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:

NMR-guided directed evolution

Abstract

Directed evolution is a powerful tool for improving existing properties and imparting completely new functionalities to proteins1,2,3,4. Nonetheless, its potential in even small proteins is inherently limited by the astronomical number of possible amino acid sequences. Sampling the complete sequence space of a 100-residue protein would require testing of 20100 combinations, which is beyond any existing experimental approach. In practice, selective modification of relatively few residues is sufficient for efficient improvement, functional enhancement and repurposing of existing proteins5. Moreover, computational methods have been developed to predict the locations and, in certain cases, identities of potentially productive mutations6,7,8,9. Importantly, all current approaches for prediction of hot spots and productive mutations rely heavily on structural information and/or bioinformatics, which is not always available for proteins of interest. Moreover, they offer a limited ability to identify beneficial mutations far from the active site, even though such changes may markedly improve the catalytic properties of an enzyme10. Machine learning methods have recently showed promise in predicting productive mutations11, but they frequently require large, high-quality training datasets, which are difficult to obtain in directed evolution experiments. Here we show that mutagenic hot spots in enzymes can be identified using NMR spectroscopy. In a proof-of-concept study, we converted myoglobin, a non-enzymatic oxygen storage protein, into a highly efficient Kemp eliminase using only three mutations. The observed levels of catalytic efficiency exceed those of proteins designed using current approaches and are similar with those of natural enzymes for the reactions that they are evolved to catalyse. Given the simplicity of this experimental approach, which requires no a priori structural or bioinformatic knowledge, we expect it to be widely applicable and to enable the full potential of directed enzyme evolution.

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: Kemp elimination.
Fig. 2: NMR-guided evolution of myoglobin.
Fig. 3: NMR-guided evolution of calmodulin.

Similar content being viewed by others

Data availability

The crystallographic data and refinement statistics were deposited in the Protein Data Bank (PDB) with the entry code 7VUC (FerrElCat), 7VUR (AlleyCat9), 7VUS (AlleyCat9 with 6-NBT), 7VUT (AlleyCat10) and 7VUU (AlleyCat10 with 6-NBT). The NMR chemical shift data are provided as part of the Supplementary Information.

References

  1. Bornscheuer, U. T. et al. Engineering the third wave of biocatalysis. Nature 485, 185–194 (2012).

    Article  ADS  CAS  PubMed  Google Scholar 

  2. Reetz, M. T. Laboratory evolution of stereoselective enzymes: a prolific source of catalysts for asymmetric reactions. Angew. Chem. Int. Ed. Engl. 50, 138–174 (2011).

    Article  CAS  PubMed  Google Scholar 

  3. Denard, C. A., Ren, H. & Zhao, H. Improving and repurposing biocatalysts via directed evolution. Curr. Opin. Chem. Biol. 25, 55–64 (2015).

    Article  CAS  PubMed  Google Scholar 

  4. Chen, K. & Arnold, F. H. Engineering new catalytic activities in enzymes. Nat. Catal. 3, 203–213 (2020).

    Article  CAS  Google Scholar 

  5. Reetz, M. T., Wilensek, S., Zha, D. & Jaeger, K. E. Directed evolution of an enantioselective enzyme through combinatorial multiple-cassette mutagenesis. Angew. Chem. Int. Ed. Engl. 40, 3589–3591 (2001).

    Article  CAS  PubMed  Google Scholar 

  6. Wijma, H. J., Floor, R. J. & Janssen, D. B. Structure- and sequence-analysis inspired engineering of proteins for enhanced thermostability. Curr. Opin. Struct. Biol. 23, 588–594 (2013).

    Article  CAS  PubMed  Google Scholar 

  7. Planas-Iglesias, J. et al. Computational design of enzymes for biotechnological applications. Biotechnol. Adv. 47, 107696 (2021).

    Article  CAS  PubMed  Google Scholar 

  8. Verma, R., Schwaneberg, U. & Roccatano, D. Computer-aided protein directed evolution: a review of web servers, databases and other computational tools for protein engineering. Comput. Struct. Biotechnol. J. 2, e201209008 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Ebert, M. C. & Pelletier, J. N. Computational tools for enzyme improvement: why everyone can - and should - use them. Curr. Opin. Chem. Biol. 37, 89–96 (2017).

    Article  CAS  PubMed  Google Scholar 

  10. Osuna, S. The challenge of predicting distal active site mutations in computational enzyme design. WIREs Comput. Mol. Sci. 11, e1502 (2021).

    Article  CAS  Google Scholar 

  11. Wu, Z., Kan, S. B. J., Lewis, R. D., Wittmann, B. J. & Arnold, F. H. Machine learning-assisted directed protein evolution with combinatorial libraries. Proc. Natl Acad. Sci. USA 116, 8852–8858 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  12. Acevedo-Rocha, C. G. et al. Pervasive cooperative mutational effects on multiple catalytic enzyme traits emerge via long-range conformational dynamics. Nat. Commun. 12, 1621 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  13. Otten, R. et al. How directed evolution reshapes the energy landscape in an enzyme to boost catalysis. Science 370, 1442–1446 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  14. Campbell, E. et al. The role of protein dynamics in the evolution of new enzyme function. Nat. Chem. Biol. 12, 944–950 (2016).

    Article  CAS  PubMed  Google Scholar 

  15. Hong, N. S. et al. The evolution of multiple active site configurations in a designed enzyme. Nat. Commun. 9, 3900 (2018).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  16. Broom, A. et al. Ensemble-based enzyme design can recapitulate the effects of laboratory directed evolution in silico. Nat. Commun. 11, 4808 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  17. Warshel, A. et al. Electrostatic basis for enzyme catalysis. Chem. Rev. 106, 3210–3235 (2006).

    Article  CAS  PubMed  Google Scholar 

  18. Kemp, D. S. & Casey, M. L. Physical organic chemistry of benzisoxazoles. II. Linearity of the Bronsted free energy relationship for the base-catalyzed decomposition of benzisoxazoles. J. Am. Chem. Soc. 95, 6670–6680 (1973).

    Article  CAS  Google Scholar 

  19. Rothlisberger, D. et al. Kemp elimination catalysts by computational enzyme design. Nature 453, 190–195 (2008).

    Article  ADS  PubMed  Google Scholar 

  20. Blomberg, R. et al. Precision is essential for efficient catalysis in an evolved Kemp eliminase. Nature 503, 418–421 (2013).

    Article  ADS  CAS  PubMed  Google Scholar 

  21. Risso, V. A. et al. De novo active sites for resurrected Precambrian enzymes. Nat. Commun. 8, 16113 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  22. Merski, M. & Shoichet, B. K. Engineering a model protein cavity to catalyze the Kemp elimination. Proc. Natl Acad. Sci. USA 109, 16179–16183 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  23. Debler, E. W., Muller, R., Hilvert, D. & Wilson, I. A. An aspartate and a water molecule mediate efficient acid–base catalysis in a tailored antibody pocket. Proc. Natl Acad. Sci. USA 106, 18539–18544 (2009).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  24. Vaissier, V., Sharma, S. C., Schaettle, K., Zhang, T. & Head-Gordon, T. Computational optimization of electric fields for improving catalysis of a designed Kemp eliminase. ACS Catal. 8, 219–227 (2018).

    Article  CAS  Google Scholar 

  25. Lamba, V. et al. Kemp eliminase activity of ketosteroid isomerase. Biochemistry 56, 582–591 (2017).

    Article  CAS  PubMed  Google Scholar 

  26. Risso, V. A. et al. Enhancing a de novo enzyme activity by computationally-focused ultra-low-throughput screening. Chem. Sci. 11, 6134–6148 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Li, A. et al. A redox-mediated Kemp eliminase. Nat. Commun. 8, 14876 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  28. Miao, Y., Metzner, R. & Asano, Y. Kemp elimination catalyzed by naturally occurring aldoxime dehydratases. ChemBioChem 18, 451–454 (2017).

    Article  CAS  PubMed  Google Scholar 

  29. Bordeaux, M., Tyagi, V. & Fasan, R. Highly diastereoselective and enantioselective olefin cyclopropanation using engineered myoglobin-based catalysts. Angew. Chem. Int. Ed. Engl. 54, 1744–1748 (2015).

    Article  CAS  PubMed  Google Scholar 

  30. Yi, J., Heinecke, J., Tan, H., Ford, P. C. & Richter-Addo, G. B. The distal pocket histidine residue in horse heart myoglobin directs the O-binding mode of nitrite to the heme iron. J. Am. Chem. Soc. 131, 18119–18128 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Wang, B. et al. Nitrosyl myoglobins and their nitrite precursors: crystal structural and quantum mechanics and molecular mechanics theoretical investigations of preferred Fe–NO ligand orientations in myoglobin distal pockets. Biochemistry 57, 4788–4802 (2018).

    Article  CAS  PubMed  Google Scholar 

  32. Korendovych, I. V. et al. Design of a switchable eliminase. Proc. Natl Acad. Sci. USA 108, 6823–6827 (2011).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  33. Moroz, O. V. et al. A single mutation in a regulatory protein produces evolvable allosterically regulated catalyst of nonnatural reaction. Angew. Chem. Int. Ed. Engl. 52, 6246–6249 (2013).

    Article  CAS  PubMed  Google Scholar 

  34. Chattopadhyaya, R., Meador, W. E., Means, A. R. & Quiocho, F. A. Calmodulin structure refined at 1.7 Å resolution. J. Mol. Biol. 228, 1177–1192 (1992).

    Article  CAS  PubMed  Google Scholar 

  35. Marshall, L. R., Zozulia, O., Lengyel-Zhand, Z. & Korendovych, I. V. Minimalist de novo design of protein catalysts. ACS Catal. 9, 9265–9275 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Casey, M. L., Kemp, D. S., Paul, K. G. & Cox, D. D. Physical organic chemistry of benzisoxazoles. I. Mechanism of base-catalyzed decomposition of benzisoxazoles. J. Org. Chem. 38, 2294–2301 (1973).

    Article  CAS  Google Scholar 

  37. Berry, E. A. & Trumpower, B. L. Simultaneous determination of hemes a, hemes b, and hemes c from pyridine hemochrome spectra. Anal. Biochem. 161, 1–15 (1987).

    Article  CAS  PubMed  Google Scholar 

  38. Barik, S. in PCR Cloning Protocols (eds. Chen, B.-Y. & Janes H.) 189–196 (Humana Press, 2002).

  39. Delaglio, F. et al. NMRPipe—a multidimensional spectral processing system based on UNIX Pipes. J. Biomol. NMR 6, 277–293 (1995).

    Article  CAS  PubMed  Google Scholar 

  40. Vranken, W. F. et al. The CCPN data model for NMR spectroscopy: development of a software pipeline. Proteins 59, 687–696 (2005).

    Article  CAS  PubMed  Google Scholar 

  41. Kannt, A., Young, S. & Bendall, D. S. The role of acidic residues of plastocyanin in its interaction with cytochrome f. Biochim. Biophys. Acta 1277, 115–126 (1996).

    Article  CAS  PubMed  Google Scholar 

  42. Storoni, L. C., McCoy, A. J. & Read, R. J. Likelihood-enhanced fast rotation functions. Acta Crystallogr. D 60, 432–438 (2004).

    Article  PubMed  Google Scholar 

  43. Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. Features and development of Coot. Acta Crystallogr. D 66, 486–501 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Adams, P. D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D 66, 213–221 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Chen, V. B. et al. MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr. D 66, 12–21 (2010).

    Article  CAS  PubMed  Google Scholar 

  46. Trott, O. & Olson, A. J. AutoDock Vina: improving the speed and accuracy of docking with a new scoring function, efficient optimization and multithreading. J. Comput. Chem. 31, 455–461 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Makhlynets, O. V. & Korendovych, I. V. Minimalist design of allosterically regulated protein catalysts. Meth. Enzymol. 580, 191–202 (2016).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Institutes of Health grant GM119634, the Japan Society for the Promotion of Science and the Alexander von Humboldt Foundation. The authors thank R. Fasan for the gift of the plasmid containing the Mb gene. J.R.H.T. thanks OpenEye Scientific Software.

Author information

Authors and Affiliations

Authors

Contributions

I.V.K., O.V.M. and A.N.V. designed the experiments. S.B., A.K., J.H.Y. and I.K. performed directed evolution, protein expression and characterization, and kinetic studies. A.D., A.N.V., O.V.M. and I.V.K. performed NMR titrations, backbone and side-chain assignment and NMR data analysis. Е.G.М., K.T. and J.R.H.T. expressed and crystallized the proteins and solved their structures. I.V.K. and O.V.M. wrote the manuscript with input from all the authors.

Corresponding authors

Correspondence to Alexander N. Volkov, Olga V. Makhlynets or Ivan V. Korendovych.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature thanks Giovanna Ghirlanda, Anthony Mittermaier and Jose Sanchez-Ruiz 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 figures and tables

Extended Data Fig. 1 Selection of negative controls for screening.

Left. Аll residues in myoglobin were sorted in bins based on their distance to the docked inhibitor (black) in FerrElCat. The residues in the van der Waals contact with the docked inhibitor were placed in bin 1 (red), the residues in direct contact with the residues in bin 1 were placed in bin 2 (orange), etc. A total of five bins were devised: red, orange, yellow, green and blue. Right. The list of the residues sorted in the five bins. Residues showing large backbone CSP and their immediate neighbors are highlighted in red and yellow, respectively. Unassigned positions and residues immediately next to unassigned stretches are shown in dark grey and light grey, respectively. Prolines are highlighted in blue. Residues showing small CSP that were selected as controls are either highlighted in green or labeled in green font (when located next to unassigned residues).

Extended Data Fig. 2 Substrate turnover by reduced Mb(L29I/H64G/V68A) (FerrElCat).

Reaction was monitored using stopped-flow at pH 8.0 for 30 s at 25 °C with 140 µM of 5-NBI and 5 nM of FerrElCat.

Extended Data Fig. 3 Catalytic parameters of Kemp eliminases evolved using directed evolution.

kcat/KM and kcat values for the evolved enzymes (top and bottom left, respectively) and improvement in kcat/KM and kcat achieved by directed evolution (top and bottom right, respectively). In cases where only kcat/KM was reported, we used 5 mM for KM, to obtain low estimate of the kcat.

Extended Data Fig. 4 Spectroelectrochemical determination of redox potentials of selected myoglobin mutants.

Mb(H64V) (left), Mb(H64G/V68A) (middle) and FerrElCat (right). The proteins were analyzed in 20 mM Tris-HCl, pH 8.0 at 20 °C in presence of the mediator (100 µM phenazine sulfate). The redox potentials (vs Ag/AgCl) are summarized in Supplementary Table 5. Data are presented as the mean values and the error bars represent standard deviations obtained from three independent measurements for Mb(H64V) and Mb(H64G/V68A). For FerrElCat, the experiment was repeated twice with essentially identical results, one set of data is shown.

Extended Data Fig. 5 Kemp elimination catalyzed by AlleyCat10 in presence of Ca2+ (black) and in absence of Ca2+ (red).

The activity of 0.1 µM AlleyCat10 was tested with 0.12-0.96 mM substrate in 20 mM HEPES, 100 mM NaCl, pH 7.0 at 22 °C with 10 mM CaCl2 (black) or 100 mM EDTA (red). Data are presented as the mean values and the error bars (most are small compared to the symbol size) represent standard deviations obtained from three independent measurements.

Extended Data Fig. 6 Absorbance spectra of Mb(H64V) (left) and Mb(L29I/H64G/V68A) (FerrElCat) (right) in the oxidized (black) and reduced (red) forms.

Spectra for the proteins (1.6 µM for Mb(H64V), 3.1 µM for FerrElCat) were collected in 20 mM HEPES, pH 7.0 in an anaerobic cuvette with a 1 cm pathlength at room temperature.

Extended Data Fig. 7 Michaelis-Menten plots of Kemp elimination catalyzed by reduced (unless otherwise stated) myoglobin mutants and by AlleyCat proteins.

Final reaction mixtures for myoglobin mutant analyses contained 1 mM L-ascorbic acid, 0.1 µM SOD, 20 nM catalase, 140-840 µM substrate, 1.5% acetonitrile in 20 mM Tris (pH 8.0). The protein concentration was 1 µM for Mb(H64V), 0.1 or 0.25 µM for Mb(H64V)-based double variants, 5 nM for Mb(H64G) or Mb(H64G)-based double or triple variants. For the AlleyCat proteins reaction mixtures contained 0.1 µM proteins with 0.12-0.96 mM substrate in 1.5% acetonitrile, 20 mM Tris, pH 8.0, 10 mM CaCl2, 100 mM NaCl. Kinetic parameters are summarized in Table 1 and Extended Data Table 1. Data are presented as the mean values and the error bars represent standard deviations obtained from at least three independent measurements (six for Mb(H64G) and reduced Mb(H64V).

Extended Data Table 1 Kinetic parameters for the Kemp elimination reaction catalyzed by myoglobin variants in the reduced state (unless otherwise stated) at pH 8.0
Extended Data Table 2 Sequences of the proteins used in this study
Extended Data Table 3 Crystallographic data collection and refinement statistics

Supplementary information

Reporting Summary

Supplementary Data 1

This file contains Supplementary Figs. 1–11 and Tables 1–6.

Supplementary Data 2

The file contains the chemical shifts values, CSP values and Z-scores for individual residues derived from NMR titrations.

Rights and permissions

Springer Nature or its licensor 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

Bhattacharya, S., Margheritis, E.G., Takahashi, K. et al. NMR-guided directed evolution. Nature 610, 389–393 (2022). https://doi.org/10.1038/s41586-022-05278-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-022-05278-9

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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