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.

Induced protein degradation: an emerging drug discovery paradigm

Key Points

  • Induced protein degradation is an emerging drug discovery platform with the potential to reduce drug exposure requirements, counteract compensatory protein expression and target the 'undruggable proteome'.

  • Selective oestrogen receptor degraders (such as fulvestrant) demonstrated some of the advantages of induced protein degradation over the traditional inhibitor-based approach.

  • Hydrophobic tagging attempts to mimic partially unfolded proteins to induce target protein degradation, although the exact mechanism of action remains unknown.

  • Proteolysis-targeting chimaeras (PROTACs) are a promising platform technology that is modular, catalytic and specific for its protein target with picomolar potencies in cell culture and demonstrated activity in mouse models.

  • An added layer of specificity over the traditional inhibition approach can be accomplished with PROTACs by utilizing different E3 ligases and conditional recruiting ligands.

  • Future development of targeted protein degradation will be focused on targeting the undruggable proteome.

Abstract

Small-molecule drug discovery has traditionally focused on occupancy of a binding site that directly affects protein function, and this approach typically precludes targeting proteins that lack such amenable sites. Furthermore, high systemic drug exposures may be needed to maintain sufficient target inhibition in vivo, increasing the risk of undesirable off-target effects. Induced protein degradation is an alternative approach that is event-driven: upon drug binding, the target protein is tagged for elimination. Emerging technologies based on proteolysis-targeting chimaeras (PROTACs) that exploit cellular quality control machinery to selectively degrade target proteins are attracting considerable attention in the pharmaceutical industry owing to the advantages they could offer over traditional small-molecule strategies. These advantages include the potential to reduce systemic drug exposure, the ability to counteract increased target protein expression that often accompanies inhibition of protein function and the potential ability to target proteins that are not currently therapeutically tractable, such as transcription factors, scaffolding and regulatory proteins.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Mechanism of induced protein degradation technologies.
Figure 2: Structures of compounds used for induced protein degradation technologies.
Figure 3: Timeline of the induced protein degradation field.

Similar content being viewed by others

References

  1. Campbell, J. et al. Large-scale profiling of kinase dependencies in cancer cell lines. Cell Rep. 14, 2490–2501 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Cowley, G. S. et al. Parallel genome-scale loss of function screens in 216 cancer cell lines for the identification of context-specific genetic dependencies. Sci. Data 1, 140035 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Wang, T. et al. Identification and characterization of essential genes in the human genome. Science 350, 1096–1101 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Barretina, J. et al. The Cancer Cell Line Encyclopedia enables predictive modelling of anticancer drug sensitivity. Nature 483, 603–607 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Hopkins, A. L. & Groom, C. R. The druggable genome. Nat. Rev. Drug Discov. 1, 727–730 (2002).

    CAS  PubMed  Google Scholar 

  6. Lazo, J. S. & Sharlow, E. R. Drugging undruggable molecular cancer targets. Annu. Rev. Pharmacol. Toxicol. 56, 23–40 (2016).

    CAS  PubMed  Google Scholar 

  7. Jin, L., Wang, W. & Fang, G. Targeting protein–protein interaction by small molecules. Annu. Rev. Pharmacol. Toxicol. 54, 435–456 (2014).

    CAS  PubMed  Google Scholar 

  8. Adjei, A. A. What is the right dose? The elusive optimal biologic dose in phase I clinical trials. J. Clin. Oncol. 24, 4054–4055 (2006).

    CAS  PubMed  Google Scholar 

  9. Cong, L. et al. Multiplex genome engineering using CRISPR/Cas systems. Science 339, 819–823 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Jinek, M. et al. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 337, 816–821 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Doudna, J. A. & Charpentier, E. The new frontier of genome engineering with CRISPR–Cas9. Science 346, 1258096 (2014).

    PubMed  Google Scholar 

  12. Hsu, P. D., Lander, E. S. & Zhang, F. Development and applications of CRISPR–Cas9 for genome engineering. Cell 157, 1262–1278 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Sander, J. D. & Joung, J. K. CRISPR–Cas systems for editing, regulating and targeting genomes. Nat. Biotechnol. 32, 347–355 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. de Smidt, P. C., Le Doan, T., de Falco, S. & van Berkel, T. J. Association of antisense oligonucleotides with lipoproteins prolongs the plasma half-life and modifies the tissue distribution. Nucleic Acids Res. 19, 4695–4700 (1991).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Geary, R. S. et al. Pharmacokinetic properties of 2′-O-(2-methoxyethyl)-modified oligonucleotide analogs in rats. J. Pharmacol. Exp. Ther. 296, 890–897 (2001).

    CAS  PubMed  Google Scholar 

  16. McMahon, B. M. et al. Pharmacokinetics and tissue distribution of a peptide nucleic acid after intravenous administration. Antisense Nucleic Acid Drug Dev. 12, 65–70 (2002).

    CAS  PubMed  Google Scholar 

  17. Marques, J. T. & Williams, B. R. Activation of the mammalian immune system by siRNAs. Nat. Biotechnol. 23, 1399–1405 (2005).

    CAS  PubMed  Google Scholar 

  18. Krieg, A. M. CpG motifs in bacterial DNA and their immune effects. Annu. Rev. Immunol. 20, 709–760 (2002).

    CAS  PubMed  Google Scholar 

  19. Dahlman, J. E., Kauffman, K. J., Langer, R. & Anderson, D. G. Nanotechnology for in vivo targeted siRNA delivery. Adv. Genet. 88, 37–69 (2014).

    CAS  PubMed  Google Scholar 

  20. Whitehead, K. A., Langer, R. & Anderson, D. G. Knocking down barriers: advances in siRNA delivery. Nat. Rev. Drug Discov. 8, 129–138 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Wilhelm, S. et al. Analysis of nanoparticle delivery to tumours. Nat. Rev. Mater. 1, 16014 (2016).

    CAS  Google Scholar 

  22. Peer, D. & Lieberman, J. Special delivery: targeted therapy with small RNAs. Gene Ther. 18, 1127–1133 (2011).

    CAS  PubMed  Google Scholar 

  23. Fedorov, Y. et al. Off-target effects by siRNA can induce toxic phenotype. RNA 12, 1188–1196 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Jackson, A. L. et al. Expression profiling reveals off-target gene regulation by RNAi. Nat. Biotechnol. 21, 635–637 (2003).

    CAS  PubMed  Google Scholar 

  25. Qiu, S., Adema, C. M. & Lane, T. A computational study of off-target effects of RNA interference. Nucleic Acids Res. 33, 1834–1847 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Burnett, J. C. & Rossi, J. J. RNA-based therapeutics: current progress and future prospects. Chem. Biol. 19, 60–71 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Melnikova, I. RNA-based therapies. Nat. Rev. Drug Discov. 6, 863–864 (2007).

    CAS  Google Scholar 

  28. Bennett, C. F. & Swayze, E. E. RNA targeting therapeutics: molecular mechanisms of antisense oligonucleotides as a therapeutic platform. Annu. Rev. Pharmacol. Toxicol. 50, 259–293 (2010).

    CAS  PubMed  Google Scholar 

  29. Perry, C. M. & Balfour, J. A. B. Fomivirsen. Drugs 57, 375–380 (1999).

    CAS  PubMed  Google Scholar 

  30. Wong, E. & Goldberg, T. Mipomersen (Kynamro): a novel antisense oligonucleotide inhibitor for the management of homozygous familial hypercholesterolemia. Pharmacy Ther. 39, 119–122 (2014).

    Google Scholar 

  31. Sinha, G. Antisense battles small molecule for slice of rare lipid disorder market. Nat. Biotechnol. 31, 179–180 (2013).

    CAS  PubMed  Google Scholar 

  32. Crews, C. M. Targeting the undruggable proteome: the small molecules of my dreams. Chem. Biol. 17, 551–555 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Overington, J. P., Al-Lazikani, B. & Hopkins, A. L. How many drug targets are there? Nat. Rev. Drug Discov. 5, 993–996 (2006).

    CAS  PubMed  Google Scholar 

  34. Duncan, J. S. et al. Dynamic reprogramming of the kinome in response to targeted MEK inhibition in triple-negative breast cancer. Cell 149, 307–321 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Visakorpi, T. et al. In vivo amplification of the androgen receptor gene and progression of human prostate cancer. Nat. Genet. 9, 401–406 (1995).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  37. Heidorn, S. J. et al. Kinase-dead BRAF and oncogenic RAS cooperate to drive tumor progression through CRAF. Cell 140, 209–221 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Rauch, J., Volinsky, N., Romano, D. & Kolch, W. The secret life of kinases: functions beyond catalysis. Cell Commun. Signal. 9, 23 (2011). An excellent review of kinase-independent functions that are unlikely to be affected by traditional kinase inhibitors but would be amendable to protein degradation.

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Tan, X., Thapa, N., Sun, Y. & Anderson, R. A. A kinase-independent role for EGF receptor in autophagy initiation. Cell 160, 145–160 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Vivanco, I. et al. A kinase-independent function of AKT promotes cancer cell survival. eLife 3, e03751 (2014).

    PubMed Central  Google Scholar 

  42. Weihua, Z. et al. Survival of cancer cells is maintained by EGFR independent of its kinase activity. Cancer Cell 13, 385–393 (2008).

    PubMed  PubMed Central  Google Scholar 

  43. Moulick, K. et al. Affinity-based proteomics reveal cancer-specific networks coordinated by Hsp90. Nat. Chem. Biol. 7, 818–826 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Kamal, A. et al. A high-affinity conformation of Hsp90 confers tumour selectivity on Hsp90 inhibitors. Nature 425, 407–410 (2003).

    CAS  PubMed  Google Scholar 

  45. Grenert, J. P. et al. The amino-terminal domain of heat shock protein 90 (hsp90) that binds geldanamycin is an ATP/ADP switch domain that regulates hsp90 conformation. J. Biol. Chem. 272, 23843–23850 (1997).

    CAS  PubMed  Google Scholar 

  46. Stebbins, C. E. et al. Crystal structure of an Hsp90–geldanamycin complex: targeting of a protein chaperone by an antitumor agent. Cell 89, 239–250 (1997).

    CAS  PubMed  Google Scholar 

  47. Prodromou, C. et al. Identification and structural characterization of the ATP/ADP-binding site in the Hsp90 molecular chaperone. Cell 90, 65–75 (1997).

    CAS  PubMed  Google Scholar 

  48. Schneider, C. et al. Pharmacologic shifting of a balance between protein refolding and degradation mediated by Hsp90. Proc. Natl Acad. Sci. USA 93, 14536–14541 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Mimnaugh, E. G., Chavany, C. & Neckers, L. Polyubiquitination and proteasomal degradation of the p185c-erbB-2 receptor protein-tyrosine kinase induced by geldanamycin. J. Biol. Chem. 271, 22796–22801 (1996).

    CAS  PubMed  Google Scholar 

  50. Samuni, Y. et al. Reactive oxygen species mediate hepatotoxicity induced by the Hsp90 inhibitor geldanamycin and its analogs. Free Radic. Biol. Med. 48, 1559–1563 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Banerji, U. et al. Phase I pharmacokinetic and pharmacodynamic study of 17-allylamino, 17-demethoxygeldanamycin in patients with advanced malignancies. J. Clin. Oncol. 23, 4152–4161 (2005).

    CAS  PubMed  Google Scholar 

  52. Taldone, T., Gozman, A., Maharaj, R. & Chiosis, G. Targeting Hsp90: small-molecule inhibitors and their clinical development. Curr. Opin. Pharmacol. 8, 370–374 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. McClellan, A. J. et al. Diverse cellular functions of the Hsp90 molecular chaperone uncovered using systems approaches. Cell 131, 121–135 (2007).

    CAS  PubMed  Google Scholar 

  54. Taipale, M., Jarosz, D. F. & Lindquist, S. HSP90 at the hub of protein homeostasis: emerging mechanistic insights. Nat. Rev. Mol. Cell Biol. 11, 515–528 (2010).

    CAS  PubMed  Google Scholar 

  55. Taipale, M. et al. Quantitative analysis of Hsp90-client interactions reveals principles of substrate recognition. Cell 150, 987–1001 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Kuduk, S. D. et al. Synthesis and evaluation of geldanamycin–testosterone hybrids. Bioorg. Med. Chem. Lett. 10, 1303–1306 (2000).

    CAS  PubMed  Google Scholar 

  57. Kuduk, S. D., Zheng, F. F., Sepp-Lorenzino, L., Rosen, N. & Danishefsky, S. J. Synthesis and evaluation of geldanamycin–estradiol hybrids. Bioorg. Med. Chem. Lett. 9, 1233–1238 (1999).

    CAS  PubMed  Google Scholar 

  58. Katerina, S. & Evangelia, P. HSP90 Inhibitors: current development and potential in cancer therapy. Recent Pat. Anticancer Drug Discov. 9, 1–20 (2014).

    Google Scholar 

  59. Jego, G., Hazoume, A., Seigneuric, R. & Garrido, C. Targeting heat shock proteins in cancer. Cancer Lett. 332, 275–285 (2013).

    CAS  PubMed  Google Scholar 

  60. Trepel, J., Mollapour, M., Giaccone, G. & Neckers, L. Targeting the dynamic HSP90 complex in cancer. Nat. Rev. Cancer 10, 537–549 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Dauvois, S., Danielian, P. S., White, R. & Parker, M. G. Antiestrogen ICI 164,384 reduces cellular estrogen receptor content by increasing its turnover. Proc. Natl Acad. Sci. USA 89, 4037–4041 (1992). The study illustrates the seminal finding that fulvestrant (ICI 164,384) induces the degradation of the ER.

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Giannetti, A. M. From experimental design to validated hits a comprehensive walk-through of fragment lead identification using surface plasmon resonance. Methods Enzymol. 493, 169–218 (2011).

    CAS  PubMed  Google Scholar 

  63. Shuker, S. B., Hajduk, P. J., Meadows, R. P. & Fesik, S. W. Discovering high-affinity ligands for proteins: SAR by NMR. Science 274, 1531–1534 (1996).

    CAS  PubMed  Google Scholar 

  64. Mashalidis, E. H., Sledz, P., Lang, S. & Abell, C. A three-stage biophysical screening cascade for fragment-based drug discovery. Nat. Protoc. 8, 2309–2324 (2013).

    CAS  PubMed  Google Scholar 

  65. Erlanson, D. A., Fesik, S. W., Hubbard, R. E., Jahnke, W. & Jhoti, H. Twenty years on: the impact of fragments on drug discovery. Nat. Rev. Drug Discov. 15, 605–619 (2016).

    CAS  PubMed  Google Scholar 

  66. Renaud, J.-P. et al. Biophysics in drug discovery: impact, challenges and opportunities. Nat. Rev. Drug Discov. 15, 679–698 (2016).

    CAS  PubMed  Google Scholar 

  67. Bleicher, K. H., Bohm, H.-J., Muller, K. & Alanine, A. I. Hit and lead generation: beyond high-throughput screening. Nat. Rev. Drug Discov. 2, 369–378 (2003).

    CAS  PubMed  Google Scholar 

  68. Macarron, R. et al. Impact of high-throughput screening in biomedical research. Nat. Rev. Drug Discov. 10, 188–195 (2011).

    CAS  PubMed  Google Scholar 

  69. Kola, I. & Landis, J. Can the pharmaceutical industry reduce attrition rates? Nat. Rev. Drug Discov. 3, 711–716 (2004).

    CAS  PubMed  Google Scholar 

  70. Bondeson, D. P. et al. Catalytic in vivo protein knockdown by small-molecule PROTACs. Nat. Chem. Biol. 11, 611–617 (2015). This paper demonstrates the first all-small-molecule VHL-based PROTAC capable of degrading ERRα and RIPK2, and it shows that PROTACs remain specific for their targets and are active in vivo.

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Lai, A. C. et al. Modular PROTAC design for the degradation of oncogenic BCR–ABL. Angew. Chem. Int. Ed. 55, 807–810 (2016). This study explores VHL- versus CRBN-based PROTACs targeting the oncogenic tyrosine kinase BCR–ABL.

    CAS  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Heldring, N. et al. Estrogen receptors: how do they signal and what are their targets. Physiol. Rev. 87, 905–931 (2007).

    CAS  PubMed  Google Scholar 

  74. Nilsson, S. et al. Mechanisms of estrogen action. Physiol. Rev. 81, 1535–1565 (2001).

    CAS  PubMed  Google Scholar 

  75. Jordan, V. C. Tamoxifen: a most unlikely pioneering medicine. Nat. Rev. Drug Discov. 2, 205–213 (2003).

    CAS  PubMed  Google Scholar 

  76. Early Breast Cancer Trialists' Collaborative Group. Tamoxifen for early breast cancer: an overview of the randomised trials. Lancet 351, 1451–1467 (1998).

  77. Martin, L. A. et al. Enhanced estrogen receptor (ER) α, ERBB2, and MAPK signal transduction pathways operate during the adaptation of MCF-7 cells to long term estrogen deprivation. J. Biol. Chem. 278, 30458–30468 (2003).

    CAS  PubMed  Google Scholar 

  78. Wakeling, A. E., Dukes, M. & Bowler, J. A potent specific pure antiestrogen with clinical potential. Cancer Res. 51, 3867–3873 (1991).

    CAS  PubMed  Google Scholar 

  79. Buzdar, A. U. & Robertson, J. F. Fulvestrant: pharmacologic profile versus existing endocrine agents for the treatment of breast cancer. Ann. Pharmacother. 40, 1572–1583 (2006).

    CAS  PubMed  Google Scholar 

  80. Wu, Y. L. et al. Structural basis for an unexpected mode of SERM-mediated ER antagonism. Mol. Cell 18, 413–424 (2005). This paper proposes a mechanism of action for SERDs by suggesting that increased surface hydrophobicity upon drug binding leads to the degradation of the ER.

    CAS  PubMed  Google Scholar 

  81. Wittmann, B. M., Sherk, A. & McDonnell, D. P. Definition of functionally important mechanistic differences among selective estrogen receptor down-regulators. Cancer Res. 67, 9549–9560 (2007).

    CAS  PubMed  Google Scholar 

  82. Kato, S. et al. Activation of the estrogen receptor through phosphorylation by mitogen-activated protein kinase. Science 270, 1491–1494 (1995).

    CAS  PubMed  Google Scholar 

  83. Connor, C. E. et al. Circumventing tamoxifen resistance in breast cancers using antiestrogens that induce unique conformational changes in the estrogen receptor. Cancer Res. 61, 2917–2922 (2001).

    CAS  PubMed  Google Scholar 

  84. Osborne, C. K., Wakeling, A. & Nicholson, R. I. Fulvestrant: an oestrogen receptor antagonist with a novel mechanism of action. Br. J. Cancer 90, S2–S6 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  85. Vergote, I. & Abram, P. Fulvestrant, a new treatment option for advanced breast cancer: tolerability versus existing agents. Ann. Oncol. 17, 200–204 (2006).

    CAS  PubMed  Google Scholar 

  86. Di Leo, A. et al. Results of the CONFIRM phase III trial comparing fulvestrant 250 mg with fulvestrant 500 mg in postmenopausal women with estrogen receptor-positive advanced breast cancer. J. Clin. Oncol. 28, 4594–4600 (2010).

    CAS  PubMed  Google Scholar 

  87. van Kruchten, M. et al. Measuring residual estrogen receptor availability during fulvestrant therapy in patients with metastatic breast cancer. Cancer Discov. 5, 72–81 (2015).

    CAS  PubMed  Google Scholar 

  88. Bross, P. F. et al. Fulvestrant in postmenopausal women with advanced breast cancer. Clin. Cancer Res. 9, 4309–4317 (2003).

    CAS  PubMed  Google Scholar 

  89. Govek, S. P. et al. Optimization of an indazole series of selective estrogen receptor degraders: tumor regression in a tamoxifen-resistant breast cancer xenograft. Bioorg. Med. Chem. Lett. 25, 5163–5167 (2015).

    CAS  PubMed  Google Scholar 

  90. Lai, A. et al. Identification of GDC-0810 (ARN-810), an orally bioavailable selective estrogen receptor degrader (SERD) that demonstrates robust activity in tamoxifen-resistant breast cancer xenografts. J. Med. Chem. 58, 4888–4904 (2015).

    CAS  PubMed  Google Scholar 

  91. Garner, F., Shomali, M., Paquin, D., Lyttle, C. R. & Hattersley, G. RAD1901: a novel, orally bioavailable selective estrogen receptor degrader that demonstrates antitumor activity in breast cancer xenograft models. Anticancer Drugs 26, 948–956 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Weir, H. M. et al. AZD9496: an oral estrogen receptor inhibitor that blocks the growth of ER-positive and ESR1 mutant breast tumours in preclinical models. Cancer Res. http://dx.doi.org/10.1158/0008-5472.can-15-2357 (2016).

  93. Matsumoto, T. et al. The androgen receptor in health and disease. Annu. Rev. Physiol. 75, 201–224 (2013).

    CAS  PubMed  Google Scholar 

  94. Crawford, E. D. et al. A controlled trial of leuprolide with and without flutamide in prostatic carcinoma. N. Engl. J. Med. 321, 419–424 (1989).

    CAS  PubMed  Google Scholar 

  95. Kolvenbag, G. J., Blackledge, G. R. & Gotting-Smith, K. Bicalutamide (Casodex) in the treatment of prostate cancer: history of clinical development. Prostate 34, 61–72 (1998).

    CAS  PubMed  Google Scholar 

  96. Linja, M. J. et al. Amplification and overexpression of androgen receptor gene in hormone-refractory prostate cancer. Cancer Res. 61, 3550–3555 (2001).

    CAS  PubMed  Google Scholar 

  97. Feldman, B. J. & Feldman, D. The development of androgen-independent prostate cancer. Nat. Rev. Cancer 1, 34–45 (2001).

    CAS  PubMed  Google Scholar 

  98. Yap, T. A. et al. Drug discovery in advanced prostate cancer: translating biology into therapy. Nat. Rev. Drug Discov. 15, 699–718 (2016).

    CAS  PubMed  Google Scholar 

  99. Scher, H. I. et al. Increased survival with enzalutamide in prostate cancer after chemotherapy. N. Engl. J. Med. 367, 1187–1197 (2012).

    CAS  PubMed  Google Scholar 

  100. Bradbury, R. H. et al. Small-molecule androgen receptor downregulators as an approach to treatment of advanced prostate cancer. Bioorg. Med. Chem. Lett. 21, 5442–5445 (2011).

    CAS  PubMed  Google Scholar 

  101. Bradbury, R. H. et al. Discovery of AZD3514, a small-molecule androgen receptor downregulator for treatment of advanced prostate cancer. Bioorg. Med. Chem. Lett. 23, 1945–1948 (2013).

    CAS  PubMed  Google Scholar 

  102. Loddick, S. A. et al. AZD3514: a small molecule that modulates androgen receptor signaling and function in vitro and in vivo. Mol. Cancer Ther. 12, 1715–1727 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  103. Omlin, A. et al. AZD3514, an oral selective androgen receptor down-regulator in patients with castration-resistant prostate cancer — results of two parallel first-in-human phase I studies. Invest. New Drugs 33, 679–690 (2015).

    CAS  PubMed  Google Scholar 

  104. Williams, R. Discontinued drugs in 2012: oncology drugs. Expert Opin. Investig. Drugs 22, 1627–1644 (2013).

    CAS  PubMed  Google Scholar 

  105. Li, H. et al. Characterization of a new class of androgen receptor antagonists with potential therapeutic application in advanced prostate cancer. Mol. Cancer Ther. 12, 2425–2435 (2013).

    CAS  PubMed  Google Scholar 

  106. Neklesa, T. K. & Crews, C. M. Chemical biology: greasy tags for protein removal. Nature 487, 308–309 (2012).

    CAS  PubMed  Google Scholar 

  107. Long, M. J., Gollapalli, D. R. & Hedstrom, L. Inhibitor mediated protein degradation. Chem. Biol. 19, 629–637 (2012). This study describes the first demonstration of the Boc 3 Arg-based HyT technology used against GST and dihydrofolate reductase.

    CAS  PubMed  PubMed Central  Google Scholar 

  108. Coffey, R. T. et al. Ubiquilin-mediated small molecule inhibition of mammalian target of rapamycin complex 1 (mTORC1) signaling. J. Biol. Chem. 291, 5221–5233 (2016).

    PubMed  PubMed Central  Google Scholar 

  109. Kubota, H. Quality control against misfolded proteins in the cytosol: a network for cell survival. J. Biochem. 146, 609–616 (2009).

    CAS  PubMed  Google Scholar 

  110. Neklesa, T. K. et al. Small-molecule hydrophobic tagging-induced degradation of HaloTag fusion proteins. Nat. Chem. Biol. 7, 538–543 (2011). The first paper to demonstrate cell culture-based and in vivo efficacy of adamantane-based HyTs for the degradation of HaloTag fusion proteins.

    CAS  PubMed  PubMed Central  Google Scholar 

  111. Los, G. V. & Wood, K. The HaloTag: a novel technology for cell imaging and protein analysis. Methods Mol. Biol. 356, 195–208 (2007).

    CAS  PubMed  Google Scholar 

  112. Tae, H. S. et al. Identification of hydrophobic tags for the degradation of stabilized proteins. Chembiochem 13, 538–541 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Neklesa, T. K. et al. A bidirectional system for the dynamic small molecule control of intracellular fusion proteins. ACS Chem. Biol. 8, 2293–2300 (2013). In contrast to HyT, a ligand separate from the HaloTag was identified in this paper that stabilizes the protein, allowing for bidirectional control of intracellular HaloTag concentrations.

    CAS  PubMed  PubMed Central  Google Scholar 

  114. Xie, T. et al. Pharmacological targeting of the pseudokinase Her3. Nat. Chem. Biol. 10, 1006–1012 (2014). This study describes the first demonstration of targeted degradation of a previously undruggable protein by turning a protein ligand into a degrader compound.

    CAS  PubMed  PubMed Central  Google Scholar 

  115. Lim, S. M. et al. Development of small molecules targeting the pseudokinase Her3. Bioorg. Med. Chem. Lett. 25, 3382–3389 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Gustafson, J. L. et al. Small-molecule-mediated degradation of the androgen receptor through hydrophobic tagging. Angew. Chem. Int. Ed. 54, 9659–9662 (2015).

    CAS  Google Scholar 

  117. Teutsch, G. et al. Non-steroidal antiandrogens: synthesis and biological profile of high-affinity ligands for the androgen receptor. J. Steroid Biochem. Mol. Biol. 48, 111–119 (1994).

    CAS  PubMed  Google Scholar 

  118. 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). This paper introduces the PROTAC technology by demonstrating proximity-induced ubiquitylation and targeted protein degradation in a Xenopus egg extract.

    CAS  PubMed  PubMed Central  Google Scholar 

  119. Yaron, A. et al. Identification of the receptor component of the IκBα-ubiquitin ligase. Nature 396, 590–594 (1998).

    CAS  PubMed  Google Scholar 

  120. Griffith, E. C. et al. Methionine aminopeptidase (type 2) is the common target for angiogenesis inhibitors AGM-1470 and ovalicin. Chem. Biol. 4, 461–471 (1997).

    CAS  PubMed  Google Scholar 

  121. Sin, N. et al. The anti-angiogenic agent fumagillin covalently binds and inhibits the methionine aminopeptidase, MetAP-2. Proc. Natl Acad. Sci. USA 94, 6099–6103 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Sakamoto, K. M. et al. Development of Protacs to target cancer-promoting proteins for ubiquitination and degradation. Mol. Cell Proteomics 2, 1350–1358 (2003).

    CAS  PubMed  Google Scholar 

  123. Hon, W. C. et al. Structural basis for the recognition of hydroxyproline in HIF-1α by pVHL. Nature 417, 975–978 (2002).

    CAS  PubMed  Google Scholar 

  124. Min, J. H. et al. Structure of an HIF-1α -pVHL complex: hydroxyproline recognition in signaling. Science 296, 1886–1889 (2002).

    CAS  PubMed  Google Scholar 

  125. Schneekloth, J. S. Jr. et al. Chemical genetic control of protein levels: selective in vivo targeted degradation. J. Am. Chem. Soc. 126, 3748–3754 (2004).

    CAS  PubMed  Google Scholar 

  126. Bargagna-Mohan, P., Baek, S.-H., Lee, H., Kim, K. & Mohan, R. Use of PROTACS as molecular probes of angiogenesis. Bioorg. Med. Chem. Lett. 15, 2724–2727 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  127. Rodriguez-Gonzalez, A. et al. Targeting steroid hormone receptors for ubiquitination and degradation in breast and prostate cancer. Oncogene 27, 7201–7211 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  128. Zhang, D., Baek, S. H., Ho, A. & Kim, K. Degradation of target protein in living cells by small-molecule proteolysis inducer. Bioorg. Med. Chem. Lett. 14, 645–648 (2004).

    CAS  PubMed  Google Scholar 

  129. Henning, R. K. et al. Degradation of Akt using protein-catalyzed capture agents. J. Pept. Sci. 22, 196–200 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  130. Chu, T.-T. et al. Specific knockdown of endogenous tau protein by peptide-directed ubiquitin-proteasome degradation. Cell Chem. Biol. 23, 453–461 (2016).

    CAS  PubMed  Google Scholar 

  131. Hines, J., Gough, J. D., Corson, T. W. & Crews, C. M. Posttranslational protein knockdown coupled to receptor tyrosine kinase activation with phosphoPROTACs. Proc. Natl Acad. Sci. USA 110, 8942–8947 (2013). By coupling the activation of receptor tyrosine kinases to the degradation of downstream effector proteins, this paper introduces the concept of conditional protein knockdown through PROTAC technology.

    CAS  PubMed  PubMed Central  Google Scholar 

  132. Cong, F., Zhang, J., Pao, W., Zhou, P. & Varmus, H. A protein knockdown strategy to study the function of β-catenin in tumorigenesis. BMC Mol. Biol. 4, 10 (2003).

    PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  134. Schneekloth, A. R., Pucheault, M., Tae, H. S. & Crews, C. M. Targeted intracellular protein degradation induced by a small molecule: en route to chemical proteomics. Bioorg. Med. Chem. Lett. 18, 5904–5908 (2008). The first all-small-molecule PROTAC is introduced in this paper by recruiting the E3 ligase MDM2 to degrade the AR.

    CAS  PubMed  Google Scholar 

  135. Ding, Q. et al. Discovery of RG7388, a potent and selective p53–MDM2 inhibitor in clinical development. J. Med. Chem. 56, 5979–5983 (2013).

    CAS  PubMed  Google Scholar 

  136. Vu, B. et al. Discovery of RG7112: a small-molecule MDM2 inhibitor in clinical development. ACS Med. Chem. Lett. 4, 466–469 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  137. Andreeff, M. et al. Results of the phase I trial of RG7112, a small-molecule MDM2 antagonist in leukemia. Clin. Cancer Res. 22, 868–876 (2016).

    CAS  PubMed  Google Scholar 

  138. Itoh, Y., Ishikawa, M., Naito, M. & Hashimoto, Y. Protein knockdown using methyl bestatin–ligand hybrid molecules: design and synthesis of inducers of ubiquitination-mediated degradation of cellular retinoic acid-binding proteins. J. Am. Chem. Soc. 132, 5820–5826 (2010). The paper introduces the first cIAP-based PROTAC that uses bestatin-mediated recruitment of cIAP to subsequently degrade CRABPI and CRABPII.

    CAS  PubMed  Google Scholar 

  139. Umezawa, H., Aoyagi, T., Suda, H., Hamada, M. & Takeuchi, T. Bestatin, an inhibitor of aminopeptidase B, produced by actinomycetes. J. Antibiot. 29, 97–99 (1976).

    CAS  Google Scholar 

  140. Orning, L., Krivi, G. & Fitzpatrick, F. A. Leukotriene A4 hydrolase. Inhibition by bestatin and intrinsic aminopeptidase activity establish its functional resemblance to metallohydrolase enzymes. J. Biol. Chem. 266, 1375–1378 (1991).

    CAS  PubMed  Google Scholar 

  141. Sekine, K. et al. Small molecules destabilize cIAP1 by activating auto-ubiquitylation. J. Biol. Chem. 283, 8961–8968 (2008).

    CAS  PubMed  Google Scholar 

  142. Itoh, Y. et al. Development of target protein-selective degradation inducer for protein knockdown. Bioorg. Med. Chem. 19, 3229–3241 (2011). This study describes an amide-type cIAP E3 ligase ligand that can recruit cIAP and also reduces autoubiquitylation of cIAP.

    CAS  PubMed  Google Scholar 

  143. Itoh, Y. et al. Double protein knockdown of cIAP1 and CRABP-II using a hybrid molecule consisting of ATRA and IAPs antagonist. Bioorg. Med. Chem. Lett. 22, 4453–4457 (2012).

    CAS  PubMed  Google Scholar 

  144. Demizu, Y. et al. Design and synthesis of estrogen receptor degradation inducer based on a protein knockdown strategy. Bioorg. Med. Chem. Lett. 22, 1793–1796 (2012).

    CAS  PubMed  Google Scholar 

  145. Itoh, Y., Kitaguchi, R., Ishikawa, M., Naito, M. & Hashimoto, Y. Design, synthesis and biological evaluation of nuclear receptor-degradation inducers. Bioorg. Med. Chem. 19, 6768–6778 (2011).

    CAS  PubMed  Google Scholar 

  146. Ohoka, N. et al. Cancer cell death induced by novel small molecules degrading the TACC3 protein via the ubiquitin–proteasome pathway. Cell Death Dis. 5, e1513 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  147. Buckley, D. L. et al. Targeting the von Hippel–Lindau E3 ubiquitin ligase using small molecules to disrupt the VHL/HIF-1α interaction. J. Am. Chem. Soc. 134, 4465–4468 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  148. Buckley, D. L. et al. Small-molecule inhibitors of the interaction between the E3 ligase VHL and HIF1α. Angew. Chem. Int. Ed. 51, 11463–11467 (2012).

    CAS  Google Scholar 

  149. Van Molle, I. et al. Dissecting fragment-based lead discovery at the von Hippel–Lindau protein:hypoxia inducible factor 1α protein–protein interface. Chem. Biol. 19, 1300–1312 (2012).

    CAS  PubMed  Google Scholar 

  150. Galdeano, C. et al. Structure-guided design and optimization of small molecules targeting the protein-protein interaction between the von Hippel–Lindau (VHL) E3 ubiquitin ligase and the hypoxia inducible factor (HIF) alpha subunit with in vitro nanomolar affinities. J. Med. Chem. 57, 8657–8663 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  151. Ito, T. et al. Identification of a primary target of thalidomide teratogenicity. Science 327, 1345–1350 (2010).

    CAS  PubMed  Google Scholar 

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

    PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  154. Gandhi, A. K. et al. Immunomodulatory agents lenalidomide and pomalidomide co-stimulate T cells by inducing degradation of T cell repressors Ikaros and Aiolos via modulation of the E3 ubiquitin ligase complex CRL4CRBN. Br. J. Haematol. 164, 811–821 (2014).

    CAS  PubMed  Google Scholar 

  155. Fischer, E. S. et al. Structure of the DDB1-CRBN E3 ubiquitin ligase in complex with thalidomide. Nature 512, 49–53 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  156. Kronke, J. et al. Lenalidomide induces ubiquitination and degradation of CK1α in del(5q) MDS. Nature 523, 183–188 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  157. Petzold, G., Fischer, E. S. & Thoma, N. H. Structural basis of lenalidomide-induced CK1α degradation by the CRL4CRBN ubiquitin ligase. Nature 532, 127–130 (2016).

    CAS  PubMed  Google Scholar 

  158. Adès, L. & Fenaux, P. Immunomodulating drugs in myelodysplastic syndromes. ASH Educ. Program Book 2011, 556–560 (2011).

    Google Scholar 

  159. Matyskiela, M. E. et al. A novel cereblon modulator recruits GSPT1 to the CRL4CRBN ubiquitin ligase. Nature 535, 252–257 (2016).

    CAS  PubMed  Google Scholar 

  160. Lu, J. et al. Hijacking the E3 ubiquitin ligase cereblon to efficiently target BRD4. Chem. Biol. 22, 755–763 (2015). This paper demonstrates a small-molecule CRBN-based PROTAC capable of degrading BRD4 that is superior to the parent inhibitor OTX-15 in Burkitt's lymphoma cell lines.

    CAS  PubMed  PubMed Central  Google Scholar 

  161. Shimamura, T. et al. Efficacy of BET bromodomain inhibition in KRAS-mutant non-small cell lung cancer. Clin. Cancer Res. 19, 6183–6192 (2013).

    CAS  PubMed  Google Scholar 

  162. Raina, K. et al. PROTAC-induced BET protein degradation as a therapy for castration-resistant prostate cancer. Proc. Natl Acad. Sci. USA 113, 7124–7129 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  163. Zengerle, M., Chan, K.-H. & Ciulli, A. Selective small molecule induced degradation of the BET bromodomain protein BRD4. ACS Chem. Biol. 10, 1770–1777 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  164. Golas, J. M. et al. SKI-606, a 4-anilino-3-quinolinecarbonitrile dual inhibitor of Src and Abl kinases, is a potent antiproliferative agent against chronic myelogenous leukemia cells in culture and causes regression of K562 xenografts in nude mice. Cancer Res. 63, 375–381 (2003).

    CAS  PubMed  Google Scholar 

  165. O'Hare, T. et al. In vitro activity of Bcr–Abl inhibitors AMN107 and BMS-354825 against clinically relevant imatinib-resistant Abl kinase domain mutants. Cancer Res. 65, 4500–4505 (2005).

    CAS  PubMed  Google Scholar 

  166. Duda, D. M. et al. Structural insights into NEDD8 activation of cullin-RING ligases: conformational control of conjugation. Cell 134, 995–1006 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  167. Fischer, E. S. et al. The molecular basis of CRL4DDB2/CSA ubiquitin ligase architecture, targeting, and activation. Cell 147, 1024–1039 (2011). This paper provides a structural rationale to the flexibility of cullin 4-based E3 ligases and explores the ubiquitylation zone of these ligases.

    CAS  PubMed  Google Scholar 

  168. Liu, J. & Nussinov, R. Flexible cullins in cullin-RING E3 ligases allosterically regulate ubiquitination. J. Biol. Chem. 286, 40934–40942 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  169. Liu, J. & Nussinov, R. The mechanism of ubiquitination in the cullin-RING E3 ligase machinery: conformational control of substrate orientation. PLoS Comput. Biol. 5, e1000527 (2009).

    PubMed  PubMed Central  Google Scholar 

  170. Petroski, M. D. & Deshaies, R. J. Function and regulation of cullin-RING ubiquitin ligases. Nat. Rev. Mol. Cell Biol. 6, 9–20 (2005).

    CAS  PubMed  Google Scholar 

  171. Deshaies, R. J. Protein degradation: prime time for PROTACs. Nat. Chem. Biol. 11, 634–635 (2015).

    CAS  PubMed  Google Scholar 

  172. Toure, M. & Crews, C. M. Small-molecule PROTACS: new approaches to protein degradation. Angew. Chem. Int. Ed. 55, 1966–1973 (2016).

    CAS  Google Scholar 

  173. Lou, K.-J. PROTAC the protein. SciBx http://dx.doi.org/10.1038/scibx.2012.514 (2012).

  174. Bouchie, A., Allison, M., Webb, S. & DeFrancesco, L. Nature Biotechnology's academic spinouts of 2013. Nat. Biotechnol. 32, 229–238 (2014).

    CAS  PubMed  Google Scholar 

  175. Chi, K. R. Drug developers delve into the cell's trash-disposal machinery. Nat. Rev. Drug Discov. 15, 295–297 (2016).

    CAS  PubMed  Google Scholar 

  176. Buckley, D. L. & Crews, C. M. Small-molecule control of intracellular protein levels through modulation of the ubiquitin proteasome system. Angew. Chem. Int. Ed. 53, 2312–2330 (2014).

    CAS  Google Scholar 

  177. Buckley, D. L. et al. HaloPROTACS: use of small molecule PROTACs to induce degradation of halotag fusion proteins. ACS Chem. Biol. 10, 1831–1837 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  178. Lipinski, C. A., Lombardo, F., Dominy, B. W. & Feeney, P. J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 46, 3–26 (2001).

    CAS  PubMed  Google Scholar 

  179. Maculins, T. et al. A generic platform for cellular screening against ubiquitin ligases. Sci. Rep. 6, 18940 (2016). This study reports an interesting platform technology that could be used to find novel E3 ligase-recruiting ligands.

    CAS  PubMed  PubMed Central  Google Scholar 

  180. Chan, C. H. et al. Pharmacological inactivation of Skp2 SCF ubiquitin ligase restricts cancer stem cell traits and cancer progression. Cell 154, 556–568 (2013).

    CAS  PubMed  Google Scholar 

  181. Orlicky, S. et al. An allosteric inhibitor of substrate recognition by the SCFCdc4 ubiquitin ligase. Nat. Biotechnol. 28, 733–737 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  182. Aghajan, M. et al. Chemical genetics screen for enhancers of rapamycin identifies a specific inhibitor of an SCF family E3 ubiquitin ligase. Nat. Biotechnol. 28, 738–742 (2010).

    CAS  PubMed  Google Scholar 

  183. Deshaies, R. J. & Joazeiro, C. A. P. RING domain E3 ubiquitin ligases. Annu. Rev. Biochem. 78, 399–434 (2009).

    CAS  PubMed  Google Scholar 

  184. DeBoer, C., Meulman, P. A., Wnuk, R. J. & Peterson, D. H. Geldanamycin, a new antibiotic. J. Antibiot. 23, 442–447 (1970).

    CAS  Google Scholar 

  185. Rinehart, K. L., Sasaki, K., Slomp, G., Grostic, M. F. & Olson, E. C. Geldanamycin. I. Structure assignment. J. Am. Chem. Soc. 92, 7591–7593 (1970).

    CAS  PubMed  Google Scholar 

  186. Supko, J. G., Hickman, R. L., Grever, M. R. & Malspeis, L. Preclinical pharmacologic evaluation of geldanamycin as an antitumor agent. Cancer Chemother. Pharmacol. 36, 305–315 (1995).

    CAS  PubMed  Google Scholar 

  187. Ramanathan, R. K. et al. Phase I pharmacokinetic-pharmacodynamic study of 17-(allylamino)-17-demethoxygeldanamycin (17AAG, NSC 330507), a novel inhibitor of heat shock protein 90, in patients with refractory advanced cancers. Clin. Cancer Res. 11, 3385–3391 (2005).

    CAS  PubMed  Google Scholar 

  188. Modi, S. et al. HSP90 inhibition is effective in breast cancer: a phase II trial of tanespimycin (17-AAG) plus trastuzumab in patients with HER2-positive metastatic breast cancer progressing on trastuzumab. Clin. Cancer Res. 17, 5132–5139 (2011).

    CAS  PubMed  Google Scholar 

  189. Demetri, G. D. et al. Final results from a phase III study of IPI-504 (retaspimycin hydrochloride) versus placebo in patients (pts) with gastrointestinal stromal tumors (GIST) following failure of kinase inhibitor therapies. ASCO 2010 Gastrointestinal Cancers Symp. http://meetinglibrary.asco.org/content/2285-72 (2010).

    Google Scholar 

  190. Lundgren, K. et al. BIIB021, an orally available, fully synthetic small-molecule inhibitor of the heat shock protein Hsp90. Mol. Cancer Ther. 8, 921–929 (2009).

    CAS  PubMed  Google Scholar 

  191. Dickson, M. A. et al. Phase II study of the HSP90-inhibitor BIIB021 in gastrointestinal stromal tumors. Ann. Oncol. 24, 252–257 (2013).

    CAS  PubMed  Google Scholar 

  192. Mitchell, P. Biogen Idec restructures, sharpens neurology focus. Nat. Biotechnol. 29, 7–8 (2011).

    CAS  PubMed  Google Scholar 

  193. Eccles, S. A. et al. NVP-AUY922: a novel heat shock protein 90 inhibitor active against xenograft tumor growth, angiogenesis, and metastasis. Cancer Res. 68, 2850–2860 (2008).

    CAS  PubMed  Google Scholar 

  194. Lin, T. Y. et al. The novel HSP90 inhibitor STA-9090 exhibits activity against Kit-dependent and -independent malignant mast cell tumors. Exp. Hematol. 36, 1266–1277 (2008).

    CAS  PubMed  Google Scholar 

  195. Ramalingam, S. et al. A randomized phase II study of ganetespib, a heat shock protein 90 inhibitor, in combination with docetaxel in second-line therapy of advanced non-small cell lung cancer (GALAXY-1). Ann. Oncol. 26, 1741–1748 (2015).

    CAS  PubMed  Google Scholar 

  196. Chatterjee, S., Bhattacharya, S., Socinski, M. A. & Burns, T. F. HSP90 inhibitors in lung cancer: promise still unfulfilled. Clin. Adv. Hematol. Oncol. 14, 346–356 (2016).

    PubMed  Google Scholar 

  197. Woodhead, A. J. et al. Discovery of (2,4-dihydroxy- 5-isopropylphenyl)-[5-(4-methylpiperazin-1-ylmethyl)-1,3-dihydroisoindol-2-yl]methanone (AT13387), a novel inhibitor of the molecular chaperone Hsp90 by fragment based drug design. J. Med. Chem. 53, 5956–5969 (2010).

    CAS  PubMed  Google Scholar 

  198. Wagner, A. J. et al. Dose-escalation study of a second-generation non-ansamycin HSP90 inhibitor, onalespib (AT13387), in combination with imatinib in patients with metastatic gastrointestinal stromal tumour. Eur. J. Cancer 61, 94–101 (2016).

    CAS  PubMed  Google Scholar 

  199. Rakhit, R., Navarro, R. & Wandless, T. J. Chemical biology strategies for posttranslational control of protein function. Chem. Biol. 21, 1238–1252 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  200. Raina, K. et al. Targeted protein destabilization reveals an estrogen-mediated ER stress response. Nat. Chem. Biol. 10, 957–962 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  201. Lytton, J., Westlin, M. & Hanley, M. R. Thapsigargin inhibits the sarcoplasmic or endoplasmic reticulum Ca-ATPase family of calcium pumps. J. Biol. Chem. 266, 17067–17071 (1991).

    CAS  PubMed  Google Scholar 

  202. Takatsuki, A., Arima, K. & Tamura, G. Tunicamycin, a new antibiotic. I. Isolation and characterization of tunicamycin. J. Antibiot. 24, 215–223 (1971).

    CAS  Google Scholar 

  203. Miyazaki, Y., Chen, L. C., Chu, B. W., Swigut, T. & Wandless, T. J. Distinct transcriptional responses elicited by unfolded nuclear or cytoplasmic protein in mammalian cells. eLife 4, e07687 (2015).

    PubMed Central  Google Scholar 

  204. Schulman, B. A. & Harper, J. W. Ubiquitin-like protein activation by E1 enzymes: the apex for downstream signalling pathways. Nat. Rev. Mol. Cell Biol. 10, 319–331 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  205. Ye, Y. & Rape, M. Building ubiquitin chains: E2 enzymes at work. Nat. Rev. Mol. Cell Biol. 10, 755–764 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  206. Winston, J. T. et al. The SCFβ-TRCP–ubiquitin ligase complex associates specifically with phosphorylated destruction motifs in IκBα and β-catenin and stimulates IκBα ubiquitination in vitro. Genes Dev. 13, 270–283 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  207. Fang, S., Jensen, J. P., Ludwig, R. L., Vousden, K. H. & Weissman, A. M. Mdm2 is a RING finger-dependent ubiquitin protein ligase for itself and p53. J. Biol. Chem. 275, 8945–8951 (2000).

    CAS  PubMed  Google Scholar 

  208. Wade, M., Li, Y.-C. & Wahl, G. M. MDM2, MDMX and p53 in oncogenesis and cancer therapy. Nat. Rev. Cancer 13, 83–96 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  209. Chau, V. et al. A multiubiquitin chain is confined to specific lysine in a targeted short-lived protein. Science 243, 1576–1583 (1989).

    CAS  PubMed  Google Scholar 

  210. Lu, Y., Lee, B. H., King, R. W., Finley, D. & Kirschner, M. W. Substrate degradation by the proteasome: a single-molecule kinetic analysis. Science 348, 1250834 (2015).

    PubMed  PubMed Central  Google Scholar 

  211. Komander, D. & Rape, M. The ubiquitin code. Annu. Rev. Biochem. 81, 203–229 (2012).

    CAS  PubMed  Google Scholar 

  212. Husnjak, K. & Dikic, I. Ubiquitin-binding proteins: decoders of ubiquitin-mediated cellular functions. Annu. Rev. Biochem. 81, 291–322 (2012).

    CAS  PubMed  Google Scholar 

  213. MacGurn, J. A., Hsu, P. C. & Emr, S. D. Ubiquitin and membrane protein turnover: from cradle to grave. Annu. Rev. Biochem. 81, 231–259 (2012).

    CAS  PubMed  Google Scholar 

  214. Lee, H., Puppala, D., Choi, E. Y., Swanson, H. & Kim, K. B. Targeted degradation of the aryl hydrocarbon receptor by the PROTAC approach: a useful chemical genetic tool. Chembiochem 8, 2058–2062 (2007).

    CAS  PubMed  Google Scholar 

  215. Tomoshige, S., Naito, M., Hashimoto, Y. & Ishikawa, M. Degradation of HaloTag-fused nuclear proteins using bestatin-HaloTag ligand hybrid molecules. Org. Biomol. Chem. 13, 9746–9750 (2015).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors thank S. Jamie-Figueroa, T. Neklesa, K. Raina and the rest of the Crews laboratory for their helpful comments. C.M.C. gratefully acknowledges the US National Institutes of Health (NIH) for its support (R35-CA197589). A.C.L acknowledges support from the NIH (MSTP NIH/NIGMS T32GM007205). C.M.C. is founder, consultant and shareholder of Arvinas, LLC.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Craig M. Crews.

Ethics declarations

Competing interests

C.M.C. is a founder, consultant and shareholder of Arvinas, LLC. A.C.L. declares no competing interests.

Related links

PowerPoint slides

Glossary

Event-driven pharmacology

A drug discovery paradigm based on identifying compounds that lead to the removal of the target protein from the system following drug binding to the target.

Sub-stoichiometric

Pertains to molecular compounds that have a greater effect than the 1:1 activity ratio seen with a stoichiometric agent.

HaloTag

A bacterial dehalogenase engineered by the Promega Corporation to covalently bind to chloroalkanes that has been used primarily for protein purification and imaging-based applications.

Erythroblastosis oncogene B3

(ERBB3). A member of the epidermal growth factor receptor (EGFR) family that has minimal kinase activity and acts primarily as a scaffold for signal transduction.

NF-κB inhibitor-α

(IκBα). Inhibitor of NF-κB that containing a destruction motif that, when phosphorylated, is recognized by the E3 ligase substrate recognition component β-transducin repeat-containing protein.

von Hippel–Lindau

(VHL). The substrate recognition portion of the E3 ligase complex VHL–elongin B/C–cullin 2 (VHL–ELOBC–CUL2) that mediates the transfer of ubiquitin onto hypoxia-inducible factor 1α (HIF1α).

Poly-D-arginine tag

An arginine-containing peptide fragment based on the HIV Tat peptide that facilitates cellular uptake of proteins.

Occupancy-driven pharmacology

A drug discovery paradigm based on identifying compounds that modulate protein function through stoichiometric drug binding and occupation of the binding site to modulate protein function.

Bromodomain-containing protein 4

(BRD4). A member of the bromodomain and extra-terminal (BET) family that recognizes acetylated lysines of histones to modulate the epigenetic code.

Breakpoint cluster region–Abelson tyrosine kinase

(BCR–ABL). An oncogenic fusion of BCR and ABL that can lead to chronic myeloid leukaemia.

Rule of five

The rule of five is a widely used guideline to predict the likelihood of poor absorption or permeability characteristics for small molecules. It is based on upper limits for physicochemical properties of the compounds, including molecular mass (for which 500 Da is the cut-off).

Degrons

Peptide segments recognized by the cellular quality control system that influence the degradation rate of the full-length protein.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lai, A., Crews, C. Induced protein degradation: an emerging drug discovery paradigm. Nat Rev Drug Discov 16, 101–114 (2017). https://doi.org/10.1038/nrd.2016.211

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrd.2016.211

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