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:

WD40 repeat domain proteins: a novel target class?

Key Points

  • WD40 repeat (WDR) domains are β-propeller domains that act as protein interaction scaffolds in multiprotein complexes.

  • WDR domain-containing proteins comprise one of the largest protein families in humans and are involved in a diverse array of cellular networks, many of which are disease-associated.

  • The central cavity of the WDR domain is structurally dynamic and diverse. Potent, selective inhibitors target the central cavity of WD repeat-containing protein 5 (WDR5) and EED, two proteins involved in chromatin complexes. An EED inhibitor has entered clinical trials in oncology.

  • Targeting WDR domains that bind to other proteins that have so far proved undruggable, such as components of the ubiquitin–proteasome system, is a promising but underexplored strategy.

  • The cellular implications of pharmacologically targeting WDR domains that are involved in multiple protein complexes are still poorly understood.

Abstract

Antagonism of protein–protein interactions (PPIs) with small molecules is becoming more feasible as a therapeutic approach. Successful PPI inhibitors tend to target proteins containing deep peptide-binding grooves or pockets rather than the more common large, flat protein interaction surfaces. Here, we review one of the most abundant PPI domains in the human proteome, the WD40 repeat (WDR) domain, which has a central peptide-binding pocket and is a member of the β-propeller domain-containing protein family. Recently, two WDR domain-containing proteins, WDR5 and EED, as well as other β-propeller domains have been successfully targeted by potent, specific, cell-active, drug-like chemical probes. Could WDR domains be a novel target class for drug discovery? Although the research is at an early stage and therefore not clinically validated, cautious optimism is justified, as WDR domain-containing proteins are involved in multiple disease-associated pathways. The druggability and structural diversity of WDR domain binding pockets suggest that understanding how to target this prevalent domain class will open up areas of disease biology that have so far resisted drug discovery efforts.

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

Figure 1: Protein interaction modes of WD40 repeat domain β-propellers.
Figure 2: WD40 repeat domain-containing proteins perform diverse cellular functions.
Figure 3: Epigenetic chemical probes against WD40 repeat domain-containing proteins.
Figure 4: Exemplar inhibitors of WD40 repeat domains and related β-propeller domains.
Figure 5: Structural dynamics of WD40 repeat domain binding pockets.

Similar content being viewed by others

References

  1. Deeks, E. D. Venetoclax: first global approval. Drugs 76, 979–987 (2016).

    Article  CAS  PubMed  Google Scholar 

  2. Roberts, A. W. et al. Targeting BCL2 with venetoclax in relapsed chronic lymphocytic leukemia. N. Engl. J. Med. 374, 311–322 (2016).

    Article  CAS  PubMed  Google Scholar 

  3. Shangary, S. & Wang, S. Small-molecule inhibitors of the MDM2–p53 protein–protein interaction to reactivate p53 function: a novel approach for cancer therapy. Annu. Rev. Pharmacol. Toxicol. 49, 223–241 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Vassilev, L. T. et al. In vivo activation of the p53 pathway by small-molecule antagonists of MDM2. Science 303, 844–848 (2004).

    Article  CAS  PubMed  Google Scholar 

  5. Delmore, J. E. et al. BET bromodomain inhibition as a therapeutic strategy to target c-Myc. Cell 146, 904–917 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Filippakopoulos, P. et al. Selective inhibition of BET bromodomains. Nature 468, 1067–1073 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Shi, J. & Vakoc, C. R. The mechanisms behind the therapeutic activity of BET bromodomain inhibition. Mol. Cell 54, 728–736 (2014).

    Article  CAS  PubMed  Google Scholar 

  8. Scott, D. E., Bayly, A. R., Abell, C. & Skidmore, J. Small molecules, big targets: drug discovery faces the protein–protein interaction challenge. Nat. Rev. Drug Discov. 15, 533–550 (2016).

