energy stress. Nat. Commun. 13, 7031 0 (2022).

    Article  Google Scholar 

  55. Alonso-Lavin, A. J., Bajic, D. & Poyatos, J. F. Tolerance to NADH/NAD+ imbalance anticipates aging and anti-aging interventions. iScience 24, 102697 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Boulanger, E. F., Sabag-Daigle, A., Thirugnanasambantham, P., Gopalan, V. & Ahmer, B. M. M. Sugar-phosphate toxicities. Microbiol Mol. Biol. Rev. 85, e0012321 (2021).

    Article  PubMed  Google Scholar 

  57. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    Article  CAS  PubMed  Google Scholar 

  58. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank members of the Thompson laboratory at MSKCC, H. Sies, M. Clasquin, A. Caudy and A. Rosebrock for discussion and feedback. We thank A. T. Truong (Agilent Technologies), B. Webb (Center for Metabolomics and Mass Spectrometry, Scripps Research) and D. Scott (Sanford Burnham Prebys Discovery Cancer Metabolism Core) for technical support. This work was supported by grants from the National Institutes of Health (R00GM121856, R03AG067301, R35GM142495 and R35GM142495-02S1 to V.C. and R01DK134675 to R.P.G.). The Seahorse XFe96 analyzer in the Saez laboratory (Scripps Research) was supported by 1S10OD16357. This is manuscript number 1072 from The Scintillon Institute.

Author information

Authors and Affiliations

Authors

Contributions

V.C. conceived the study. X.P., M.L.H. and E.N.A. performed all experiments with assistance from A.L.Z. and C.Y. C.Y. performed enzyme kinetics experiments. S.V. and J.R.C. performed LC–MS experiments. X.P. performed imaging experiments and analyzed RNA-seq data. N.S. and R.P.G. performed experiments in mice. V.C., X.P. and M.L.H. wrote the manuscript with input from all the authors. All authors contributed to editing the manuscript and approved the final version.

Corresponding author

Correspondence to Valentin Cracan.

Ethics declarations

Competing interests

V.C. is listed as an inventor on a patent application on the therapeutic uses of LbNOX and TPNOX (US patent application US20190017034A1). The remaining authors declare no competing interests.

Peer review

Peer review information

Nature Chemical Biology thanks Melanie McReynolds, Yi Yang and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 SoNar imaging of HeLa cells expressing EcSTH and mitoEcSTH.

(a) Left panel is a schematic representation of live cell imaging using genetically encoded biosensor SoNar. The fluorescence intensity of the SoNar channel is represented in pseudocolor. The ratio of fluorescent intensities from excitation at 400 and 488 nm reflects cellular NADH/NAD+ ratio. High F400/488 ratio corresponds to high cellular NADH/NAD+ ratio and vice versa. To the right are widefield images of HeLa cells with lentivirus mediated LUC, LbNOX, EcSTH, mitoEcSTH expression under Dox control transiently expressing SoNar, in basal medium (DMEM without pyruvate, fluorescent vitamins and phenol red with 5 mM glucose, 25 mM HEPES, pH 7.4 and 1% dialyzed FBS), or after addition of 1 μM Ant A, or when cells were switched to the basal medium without glucose but with 10 mM pyruvate. Scale bars: 50 µm. (b-c) Quantification of the time course measurements of the fluorescence ratio (F400/488) for HeLa cells with lentivirus mediated LUC, LbNOX, EcSTH, mitoEcSTH expression under Dox control transiently expressing SoNar for conditions shown in (a). Values are mean ± s.d.; n = 8, 8, 10, 10 in (b), n = 7, 6, 6, 7 in (c) biologically independent samples.

Source data

Extended Data Fig. 2 iNAP1 imaging of HeLa cells expressing EcSTH and mitoEcSTH.

(a) Left panel is a schematic representation of principle of live cell imaging using genetically encoded biosensor iNAP1. The fluorescence intensity of the iNAP1 channel is represented in pseudocolor. The ratio of fluorescent intensities from excitations at 400 and 488 reflects cellular levels of NADPH. High F400/488 ratio corresponds to high NADPH and vice versa. To the right are widefield images of HeLa cells with lentivirus mediated LUC, LbNOX, EcSTH, mitoEcSTH expression under Dox control transiently expressing iNAP1, in basal medium (DMEM without pyruvate, fluorescent vitamins and phenol red with 5 mM glucose, 25 mM HEPES, pH 7.4 and 1% dialyzed FBS), or when cells were switched to the basal medium without glucose but with 10 mM pyruvate. Scale bars: 50 µm. Quantification of the time course measurements of the fluorescence ratio (F400/488) for HeLa cells with lentivirus mediated LUC, LbNOX, EcSTH, mitoEcSTH expression under Dox control transiently expressing iNAP1 in the basal medium followed by replacement with basal medium without glucose but with 10 mM pyruvate (b), after addition of G6PDi (c), after addition of 6-AN (d), or after addition of diamide (e). Values are mean ± s.d.; n = 9, 8, 8, 8 in (b), n = 8, 8, 8, 8 in (c), n = 8, 8, 8, 8 in (d), n = 8, 9, 7, 8 in (e) biologically independent samples.

Source data

Extended Data Fig. 3 Impact of EcSTH and mitoEcSTH expression on oxygen consumption and acidification rates.

Oxygen consumption rate (OCR) (a) and extracellular acidification rate (ECAR) (b) of HeLa cells expressing EcSTH and mitoEcSTH before and after addition of 1 µM antimycin A (ANT) measured in pyruvate free HEPES/DMEM+dFBS media. Values are mean ± s.d.; n = 12, 12, 12, 10 for OCR traces in (a), n = 4 for bar graphs depicting OCR quantification in (a), n = 12, 12, 12, 10 for ECAR tracers in (b), n = 4 for bar graphs depicting ECAR quantification in (b) biologically independent samples. The statistical significance indicated for (a-b) represents a One-Way ANOVA followed by uncorrected Fisher′s least significant difference test.

Source data

Extended Data Fig. 4 Activity of EcSTH and mitoEcSTH in HeLa cells with inhibited ETC or under hypoxia conditions.

The total NADH/NAD+ (a) and NADPH/NADP+ (b) ratios measured in HeLa cells expressing EcSTH and mitoEcSTH three hours after changing to fresh pyruvate-free DMEM+dFBS with 1 μM antimycin A (ANT), 1 μM piericidin A (Pier), or 1 mM pyruvate (PYR). The total NADH/NAD+ (c) and NADPH/NADP+ (d) ratios measured in HeLa cells expressing EcSTH and mitoEcSTH incubated for 24 hours at 5%CO2/1%O2 hypoxia. Values are mean ± s.d.; n = 8, 4, 4, 3, 4, 4, 4, 3, 8, 4, 4, 3 in (a), n = 4 in (b), n = 8 in (c, d) biologically independent samples. The statistical significance indicated for (a-d) represents a One-Way ANOVA followed by uncorrected Fisher′s least significant difference test. NS, no significant difference.

Source data

Extended Data Fig. 5 Metabolic features of LbNOX, EcSTH and mitoEcSTH expression in HeLa cells.

Targeted metabolomics in HeLa cells expressing LbNOX (a), EcSTH (b) and mitoEcSTH (c). LUC expressing HeLa cells were used as a control in (a-c). The statistical significance indicated for (a-c) represents p value cutoff = 0.05, fold change cutoff = 1.5 and Welch t-test (FDR corrected).

Extended Data Fig. 6 Representative redox couples in HeLa cells expressing EcSTH and mitoEcSTH.

Intracellular lactate/pyruvate (a), 2-hydroxybutyrate/2-ketobutyrate (2HB/2KB) (b), 3-hydroxybutyrate/acetoacetate (3HB/AcAc) (c), glycerol 3-phosphate/dihydroxyacetone phosphate (Gro3P/DHAP) (d), C16:1/C16:0 (e) and AMP/ATP (f) ratios measured in HeLa cells expressing EcSTH and mitoEcSTH. In (a-f) values are mean ± s.d.; n = 3 biologically independent samples. Statistically significant differences were calculated by using a Welch ANOVA followed by unpaired t-test. NS, no significant difference.

Source data

Extended Data Fig. 7 Rescue of the anti-proliferative effect of EcSTH expression with extracellular electron acceptors.

Proliferation of HeLa cells with EcSTH and mitoEcSTH expression in DMEM+dFBS when supplemented with exogenous electron acceptors (10 mM pyruvate (PYR), 10 mM oxaloacetate (OAA) or 10 mM α-ketobutyrate (2KB)). Values are mean ± s.d.; n = 3 biologically independent samples.

Source data

Extended Data Fig. 8 Analysis of GO terms linked to NAD(P) redox metabolism.

(a-b) Venn diagrams displaying the similarities and differences of genes from indicated gene sets. Yellow circles include all genes in the GO term ‘oxidation-reduction’ pathways (see Fig. 4e). The salmon circle contains all genes in the GO term ‘cholesterol biosynthetic’ pathways (see Fig. 4e). The light brown circle indicates the NAD(P)ome (genes in humans that encode NAD+ or NADP+ dependent enzymes as defined in27). The up-regulated genes are marked in red and down-regulated genes are marked in blue.

Extended Data Fig. 9 Impact of EcSTH and mitoEcSTH expression on NAD(P)H/NAD(P)+ ratios and proliferation in different cell lines.

Western blot analysis of EcSTH and mitoEcSTH expression in A549 lung adenocarcinoma cells (a), MIA PaCa-2 epithelial tumor tissue of the pancreas (e) and PANC-1 pancreatic duct epithelioid carcinoma (i). (a, e, i) Representative blots are shown. The total NADH/NAD+ ratios in A549 (b), MIA PaCa-2 (f) and PANC-1 (j) cells expressing EcSTH and mitoEcSTH. The total NADPH/NADP+ ratios in A549 (c), MIA PaCa-2 (g) and PANC-1 (k) cells expressing EcSTH and mitoEcSTH. Proliferation of A549 (d), MIA PaCa-2 (h) and PANC-1 (l) cells expressing EcSTH and mitoEcSTH in pyruvate-free DMEM+dFBS. LUC and LbNOX expressing HeLa cells were used as controls in (a-l). Values are mean ± s.d.; n = 8 in (b-c), n = 6 in (f, g, j, k) biologically independent samples. The statistical significance indicated for (b, c, f, g, j, k) represents a Welch ANOVA followed by unpaired t-test. NS, no significant difference. For growth curves (d, h, l) error bars represent mean ± s.d.; n = 3 biologically independent samples.

Source data

Extended Data Fig. 10 Untargeted metabolomics of C2C12 cells expressing EcSTH and mitoEcSTH.

Untargeted metabolomics in C2C12 cells expressing LbNOX (a), EcSTH (b) and mitoEcSTH (c). LUC expressing C2C12 cells were used as a control in (ac). The statistical significance indicated for (a-c) represents p value cutoff = 0.05, fold change cutoff = 1.5 and Welch t-test (FDR corrected).

Supplementary information

Supplementary Information

Supplementary Figs. 1–17.

Reporting Summary

Supplementary Tables 1–12

Relevant to Methods (Tables 1–5), raw metabolomics peak areas (Tables 6–7) and RNA-seq (Tables 8–12).

Supplementary Data 1

Source data for Supplementary Fig. 1.

Supplementary Data 2

Source data for Supplementary Fig. 2.

Supplementary Data 3

Source data for Supplementary Fig. 3.

Supplementary Data 4

Source data for Supplementary Fig. 5.

Supplementary Data 5

Source data for Supplementary Fig. 6.

Supplementary Data 6

Source data for Supplementary Fig. 7.

Supplementary Data 7

Source data for Supplementary Fig. 9.

Supplementary Data 8

Source data for Supplementary Fig. 10.

Supplementary Data 9

Source data for Supplementary Fig. 11.

Supplementary Data 10

Source data for Supplementary Fig. 12.

Supplementary Data 11

Source data for Supplementary Fig. 13.

Supplementary Data 12

Source data for Supplementary Fig. 14.

Supplementary Data 13

Source data for Supplementary Fig. 17.

Source data

Source Data Fig. 1

Source data.

Source Data Fig. 2

Source data.

Source Data Fig. 5

Source data.

Source Data Fig. 6

Source data.

Source Data Fig. 1

Uncut western blot for Fig. 1b part 1.

Source Data Fig. 1

Uncut western blot for Fig. 1b part 2.

Source Data Fig. 2

Uncut SDS–PAGE.

Source Data Fig. 5

Uncut western blot for Fig. 5a.

Source Data Fig. 5

Uncut western blot for Fig. 5e.

Source Data Fig. 5

Uncut western blot for Fig. 5i.

Source Data Extended Data Fig. 1

Source data.

Source Data Extended Data Fig. 2

Source data.

Source Data Extended Data Fig. 3

Source data.

Source Data Extended Data Fig. 4

Source data.

Source Data Extended Data Fig. 6

Source data.

Source Data Extended Data Fig. 7

Source data.

Source Data Extended Data Fig. 9

Source data.

Source Data Extended Data Fig. 9

Uncut western blot for Extended Data Fig. 9a.

Source Data Extended Data Fig. 9

Uncut western blot for Extended Data Fig. 9e.

Source Data Extended Data Fig. 9

Uncut western blot for Extended Data Fig. 9i.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pan, X., Heacock, M.L., Abdulaziz, E.N. et al. A genetically encoded tool to increase cellular NADH/NAD+ ratio in living cells. Nat Chem Biol (2023). https://doi.org/10.1038/s41589-023-01460-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1038/s41589-023-01460-w

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