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:

Discovery of small-molecule enzyme activators by activity-based protein profiling

Abstract

Activity-based protein profiling (ABPP) has been used extensively to discover and optimize selective inhibitors of enzymes. Here, we show that ABPP can also be implemented to identify the converse—small-molecule enzyme activators. Using a kinetically controlled, fluorescence polarization-ABPP assay, we identify compounds that stimulate the activity of LYPLAL1—a poorly characterized serine hydrolase with complex genetic links to human metabolic traits. We apply ABPP-guided medicinal chemistry to advance a lead into a selective LYPLAL1 activator suitable for use in vivo. Structural simulations coupled to mutational, biochemical and biophysical analyses indicate that this compound increases LYPLAL1’s catalytic activity likely by enhancing the efficiency of the catalytic triad charge-relay system. Treatment with this LYPLAL1 activator confers beneficial effects in a mouse model of diet-induced obesity. These findings reveal a new mode of pharmacological regulation for this large enzyme family and suggest that ABPP may aid discovery of activators for additional enzyme classes.

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: Structure of a small-molecule activator of LYPLAL1.
Fig. 2: Characterization of a small-molecule activator of LYPLAL1.
Fig. 3: Summary of findings from structure–activity study.
Fig. 4: Optimization of LYPLAL1 activators.
Fig. 5: Key noncatalytic residues regulate LYPLAL1 activity and pharmacological activator efficacy.
Fig. 6: LYPLAL1 activation imparts metabolic benefits in vivo.

Similar content being viewed by others

Data availability

All data generated or analyzed during this study are included in this article and its Supplementary Information files, or are available from the corresponding authors upon request. Source data for Figs. 2 and 46 and Extended Data Figs. 16 are provided with the paper. Structures of 4 (CCDC 1825320), 12 (CCDC 1825321), 34 (CCDC 1825319), 37 (CCDC 1825322) and 78 (CCDC 1825323) established by single-crystal X-ray structure determinations and accompanying data have been deposited in the Cambridge Crystallographic Data Center (CCDC). Data and reagents requests should be addressed to E. Saez (esaez@scripps.edu) or D.L. Boger (boger@scripps.edu).

References

  1. Bachovchin, D. A. & Cravatt, B. F. The pharmacological landscape and therapeutic potential of serine hydrolases. Nat. Rev. Drug Discov. 11, 52–68 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Long, J. Z. & Cravatt, B. F. The metabolic serine hydrolases and their functions in mammalian physiology and disease. Chem. Rev. 111, 6022–6063 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Xu, J. et al. Discovery of potent, selective, and orally bioavailable pyridone-based dipeptidyl peptidase-4 inhibitors. Bioorg. Med. Chem. Lett. 16, 1346–1349 (2006).

    Article  CAS  PubMed  Google Scholar 

  4. Zhi, J. et al. Retrospective population-based analysis of the dose-response (fecal fat excretion) relationship of orlistat in normal and obese volunteers. Clin. Pharmacol. Ther. 56, 82–85 (1994).

    Article  CAS  PubMed  Google Scholar 

  5. Cravatt, B. F., Wright, A. T. & Kozarich, J. W. Activity-based protein profiling: from enzyme chemistry to proteomic chemistry. Annu. Rev. Biochem. 77, 383–414 (2008).

    Article  CAS  PubMed  Google Scholar 

  6. Sadaghiani, A. M., Verhelst, S. H. & Bogyo, M. Tagging and detection strategies for activity-based proteomics. Curr. Opin. Chem. Biol. 11, 20–28 (2007).

    Article  CAS  PubMed  Google Scholar 

  7. Heal, W. P., Dang, T. H. & Tate, E. W. Activity-based probes: discovering new biology and new drug targets. Chem. Soc. Rev. 40, 246–257 (2011).

    Article  CAS  PubMed  Google Scholar 

  8. Jessani, N. & Cravatt, B. F. The development and application of methods for activity-based protein profiling. Curr. Opin. Chem. Biol. 8, 54–59 (2004).

    Article  CAS  PubMed  Google Scholar 

  9. Jessani, N., Liu, Y., Humphrey, M. & Cravatt, B. F. Enzyme activity profiles of the secreted and membrane proteome that depict cancer invasiveness. Proc. Natl Acad. Sci. USA 99, 10335–10340 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Jessani, N. et al. A streamlined platform for high-content functional proteomics of primary human specimens. Nat. Methods 2, 691–697 (2005).