    Article  CAS  PubMed  Google Scholar 

  9. Niesen, F. H., Berglund, H. & Vedadi, M. The use of differential scanning fluorimetry to detect ligand interactions that promote protein stability. Nat. Protoc. 2, 2212–2221 (2007).

    Article  CAS  PubMed  Google Scholar 

  10. Annis, D. A., Nickbarg, E., Yang, X., Ziebell, M. R. & Whitehurst, C. E. Affinity selection-mass spectrometry screening techniques for small molecule drug discovery. Curr. Opin. Chem. Biol. 11, 518–526 (2007).

    Article  CAS  PubMed  Google Scholar 

  11. Clark, M. A. et al. Design, synthesis and selection of DNA-encoded small-molecule libraries. Nat. Chem. Biol. 5, 647–654 (2009).

    Article  CAS  PubMed  Google Scholar 

  12. Pearce, N. M. et al. A multi-crystal method for extracting obscured crystallographic states from conventionally uninterpretable electron density. Nat. Commun. 8, 15123 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Lea, W. A. & Simeonov, A. Fluorescence polarization assays in small molecule screening. Expert Opin. Drug Discov. 6, 17–32 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Juhasz, T., Szeltner, Z., Fulop, V. & Polgar, L. Unclosed β-propellers display stable structures: implications for substrate access to the active site of prolyl oligopeptidase. J. Mol. Biol. 346, 907–917 (2005).

    Article  CAS  PubMed  Google Scholar 

  15. Fulop, V., Bocskei, Z. & Polgar, L. Prolyl oligopeptidase: an unusual β-propeller domain regulates proteolysis. Cell 94, 161–170 (1998).

    Article  CAS  PubMed  Google Scholar 

  16. Stirnimann, C. U., Petsalaki, E., Russell, R. B. & Muller, C. W. WD40 proteins propel cellular networks. Trends Biochem. Sci. 35, 565–574 (2010). This article discusses the prevalence of WDR domains in protein networks and the structural arrangement of WDR domain-containing complexes.

    Article  CAS  PubMed  Google Scholar 

  17. Margueron, R. et al. Role of the polycomb protein EED in the propagation of repressive histone marks. Nature 461, 762–767 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Migliori, V. et al. Symmetric dimethylation of H3R2 is a newly identified histone mark that supports euchromatin maintenance. Nat. Struct. Mol. Biol. 19, 136–144 (2012).

    Article  CAS  PubMed  Google Scholar 

  19. Hao, B., Oehlmann, S., Sowa, M. E., Harper, J. W. & Pavletich, N. P. Structure of a Fbw7–Skp1–cyclin E complex: multisite-phosphorylated substrate recognition by SCF ubiquitin ligases. Mol. Cell 26, 131–143 (2007).

    Article  CAS  PubMed  Google Scholar 

  20. Orlicky, S., Tang, X., Willems, A., Tyers, M. & Sicheri, F. Structural basis for phosphodependent substrate selection and orientation by the SCFCdc4 ubiquitin ligase. Cell 112, 243–256 (2003).

    Article  CAS  PubMed  Google Scholar 

  21. Wang, Y. et al. WDSPdb: a database for WD40-repeat proteins. Nucleic Acids Res. 43, D339–D344 (2015).

    Article  CAS  PubMed  Google Scholar 

  22. Wang, Y., Jiang, F., Zhuo, Z., Wu, X. H. & Wu, Y. D. A method for WD40 repeat detection and secondary structure prediction. PLoS ONE 8, e65705 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Chen, C. K., Chan, N. L. & Wang, A. H. The many blades of the β-propeller proteins: conserved but versatile. Trends Biochem. Sci. 36, 553–561 (2011).

    Article  CAS  PubMed  Google Scholar 

  24. Croft, D. et al. The Reactome pathway knowledgebase. Nucleic Acids Res. 42, D472–D477 (2014).

    Article  CAS  PubMed  Google Scholar 

  25. Grebien, F. et al. Pharmacological targeting of the Wdr5–MLL interaction in C/EBPα N-terminal leukemia. Nat. Chem. Biol. 11, 571–578 (2015). This article reports on a nanomolar inhibitor of a WDR domain.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. He, Y. et al. The EED protein-protein interaction inhibitor A-395 inactivates the PRC2 complex. Nat. Chem. Biol. 13, 389–395 (2017).

    Article  CAS  PubMed  Google Scholar 

  27. Qi, W. et al. An allosteric PRC2 inhibitor targeting the H3K27me3 binding pocket of EED. Nat. Chem. Biol. 13, 381–388 (2017). Refs 26 and 27 describe inhibitors that target a WDR domain of PRC2 and are active against cancer cells that are resistant to catalytic inhibitors.

    Article  CAS  PubMed  Google Scholar 

  28. Davis, R. J., Welcker, M. & Clurman, B. E. Tumor suppression by the Fbw7 ubiquitin ligase: mechanisms and opportunities. Cancer Cell 26, 455–464 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Laskowski, R. A. et al. Integrating population variation and protein structural analysis to improve clinical interpretation of missense variation: application to the WD40 domain. Hum. Mol. Genet. 25, 927–935 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Dawson, M. A. The cancer epigenome: concepts, challenges, and therapeutic opportunities. Science 355, 1147–1152 (2017).

    Article  CAS  PubMed  Google Scholar 

  31. Pfister, S. X. & Ashworth, A. Marked for death: targeting epigenetic changes in cancer. Nat. Rev. Drug Discov. 16, 241–263 (2017).

    Article  CAS  PubMed  Google Scholar 

  32. Shortt, J., Ott, C. J., Johnstone, R. W. & Bradner, J. E. A chemical probe toolbox for dissecting the cancer epigenome. Nat. Rev. Cancer 17, 160–183 (2017).

    Article  CAS  PubMed  Google Scholar 

  33. Arrowsmith, C. H., Bountra, C., Fish, P. V., Lee, K. & Schapira, M. Epigenetic protein families: a new frontier for drug discovery. Nat. Rev. Drug Discov. 11, 384–400 (2012).

    Article  CAS  PubMed  Google Scholar 

  34. Zhu, J. et al. Gain-of-function p53 mutants co-opt chromatin pathways to drive cancer growth. Nature 525, 206–211 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Thiel, A. T. et al. MLL–AF9-induced leukemogenesis requires coexpression of the wild-type Mll allele. Cancer Cell 17, 148–159 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Cao, F. et al. Targeting MLL1 H3K4 methyltransferase activity in mixed-lineage leukemia. Mol. Cell 53, 247–261 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Li, Y. et al. Structural basis for activity regulation of MLL family methyltransferases. Nature 530, 447–452 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Garapaty-Rao, S. et al. Identification of EZH2 and EZH1 small molecule inhibitors with selective impact on diffuse large B cell lymphoma cell growth. Chem. Biol. 20, 1329–1339 (2013).

    Article  CAS  PubMed  Google Scholar 

  39. McCabe, M. T. et al. EZH2 inhibition as a therapeutic strategy for lymphoma with EZH2-activating mutations. Nature 492, 108–112 (2012).

    Article  CAS  PubMed  Google Scholar 

  40. Konze, K. D. et al. An orally bioavailable chemical probe of the lysine methyltransferases EZH2 and EZH1. ACS Chem. Biol. 8, 1324–1334 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Knutson, S. K. et al. A selective inhibitor of EZH2 blocks H3K27 methylation and kills mutant lymphoma cells. Nat. Chem. Biol. 8, 890–896 (2012).

    Article  CAS  PubMed  Google Scholar 

  42. Kim, K. H. et al. SWI/SNF-mutant cancers depend on catalytic and non-catalytic activity of EZH2. Nat. Med. 21, 1491–1496 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Kim, W. et al. Targeted disruption of the EZH2–EED complex inhibits EZH2-dependent cancer. Nat. Chem. Biol. 9, 643–650 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Wolffe, A. P., Urnov, F. D. & Guschin, D. Co-repressor complexes and remodelling chromatin for repression. Biochem. Soc. Trans. 28, 379–386 (2000).

    Article  CAS  PubMed  Google Scholar 

  45. Kitange, G. J. et al. Retinoblastoma binding protein 4 modulates temozolomide sensitivity in glioblastoma by regulating DNA repair proteins. Cell Rep. 14, 2587–2598 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Chan-Penebre, E. et al. A selective inhibitor of PRMT5 with in vivo and in vitro potency in MCL models. Nat. Chem. Biol. 11, 432–437 (2015).

    Article  CAS  PubMed  Google Scholar 

  47. Kryukov, G. V. et al. MTAP deletion confers enhanced dependency on the PRMT5 arginine methyltransferase in cancer cells. Science 351, 1214–1218 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Mavrakis, K. J. et al. Disordered methionine metabolism in MTAP/CDKN2A-deleted cancers leads to dependence on PRMT5. Science 351, 1208–1213 (2016).

    Article  CAS  PubMed  Google Scholar 

  49. Chen, H., Lorton, B., Gupta, V. & Shechter, D. A. TGFβ–PRMT5–MEP50 axis regulates cancer cell invasion through histone H3 and H4 arginine methylation coupled transcriptional activation and repression. Oncogene 36, 373–386 (2017).

    Article  CAS  PubMed  Google Scholar 

  50. Peters, J. M. The anaphase promoting complex/cyclosome: a machine designed to destroy. Nat. Rev. Mol. Cell Biol. 7, 644–656 (2006).

    Article  CAS  PubMed  Google Scholar 

  51. Kidokoro, T. et al. CDC20, a potential cancer therapeutic target, is negatively regulated by p53. Oncogene 27, 1562–1571 (2008).

    Article  CAS  PubMed  Google Scholar 

  52. Mao, D. D. et al. A CDC20–APC/SOX2 signaling axis regulates human glioblastoma stem-like cells. Cell Rep. 11, 1809–1821 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Xie, Q. et al. CDC20 maintains tumor initiating cells. Oncotarget 6, 13241–13254 (2015).

    PubMed  PubMed Central  Google Scholar 

  54. Marucci, G. et al. Gene expression profiling in glioblastoma and immunohistochemical evaluation of IGFBP-2 and CDC20. Virchows Arch. 453, 599–609 (2008).

    Article  CAS  PubMed  Google Scholar 

  55. Van Roosbroeck, K. et al. JAK2 rearrangements, including the novel SEC31A–JAK2 fusion, are recurrent in classical Hodgkin lymphoma. Blood 117, 4056–4064 (2011).

    Article  CAS  PubMed  Google Scholar 

  56. Bedwell, C. et al. Cytogenetically complex SEC31A–ALK fusions are recurrent in ALK-positive large B-cell lymphomas. Haematologica 96, 343–346 (2011).

    Article  CAS  PubMed  Google Scholar 

  57. Van Roosbroeck, K. et al. ALK-positive large B-cell lymphomas with cryptic SEC31A–ALK and NPM1–ALK fusions. Haematologica 95, 509–513 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Xia, B. et al. Fanconi anemia is associated with a defect in the BRCA2 partner PALB2. Nat. Genet. 39, 159–161 (2007).

    Article  CAS  PubMed  Google Scholar 

  59. Reid, S. et al. Biallelic mutations in PALB2 cause Fanconi anemia subtype FA-N and predispose to childhood cancer. Nat. Genet. 39, 162–164 (2007).

    Article  CAS  PubMed  Google Scholar 

  60. Rahman, N. et al. PALB2, which encodes a BRCA2-interacting protein, is a breast cancer susceptibility gene. Nat. Genet. 39, 165–167 (2007).

    Article  CAS  PubMed  Google Scholar 

  61. Xia, B. et al. Control of BRCA2 cellular and clinical functions by a nuclear partner, PALB2. Mol. Cell 22, 719–729 (2006).

    Article  CAS  PubMed  Google Scholar 

  62. Erkko, H. et al. A recurrent mutation in PALB2 in Finnish cancer families. Nature 446, 316–319 (2007).

    Article  CAS  PubMed  Google Scholar 

  63. Smith, M. A. et al. Initial testing (stage 1) of the PARP inhibitor BMN 673 by the pediatric preclinical testing program: PALB2 mutation predicts exceptional in vivo response to BMN 673. Pediatr. Blood Cancer 62, 91–98 (2015).

    Article  CAS  PubMed  Google Scholar 

  64. Martini, E., Roche, D. M., Marheineke, K., Verreault, A. & Almouzni, G. Recruitment of phosphorylated chromatin assembly factor 1 to chromatin after UV irradiation of human cells. J. Cell Biol. 143, 563–575 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Cheloufi, S. et al. The histone chaperone CAF-1 safeguards somatic cell identity. Nature 528, 218–224 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Zhang, H. et al. MLL1 inhibition reprograms epiblast stem cells to naive pluripotency. Cell Stem Cell 18, 481–494 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Tanaka, T. et al. Lis1 and doublecortin function with dynein to mediate coupling of the nucleus to the centrosome in neuronal migration. J. Cell Biol. 165, 709–721 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Haverfield, E. V., Whited, A. J., Petras, K. S., Dobyns, W. B. & Das, S. Intragenic deletions and duplications of the LIS1 and DCX genes: a major disease-causing mechanism in lissencephaly and subcortical band heterotopia. Eur. J. Hum. Genet. 17, 911–918 (2009).

    Article  CAS  PubMed  Google Scholar 

  69. Reiner, O. et al. Isolation of a Miller-Dieker lissencephaly gene containing G protein β-subunit-like repeats. Nature 364, 717–721 (1993).

    Article  CAS  PubMed  Google Scholar 

  70. Bi, W. et al. Increased LIS1 expression affects human and mouse brain development. Nat. Genet. 41, 168–177 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Healy, D. G. et al. Phenotype, genotype, and worldwide genetic penetrance of LRRK2-associated Parkinson's disease: a case-control study. Lancet Neurol. 7, 583–590 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Smith, W. W. et al. Kinase activity of mutant LRRK2 mediates neuronal toxicity. Nat. Neurosci. 9, 1231–1233 (2006).

    Article  CAS  PubMed  Google Scholar 

  73. West, A. B. et al. Parkinson's disease-associated mutations in leucine-rich repeat kinase 2 augment kinase activity. Proc. Natl Acad. Sci. USA 102, 16842–16847 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Li, T. et al. Novel LRRK2 GTP-binding inhibitors reduced degeneration in Parkinson's disease cell and mouse models. Hum. Mol. Genet. 23, 6212–6222 (2014).

    Article  CAS  PubMed  Google Scholar 

  75. Deng, X. et al. Characterization of a selective inhibitor of the Parkinson's disease kinase LRRK2. Nat. Chem. Biol. 7, 203–205 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Lee, B. D. et al. Inhibitors of leucine-rich repeat kinase-2 protect against models of Parkinson's disease. Nat. Med. 16, 998–1000 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Kett, L. R. et al. LRRK2 Parkinson disease mutations enhance its microtubule association. Hum. Mol. Genet. 21, 890–899 (2012).

    Article  CAS  PubMed  Google Scholar 

  78. Jorgensen, N. D. et al.The WD40 domain is required for LRRK2 neurotoxicity. PLoS ONE 4, e8463 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Klenke, S., Kussmann, M. & Siffert, W. The GNB3 C825T polymorphism as a pharmacogenetic marker in the treatment of hypertension, obesity, and depression. Pharmacogenet. Genomics 21, 594–606 (2011).

    Article  CAS  PubMed  Google Scholar 

  80. Siffert, W. et al. Association of a human G-protein β3 subunit variant with hypertension. Nat. Genet. 18, 45–48 (1998).

    Article  CAS  PubMed  Google Scholar 

  81. Goldlust, I. S. et al. Mouse model implicates GNB3 duplication in a childhood obesity syndrome. Proc. Natl Acad. Sci. USA 110, 14990–14994 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Nilsson, J., Sengupta, J., Frank, J. & Nissen, P. Regulation of eukaryotic translation by the RACK1 protein: a platform for signalling molecules on the ribosome. EMBO Rep. 5, 1137–1141 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. McCahill, A., Warwicker, J., Bolger, G. B., Houslay, M. D. & Yarwood, S. J. The RACK1 scaffold protein: a dynamic cog in cell response mechanisms. Mol. Pharmacol. 62, 1261–1273 (2002).

    Article  CAS  PubMed  Google Scholar 

  84. Boyle, W. J., Simonet, W. S. & Lacey, D. L. Osteoclast differentiation and activation. Nature 423, 337–342 (2003).

    Article  CAS  PubMed  Google Scholar 

  85. Lin, J., Lee, D., Choi, Y. & Lee, S. Y. The scaffold protein RACK1 mediates the RANKL-dependent activation of p38 MAPK in osteoclast precursors. Sci. Signal 8, ra54 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Majzoub, K. et al. RACK1 controls IRES-mediated translation of viruses. Cell 159, 1086–1095 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Kim, K. et al. VprBP has intrinsic kinase activity targeting histone H2A and represses gene transcription. Mol. Cell 52, 459–467 (2013).

    Article  CAS  PubMed  Google Scholar 

  88. Angers, S. et al. Molecular architecture and assembly of the DDB1–CUL4A ubiquitin ligase machinery. Nature 443, 590–593 (2006).

    Article  CAS  PubMed  Google Scholar 

  89. Wu, Y. et al. The DDB1–DCAF1–Vpr–UNG2 crystal structure reveals how HIV-1 Vpr steers human UNG2 toward destruction. Nat. Struct. Mol. Biol. 23, 933–940 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Schwefel, D. et al. Structural basis of lentiviral subversion of a cellular protein degradation pathway. Nature 505, 234–238 (2014).

    Article  CAS  PubMed  Google Scholar 

  91. Li, T., Chen, X., Garbutt, K. C., Zhou, P. & Zheng, N. Structure of DDB1 in complex with a paramyxovirus V protein: viral hijack of a propeller cluster in ubiquitin ligase. Cell 124, 105–117 (2006).

    Article  CAS  PubMed  Google Scholar 

  92. Orlicky, S. et al. An allosteric inhibitor of substrate recognition by the SCFCdc4 ubiquitin ligase. Nat. Biotechnol. 28, 733–737 (2010). The inhibitor in this article binds at a side pocket of a WDR domain and allosterically occludes the central cavity.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Sackton, K. L. et al. Synergistic blockade of mitotic exit by two chemical inhibitors of the APC/C. Nature 514, 646–649 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Bolshan, Y. et al. Synthesis, optimization, and evaluation of novel small molecules as antagonists of WDR5–MLL interaction. ACS Med. Chem. Lett. 4, 353–357 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Getlik, M. et al. Structure-based optimization of a small molecule antagonist of the interaction between WD repeat-containing protein 5 (WDR5) and mixed-lineage leukemia 1 (MLL1). J. Med. Chem. 59, 2478–2496 (2016).

    Article  CAS  PubMed  Google Scholar 

  96. Senisterra, G. et al. Small-molecule inhibition of MLL activity by disruption of its interaction with WDR5. Biochem. J. 449, 151–159 (2013).

    Article  CAS  PubMed  Google Scholar 

  97. Curtin, M. L. et al. SAR of amino pyrrolidines as potent and novel protein-protein interaction inhibitors of the PRC2 complex through EED binding. Bioorg. Med. Chem. Lett. 27, 1576–1583 (2017).

    Article  CAS  PubMed  Google Scholar 

  98. Huang, Y. et al. Discovery of first-in-class, potent, and orally bioavailable embryonic ectoderm development (EED) inhibitor with robust anticancer efficacy. J. Med. Chem. 60, 2215–2226 (2017).

    Article  CAS  PubMed  Google Scholar 

  99. Li, L. et al. Discovery and molecular basis of a diverse set of polycomb repressive complex 2 inhibitors recognition by EED. PLoS ONE 12, e0169855 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Csizmok, V. et al. An allosteric conduit facilitates dynamic multisite substrate recognition by the SCFCdc4 ubiquitin ligase. Nat. Commun. 8, 13943 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Davies, T. G. et al. Monoacidic inhibitors of the kelch-like ech-associated protein 1: nuclear factor erythroid 2-related factor 2 (KEAP1:NRF2) protein–protein interaction with high cell potency identified by fragment-based discovery. J. Med. Chem. 59, 3991–4006 (2016). This article describes a fragment-based approach to target the central cavity of a β-propeller domain.

    Article  CAS  PubMed  Google Scholar 

  102. Kerry, P. S. et al. Structural basis for a class of nanomolar influenza A neuraminidase inhibitors. Sci. Rep. 3, 2871 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  103. von Kleist, L. et al. Role of the clathrin terminal domain in regulating coated pit dynamics revealed by small molecule inhibition. Cell 146, 471–484 (2011).

    Article  CAS  PubMed  Google Scholar 

  104. Korczynska, M., Mukhtar, T. A., Wright, G. D. & Berghuis, A. M. Structural basis for streptogramin B resistance in Staphylococcus aureus by virginiamycin B lyase. Proc. Natl Acad. Sci. USA 104, 10388–10393 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Springer, T. A., Zhu, J. & Xiao, T. Structural basis for distinctive recognition of fibrinogen γC peptide by the platelet integrin αIIbβ3 . J. Cell Biol. 182, 791–800 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Tan, X. et al. Mechanism of auxin perception by the TIR1 ubiquitin ligase. Nature 446, 640–645 (2007). This study reports on a small molecule that activates a β-propeller domain.

    Article  CAS  PubMed  Google Scholar 

  107. Sheard, L. B. et al. Jasmonate perception by inositol-phosphate-potentiated COI1–JAZ co-receptor. Nature 468, 400–405 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Avdic, V. et al. Structural and biochemical insights into MLL1 core complex assembly. Structure 19, 101–108 (2011).

    Article  CAS  PubMed  Google Scholar 

  109. Baker, T. et al. Acquisition of a single EZH2 D1 domain mutation confers acquired resistance to EZH2-targeted inhibitors. Oncotarget 6, 32646–32655 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  110. Gibaja, V. et al. Development of secondary mutations in wild-type and mutant EZH2 alleles cooperates to confer resistance to EZH2 inhibitors. Oncogene 35, 558–566 (2016).

    Article  CAS  PubMed  Google Scholar 

  111. Edwards, A. M. et al. Preclinical target validation using patient-derived cells. Nat. Rev. Drug Discov. 14, 149–150 (2015).

    Article  CAS  PubMed  Google Scholar 

  112. Bunnage, M. E., Chekler, E. L. & Jones, L. H. Target validation using chemical probes. Nat. Chem. Biol. 9, 195–199 (2013).

    Article  CAS  PubMed  Google Scholar 

  113. Frye, S. V. The art of the chemical probe. Nat. Chem. Biol. 6, 159–161 (2010).

    Article  CAS  PubMed  Google Scholar 

  114. Kronke, J. et al. Lenalidomide causes selective degradation of IKZF1 and IKZF3 in multiple myeloma cells. Science 343, 301–305 (2014).

    Article  CAS  PubMed  Google Scholar 

  115. Lu, G. et al. The myeloma drug lenalidomide promotes the cereblon-dependent destruction of Ikaros proteins. Science 343, 305–309 (2014).

    Article  CAS  PubMed  Google Scholar 

  116. Winter, G. E. et al. Phthalimide conjugation as a strategy for in vivo target protein degradation. Science 348, 1376–1381 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Sakamoto, K. M. et al. Protacs: chimeric molecules that target proteins to the Skp1–Cullin-F box complex for ubiquitination and degradation. Proc. Natl Acad. Sci. USA 98, 8554–8559 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Saenz, D. T. et al. Novel BET protein proteolysis-targeting chimera exerts superior lethal activity than bromodomain inhibitor (BETi) against post-myeloproliferative neoplasm secondary (s) AML cells. Leukemia http://dx.doi.org/10.1038/leu.2016.393 (2017).

  119. Uehara, T. et al. Selective degradation of splicing factor CAPERα by anticancer sulfonamides. Nat. Chem. Biol. (2017).

  120. Han, T. et al. Anticancer sulfonamides target splicing by inducing RBM39 degradation via recruitment to DCAF15. Science 356, eaal3755 (2017). Refs 119 and 120 describe a WDR-containing protein as the target of a Protac small molecule, resulting in ubiquitylation and degradation of an oncoprotein.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The Structural Genomics Consortium (SGC) is a registered charity (number 1097737) that receives funds from AbbVie, Bayer Pharma, Boehringer Ingelheim, Canada Foundation for Innovation, Eshelman Institute for Innovation, Genome Canada through Ontario Genomics Institute (OGI-055), Innovative Medicines Initiative (EU/EFPIA) (ULTRA-DD grant no. 115766), Janssen, Merck & Co., Novartis Pharma AG, Ontario Ministry of Research, Innovation and Science (MRIS), Pfizer, São Paulo Research Foundation-FAPESP, Takeda and the Wellcome Trust. C.H.A. is supported by the Canadian Institute for Health Research (CIHR) (FRN-125792) and holds a Canada Research Chair in Structural Genomics. M.Ty. is supported by grants from the CIHR (MOP 126129), the Canadian Cancer Society Research Institute (703906), Genome Canada, Genome Quebec and National Institutes of Health (R01OD010929), and holds a Canada Research Chair in Systems and Synthetic Biology. The authors acknowledge Z.D. Wang for his contribution in generating the list of 361 WDR-containing proteins and C. Jakob for her careful review of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Matthieu Schapira.

Ethics declarations

Competing interests

M.To. is an employee of AbbVie and owns AbbVie stock. M.S., M.Ty. and C.H.A. declare no competing interests.

Supplementary information

Supplementary information

Supplementary Tables (PDF 2958 kb)

PowerPoint slides

Glossary

Thermal stabilization

A thermodynamic property used to screen chemical libraries, and which relies on the fact that the melting point of a protein often increases upon the binding of a small molecule.

Affinity selection–mass spectrometry

A screening technique whereby a protein is exposed to small molecules, receptor–ligand complexes are isolated and bound ligands are identified by mass spectrometry.

DNA-encoded libraries

Chemical libraries in which each small molecule is linked to a unique DNA fragment that serves as a barcode. These typically very large libraries can be interrogated by affinity selection.

High-throughput fragment screening

The screening of chemical libraries composed of chemical fragments (molecules smaller than 300 Da) typically at high concentration using direct binding rather than enzymatic assays.

Fluorescence polarization

Fluorescence-based binding assay that relies on the fact that free molecules rotate faster than large molecular complexes.

Apo state

The unbound state of a protein or other macromolecule, as opposed to the 'holo' state, when complexed to a ligand.

Bidentate ligands

Compounds with two chemical groups carrying distinct functions, for example each binding to a different protein interaction partner.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Schapira, M., Tyers, M., Torrent, M. et al. WD40 repeat domain proteins: a novel target class?. Nat Rev Drug Discov 16, 773–786 (2017). https://doi.org/10.1038/nrd.2017.179

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrd.2017.179

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer