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.

  • Opinion
  • Published:

Targeting tumour-supportive cellular machineries in anticancer drug development

Abstract

Traditional anticancer chemotherapeutics targeting DNA replication and cell division have severe side effects, but they have proved to be highly successful in treating some cancers. Drugs targeting signalling oncoproteins that have gained tumour-driving functions through mutations or overexpression were subsequently developed to increase specificity and thus reduce side effects, but have limitations such as the development of resistance. Now, a new wave of small-molecule anticancer agents is emerging, targeting complex multicomponent cellular machineries — including chromatin modifiers, heat shock protein chaperones and the proteasome — which thus interfere with those support systems that are more essential for cancer cells than for normal cells. Here, we provide our perspective on the advantages and limitations of agents that target tumour-supportive cellular machineries (other than those involving DNA replication), comparing them with agents that target signalling intermediates.

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
Figure 2: Cellular multicomponent machineries as current and future targets for anticancer drugs.

Similar content being viewed by others

References

  1. Goodman, L. S. et al. Nitrogen mustard therapy; use of methyl-bis (beta-chloroethyl) amine hydrochloride and tris (beta-chloroethyl) amine hydrochloride for Hodgkin's disease, lymphosarcoma, leukemia and certain allied and miscellaneous disorders. JAMA 132, 126–132 (1946).

    CAS  Google Scholar 

  2. Gilman, A. & Philips, F. S. The biological actions and therapeutic applications of the B-chloroethyl amines and sulfides. Science 103, 409–415 (1946).

    CAS  PubMed  Google Scholar 

  3. Mukherjee, S. The Emperor of All Maladies: A Biography of Cancer (Scribner, 2010). This summarizes the remarkable history of cancer treatment, explaining how the first wave of chemotherapeutics was developed.

    Google Scholar 

  4. Jackson, J. R., Patrick, D. R., Dar, M. M. & Huang, P. S. Targeted anti-mitotic therapies: can we improve on tubulin agents? Nature Rev. Cancer 7, 107–117 (2007).

    CAS  Google Scholar 

  5. Armstrong, G. T. et al. Occurrence of multiple subsequent neoplasms in long-term survivors of childhood cancer: a report from the childhood cancer survivor study. J. Clin. Oncol. 29, 3056–3064 (2011).

    PubMed  PubMed Central  Google Scholar 

  6. Travis, L. B. et al. Second cancers among 40,576 testicular cancer patients: focus on long-term survivors. J. Natl Cancer Inst. 97, 1354–1365 (2005).

    PubMed  Google Scholar 

  7. Strebhardt, K. & Ullrich, A. Paul Ehrlich's magic bullet concept: 100 years of progress. Nature Rev. Cancer 8, 473–480 (2008).

    CAS  Google Scholar 

  8. Luo, J., Solimini, N. L. & Elledge, S. J. Principles of cancer therapy: oncogene and non-oncogene addiction. Cell 136, 823–837 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Felsher, D. W. MYC inactivation elicits oncogene addiction through both tumor cell-intrinsic and host-dependent mechanisms. Genes Cancer 1, 597–604 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Capdeville, R., Buchdunger, E., Zimmermann, J. & Matter, A. Glivec (STI571, imatinib), a rationally developed, targeted anticancer drug. Nature Rev. Drug Discov. 1, 493–502 (2002).

    CAS  Google Scholar 

  11. Vogel, C. L. et al. Efficacy and safety of trastuzumab as a single agent in first-line treatment of HER2-overexpressing metastatic breast cancer. J. Clin. Oncol. 20, 719–726 (2002).

    CAS  PubMed  Google Scholar 

  12. Ashwell, S. & Zabludoff, S. DNA damage detection and repair pathways — recent advances with inhibitors of checkpoint kinases in cancer therapy. Clin. Cancer Res. 14, 4032–4037 (2008).

    CAS  PubMed  Google Scholar 

  13. Guertin, D. A. & Sabatini, D. M. Defining the role of mTOR in cancer. Cancer Cell 12, 9–22 (2007).

    CAS  PubMed  Google Scholar 

  14. Sierra, J. R., Cepero, V. & Giordano, S. Molecular mechanisms of acquired resistance to tyrosine kinase targeted therapy. Mol. Cancer 9, 75 (2010).

    PubMed  PubMed Central  Google Scholar 

  15. Milojkovic, D. & Apperley, J. Mechanisms of resistance to imatinib and second-generation tyrosine inhibitors in chronic myeloid leukemia. Clin. Cancer Res. 15, 7519–7527 (2009).

    CAS  PubMed  Google Scholar 

  16. Druker, B. J. et al. Five-year follow-up of patients receiving imatinib for chronic myeloid leukemia. N. Engl. J. Med. 355, 2408–2417 (2006).

    CAS  PubMed  Google Scholar 

  17. Grebien, F. et al. Targeting the SH2-kinase interface in Bcr-Abl inhibits leukemogenesis. Cell 147, 306–319 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Chan, W. W. et al. Conformational control inhibition of the BCR-ABL1 tyrosine kinase, including the gatekeeper T315I mutant, by the switch-control inhibitor DCC-2036. Cancer Cell 19, 556–568 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Zhang, J. et al. Targeting Bcr-Abl by combining allosteric with ATP-binding-site inhibitors. Nature 463, 501–506 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Gregory, M. A. et al. Wnt/Ca2+/NFAT signaling maintains survival of Ph+ leukemia cells upon inhibition of Bcr-Abl. Cancer Cell 18, 74–87 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Naka, K. et al. TGF-β-FOXO signalling maintains leukaemia-initiating cells in chronic myeloid leukaemia. Nature 463, 676–680 (2010). References 20 and 21 describe the mechanisms by which collateral pathways overcome the inhibition of BCR–ABL.

    CAS  PubMed  Google Scholar 

  22. Villamor, N., Montserrat, E. & Colomer, D. Mechanism of action and resistance to monoclonal antibody therapy. Seminars Oncol. 30, 424–433 (2003).

    CAS  Google Scholar 

  23. Berns, K. et al. A functional genetic approach identifies the PI3K pathway as a major determinant of trastuzumab resistance in breast cancer. Cancer Cell 12, 395–402 (2007).

    CAS  PubMed  Google Scholar 

  24. Pohlmann, P. R., Mayer, I. A. & Mernaugh, R. Resistance to trastuzumab in breast cancer. Clin. Cancer Res. 15, 7479–7491 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Nahta, R., Yu, D., Hung, M. C., Hortobagyi, G. N. & Esteva, F. J. Mechanisms of disease: understanding resistance to HER2-targeted therapy in human breast cancer. Nature Clin. Pract. Oncol. 3, 269–280 (2006).

    CAS  Google Scholar 

  26. Knight, Z. A., Lin, H. & Shokat, K. M. Targeting the cancer kinome through polypharmacology. Nature Rev. Cancer 10, 130–137 (2010).

    CAS  Google Scholar 

  27. Apsel, B. et al. Targeted polypharmacology: discovery of dual inhibitors of tyrosine and phosphoinositide kinases. Nature Chem. Biol. 4, 691–699 (2008).

    CAS  Google Scholar 

  28. Furge, K. A., MacKeigan, J. P. & Teh, B. T. Kinase targets in renal-cell carcinomas: reassessing the old and discovering the new. Lancet Oncol. 11, 571–578 (2010).

    CAS  PubMed  Google Scholar 

  29. Flaherty, K. T. et al. Combined BRAF and MEK inhibition in melanoma with BRAF V600 mutations. N. Engl. J. Med. 367, 1694–1703 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Spigel, D. R. et al. Randomized, double-blind, placebo-controlled, phase II trial of sorafenib and erlotinib or erlotinib alone in previously treated advanced non-small-cell lung cancer. J. Clin. Oncol. 29, 2582–2589 (2011).

    CAS  PubMed  Google Scholar 

  31. Sequist, L. V. et al. Randomized phase II study of erlotinib plus tivantinib versus erlotinib plus placebo in previously treated non-small-cell lung cancer. J. Clin. Oncol. 29, 3307–3315 (2011).

    CAS  PubMed  Google Scholar 

  32. Vogelstein, B. et al. Cancer genome landscapes. Science 339, 1546–1558 (2013). This is a summary and critical assessment of what has been learnt from cancer genome sequencing.

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Hanahan, D. & Weinberg, R. A. The hallmarks of cancer. Cell 100, 57–70 (2000).

    CAS  PubMed  Google Scholar 

  34. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell 144, 646–674 (2011). Along with reference 8, references 33 and 34 are seminal reviews describing the hallmarks of cancer and their exploitation for therapy.

    CAS  PubMed  Google Scholar 

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

    CAS  Google Scholar 

  36. Dai, C., Dai, S. & Cao, J. Proteotoxic stress of cancer: implication of the heat-shock response in oncogenesis. J. Cell. Physiol. 227, 2982–2987 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Stadtman, E. R. Protein oxidation in aging and age-related diseases. Ann. NY Acad. Sci. 928, 22–38 (2001).

    CAS  PubMed  Google Scholar 

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

    CAS  Google Scholar 

  39. Terzian, T. et al. The inherent instability of mutant p53 is alleviated by Mdm2 or p16INK4a loss. Genes Dev. 22, 1337–1344 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Li, D., Marchenko, N. D. & Moll, U. M. SAHA shows preferential cytotoxicity in mutant p53 cancer cells by destabilizing mutant p53 through inhibition of the HDAC6–Hsp90 chaperone axis. Cell Death Differ. 18, 1904–1913 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Li, D. et al. Functional inactivation of endogenous MDM2 and CHIP by HSP90 causes aberrant stabilization of mutant p53 in human cancer cells. Mol. Cancer Res. 9, 577–588 (2011). References 40 and 41 describe a stabilized mutant p53 protein as a degradation target for anti-HDAC6 or HSP90 therapy.

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Sato, S., Fujita, N. & Tsuruo, T. Modulation of Akt kinase activity by binding to Hsp90. Proc. Natl Acad. Sci. USA 97, 10832–10837 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Whitesell, L. & Lindquist, S. L. HSP90 and the chaperoning of cancer. Nature Rev. Cancer 5, 761–772 (2005).

    CAS  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. De Raedt, T. et al. Exploiting cancer cell vulnerabilities to develop a combination therapy for ras-driven tumors. Cancer Cell 20, 400–413 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Chen, G., Bradford, W. D., Seidel, C. W. & Li, R. Hsp90 stress potentiates rapid cellular adaptation through induction of aneuploidy. Nature 482, 246–250 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Cerchietti, L. C. et al. A purine scaffold Hsp90 inhibitor destabilizes BCL-6 and has specific antitumor activity in BCL-6-dependent B cell lymphomas. Nature Med. 15, 1369–1376 (2009).

    CAS  PubMed  Google Scholar 

  48. Schulz, R. et al. Inhibiting the HSP90 chaperone destabilizes macrophage migration inhibitory factor and thereby inhibits breast tumor progression. J. Exp. Med. 209, 275–289 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. 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 

  50. Leu, J. I., Pimkina, J., Frank, A., Murphy, M. E. & George, D. L. A small molecule inhibitor of inducible heat shock protein 70. Mol. Cell 36, 15–27 (2009). This paper describes HSP70 as a target for therapy.

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Dorard, C. et al. Expression of a mutant HSP110 sensitizes colorectal cancer cells to chemotherapy and improves disease prognosis. Nature Med. 17, 1283–1289 (2011).

    CAS  PubMed  Google Scholar 

  52. Adams, J. The proteasome: a suitable antineoplastic target. Nature Rev. Cancer 4, 349–360 (2004).

    CAS  Google Scholar 

  53. Inuzuka, H. et al. SCF(FBW7) regulates cellular apoptosis by targeting MCL1 for ubiquitylation and destruction. Nature 471, 104–109 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Wertz, I. E. et al. Sensitivity to antitubulin chemotherapeutics is regulated by MCL1 and FBW7. Nature 471, 110–114 (2011).

    CAS  PubMed  Google Scholar 

  55. Nijhawan, D. et al. Cancer vulnerabilities unveiled by genomic loss. Cell 150, 842–854 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Luo, J. et al. A genome-wide RNAi screen identifies multiple synthetic lethal interactions with the Ras oncogene. Cell 137, 835–848 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Groll, M., Berkers, C. R., Ploegh, H. L. & Ovaa, H. Crystal structure of the boronic acid-based proteasome inhibitor bortezomib in complex with the yeast 20S proteasome. Structure 14, 451–456 (2006).

    CAS  PubMed  Google Scholar 

  58. Parlati, F. et al. Carfilzomib can induce tumor cell death through selective inhibition of the chymotrypsin-like activity of the proteasome. Blood 114, 3439–3447 (2009).

    CAS  PubMed  Google Scholar 

  59. Richardson, P. G. et al. Extended follow-up of a phase 3 trial in relapsed multiple myeloma: final time-to-event results of the APEX trial. Blood 110, 3557–3560 (2007).

    CAS  PubMed  Google Scholar 

  60. Vij, R. et al. An open-label, single-arm, phase 2 (PX-171-004) study of single-agent carfilzomib in bortezomib-naive patients with relapsed and/or refractory multiple myeloma. Blood 119, 5661–5670 (2012).

    PubMed  PubMed Central  Google Scholar 

  61. D'Arcy, P. et al. Inhibition of proteasome deubiquitinating activity as a new cancer therapy. Nature Med. 17, 1636–1640 (2011).

    CAS  PubMed  Google Scholar 

  62. Lee, B. H. et al. Enhancement of proteasome activity by a small-molecule inhibitor of USP14. Nature 467, 179–184 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Ceccarelli, D. F. et al. An allosteric inhibitor of the human Cdc34 ubiquitin-conjugating enzyme. Cell 145, 1075–1087 (2011).

    CAS  PubMed  Google Scholar 

  64. Soucy, T. A. et al. An inhibitor of NEDD8-activating enzyme as a new approach to treat cancer. Nature 458, 732–736 (2009).

    CAS  PubMed  Google Scholar 

  65. Chauhan, D. et al. A small molecule inhibitor of ubiquitin-specific protease-7 induces apoptosis in multiple myeloma cells and overcomes bortezomib resistance. Cancer Cell 22, 345–358 (2012). References 61–65 describe novel approaches to target the ubiquitin–proteasome system.

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Stresemann, C. & Lyko, F. Modes of action of the DNA methyltransferase inhibitors azacytidine and decitabine. Int. J. Cancer 123, 8–13 (2008).

    CAS  PubMed  Google Scholar 

  67. Issa, J. P., Kantarjian, H. M. & Kirkpatrick, P. Azacitidine. Nature Rev. Drug Discov. 4, 275–276 (2005).

    CAS  Google Scholar 

  68. Kantarjian, H. et al. Decitabine improves patient outcomes in myelodysplastic syndromes: results of a phase III randomized study. Cancer 106, 1794–1803 (2006).

    CAS  PubMed  Google Scholar 

  69. Bolden, J. E., Peart, M. J. & Johnstone, R. W. Anticancer activities of histone deacetylase inhibitors. Nature Rev. Drug Discov. 5, 769–784 (2006).

    CAS  Google Scholar 

  70. Minucci, S. & Pelicci, P. G. Histone deacetylase inhibitors and the promise of epigenetic (and more) treatments for cancer. Nature Rev. Cancer 6, 38–51 (2006).

    CAS  Google Scholar 

  71. Mann, B. S., Johnson, J. R., Cohen, M. H., Justice, R. & Pazdur, R. FDA approval summary: vorinostat for treatment of advanced primary cutaneous T-cell lymphoma. Oncol. 12, 1247–1252 (2007).

    CAS  Google Scholar 

  72. Coiffier, B. et al. Results from a pivotal, open-label, phase II study of romidepsin in relapsed or refractory peripheral T-cell lymphoma after prior systemic therapy. J. Clin. Oncol. 30, 631–636 (2012).

    CAS  PubMed  Google Scholar 

  73. Beyer, U., Moll-Rocek, J., Moll, U. M. & Dobbelstein, M. Endogenous retrovirus drives hitherto unknown proapoptotic p63 isoforms in the male germ line of humans and great apes. Proc. Natl Acad. Sci. USA 108, 3624–3629 (2011). This paper describes the massively augmented, LTR-driven, Tp63 expression as a result of HDAC inhibition.

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Dawson, M. A. et al. Inhibition of BET recruitment to chromatin as an effective treatment for MLL-fusion leukaemia. Nature 478, 529–533 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  76. Mertz, J. A. et al. Targeting MYC dependence in cancer by inhibiting BET bromodomains. Proc. Natl Acad. Sci. USA 108, 16669–16674 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Zuber, J. et al. RNAi screen identifies Brd4 as a therapeutic target in acute myeloid leukaemia. Nature 478, 524–528 (2011). References 74–78 establish bromodomains as cancer drug targets.

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Albert, M. & Helin, K. Histone methyltransferases in cancer. Seminars Cell Dev. Biol. 21, 209–220 (2010).

    CAS  Google Scholar 

  80. Varambally, S. et al. Genomic loss of microRNA-101 leads to overexpression of histone methyltransferase EZH2 in cancer. Science 322, 1695–1699 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  83. Qi, W. et al. Selective inhibition of Ezh2 by a small molecule inhibitor blocks tumor cells proliferation. Proc. Natl Acad. Sci. USA 109, 21360–21365 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  84. Knutson, S. K. et al. Durable tumor regression in genetically altered malignant rhabdoid tumors by inhibition of methyltransferase EZH2. Proc. Natl Acad. Sci. USA 110, 7922–7927 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  85. Wilson, B. G. et al. Epigenetic antagonism between polycomb and SWI/SNF complexes during oncogenic transformation. Cancer Cell 18, 316–328 (2010). References 80–85 describe the establishment of the PRC2 component EZH2 as a potential cancer drug target.

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Sharma, S. V. et al. A chromatin-mediated reversible drug-tolerant state in cancer cell subpopulations. Cell 141, 69–80 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  87. Zender, L. et al. An oncogenomics-based in vivo RNAi screen identifies tumor suppressors in liver cancer. Cell 135, 852–864 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Zender, L. et al. Identification and validation of oncogenes in liver cancer using an integrative oncogenomic approach. Cell 125, 1253–1267 (2006). References 87 and 88 describe in vivo loss-of-function screens to identify high-priority cancer drug targets, including members of cellular machineries.

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Pepperkok, R. & Ellenberg, J. High-throughput fluorescence microscopy for systems biology. Nature Rev. Mol. Cell Biol. 7, 690–696 (2006).

    CAS  Google Scholar 

  90. Scheffner, M., Werness, B. A., Huibregtse, J. M., Levine, A. J. & Howley, P. M. The E6 oncoprotein encoded by human papillomavirus types 16 and 18 promotes the degradation of p53. Cell 63, 1129–1136 (1990).

    CAS  PubMed  Google Scholar 

  91. Blachere, N. E. et al. Heat shock protein-peptide complexes, reconstituted in vitro, elicit peptide-specific cytotoxic T lymphocyte response and tumor immunity. J. Exp. Med. 186, 1315–1322 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Mann, J. Natural products in cancer chemotherapy: past, present and future. Nature Rev. Cancer 2, 143–148 (2002).

    CAS  Google Scholar 

  93. Pommier, Y. & Marchand, C. Interfacial inhibitors: targeting macromolecular complexes. Nature Rev. Drug Discov. 11, 25–36 (2012).

    CAS  Google Scholar 

  94. Sali, A., Glaeser, R., Earnest, T. & Baumeister, W. From words to literature in structural proteomics. Nature 422, 216–225 (2003).

    CAS  PubMed  Google Scholar 

  95. Will, C. L. & Lührmann, R. Spliceosome structure and function. Cold Spring Harb. Perspect. Biol. 3, a003707 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  96. Bonnal, S., Vigevani, L. & Valcarcel, J. The spliceosome as a target of novel antitumour drugs. Nature Rev. Drug Discov. 11, 847–859 (2012).

    CAS  Google Scholar 

  97. Schreiber, S. L. et al. Towards patient-based cancer therapeutics. Nature Biotech. 28, 904–906 (2010).

    CAS  Google Scholar 

  98. Chen, Z. et al. A murine lung cancer co-clinical trial identifies genetic modifiers of therapeutic response. Nature 483, 613–617 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  99. Goldman, J. W. et al. A first in human, safety, pharmacokinetics, and clinical activity phase I study of once weekly administration of the Hsp90 inhibitor ganetespib (STA-9090) in patients with solid malignancies. BMC Cancer 13, 152 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  100. 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 

  101. Grem, J. L. et al. Phase I and pharmacologic study of 17-(allylamino)-17-demethoxygeldanamycin in adult patients with solid tumors. J. Clin. Oncol. 23, 1885–1893 (2005).

    CAS  PubMed  Google Scholar 

  102. Richardson, P. G. et al. Bortezomib or high-dose dexamethasone for relapsed multiple myeloma. N. Engl. J. Med. 352, 2487–2498 (2005).

    CAS  PubMed  Google Scholar 

  103. Chen, J. et al. A restricted cell population propagates glioblastoma growth after chemotherapy. Nature 488, 522–526 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  104. Balko, J. M. et al. Profiling of residual breast cancers after neoadjuvant chemotherapy identifies DUSP4 deficiency as a mechanism of drug resistance. Nature Med. 18, 1052–1059 (2012).

    CAS  PubMed  Google Scholar 

  105. Gupta, P. B. et al. Identification of selective inhibitors of cancer stem cells by high-throughput screening. Cell 138, 645–659 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  106. Gupta, P. B., Chaffer, C. L. & Weinberg, R. A. Cancer stem cells: mirage or reality? Nature Med. 15, 1010–1012 (2009).

    CAS  PubMed  Google Scholar 

  107. Garnett, M. J. et al. Systematic identification of genomic markers of drug sensitivity in cancer cells. Nature 483, 570–575 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  108. 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 

  109. Tredan, O., Galmarini, C. M., Patel, K. & Tannock, I. F. Drug resistance and the solid tumor microenvironment. J. Natl Cancer Inst. 99, 1441–1454 (2007).

    CAS  PubMed  Google Scholar 

  110. Provenzano, P. P. et al. Enzymatic targeting of the stroma ablates physical barriers to treatment of pancreatic ductal adenocarcinoma. Cancer Cell 21, 418–429 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  111. Olive, K. P. et al. Inhibition of Hedgehog signaling enhances delivery of chemotherapy in a mouse model of pancreatic cancer. Science 324, 1457–1461 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Yauch, R. L. et al. A paracrine requirement for hedgehog signalling in cancer. Nature 455, 406–410 (2008).

    CAS  PubMed  Google Scholar 

  113. Ahmed, F., Steele, J. C., Herbert, J. M., Steven, N. M. & Bicknell, R. Tumor stroma as a target in cancer. Curr. Cancer Drug Targets 8, 447–453 (2008).

    CAS  PubMed  Google Scholar 

  114. Kelloff, G. J. & Sigman, C. C. Cancer biomarkers: selecting the right drug for the right patient. Nature Rev. Drug Discov. 11, 201–214 (2012).

    CAS  Google Scholar 

  115. Vogelstein, B. & Kinzler, K. W. Cancer genes and the pathways they control. Nature Med. 10, 789–799 (2004). This study describes the major signalling pathways affected by oncogenic mutations.

    CAS  PubMed  Google Scholar 

  116. Walther, A. et al. Genetic prognostic and predictive markers in colorectal cancer. Nature Rev. Cancer 9, 489–499 (2009).

    CAS  Google Scholar 

  117. McShane, L. M. et al. REporting recommendations for tumor MARKer prognostic studies (REMARK). Nature Clin. Pract. Urol. 2, 416–422 (2005).

    CAS  Google Scholar 

  118. Kastan, M. B. & Bartek, J. Cell-cycle checkpoints and cancer. Nature 432, 316–323 (2004).

    CAS  PubMed  Google Scholar 

  119. Zhou, B. B. & Bartek, J. Targeting the checkpoint kinases: chemosensitization versus chemoprotection. Nature Rev. Cancer 4, 216–225 (2004).

    CAS  Google Scholar 

  120. Sabatini, D. M. mTOR and cancer: insights into a complex relationship. Nature Rev. Cancer 6, 729–734 (2006).

    CAS  Google Scholar 

  121. Garcia, M. A. et al. Impact of protein kinase PKR in cell biology: from antiviral to antiproliferative action. Microbiol. Mol. Biol. Rev. 70, 1032–1060 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Farassati, F., Yang, A. D. & Lee, P. W. Oncogenes in Ras signalling pathway dictate host-cell permissiveness to herpes simplex virus 1. Nature Cell Biol. 3, 745–750 (2001).

    CAS  PubMed  Google Scholar 

  123. Strong, J. E., Coffey, M. C., Tang, D., Sabinin, P. & Lee, P. W. The molecular basis of viral oncolysis: usurpation of the Ras signaling pathway by reovirus. EMBO J. 17, 3351–3362 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  124. Chlebowski, R. T. et al. Breast cancer after use of estrogen plus progestin in postmenopausal women. N. Engl. J. Med. 360, 573–587 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  125. Levine, A. J. & Oren, M. The first 30 years of p53: growing ever more complex. Nature Rev. Cancer 9, 749–758 (2009).

    CAS  Google Scholar 

  126. Bandyopadhyay, D., Mishra, A. & Medrano, E. E. Overexpression of histone deacetylase 1 confers resistance to sodium butyrate-mediated apoptosis in melanoma cells through a p53-mediated pathway. Cancer Res. 64, 7706–7710 (2004).

    CAS  PubMed  Google Scholar 

  127. Fantin, V. R. & Richon, V. M. Mechanisms of resistance to histone deacetylase inhibitors and their therapeutic implications. Clin. Cancer Res. 13, 7237–7242 (2007).

    CAS  PubMed  Google Scholar 

  128. Oerlemans, R. et al. Molecular basis of bortezomib resistance: proteasome subunit beta5 (PSMB5) gene mutation and overexpression of PSMB5 protein. Blood 112, 2489–2499 (2008).

    CAS  PubMed  Google Scholar 

  129. Kaelin, W. G. Jr. The concept of synthetic lethality in the context of anticancer therapy. Nature Rev. Cancer 5, 689–698 (2005). This paper describes the principle of synthetic lethality, extended from genetics to cancer therapy.

    CAS  Google Scholar 

  130. Davies, J. & Davies, D. Origins and evolution of antibiotic resistance. Microbiol. Mol. Biol. Rev. 74, 417–433 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  131. Lukas, J., Lukas, C. & Bartek, J. More than just a focus: the chromatin response to DNA damage and its role in genome integrity maintenance. Nature Cell Biol. 13, 1161–1169 (2011).

    CAS  PubMed  Google Scholar 

  132. Luijsterburg, M. S. & van Attikum, H. Chromatin and the DNA damage response: the cancer connection. Mol. Oncol. 5, 349–367 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Prenzel, T. et al. Estrogen-dependent gene transcription in human breast cancer cells relies upon proteasome-dependent monoubiquitination of histone H2B. Cancer Res. 71, 5739–5753 (2011).

    CAS  PubMed  Google Scholar 

  134. Fotheringham, S. et al. Genome-wide loss-of-function screen reveals an important role for the proteasome in HDAC inhibitor-induced apoptosis. Cancer Cell 15, 57–66 (2009).

    CAS  PubMed  Google Scholar 

  135. Kovacs, J. J. et al. HDAC6 regulates Hsp90 acetylation and chaperone-dependent activation of glucocorticoid receptor. Mol. Cell 18, 601–607 (2005). This paper describes HDAC6 as a regulator of HSP90.

    CAS  PubMed  Google Scholar 

  136. Bouwman, P. & Jonkers, J. The effects of deregulated DNA damage signalling on cancer chemotherapy response and resistance. Nature Rev. Cancer 12, 587–598 (2012).

    CAS  Google Scholar 

  137. Savage, P., Stebbing, J., Bower, M. & Crook, T. Why does cytotoxic chemotherapy cure only some cancers? Nature Clin. Pract. Oncol. 6, 43–52 (2009).

    CAS  Google Scholar 

  138. Edwards, S. L. et al. Resistance to therapy caused by intragenic deletion in BRCA2. Nature 451, 1111–1115 (2008).

    CAS  PubMed  Google Scholar 

  139. Gibson, B. A. & Kraus, W. L. New insights into the molecular and cellular functions of poly(ADP-ribose) and PARPs. Nature Rev. Mol. Cell Biol. 13, 411–424 (2012).

    CAS  Google Scholar 

  140. Rouleau, M., Patel, A., Hendzel, M. J., Kaufmann, S. H. & Poirier, G. G. PARP inhibition: PARP1 and beyond. Nature Rev. Cancer 10, 293–301 (2010).

    CAS  Google Scholar 

  141. Bryant, H. E. et al. Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature 434, 913–917 (2005).

    CAS  PubMed  Google Scholar 

  142. Johnson, N. et al. Compromised CDK1 activity sensitizes BRCA-proficient cancers to PARP inhibition. Nature Med. 17, 875–882 (2011).

    CAS  PubMed  Google Scholar 

  143. Curtin, N. J. DNA repair dysregulation from cancer driver to therapeutic target. Nature Rev. Cancer 12, 801–817 (2012).

    CAS  Google Scholar 

  144. Kopper, F. et al. Damage-induced DNA replication stalling relies on MAPK-activated protein kinase 2 activity. Proc. Natl Acad. Sci. USA 110, 16856–16861 (2013).

    PubMed  PubMed Central  Google Scholar 

  145. Kaestner, P. & Bastians, H. Mitotic drug targets. J. Cell. Biochem. 111, 258–265 (2010).

    CAS  PubMed  Google Scholar 

  146. Jordan, M. A. & Wilson, L. Microtubules as a target for anticancer drugs. Nature Rev. Cancer 4, 253–265 (2004).

    CAS  Google Scholar 

  147. Domingo-Domenech, J. et al. Suppression of acquired docetaxel resistance in prostate cancer through depletion of Notch- and Hedgehog-dependent tumor-initiating cells. Cancer Cell 22, 373–388 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  148. Zeng, X. et al. Pharmacologic inhibition of the anaphase-promoting complex induces a spindle checkpoint-dependent mitotic arrest in the absence of spindle damage. Cancer Cell 18, 382–395 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  149. Huang, H. C., Shi, J., Orth, J. D. & Mitchison, T. J. Evidence that mitotic exit is a better cancer therapeutic target than spindle assembly. Cancer Cell 16, 347–358 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  150. Kwon, M. et al. Mechanisms to suppress multipolar divisions in cancer cells with extra centrosomes. Genes Dev. 22, 2189–2203 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  151. Janssen, A., Kops, G. J. & Medema, R. H. Elevating the frequency of chromosome mis-segregation as a strategy to kill tumor cells. Proc. Natl Acad. Sci. USA 106, 19108–19113 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  152. Bantscheff, M. et al. Quantitative chemical proteomics reveals mechanisms of action of clinical ABL kinase inhibitors. Nature Biotech. 25, 1035–1044 (2007).

    CAS  Google Scholar 

  153. Sliwkowski, M. X. & Mellman, I. Antibody therapeutics in cancer. Science 341, 1192–1198 (2013).

    CAS  PubMed  Google Scholar 

  154. Scott, A. M., Wolchok, J. D. & Old, L. J. Antibody therapy of cancer. Nature Rev. Cancer 12, 278–287 (2012).

    CAS  Google Scholar 

  155. Verma, S. et al. Trastuzumab emtansine for HER2-positive advanced breast cancer. N. Engl. J. Med. 367, 1783–1791 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  156. Younes, A. et al. Results of a pivotal phase II study of brentuximab vedotin for patients with relapsed or refractory Hodgkin's lymphoma. J. Clin. Oncol. 30, 2183–2189 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  157. Chambers, C. A., Kuhns, M. S., Egen, J. G. & Allison, J. P. CTLA-4-mediated inhibition in regulation of T cell responses: mechanisms and manipulation in tumor immunotherapy. Annu. Rev. Immunol. 19, 565–594 (2001).

    CAS  PubMed  Google Scholar 

  158. Topalian, S. L., Drake, C. G. & Pardoll, D. M. Targeting the PD-1/B7-H1(PD-L1) pathway to activate anti-tumor immunity. Curr. Opin. Immunol. 24, 207–212 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  159. Hodi, F. S. et al. Improved survival with ipilimumab in patients with metastatic melanoma. N. Engl. J. Med. 363, 711–723 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  160. Topalian, S. L. et al. Safety, activity, and immune correlates of anti-PD-1 antibody in cancer. N. Engl. J. Med. 366, 2443–2454 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  161. Osborne, R. J. et al. Phase III trial of weekly methotrexate or pulsed dactinomycin for low-risk gestational trophoblastic neoplasia: a gynecologic oncology group study. J. Clin. Oncol. 29, 825–831 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  162. Bywater, M. J. et al. Inhibition of RNA polymerase I as a therapeutic strategy to promote cancer-specific activation of p53. Cancer Cell 22, 51–65 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  163. Lum, P. Y. et al. Discovering modes of action for therapeutic compounds using a genome-wide screen of yeast heterozygotes. Cell 116, 121–137 (2004).

    CAS  PubMed  Google Scholar 

  164. Rubbi, C. P. & Milner, J. Disruption of the nucleolus mediates stabilization of p53 in response to DNA damage and other stresses. EMBO J. 22, 6068–6077 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  165. Kurki, S. et al. Nucleolar protein NPM interacts with HDM2 and protects tumor suppressor protein p53 from HDM2-mediated degradation. Cancer Cell 5, 465–475 (2004).

    CAS  PubMed  Google Scholar 

  166. Sasaki, M. et al. Regulation of the MDM2-P53 pathway and tumor growth by PICT1 via nucleolar RPL11. Nature Med. 17, 944–951 (2011).

    CAS  PubMed  Google Scholar 

  167. Hein, N., Hannan, K. M., George, A. J., Sanij, E. & Hannan, R. D. The nucleolus: an emerging target for cancer therapy. Trends Mol. Med. 19, 643–654 (2013).

    CAS  PubMed  Google Scholar 

  168. Kaida, D. et al. Spliceostatin A targets SF3b and inhibits both splicing and nuclear retention of pre-mRNA. Nature Chem. Biol. 3, 576–583 (2007).

    CAS  Google Scholar 

  169. Roybal, G. A. & Jurica, M. S. Spliceostatin A inhibits spliceosome assembly subsequent to prespliceosome formation. Nucleic Acids Res. 38, 6664–6672 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  170. Nakajima, H. et al. New antitumor substances, FR901463, FR901464 and FR901465. II. Activities against experimental tumors in mice and mechanism of action. J. Antibiot. 49, 1204–1211 (1996).

    CAS  Google Scholar 

  171. Filipowicz, W., Bhattacharyya, S. N. & Sonenberg, N. Mechanisms of post-transcriptional regulation by microRNAs: are the answers in sight? Nature Rev. Genetics 9, 102–114 (2008).

    CAS  Google Scholar 

  172. Calin, G. A. & Croce, C. M. MicroRNA-cancer connection: the beginning of a new tale. Cancer Res. 66, 7390–7394 (2006).

    CAS  PubMed  Google Scholar 

  173. Tan, G. S. et al. Small molecule inhibition of RISC loading. ACS Chem. Biol. 7, 403–410 (2012).

    CAS  PubMed  Google Scholar 

  174. Pare, J. M. et al. Hsp90 regulates the function of argonaute 2 and its recruitment to stress granules and P-bodies. Mol. Biol. Cell 20, 3273–3284 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  175. Kumar, M. S. et al. Dicer1 functions as a haploinsufficient tumor suppressor. Genes Dev. 23, 2700–2704 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  176. Davis-Dusenbery, B. N. & Hata, A. MicroRNA in cancer: the involvement of aberrant microRNA biogenesis regulatory pathways. Genes Cancer 1, 1100–1114 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  177. Ruggero, D. & Pandolfi, P. P. Does the ribosome translate cancer? Nature Rev. Cancer 3, 179–192 (2003).

    CAS  Google Scholar 

  178. Wetzler, M. & Segal, D. Omacetaxine as an anticancer therapeutic: what is old is new again. Curr. Pharm. Design 17, 59–64 (2011).

    CAS  Google Scholar 

  179. Cortes, J. et al. Phase 2 study of subcutaneous omacetaxine mepesuccinate after TKI failure in patients with chronic-phase CML with T315I mutation. Blood 120, 2573–2580 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  180. Culjkovic-Kraljacic, B., Baguet, A., Volpon, L., Amri, A. & Borden, K. L. The oncogene eIF4E reprograms the nuclear pore complex to promote mRNA export and oncogenic transformation. Cell Rep. 2, 207–215 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  181. Kentsis, A., Topisirovic, I., Culjkovic, B., Shao, L. & Borden, K. L. Ribavirin suppresses eIF4E-mediated oncogenic transformation by physical mimicry of the 7-methyl guanosine mRNA cap. Proc. Natl Acad. Sci. USA 101, 18105–18110 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  182. Assouline, S. et al. Molecular targeting of the oncogene eIF4E in acute myeloid leukemia (AML): a proof-of-principle clinical trial with ribavirin. Blood 114, 257–260 (2009).

    CAS  PubMed  Google Scholar 

  183. Pettersson, F. et al. Ribavirin treatment effects on breast cancers overexpressing eIF4E, a biomarker with prognostic specificity for luminal B-type breast cancer. Clin. Cancer Res. 17, 2874–2884 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  184. Aloisi, A. et al. BCR-ABL nuclear entrapment kills human CML cells: ex vivo study on 35 patients with the combination of imatinib mesylate and leptomycin B. Blood 107, 1591–1598 (2006).

    CAS  PubMed  Google Scholar 

  185. Azmi, A. S. et al. Selective inhibitors of nuclear export block pancreatic cancer cell proliferation and reduce tumor growth in mice. Gastroenterology 144, 447–456 (2012).

    PubMed  Google Scholar 

  186. Mutka, S. C. et al. Identification of nuclear export inhibitors with potent anticancer activity in vivo. Cancer Res. 69, 510–517 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  187. Etchin, J. et al. Antileukemic activity of nuclear export inhibitors that spare normal hematopoietic cells. Leukemia 27, 66–74 (2013).

    CAS  PubMed  Google Scholar 

  188. Vander Heiden, M. G., Cantley, L. C. & Thompson, C. B. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science 324, 1029–1033 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  189. Yang, W. et al. ERK1/2-dependent phosphorylation and nuclear translocation of PKM2 promotes the Warburg effect. Nature Cell Biol. 15, 124 (2012).

    Google Scholar 

  190. Christofk, H. R., Vander Heiden, M. G., Wu, N., Asara, J. M. & Cantley, L. C. Pyruvate kinase M2 is a phosphotyrosine-binding protein. Nature 452, 181–186 (2008).

    CAS  PubMed  Google Scholar 

  191. Hitosugi, T. et al. Tyrosine phosphorylation inhibits PKM2 to promote the Warburg effect and tumor growth. Sci. Signal. 2, ra73 (2009).

    PubMed  PubMed Central  Google Scholar 

  192. Schnabel, J. Targeting tumour metabolism. Nature Rev. Drug Discov. 9, 503–504 (2010).

    Google Scholar 

  193. Teicher, B. A., Linehan, W. M. & Helman, L. J. Targeting cancer metabolism. Clin. Cancer Res. 18, 5537–5545 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  194. Colleoni, M. et al. Classical cyclophosphamide, methotrexate, and fluorouracil chemotherapy is more effective in triple-negative, node-negative breast cancer: results from two randomized trials of adjuvant chemoendocrine therapy for node-negative breast cancer. J. Clin. Oncol. 28, 2966–2973 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  195. Wang, F. et al. Targeted inhibition of mutant IDH2 in leukemia cells induces cellular differentiation. Science 340, 622–626 (2013).

    CAS  PubMed  Google Scholar 

  196. Rohle, D. et al. An inhibitor of mutant IDH1 delays growth and promotes differentiation of glioma cells. Science 340, 626–630 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  197. Danial, N. N. et al. BAD and glucokinase reside in a mitochondrial complex that integrates glycolysis and apoptosis. Nature 424, 952–956 (2003).

    CAS  PubMed  Google Scholar 

  198. Raj, L. et al. Selective killing of cancer cells by a small molecule targeting the stress response to ROS. Nature 475, 231–234 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  199. Adams, D. J. et al. Synthesis, cellular evaluation, and mechanism of action of piperlongumine analogs. Proc. Natl Acad. Sci. USA 109, 15115–15120 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  200. Vaseva, A. V. et al. p53 opens the mitochondrial permeability transition pore to trigger necrosis. Cell 149, 1536–1548 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  201. Ni Chonghaile, T. & Letai, A. Mimicking the BH3 domain to kill cancer cells. Oncogene 27 (Suppl. 1), 149–157 (2008).

    Google Scholar 

  202. Ni Chonghaile, T. et al. Pretreatment mitochondrial priming correlates with clinical response to cytotoxic chemotherapy. Science 334, 1129–1133 (2011).

    PubMed  Google Scholar 

  203. Ready, N. et al. Double-blind, placebo-controlled, randomized phase 2 study of the proapoptotic agent AT-101 plus docetaxel, in second-line non-small cell lung cancer. J. Thorac. Oncol. 6, 781–785 (2011).

    PubMed  Google Scholar 

  204. Sonpavde, G. et al. Randomized phase II trial of docetaxel plus prednisone in combination with placebo or AT-101, an oral small molecule Bcl-2 family antagonist, as first-line therapy for metastatic castration-resistant prostate cancer. Ann. Oncol. 23, 1803–1808 (2012).

    CAS  PubMed  Google Scholar 

  205. Kim, I., Xu, W. & Reed, J. C. Cell death and endoplasmic reticulum stress: disease relevance and therapeutic opportunities. Nature Rev. Drug Discov. 7, 1013–1030 (2008).

    CAS  Google Scholar 

  206. Denmeade, S. R. et al. Engineering a prostate-specific membrane antigen-activated tumor endothelial cell prodrug for cancer therapy. Sci. Transl. Med. 4, 140ra186 (2012).

    Google Scholar 

  207. Rubinsztein, D. C., Gestwicki, J. E., Murphy, L. O. & Klionsky, D. J. Potential therapeutic applications of autophagy. Nature Rev. Drug Discov. 6, 304–312 (2007).

    CAS  Google Scholar 

  208. Janku, F., McConkey, D. J., Hong, D. S. & Kurzrock, R. Autophagy as a target for anticancer therapy. Nature Rev. Clin. Oncol. 8, 528–539 (2011).

    CAS  Google Scholar 

  209. Kaufmann, A. M. & Krise, J. P. Lysosomal sequestration of amine-containing drugs: analysis and therapeutic implications. J. Pharm. Sci. 96, 729–746 (2007).

    CAS  PubMed  Google Scholar 

  210. Goldberg, S. B. et al. A phase I study of erlotinib and hydroxychloroquine in advanced non-small-cell lung cancer. J. Thorac. Oncol. 7, 1602–1608 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  211. Saftig, P. & