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.

  • Review Article
  • Published:

Discovering new biology with drug-resistance alleles

Abstract

Small molecule drugs form the backbone of modern medicine’s therapeutic arsenal. Often less appreciated is the role that small molecules have had in advancing basic biology. In this Review, we highlight how resistance mutations have unlocked the potential of small molecule chemical probes to discover new biology. We describe key instances in which resistance mutations and related genetic variants yielded foundational biological insight and categorize these examples on the basis of their role in the discovery of novel molecular mechanisms, protein allostery, physiology and cell signaling. Next, we suggest ways in which emerging technologies can be leveraged to systematically introduce and characterize resistance mutations to catalyze basic biology research and drug discovery. By recognizing how resistance mutations have propelled biological discovery, we can better harness new technologies and maximize the potential of small molecules to advance our understanding of biology and improve human health.

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: Resistance mutations identify small molecule targets and reveal new biology.
Fig. 2: Resistance mutations unveil protein structural information and mechanisms of protein allosteric regulation.
Fig. 3: Resistance mutations elucidate physiology and provide insight into species differences in small molecule mechanism of action.
Fig. 4: Allele-selective kinase inhibitors and resistance mutations uncover the importance of BRAF dimerization in the MAPK signaling pathway.
Fig. 5: Resistance mutations uncover the mechanism of action of LSD1 inhibitors in AML.
Fig. 6: Mutagenesis strategies enhance resistance mutation identification and enable novel methods to investigate the small molecule–protein interface.

Similar content being viewed by others

References

  1. Tomasz, A. From penicillin-binding proteins to the lysis and death of bacteria: a 1979 view. Rev. Infect. Dis. 1, 434–467 (1979).

    Article  CAS  PubMed  Google Scholar 

  2. Borisy, G. G. & Taylor, E. W. The mechanism of action of colchicine. Binding of colchincine-3H to cellular protein. J. Cell Biol. 34, 525–533 (1967).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Dixon, S. J. et al. Ferroptosis: an iron-dependent form of nonapoptotic cell death. Cell 149, 1060–1072 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Schenone, M., Dančík, V., Wagner, B. K. & Clemons, P. A. Target identification and mechanism of action in chemical biology and drug discovery. Nat. Chem. Biol. 9, 232–240 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Shalem, O. et al. Genome-scale CRISPR–Cas9 knockout screening in human cells. Science 343, 84–87 (2014).

    Article  CAS  PubMed  Google Scholar 

  6. Wang, T., Wei, J. J., Sabatini, D. M. & Lander, E. S. Genetic screens in human cells using the CRISPR–Cas9 system. Science 343, 80–84 (2014).

    Article  CAS  PubMed  Google Scholar 

  7. Tochini-Valentini, G. P., Marino, P. & Colvill, A. J. Mutant of E. coli containing an altered DNA-dependent RNA polymerase. Nature 220, 275–276 (1968).

    Article  Google Scholar 

  8. Ezekial, D. H. & Hutchins, J. E. Mutations affecting RNA polymerase associated with rifampicin resistance in Escherichia coli. Nature 220, 276–277 (1968).

    Article  Google Scholar 

  9. Ozaki, M., Mizushima, S. & Nomura, M. Identification and functional characterization of the protein controlled by the streptomycin-resistant locus in E. coli. Nature 222, 333–339 (1969).

    Article  CAS  PubMed  Google Scholar 

  10. Branscomb, E. W. & Galas, D. J. Progressive decrease in protein synthesis accuracy induced by streptomycin in Escherichia coli. Nature 254, 161–163 (1975).

    Article  CAS  PubMed  Google Scholar 

  11. Galas, D. J. & Branscomb, E. W. Ribosome slowed by mutation to streptomycin resistance. Nature 262, 617–619 (1976). This study showed that streptomycin resistance mutations, which were known to increase the accuracy of translation, also led to decreased translational kinetics. This finding demonstrated that the accuracy of translation was in part dictated by the speed of translation.

    Article  CAS  PubMed  Google Scholar 

  12. Crick, F. H. C., Griffith, J. S. & Orgel, L. E. Codes without commas. Proc. Natl Acad. Sci. USA 43, 416–421 (1957).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Heitman, J., Movva, N. R. & Hall, M. N. Targets for cell cycle arrest by the immunosuppressant rapamycin in yeast. Science 253, 905–909 (1991). This paper reported the first identification of a TOR gene in any organism.

    Article  CAS  PubMed  Google Scholar 

  14. Harding, M. W., Galat, A., Uehlingt, D. E. & Schreibert, S. L. A receptor for the immuno-suppressant FK506 is a cistrans peptidyl-prolyl isomerase. Nature 341, 758–760 (1989).

    Article  CAS  PubMed  Google Scholar 

  15. Brown, E. J. et al. A mammalian protein targeted by G1-arresting rapamycin–receptor complex. Nature 369, 756–758 (1994).

    Article  CAS  PubMed  Google Scholar 

  16. Sabatini, D. M., Erdjument-Bromage, H., Lui, M., Tempst, P. & Snyder, S. H. RAFT1: a mammalian protein that binds to FKBP12 in a rapamycin-dependent fashion and is homologous to yeast TORs. Cell 78, 35–43 (1994).

    Article  CAS  PubMed  Google Scholar 

  17. Sabers, C. J. et al. Isolation of a protein target of the FKBP12–rapamycin complex in mammalian cells. J. Biol. Chem. 270, 815–822 (1995).

    Article  CAS  PubMed  Google Scholar 

  18. Liu, J. et al. Calcineurin is a common target of cyclophilin–cyclosporin A and FKBP–FK506 complexes. Cell 66, 807–815 (1991).

    Article  CAS  PubMed  Google Scholar 

  19. Schreiber, S. L. The rise of molecular glues. Cell 184, 3–9 (2021).

    Article  CAS  PubMed  Google Scholar 

  20. Kasap, C., Elemento, O. & Kapoor, T. M. DrugTargetSeqR: a genomics- and CRISPR–Cas9-based method to analyze drug targets. Nat. Chem. Biol. 10, 626–628 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Han, T. et al. Anticancer sulfonamides target splicing by inducing RBM39 degradation via recruitment to DCAF15. Science 356, eaal3755 (2017).

    Article  PubMed  CAS  Google Scholar 

  22. Uehara, T. et al. Selective degradation of splicing factor CAPERα by anticancer sulfonamides. Nat. Chem. Biol. 13, 675–680 (2017).

    Article  CAS  PubMed  Google Scholar 

  23. Noda, M., Suzuki, H., Numa, S. & Stiihmer, W. A single point mutation confers tetrodotoxin and saxitoxin insensitivity on the sodium channel II. FEBS Lett. 259, 213–216 (1989).

    Article  CAS  PubMed  Google Scholar 

  24. Terlau, H. et al. Mapping the site of block by tetrodotoxin and saxitoxin of sodium channel H. FEBS Lett. 293, 93–96 (1991).

    Article  CAS  PubMed  Google Scholar 

  25. Heinemann, S. H., Terlau, H., Stuhmer, W., Imoto, K. & Numa, S. Calcium channel characteristics conferred on the sodium channel by single mutations. Nature 356, 441–443 (1992).

    Article  CAS  PubMed  Google Scholar 

  26. Gerber, H. et al. RNA polymerase II C-terminal domain required for enhancer-driven transcription. Nature 374, 660–662 (1995).

    Article  CAS  PubMed  Google Scholar 

  27. Bartolomei, M. S., Halden, N. F., Cullen, C. R. & Corden, J. L. Genetic analysis of the repetitive carboxyl-terminal domain of the largest subunit of mouse RNA polymerase II. Mol. Cell. Biol. 8, 330–339 (1988).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Azam, M., Latek, R. R. & Daley, G. Q. Mechanisms of autoinhibition and STI-571/imatinib resistance revealed by mutagenesis of BCR-ABL. Cell 112, 831–843 (2003). This report identified a novel mechanism of autoinhibition for ABL kinase through characterization of imatinib resistance mutations outside of the drug-binding site.

    Article  CAS  PubMed  Google Scholar 

  29. Nardi, V., Azam, M. & Daley, G. Q. Mechanisms and implications of imatinib resistance mutations in BCR-ABL. Curr. Opin. Hematol. 11, 35–43 (2004).

    Article  CAS  PubMed  Google Scholar 

  30. Nagar, B. et al. Structural basis for the autoinhibition of c-Abl tyrosine kinase. Cell 112, 859–871 (2003).

    Article  CAS  PubMed  Google Scholar 

  31. Hantschel, O. et al. A myristoyl/phosphotyrosine switch regulates c-Abl. Cell 112, 845–857 (2003).

    Article  CAS  PubMed  Google Scholar 

  32. Adrián, F. J. et al. Allosteric inhibitors of Bcr-Abl-dependent cell proliferation. Nat. Chem. Biol. 2, 95–102 (2006).

    Article  PubMed  CAS  Google Scholar 

  33. Wylie, A. A. et al. The allosteric inhibitor ABL001 enables dual targeting of BCR-ABL1. Nature 543, 733–737 (2017).

    Article  CAS  PubMed  Google Scholar 

  34. Melnikov, A., Rogov, P., Wang, L., Gnirke, A. & Mikkelsen, T. S. Comprehensive mutational scanning of a kinase in vivo reveals substrate-dependent fitness landscapes. Nucleic Acids Res. 42, e112 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Persky, N. S. et al. Defining the landscape of ATP-competitive inhibitor resistance residues in protein kinases. Nat. Struct. Mol. Biol. 27, 92–104 (2020).

    Article  CAS  PubMed  Google Scholar 

  36. Rudolph, U. et al. Benzodiazepine actions mediated by specific γ-aminobutyric acidA receptor subtypes. Nature 401, 796–800 (1999). Using mouse models with resistance mutations in specific GABA receptor subtypes, this report identified the specific GABA receptor subtypes and thus brain regions that mediated different aspects of the physiological response to benzodiazepines.

    Article  CAS  PubMed  Google Scholar 

  37. McKernan, R. M. et al. Sedative but not anxiolytic properties of benzodiazepines are mediated by the GABAA receptor α1 subtype. Nat. Neurosci. 3, 587–592 (2000).

    Article  CAS  PubMed  Google Scholar 

  38. Low, K. et al. Molecular and neuronal substrate for the selective attenuation of anxiety. Science 290, 131–134 (2000).

    Article  CAS  PubMed  Google Scholar 

  39. Chen, X., Gerven, J., van, Cohen, A. & Jacobs, G. Human pharmacology of positive GABA-A subtype-selective receptor modulators for the treatment of anxiety. Acta Pharm. Sin. 40, 571–582 (2019).

    Article  CAS  Google Scholar 

  40. Mallal, S. et al. Association between presence of HLA-B*5701, HLA-DR7, and HLA-DQ3 and hypersensitivity to HIV-1 reverse-transcriptase inhibitor abacavir. Lancet 359, 727–732 (2002).

    Article  CAS  PubMed  Google Scholar 

  41. Illing, P. T. et al. Immune self-reactivity triggered by drug-modified HLA–peptide repertoire. Nature 486, 554–558 (2012).

    Article  CAS  PubMed  Google Scholar 

  42. Mcbride, W. G. Thalidomide and congenital abnormalities. Lancet 278, 1358 (1961).

    Article  Google Scholar 

  43. Ito, T. et al. Identification of a primary target of thalidomide teratogenicity. Science 327, 1345–1350 (2010). This study identified CRBN as the target of thalidomide. CRBN resistance mutations were used to show that thalidomide binding to CRBN was necessary for thalidomide’s teratogenicity in zebrafish.

    Article  CAS  PubMed  Google Scholar 

  44. Wu, T. et al. Targeted protein degradation as a powerful research tool in basic biology and drug target discovery. Nat. Struct. Mol. Biol. 27, 605–614 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Fratta, I. D., Sigg, E. B., Maiorana, K. & Davies, S. Teratogenic effects of thalidomide in rabbits, rats, hamsters and mice. Toxicol. Appl. Pharmacol. 7, 268–286 (1965).

    Article  CAS  PubMed  Google Scholar 

  46. Krönke, J. et al. Lenalidomide induces ubiquitination and degradation of CK1α in del(5q) MDS. Nature 523, 183–188 (2015). This paper identified the genetic variation in CRBN responsible for the difference in the antineoplastic activity of thalidomide and its analogs between mice and humans.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. Donovan, K. A. et al. Thalidomide promotes degradation of SALL4, a transcription factor implicated in Duane radial ray syndrome. eLife 7, e38430 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  48. Matyskiela, M. E. et al. Crystal structure of the SALL4–pomalidomide–cereblon–DDB1 complex. Nat. Struct. Mol. Biol. 27, 319–322 (2020).

    Article  CAS  PubMed  Google Scholar 

  49. Fink, E. C. et al. CrbnI391V is sufficient to confer in vivo sensitivity to thalidomide and its derivatives in mice. Blood 132, 1535–1544 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Kohlhase, J. et al. Mutations at the SALL4 locus on chromosome 20 result in a range of clinically overlapping phenotypes, including Okihiro syndrome, Holt–Oram syndrome, acro-renal-ocular syndrome, and patients previously reported to represent thalidomide embryopathy. Hum. Mol. Genet. 40, 473–478 (2003).

    CAS  Google Scholar 

  51. Greek, R., Shanks, N. & Rice, M. J. The history and implications of testing thalidomide on animals. J. Philos. Sci. Law 11, 1–32 (2011).

    Article  Google Scholar 

  52. Tsai, J. et al. Discovery of a selective inhibitor of oncogenic B-Raf kinase with potent antimelanoma activity. Proc. Natl Acad. Sci. USA 105, 3041–3046 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Bollag, G. et al. Clinical efficacy of a RAF inhibitor needs broad target blockade in BRAF-mutant melanoma. Nature 467, 596–599 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Dummer, R. et al. AZD6244 (ARRY-142886) vs temozolomide (TMZ) in patients (pts) with advanced melanoma: an open-label, randomized, multicenter, phase II study. J. Clin. Oncol. 26, 9033 (2008).

    Article  Google Scholar 

  55. Karoulia, Z., Gavathiotis, E. & Poulikakos, P. I. New perspectives for targeting RAF kinase in human cancer. Nat. Rev. Cancer 17, 676–691 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Poulikakos, P. I., Zhang, C., Bollag, G., Shokat, K. M. & Rosen, N. RAF inhibitors transactivate RAF dimers and ERK signalling in cells with wild-type BRAF. Nature 464, 427–430 (2010). This study used allele-selective kinase inhibitors to identify the mechanism underlying paradoxical activation by canonical BRAF inhibitors such as vemurafenib.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Hatzivassiliou, G. et al. RAF inhibitors prime wild-type RAF to activate the MAPK pathway and enhance growth. Nature 464, 431–435 (2010).

    Article  CAS  PubMed  Google Scholar 

  58. Bishop, A. C. et al. Generation of monospecific nanomolar tyrosine kinase inhibitors via a chemical genetic approach. J. Am. Chem. Soc. 121, 627–631 (1999).

    Article  CAS  Google Scholar 

  59. Bishop, A. C. et al. A chemical switch for inhibitor-sensitive alleles. Nature 408, 961–964 (2000).

    CAS  Google Scholar 

  60. Blair, J. A. et al. Structure-guided development of affinity probes for tyrosine kinases using chemical genetics. Nat. Chem. Biol. 3, 229–238 (2007).

    Article  CAS  PubMed  Google Scholar 

  61. Islam, K. The bump-and-hole tactic: expanding the scope of chemical genetics. Cell Chem. Biol. 25, 1171–1184 (2018). This review provides an excellent overview of bump–hole techniques and their applications across a variety of fields.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Rajakulendran, T., Sahmi, M., Lefrançois, M., Sicheri, F. & Therrien, M. A dimerization-dependent mechanism drives RAF catalytic activation. Nature 461, 542–545 (2009).

    Article  CAS  PubMed  Google Scholar 

  63. Karoulia, Z. et al. An integrated model of RAF inhibitor action predicts inhibitor activity against oncogenic BRAF signaling. Cancer Cell 30, 485–498 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Kondo, Y. et al. Cryo-EM structure of a dimeric B-Raf:14-3-3 complex reveals asymmetry in the active sites of B-Raf kinases. Science 366, 109–115 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Poulikakos, P. I. et al. RAF inhibitor resistance is mediated by dimerization of aberrantly spliced BRAFV600E. Nature 480, 387–390 (2011). This study identified and characterized the BRAF p61 isoform as resistant to canonical BRAF inhibitors such as vemurafenib. The BRAF p61 isoform both confirmed the mechanism of action of first-generation BRAF inhibitors such as vemurafenib and was used as a tool in later studies to identify next-generation therapeutic candidates targeting the MAPK pathway.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Corcoran, R. B. et al. EGFR-mediated reactivation of MAPK signaling contributes to insensitivity of BRAF-mutant colorectal cancers to RAF inhibition with vemurafenib. Cancer Discov. 2, 227–235 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Cotto-Rios, X. M. et al. Inhibitors of BRAF dimers using an allosteric site. Nat. Commun. 11, 4370 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Adamopoulos, C. et al. Exploiting allosteric properties of RAF and MEK inhibitors to target therapy-resistant tumors driven by oncogenic BRAF signaling. Cancer Discov. 11, 1716–1735 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Vinyard, M. et al. CRISPR-suppressor scanning reveals a nonenzymatic role of LSD1 in AML. Nat. Chem. Biol. 15, 529–539 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Maiques-Diaz, A. et al. Enhancer activation by pharmacologic displacement of LSD1 from GFI1 induces differentiation in acute myeloid leukemia. Cell Rep. 22, 3641–3659 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Hertz, N. T. et al. A neo-substrate that amplifies catalytic activity of Parkinson’s-disease- related kinase PINK1. Cell 154, 737–747 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Suh, J. L. et al. Discovery of selective activators of PRC2 mutant EED-I363M. Sci. Rep. 9, 6524 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  73. Słabicki, M. et al. The CDK inhibitor CR8 acts as a molecular glue degrader that depletes cyclin K. Nature 585, 293–297 (2020).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  74. Słabicki, M. et al. Small-molecule-induced polymerization triggers degradation of BCL6. Nature 588, 164–168 (2020).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  75. Hietpas, R., Roscoe, B., Jiang, L. & Bolon, D. N. A. Fitness analyses of all possible point mutations for regions of genes in yeast. Nat. Protoc. 7, 1382–1396 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Huang, Z. et al. A functional variomics tool for discovering drug-resistance genes and drug targets. Cell Rep. 3, 577–585 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Wu, T. J. et al. Identification of a non-Gatekeeper hot spot for drug-resistant mutations in mTOR kinase. Cell Rep. 11, 446–459 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Ting, T. C. et al. Aryl sulfonamides degrade RBM39 and RBM23 by recruitment to CRL4–DCAF15. Cell Rep. 29, 1499–1510 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Brenan, L. et al. Phenotypic characterization of a comprehensive set of MAPK1/ERK2 missense mutants. Cell Rep. 17, 1171–1183 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Sievers, Q. L. et al. Defining the human C2H2 zinc finger degrome targeted by thalidomide analogs through CRBN. Science 362, eaat0572 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  81. Giacomelli, A. O. et al. Mutational processes shape the landscape of TP53 mutations in human cancer. Nat. Genet. 50, 1381–1387 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Donovan, K. F. et al. Creation of novel protein variants with CRISPR/Cas9-mediated mutagenesis: turning a screening by-product into a discovery tool. PLoS ONE 12, e0170445 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  83. Ipsaro, J. J. et al. Rapid generation of drug-resistance alleles at endogenous loci using CRISPR–Cas9 indel mutagenesis. PLoS ONE 12, e0172177 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  84. Neggers, J. E. et al. Target identification of small molecules using large-scale CRISPR–Cas mutagenesis scanning of essential genes. Nat. Commun. 9, 502 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  85. Findlay, G. M., Boyle, E. A., Hause, R. J., Klein, J. C. & Shendure, J. Saturation editing of genomic regions by multiplex homology-directed repair. Nature 513, 120–123 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Zyryanova, A. F. et al. Binding of ISRIB reveals a regulatory site in the nucleotide exchange factor eIF2B. Science 359, 1533–1536 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Ma, L. et al. CRISPR–Cas9-mediated saturated mutagenesis screen predicts clinical drug resistance with improved accuracy. Proc. Natl Acad. Sci. USA 114, 11751–11756 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Komor, A. C., Kim, Y. B., Packer, M. S., Zuris, J. A. & Liu, D. R. Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage. Nature 533, 420–424 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Gaudelli, N. M. et al. Programmable base editing of A•T to G•C in genomic DNA without DNA cleavage. Nature 551, 464–471 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Ma, Y. et al. Targeted AID-mediated mutagenesis (TAM) enables efficient genomic diversification in mammalian cells. Nat. Methods 13, 1029–1035 (2016).

    Article  CAS  PubMed  Google Scholar 

  91. Hess, G. T. et al. Directed evolution using dCas9-targeted somatic hypermutation in mammalian cells. Nat. Methods 13, 1036–1042 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Moore, C. L., Papa, L. J. & Shoulders, M. D. A processive protein chimera introduces mutations across defined DNA regions in vivo. J. Am. Chem. Soc. 140, 11560–11564 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Chen, H. et al. Efficient, continuous mutagenesis in human cells using a pseudo-random DNA editor. Nat. Biotechnol. 38, 165–168 (2020).

    Article  CAS  PubMed  Google Scholar 

  94. Hanna, R. E. et al. Massively parallel assessment of human variants with base editor screens. Cell 184, 1064–1080 (2021).

    Article  CAS  PubMed  Google Scholar 

  95. Cuella-Martin, R. et al. Functional interrogation of DNA damage response variants with base editing screens. Cell 184, 1081–1097 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Anzalone, A. V. et al. Search-and-replace genome editing without double-strand breaks or donor DNA. Nature 576, 149–157 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Erwood, S. et al. Saturation variant interpretation using CRISPR prime editing. Preprint at bioRxiv https://doi.org/10.1101/2021.05.11.443710 (2021).

  98. Cupido, T., Pisa, R., Kelley, M. E. & Kapoor, T. M. Designing a chemical inhibitor for the AAA protein spastin using active site mutations. Nat. Chem. Biol. 15, 444–452 (2019). This report demonstrated the successful use of RADD to develop a selective inhibitor of the AAA protein spastin.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Pisa, R., Cupido, T., Steinman, J. B., Jones, N. H. & Kapoor, T. M. Analyzing resistance to design selective chemical inhibitors for AAA proteins. Cell Chem. Biol. 26, 1263–1273 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank members of the Liau laboratory, especially A. Siegenfeld, H. S. Kwok, P. Gosavi, N. Lue, S. Roseman, E. M. Garcia and Y. Koga, for discussions and comments on the manuscript. We also thank S. M. Kissler, A. Choudhary, J. Kim, M. D. Shair and D. Kahne for discussions. Figures were created with BioRender.com. A.M.F. was supported by award no. T32GM007753 from the National Institute of General Medical Sciences. This work was supported by award no. 1DP2GM137494 from the National Institute of General Medical Sciences.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Brian B. Liau.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Chemical Biology thanks Cory Johannessen, Jun Liu and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Freedy, A.M., Liau, B.B. Discovering new biology with drug-resistance alleles. Nat Chem Biol 17, 1219–1229 (2021). https://doi.org/10.1038/s41589-021-00865-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-021-00865-9

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