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:

Shifting redox reaction equilibria on demand using an orthogonal redox cofactor

Abstract

Nature’s two redox cofactors, nicotinamide adenine dinucleotide (NAD+) and nicotinamide adenine dinucleotide phosphate (NADP+), are held at different reduction potentials, driving catabolism and anabolism in opposite directions. In biomanufacturing, there is a need to flexibly control redox reaction direction decoupled from catabolism and anabolism. We established nicotinamide mononucleotide (NMN+) as a noncanonical cofactor orthogonal to NAD(P)+. Here we present the development of Nox Ortho, a reduced NMN+ (NMNH)-specific oxidase, that completes the toolkit to modulate NMNH:NMN+ ratio together with an NMN+-specific glucose dehydrogenase (GDH Ortho). The design principle discovered from Nox Ortho engineering and modeling is facilely translated onto six different enzymes to create NMN(H)-orthogonal biocatalysts with a consistent ~103–106-fold cofactor specificity switch from NAD(P)+ to NMN+. We assemble these enzymes to produce stereo-pure 2,3-butanediol in cell-free systems and in Escherichia coli, enabled by NMN(H)’s distinct redox ratio firmly set by its designated driving forces, decoupled from both NAD(H) and NADP(H).

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: Overview of the four proposed BDO stereo-upgrading systems for testing the crosstalk of two orthogonal cofactors.
Fig. 2: Directed evolution of Ll Nox to exclude NADH.
Fig. 3: Engineering Bdhs to use NMN(H).
Fig. 4: Orthogonal redox driving forces enable BDO stereo-upgrading in four cell-free systems.
Fig. 5: Resting cell stereo-upgrading of m-BDO to (SS)-BDO in E. coli.
Fig. 6: Visualizing emerging design principles for noncanonical redox cofactor-specific enzyme design.

Similar content being viewed by others

Data availability

Information on the strains and plasmids used in this study is available in Supplementary Data. Accession codes for genes investigated in this study can be found in Supplementary Tables 4 and 5. All DNA sequence and structural information used is publicly available (Ser Bdh2, UniProt A0A7U3Z3T2; Kp Dar, UniProt D7RP28; Bs BdhA, UniProt O34788; Cs Bdh, UniProt M1MWX5; Bs Gdh, UniProt P12310; Ll Nox, UniProt A2RIB7; Lb Nox, UniProt Q03Q85, PDB 5VN0; Tp Nox (Lb Nox G159A;D177A;A178R;M179S;P184R); Ecl Dar, GenBank JN035909; Kp BudC, PDB 1GEG; Ka BudC, National Center for Biotechnology Information (NCBI) Ref WP_015366942; Ka Dar, GenBank VEC79507; Ko BudC, GenBank AEX06195; As Bdh, GenBank NLH91458; Cr Bdh, GenBank AEI90716; Ca Bdh, NCBI Ref WP_073006451; Pb Bdh, NCBI Ref WP_148550966; Cbo Bdh, NCBI Ref WP_075141790; Km Bdh, NCBI Ref WP_102401827; Cbu Bdh, NCBI Ref WP_104675707; Zp Bdh, NCBI Ref WP_027704711). All data generated in this study are provided in the Supplementary Information. Source data are provided with this paper.

Code availability

The input files and source code for the Rosetta molecular simulation and MD simulations are publicly accessible on Zenodo (https://doi.org/10.5281/zenodo.11478967)80.

References

  1. Liu, Y. et al. Biofuels for a sustainable future. Cell 184, 1636–1647 (2021).

    Article  CAS  PubMed  Google Scholar 

  2. Liao, J. C., Mi, L., Pontrelli, S. & Luo, S. Fuelling the future: microbial engineering for the production of sustainable biofuels. Nat. Rev. Microbiol. 14, 288–304 (2016).

    Article  CAS  PubMed  Google Scholar 

  3. Opgenorth, P. H., Korman, T. P. & Bowie, J. U. A synthetic biochemistry molecular purge valve module that maintains redox balance. Nat. Commun. 5, 4113 (2014).

    Article  CAS  PubMed  Google Scholar 

  4. Cresko, J. et al. Industrial Decarbonization Roadmap (US Department of Energy, 2022).

  5. Lee, S. Y. & Kim, H. U. Systems strategies for developing industrial microbial strains. Nat. Biotechnol. 33, 1061–1072 (2015).

    Article  CAS  PubMed  Google Scholar 

  6. 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 

  7. Zachos, I., Döring, M., Tafertshofer, G., Simon, R. C. & Sieber, V. carba nicotinamide adenine dinucleotide phosphate: robust cofactor for redox biocatalysis. Angew. Chem. Int. Ed. Engl. 60, 14701–14706 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Beier, A. et al. Switch in cofactor specificity of a Baeyer–Villiger monooxygenase. ChemBioChem 17, 2312–2315 (2016).

    Article  CAS  PubMed  Google Scholar 

  9. Chánique, A. M. & Parra, L. P. Protein engineering for nicotinamide coenzyme specificity in oxidoreductases: attempts and challenges. Front. Microbiol. 9, 194 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  10. Rollin, J. A., Tam, T. K. & Zhang, Y.-H. P. New biotechnology paradigm: cell-free biosystems for biomanufacturing. Green Chem. 15, 1708 (2013).

    Article  CAS  Google Scholar 

  11. Mampel, J., Buescher, J. M., Meurer, G. & Eck, J. Coping with complexity in metabolic engineering. Trends Biotechnol. 31, 52–60 (2013).

    Article  CAS  PubMed  Google Scholar 

  12. Black, W. B. et al. Engineering a nicotinamide mononucleotide redox cofactor system for biocatalysis. Nat. Chem. Biol. 16, 87–94 (2020).

    Article  CAS  PubMed  Google Scholar 

  13. Pandit, A. V., Srinivasan, S. & Mahadevan, R. Redesigning metabolism based on orthogonality principles. Nat. Commun. 8, 1–11 (2017).

    Article  Google Scholar 

  14. Wang, X. et al. Creating enzymes and self-sufficient cells for biosynthesis of the non-natural cofactor nicotinamide cytosine dinucleotide. Nat. Commun. 12, 2116 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Richardson, K. N., Black, W. B. & Li, H. Aldehyde production in crude lysate- and whole cell-based biotransformation using a noncanonical redox cofactor system. ACS Catal. 10, 8898–8903 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Bat-Erdene, U. et al. Cell-free total biosynthesis of plant terpene natural products using an orthogonal cofactor regeneration system. ACS Catal. 11, 9898–9903 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Milo, R., Jorgensen, P., Moran, U., Weber, G. & Springer, M. BioNumbers—the database of key numbers in molecular and cell biology. Nucleic Acids Res. 38, D750–D753 (2010).

    Article  CAS  PubMed  Google Scholar 

  18. Pan, X. et al. A genetically encoded tool to increase cellular NADH/NAD+ ratio in living cells. Nat. Chem. Biol. 20, 594–604 (2024).

    Article  CAS  PubMed  Google Scholar 

  19. Voss, C. V. et al. Orchestration of concurrent oxidation and reduction cycles for stereoinversion and deracemisation of sec-alcohols. J. Am. Chem. Soc. 130, 13969–13972 (2008).

    Article  CAS  PubMed  Google Scholar 

  20. Liu, H. & Bowie, J. U. Cell-free synthetic biochemistry upgrading of ethanol to 1,3 butanediol. Sci. Rep. 11, 9449 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Knaus, T., Cariati, L., Masman, M. F. & Mutti, F. G. In vitro biocatalytic pathway design: orthogonal network for the quantitative and stereospecific amination of alcohols. Org. Biomol. Chem. 15, 8313–8325 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Drenth, J., Yang, G., Paul, C. E. & Fraaije, M. W. A tailor-made deazaflavin-mediated recycling system for artificial nicotinamide cofactor biomimetics. ACS Catal. 11, 11561–11569 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Knaus, T. et al. Better than nature: nicotinamide biomimetics that outperform natural coenzymes. J. Am. Chem. Soc. 138, 1033–1039 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Zachos, I. et al. Hot flows: evolving an archaeal glucose dehydrogenase for ultrastable carba-NADP+ using microfluidics at elevated temperatures. ACS Catal. 12, 1841–1846 (2022).