    Article  CAS  PubMed  Google Scholar 

  11. Liu, Y., Patricelli, M. P. & Cravatt, B. F. Activity-based protein profiling: the serine hydrolases. Proc. Natl Acad. Sci. USA 96, 14694–14699 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Patricelli, M. P., Giang, D. K., Stamp, L. M. & Burbaum, J. J. Direct visualization of serine hydrolase activities in complex proteomes using fluorescent active site-directed probes. Proteomics 1, 1067–1071 (2001).

    Article  CAS  PubMed  Google Scholar 

  13. Jessani, N. et al. Carcinoma and stromal enzyme activity profiles associated with breast tumor growth in vivo. Proc. Natl Acad. Sci. USA 101, 13756–13761 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Dominguez, E. et al. Integrated phenotypic and activity-based profiling links Ces3 to obesity and diabetes. Nat. Chem. Biol. 10, 113–121 (2014).

    Article  CAS  PubMed  Google Scholar 

  15. Li, W., Blankman, J. L. & Cravatt, B. F. A functional proteomic strategy to discover inhibitors for uncharacterized hydrolases. J. Am. Chem. Soc. 129, 9594–9595 (2007).

    Article  CAS  PubMed  Google Scholar 

  16. Leung, D., Hardouin, C., Boger, D. L. & Cravatt, B. F. Discovering potent and selective reversible inhibitors of enzymes in complex proteomes. Nat. Biotechnol. 21, 687–691 (2003).

    Article  CAS  PubMed  Google Scholar 

  17. Bachovchin, D. A., Brown, S. J., Rosen, H. & Cravatt, B. F. Identification of selective inhibitors of uncharacterized enzymes by high-throughput screening with fluorescent activity-based probes. Nat. Biotechnol. 27, 387–394 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Bachovchin, D. A. et al. Superfamily-wide portrait of serine hydrolase inhibition achieved by library-versus-library screening. Proc. Natl Acad. Sci. USA 107, 20941–20946 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Manning, A. K. et al. A genome-wide approach accounting for body mass index identifies genetic variants influencing fasting glycemic traits and insulin resistance. Nat. Genet. 44, 659–669 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Scott, R. A. et al. Large-scale association analyses identify new loci influencing glycemic traits and provide insight into the underlying biological pathways. Nat. Genet. 44, 991–1005 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Klarin, D. et al. Genetics of blood lipids among ~300,000 multi-ethnic participants of the Million Veteran Program. Nat. Genet. 50, 1514–1523 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Xue, A. et al. Genome-wide association analyses identify 143 risk variants and putative regulatory mechanisms for type 2 diabetes. Nat. Commun. 9, 2941 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Graff, M. et al. Genome-wide physical activity interactions in adiposity—a meta-analysis of 200,452 adults. PLoS Genet. 13, e1006528 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Winkler, T. W. et al. The influence of age and sex on genetic associations with adult body size and shape: a large-scale genome-wide interaction study. PLoS Genet. 11, e1005378 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Shungin, D. et al. New genetic loci link adipose and insulin biology to body fat distribution. Nature 518, 187–196 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Berndt, S. I. et al. Genome-wide meta-analysis identifies 11 new loci for anthropometric traits and provides insights into genetic architecture. Nat. Genet. 45, 501–512 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Speliotes, E. K. et al. Association analyses of 249,796 individuals reveal 18 new loci associated with body mass index. Nat. Genet. 42, 937–948 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Speliotes, E. K. et al. Genome-wide association analysis identifies variants associated with nonalcoholic fatty liver disease that have distinct effects on metabolic traits. PLoS Genet. 7, e1001324 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Lindgren, C. M. et al. Genome-wide association scan meta-analysis identifies three loci influencing adiposity and fat distribution. PLoS Genet. 5, e1000508 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Dastani, Z. et al. Novel loci for adiponectin levels and their influence on type 2 diabetes and metabolic traits: a multi-ethnic meta-analysis of 45,891 individuals. PLoS Genet. 8, e1002607 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Heid, I. M. et al. Meta-analysis identifies 13 new loci associated with waist-hip ratio and reveals sexual dimorphism in the genetic basis of fat distribution. Nat. Genet. 42, 949–960 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Lei, X., Callaway, M., Zhou, H., Yang, Y. & Chen, W. Obesity associated Lyplal1 gene is regulated in diet induced obesity but not required for adipocyte differentiation. Mol. Cell Endocrinol. 411, 207–213 (2015).

    Article  CAS  PubMed  Google Scholar 

  33. Watson, R. A. et al. Lyplal1 is dispensable for normal fat deposition in mice. Dis. Model Mech. 10, 1481–1488 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Norheim, F. et al. Gene-by-sex interactions in mitochondrial functions and cardio-metabolic traits. Cell Metab. 29, 932–949.e4 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Ahn, K. et al. Discovery of a selective covalent inhibitor of lysophospholipase-like 1 (LYPLAL1) as a tool to evaluate the role of this serine hydrolase in metabolism. ACS Chem. Biol. 11, 2529–2540 (2016).

    Article  CAS  PubMed  Google Scholar 

  36. Simon, G. M. & Cravatt, B. F. Activity-based proteomics of enzyme superfamilies: serine hydrolases as a case study. J. Biol. Chem. 285, 11051–11055 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Burger, M. et al. Crystal structure of the predicted phospholipase LYPLAL1 reveals unexpected functional plasticity despite close relationship to acyl protein thioesterases. J. Lipid Res. 53, 43–50 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Won, S. J. et al. Molecular mechanism for isoform-selective inhibition of acyl protein thioesterases 1 and 2 (APT1 and APT2). ACS Chem. Biol. 11, 3374–3382 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Hameed, A. et al. Small molecules as activators in medicinal chemistry (2000–2016). Expert Opin. Ther. Pat. 27, 1089–1110 (2017).

    Article  CAS  PubMed  Google Scholar 

  40. Lass, A. et al. Adipose triglyceride lipase-mediated lipolysis of cellular fat stores is activated by CGI-58 and defective in Chanarin–Dorfman Syndrome. Cell Metab. 3, 309–319 (2006).

    Article  CAS  PubMed  Google Scholar 

  41. Yang, P. & Liu, K. Activity-based protein profiling: recent advances in probe development and applications. Chem. Bio. Chem. 16, 712–724 (2015).

    Article  CAS  PubMed  Google Scholar 

  42. Chen, Z. et al. Functional screening of candidate causal genes for insulin resistance in human preadipocytes and adipocytes. Circ. Res. 126, 330–346 (2020).

    Article  CAS  PubMed  Google Scholar 

  43. Bai, N., Roder, H., Dickson, A. & Karanicolas, J. Isothermal analysis of thermofluor data can readily provide quantitative binding affinities. Sci. Rep. 9, 2650 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  44. Jo, S., Kim, T., Iyer, V. G. & Im, W. CHARMM‐GUI: a web‐based graphical user interface for CHARMM. J. Comput. Chem. 29, 1859–1865 (2008).

    Article  CAS  PubMed  Google Scholar 

  45. Brooks, B. R. et al. CHARMM: the biomolecular simulation program. J. Comput. Chemistry 30, 1545–1614 (2009).

    Article  CAS  Google Scholar 

  46. Phillips, J. C. et al. Scalable molecular dynamics with NAMD. J. Comput. Chem. 26, 1781–1802 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  Google Scholar 

  48. Essmann, U. et al. A smooth particle mesh Ewald method. J. Chem. Phys. 103, 8577–8593 (1998).

    Article  Google Scholar 

  49. Ryckaert, J.-P., Ciccotti, G. & Berendsen, H. J. Numerical integration of the Cartesian equations of motion of a system with constraints: molecular dynamics of n-alkanes. J. Comput. Phys. 23, 327–341 (1977).

    Article  CAS  Google Scholar 

  50. Gowers, R.J. et al. MDAnalysis: a Python package for the rapid analysis of molecular dynamics simulations (Los Alamos National Laboratory, 2019).

  51. Michaud-Agrawal, N., Denning, E. J., Woolf, T. B. & Beckstein, O. MDAnalysis: a toolkit for the analysis of molecular dynamics simulations. J. Comput. Chem. 32, 2319–2327 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Tan, Y. S., Spring, D. R., Abell, C. & Verma, C. The use of chlorobenzene as a probe molecule in molecular dynamics simulations. J. Chem. Inform. Model. 54, 1821–1827 (2014).

    Article  CAS  Google Scholar 

  53. Vanommeslaeghe, K. & MacKerell, A. D. Jr Automation of the CHARMM general force field (CGenFF) I: bond perception and atom typing. J. Chem. Inform. Model. 52, 3144–3154 (2012).

    Article  CAS  Google Scholar 

  54. Vanommeslaeghe, K., Raman, E. P. & MacKerell, A. D. Jr Automation of the CHARMM general force field (CGenFF) II: assignment of bonded parameters and partial atomic charges. J. Chem. Inform. Model. 52, 3155–3168 (2012).

    Article  CAS  Google Scholar 

  55. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Yu, W., He, X., Vanommeslaeghe, K. & MacKerell, A. D. Jr Extension of the CHARMM general force field to sulfonyl‐containing compounds and its utility in biomolecular simulations. J. Comput. Chem. 33, 2451–2468 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Walt, Svd, Colbert, S. C. & Varoquaux, G. The NumPy Array: a structure for efficient numerical computation. Comput. Sci. Eng. 13, 22–30 (2011).

    Article  Google Scholar 

  58. Ghanakota, P. & Carlson, H. A. Moving beyond active-site detection: MixMD applied to allosteric systems. J. Phys. Chem. B. 120, 8685–8695 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank C. Moore and A. Rheingold of the Crystallography Facility at the University of California, San Diego for X-ray structure determinations of 4, 12, 34, 37 and 78, C. Vernochet, T.V. Magee and C. Hong at Pfizer for compounds C11 and C12, and M. Petrassi at Calibr for discussions. This work was supported by National Institutes of Health grant nos. DA015648 (to D.L.B.), GM069832 (to S.F.), DK099810 (to E.S. and B.F.C.) and DK114785 (to E.S. and B.F.C.). B.P.K. was supported in part by fellowship no. 15POST25100007 from the American Heart Association.

Author information

Authors and Affiliations

Authors

Contributions

E.S. conceived the project. D.W.W., S.F., B.F.C., D.L.B. and E.S. designed and directed research. T.J. and J.X. purified recombinant proteins. T.J., W.K., B.P.K. and C.G. developed and ran Fluopol-ABPP HTS screen, completed initial activator characterization and supported SAR studies. S.G. and S.C. carried out medicinal chemistry and SAR-guided synthesis. B.P.K., W.K., T.J., A.S., S.M.K. and C.G. ran ABPP assays, mutagenesis experiments and protein analyses. S.K. performed nanoDSF assays. B.P.K., T.J. and A.S. performed substrate hydrolysis assays with help from J.X. J.E. ran MD simulations. D.O. carried out proteomic studies. M.C. performed in vitro ADME studies. B.P.K., A.G., A.S., S.M.K. and J.M.B. performed mouse studies and cell-based assays. All authors analyzed the data. D.L.B. and E.S. wrote the manuscript with input from other authors.

Corresponding authors

Correspondence to Dale L. Boger or Enrique Saez.

Ethics declarations

Competing interests

The authors declare no competing interests.

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 Screen to identify modulators of LYPLAL1 activity.

(a) ABPP-based fluorescence polarization assay (Fluopol-ABPP) amenable for high-throughput screening. Purified mLYPLAL1 is incubated with compounds prior to addition of the FP-rhodamine (FP-Rh) probe. FP-biotin probe (a non-fluorescent FP probe, 25 μM) serves as a control for inhibition, while the PPARγ ligand rosiglitazone (25 μM) is used as an inactive control. Data are shown as mean ± s.d., n = 12, where n represents independent samples. (b) A 16,000-compound screen (single point, 10 μM) identified hits that reduced the fluorescence polarization signal of FP-rhodamine labeling of mLYPLAL1 (putative inhibitors) and some that increased it (potential activators). Percent mLYPLAL1 activity was calculated relative to DMSO (100% activity) and FP-biotin (0% activity) wells. 4 is shown as a red dot.

Source data

Extended Data Fig. 2 Validation of primary HTS hits.

Candidate LYPLAL1 inhibitors and activators were analyzed by gel-based ABPP. Hit picks (10 μM) were incubated with 100 nM of purified mLYPLAL1 for 1.5 h at 37 °C prior to the addition of FP-rhodamine probe (1 μM). After 1 h at 37 °C, reactions were quenched, separated by SDS–PAGE and in-gel fluorescence scanned. LYPLAL1 activity was calculated relative to DMSO (100%). Designations above lanes correspond to compound location on the plate. Red asterisk denotes 4. Representative results from two independent experiments; similar results were obtained in both experiments. Uncropped gels are shown in Supplementary Fig. 6.

Source data

Extended Data Fig. 3 Derivatized thermal melt curves for compound-treated mLYPLAL1.

Temperature-dependent fluorescence shifts in purified mLYPLAL1 incubated with increasing concentrations of 4 and 12. Increased protein flexibility is noted upon compound interaction. In contrast, an inactive compound, 80, does not shift the derivatized thermal melt curve of mLYPLAL1. Representative results from three independent experiments; similar results were obtained in all experiments.

Source data

Extended Data Fig. 4 Evolution of the global Root Mean Square Deviation (RMSD) during Molecular Dynamics (MD) trajectories.

(a) WT hLYPLAL1 and R80 mutants, and (b) WT hLYPLAL1 with chlorobenzene as cosolvent. The RMSD was calculated considering only the backbone atoms and using the starting frames as the reference structures. Results from 10 independent simulations.

Source data

Extended Data Fig. 5 Effects of chronic treatment of DIO mice with a pharmacological LYPLAL1 activator.

DIO mice were treated with 12 for 28 days. (a) Body weights (left panel: n = 8 for veh, n = 9 for 12; right panel: n = 7 for veh, 12 and 12 + C11 groups, n = 8 for C11) prior to OGTT (left panel: d 11, right panel: d 8) and ITT (left panel: d 15, right panel: d 12); cumulative food intake per mouse (left panel: n = 3 cages for veh and 12, right panel: n = 2 cages for all groups) of vehicle and compound-treated DIO mice. (b) Indicated tissues from DIO mice were analyzed by western blot. The intensity of the LYPLAL1 signal was normalized to that of HSP90, the loading control. Quantification is shown on the right; n = 8 for veh and 7 for 12. n represents individual mice (body weights and western blot) or cages (food intake). Error bars represent mean ± s.e.m. Representative results from two independent experiments; similar results were obtained in both experiments. Uncropped blots for b are shown in Supplementary Fig. 6.

Source data

Extended Data Fig. 6 Glucose production in primary mouse hepatocytes treated with LYPLAL1 modulators.

Basal (20 mM lactate/2 mM pyruvate) (left panel) and glucagon-stimulated (middle panel) glucose production was measured in primary mouse hepatocytes treated for 6 h with 12, C11, or both compounds in combination; results were normalized to protein content. Gel-based ABPP analysis of 50 nM purified mLYPLAL1 treated for 1 h with 12, C11, and the combination (right panel). Protein levels are shown in the bottom panel. Error bars represent mean ± s.d.; n = 5 per group where n=biologically independent samples; *p = 0.02, **p = 0.003, ***p = 0.0003 by two-tailed t-test. Representative results from two independent experiments; similar results were obtained in each experiment. Uncropped gels/blots are shown in Supplementary Fig. 6.

Source data

Supplementary information

Supplementary information

Supplementary Tables 1–8, Figs. 1–6 and Notes.

Reporting Summary

Source data

Source data Fig. 2

Statistical source data.

Source data Fig. 2

Unprocessed gels.

Source data Fig. 4

Statistical source data.

Source data Fig. 4

Unprocessed gels and western blots.

Source data Fig. 5

Statistical source data.

Source data Fig. 5

Unprocessed gels and western blots.

Source data Fig. 6

Statistical source data.

Source data Fig. 6

Unprocessed gels.

Extended Data Fig. 1

Statistical source data.

Extended Data Fig. 2

Unprocessed gels.

Extended Data Fig. 3

Statistical source data.

Extended Data Fig. 4

Statistical source data.

Extended Data Fig. 5

Statistical source data.

Extended Data Fig. 5

Unprocessed western blots.

Extended Data Fig. 6

Statistical source data.

Extended Data Fig. 6

Unprocessed gels.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kok, B.P., Ghimire, S., Kim, W. et al. Discovery of small-molecule enzyme activators by activity-based protein profiling. Nat Chem Biol 16, 997–1005 (2020). https://doi.org/10.1038/s41589-020-0555-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-020-0555-4

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