    Article  CAS  Google Scholar 

  25. Campbell, E., Meredith, M., Minteer, S. D. & Banta, S. Enzymatic biofuel cells utilizing a biomimetic cofactor. Chem. Commun. 48, 1898 (2012).

    Article  CAS  Google Scholar 

  26. Guarneri, A. et al. Flavoenzyme‐mediated regioselective aromatic hydroxylation with coenzyme biomimetics. ChemCatChem 12, 1368–1375 (2020).

    Article  CAS  Google Scholar 

  27. Meng, D. et al. Coenzyme engineering of glucose-6-phosphate dehydrogenase on a nicotinamide-based biomimic and its application as a glucose biosensor. ACS Catal. 13, 1983–1998 (2023).

    Article  CAS  Google Scholar 

  28. King, E. et al. Orthogonal glycolytic pathway enables directed evolution of noncanonical cofactor oxidase. Nat. Commun. 13, 7282 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Zhang, L. et al. Directed evolution of phosphite dehydrogenase to cycle noncanonical redox cofactors via universal growth selection platform. Nat. Commun. 13, 5021 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. King, E. et al. Engineering Embden–Meyerhof–Parnas glycolysis to generate noncanonical reducing power. ACS Catal. 12, 8582–8592 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Huang, R., Chen, H., Upp, D. M., Lewis, J. C. & Zhang, Y.-H. P. J. A high-throughput method for directed evolution of NAD(P)+-dependent dehydrogenases for the reduction of biomimetic nicotinamide analogues. ACS Catal. 9, 11709–11719 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Liang, K. & Shen, C. R. Selection of an endogenous 2,3-butanediol pathway in Escherichia coli by fermentative redox balance. Metab. Eng. 39, 181–191 (2017).

    Article  CAS  PubMed  Google Scholar 

  33. Maina, S. et al. Prospects on bio-based 2,3-butanediol and acetoin production: recent progress and advances. Biotechnol. Adv. 54, 107783 (2022).

    Article  CAS  PubMed  Google Scholar 

  34. Xiao, Z. et al. A novel whole-cell biocatalyst with NAD+ regeneration for production of chiral chemicals. PLoS ONE 5, e8860 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  35. He, Y. et al. Efficient (3S)-acetoin and (2S,3S)-2,3-butanediol production from meso-2,3-butanediol using whole-cell biocatalysis. Molecules 23, 691 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Zu, H. et al. Highly enantioselective synthesis of (R)-1,3-butanediol via deracemization of the corresponding racemate by a whole-cell stereoinverting cascade system. Microb. Cell Fact. 19, 125 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. 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 

  38. Sellés Vidal, L., Kelly, C. L., Mordaka, P. M. & Heap, J. T. Review of NAD(P)H-dependent oxidoreductases: properties, engineering and application. Biochim. Biophys. Acta Proteins Proteom. 1866, 327–347 (2018).

    Article  PubMed  Google Scholar 

  39. Jan, J., Martinez, I., Wang, Y., Bennett, G. N. & San, K. Y. Metabolic engineering and transhydrogenase effects on NADPH availability in Escherichia coli. Biotechnol. Prog. 29, 1124–1130 (2013).

    Article  CAS  PubMed  Google Scholar 

  40. Weckbecker, A. & Hummel, W. Improved synthesis of chiral alcohols with Escherichia coli cells co-expressing pyridine nucleotide transhydrogenase, NADP+-dependent alcohol dehydrogenase and NAD+-dependent formate dehydrogenase. Biotechnol. Lett. 26, 1739–1744 (2004).

    Article  CAS  PubMed  Google Scholar 

  41. 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 

  42. Krebs, H. A. & Veech, R. L. Equilibrium relations between pyridine nucleotides and adenine nucleotides and their roles in the regulation of metabolic processes. Adv. Enzym. Regul. 7, 397–413 (1969).

    Article  CAS  Google Scholar 

  43. Wu, J. T., Wu, L. H. & Knight, J. A. Stability of NADPH: effect of various factors on the kinetics of degradation. Clin. Chem. 32, 314–319 (1986).

    Article  CAS  PubMed  Google Scholar 

  44. Guarneri, A., van Berkel, W. J. & Paul, C. E. Alternative coenzymes for biocatalysis. Curr. Opin. Biotechnol. 60, 63–71 (2019).

    Article  CAS  PubMed  Google Scholar 

  45. Nowak, C., Pick, A., Csepei, L.-I. & Sieber, V. Characterization of biomimetic cofactors according to stability, redox potentials, and enzymatic conversion by NADH oxidase from Lactobacillus pentosus. ChemBioChem 18, 1944–1949 (2017).

    Article  CAS  PubMed  Google Scholar 

  46. Nowak, C. et al. A water-forming NADH oxidase from Lactobacillus pentosus suitable for the regeneration of synthetic biomimetic cofactors. Front. Microbiol. 6, 957 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  47. Zachos, I., Nowak, C. & Sieber, V. Biomimetic cofactors and methods for their recycling. Curr. Opin. Chem. Biol. 49, 59–66 (2019).

    Article  CAS  PubMed  Google Scholar 

  48. Rasor, B. J., Yi, X., Brown, H., Alper, H. S. & Jewett, M. C. An integrated in vivo/in vitro framework to enhance cell-free biosynthesis with metabolically rewired yeast extracts. Nat. Commun. 12, 5139 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Sudar, M., Findrik, Z., Domanovac, M. V. & Vasić-Rački, D. Coenzyme regeneration catalyzed by NADH oxidase from Lactococcus lactis. Biochem. Eng. J. 88, 12–18 (2014).

    Article  CAS  Google Scholar 

  50. Kim, S. & Hahn, J. S. Efficient production of 2,3-butanediol in Saccharomyces cerevisiae by eliminating ethanol and glycerol production and redox rebalancing. Metab. Eng. 31, 94–101 (2015).

    Article  CAS  PubMed  Google Scholar 

  51. Richter, F., Leaver-Fay, A., Khare, S. D., Bjelic, S. & Baker, D. De novo enzyme design using Rosetta3. PLoS ONE 6, e19230 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Bennett, B. D. et al. Absolute metabolite concentrations and implied enzyme active site occupancy in Escherichia coli. Nat. Chem. Biol. 5, 593–599 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Wang, J. et al. Engineering formaldehyde dehydrogenase from Pseudomonas putida to favor nicotinamide cytosine dinucleotide. ChemBioChem 23, 6–11 (2022).

    Article  Google Scholar 

  54. Liu, Y. et al. Structural insights into phosphite dehydrogenase variants favoring a non-natural redox cofactor. ACS Catal. 9, 1883–1887 (2019).

    Article  CAS  Google Scholar 

  55. Li, J. X. et al. Enhanced production of optical (S)-acetoin by a recombinant Escherichia coli whole-cell biocatalyst with NADH regeneration. RSC Adv. 8, 30512–30519 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Otagiri, M. et al. Crystal structure of meso-2,3-butanediol dehydrogenase in a complex with NAD+ and inhibitor mercaptoethanol at 1.7 Å resolution for understanding of chiral substrate recognition mechanisms. J. Biochem. 129, 205–208 (2001).

    Article  CAS  PubMed  Google Scholar 

  57. Li, L. et al. Biocatalytic production of (2S,3S)-2,3-butanediol from diacetyl using whole cells of engineered Escherichia coli. Bioresour. Technol. 115, 111–116 (2012).

    Article  CAS  PubMed  Google Scholar 

  58. Zhang, L. et al. Mechanism of 2,3-butanediol stereoisomers formation in a newly isolated Serratia sp. T241. Sci. Rep. 6, 19257 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Yan, Y., Lee, C.-C. & Liao, J. C. Enantioselective synthesis of pure (R,R)-2,3-butanediol in Escherichia coli with stereospecific secondary alcohol dehydrogenases. Org. Biomol. Chem. 7, 3914 (2009).

    Article  CAS  PubMed  Google Scholar 

  60. Jones, D. P. & Sies, H. The redox code. Antioxid. Redox Signal. 23, 734–746 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Nielsen, D. R., Yoon, S.-H., Yuan, C. J. & Prather, K. L. J. Metabolic engineering of acetoin and meso-2,3-butanediol biosynthesis in E. coli. Biotechnol. J. 5, 274–284 (2010).

    Article  CAS  PubMed  Google Scholar 

  62. Liang, K. & Shen, C. R. Engineering cofactor flexibility enhanced 2,3-butanediol production in Escherichia coli. J. Ind. Microbiol. Biotechnol. 44, 1605–1612 (2017).

    Article  CAS  PubMed  Google Scholar 

  63. Black, W. B. et al. Metabolic engineering of Escherichia coli for optimized biosynthesis of nicotinamide mononucleotide, a noncanonical redox cofactor. Microb. Cell Fact. 19, 1–10 (2020).

    Article  Google Scholar 

  64. Toledo-Patiño, S., Pascarelli, S., Uechi, G. & Laurino, P. Insertions and deletions mediated functional divergence of Rossmann fold enzymes. Proc. Natl Acad. Sci. USA 119, e2207965119 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Rao, S. T. & Rossmann, M. G. Comparison of super-secondary structures in proteins. J. Mol. Biol. 76, 241–256 (1973).

    Article  CAS  PubMed  Google Scholar 

  66. Medvedev, K. E., Kinch, L. N., Dustin Schaeffer, R., Pei, J. & Grishin, N. V. A fifth of the protein world: Rossmann-like proteins as an evolutionarily successful structural unit. J. Mol. Biol. 433, 166788 (2021).

    Article  CAS  PubMed  Google Scholar 

  67. Sillitoe, I. et al. CATH: increased structural coverage of functional space. Nucleic Acids Res. 49, D266–D273 (2021).

    Article  CAS  PubMed  Google Scholar 

  68. Bücher, T. & Klingenberg, M. Wege des Wasserstoffs in der lebendigen organisation. Angew. Chem. 70, 552–570 (1958).

    Article  Google Scholar 

  69. Agapakis, C. M. & Silver, P. A. Modular electron transfer circuits for synthetic biology. Bioeng. Bugs 1, 413–418 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  70. Rutter, J., Reick, M., Wu, L. C. & McKnight, S. L. Regulation of crock and NPAS2 DNA binding by the redox state of NAD cofactors. Science 293, 510–514 (2001).

    Article  CAS  PubMed  Google Scholar 

  71. Atkinson, J. T. et al. Metalloprotein switches that display chemical-dependent electron transfer in cells. Nat. Chem. Biol. 15, 189–195 (2019).

    Article  CAS  PubMed  Google Scholar 

  72. Huang, G. et al. Circadian oscillations of NADH redox state using a heterologous metabolic sensor in mammalian cells. J. Biol. Chem. 291, 23906–23914 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Gibson, D. G. et al. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat. Methods 6, 343–345 (2009).

    Article  CAS  PubMed  Google Scholar 

  74. Maxel, S. et al. A growth-based, high-throughput selection platform enables remodeling of 4-hydroxybenzoate hydroxylase active site. ACS Catal. 10, 6969–6974 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Jumper, J. et al. Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Tyka, M. D. et al. Alternate states of proteins revealed by detailed energy landscape mapping. J. Mol. Biol. 405, 607–618 (2011).

    Article  CAS  PubMed  Google Scholar 

  77. Wallen, J. R. et al. Structural analysis of Streptococcus pyogenes NADH oxidase: conformational dynamics involved in formation of the C(4a)-peroxyflavin intermediate. Biochemistry 54, 6815–6829 (2015).

    Article  CAS  PubMed  Google Scholar 

  78. Otagiri, M. et al. Structural basis for chiral substrate recognition by two 2,3-butanediol dehydrogenases. FEBS Lett. 584, 219–223 (2010).

    Article  CAS  PubMed  Google Scholar 

  79. Fleishman, S. J. et al. RosettaScripts: a scripting language interface to the Rosetta macromolecular modeling suite. PLoS ONE 6, e20161 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Zhu, Q. Rosetta files related to the engineering of Nox and Bdh enzymes. Zenodo https://doi.org/10.5281/zenodo.11478967 (2024).

  81. Huang, J. et al. CHARMM36m: an improved force field for folded and intrinsically disordered proteins. Nat. Methods 14, 71–73 (2017).

    Article  CAS  PubMed  Google Scholar 

  82. Vanommeslaeghe, K. et al. CHARMM general force field: a force field for drug‐like molecules compatible with the CHARMM all‐atom additive biological force fields. J. Comput. Chem. 31, 671–690 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Jorgensen, W. L., Chandrasekhar, J., Madura, J. D., Impey, R. W. & Klein, M. L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 79, 926–935 (1983).

    Article  CAS  Google Scholar 

  84. Feller, S. E., Zhang, Y., Pastor, R. W. & Brooks, B. R. Constant pressure molecular dynamics simulation: the Langevin piston method. J. Chem. Phys. 103, 4613–4621 (1995).

    Article  CAS  Google Scholar 

  85. Darden, T., York, D. & Pedersen, L. Particle mesh Ewald: An Nlog(N) method for Ewald sums in large systems. J. Chem. Phys. 98, 10089–10092 (1993).

    Article  CAS  Google Scholar 

  86. Allen, M. P. & Tildesley, D. J. Computer Simulation of Liquids 2nd edn (Oxford University Press, 2017).

  87. Phillips, J. C. et al. Scalable molecular dynamics on CPU and GPU architectures with NAMD. J. Chem. Phys. 153, 044130 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Humphrey, W., Dalke, A. & Schulten, K. VMD: visual molecular dynamics. J. Mol. Graph. 14, 33–38 (1996).

    Article  CAS  PubMed  Google Scholar 

  89. Zhang, Y. TM-align: a protein structure alignment algorithm based on the TM-score. Nucleic Acids Res. 33, 2302–2309 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Virtanen, P. et al. SciPy 1.0: fundamental algorithms for scientific computing in Python. Nat. Methods 17, 261–272 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

H.L. acknowledges support from the University of California, Irvine, the National Science Foundation (NSF) (award nos. 1847705 and MCB-2328145), the National Institutes of Health (NIH) (award no. DP2 GM137427), an Alfred Sloan research fellowship and the Advanced Research Projects Agency—Energy (award no. DE-AR0001508). Y.C., E.L. and J.B.S. acknowledge the funding of the National Institute of Environmental Health Sciences (grant no. P42ES004699), the NIH (grant no. R01 GM 076324-11) and the NSF (grant nos. 1627539, 1805510 and 1827246). R.L. acknowledges support from the NIH (award no. R35 GM130367). D.A acknowledges support from the NSF Graduate Research Fellowship Program (grant no. DGE1839285). We acknowledge valuable technical support provided by AAT Bioquest for the redox ratio kits.

Author information

Authors and Affiliations

Authors

Contributions

D.A., Y.Z., E.K. and H.L. designed the experiments. Y.C. and E.L. performed Rosetta modeling. E.L. and D.A. performed bioinformatic analysis of the NAD(P)-binding Rossmann-like domain superfamily. E.K. and Y.Z. performed rational protein engineering experiments. Y.Z. performed growth-based selection of Nox. Q.Z., Y.W. and R.L. performed MD simulations and analyzed the results. D.A. and E.K. bioprospected Bdh homologs. D.A. performed the Michaelis–Menten kinetic experiments for Bdhs. Y.Z. performed the Michaelis–Menten kinetic experiments for Nox. D.A. and S.P. performed cell-free biotransformation experiments. D.A. performed the resting cell biotransformation experiments. W.B.B. and D.A. reformulated the colorimetric redox ratio assay. D.A. performed the colorimetric redox ratio assays. Y.C., E.L. and J.B.S. analyzed the modeling results. All authors analyzed the data and wrote the paper.

Corresponding author

Correspondence to Han Li.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Chemical Biology thanks Zhen Chen, Steffen Lindner and Carine Vergne-Vaxelaire 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.

Supplementary information

Supplementary Information

Supplementary Methods, Tables 1–5 and Figs. 1–12.

Reporting Summary

Supplementary Data

Key resources table.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

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

Aspacio, D., Zhang, Y., Cui, Y. et al. Shifting redox reaction equilibria on demand using an orthogonal redox cofactor. Nat Chem Biol (2024). https://doi.org/10.1038/s41589-024-01702-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1038/s41589-024-01702-5

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing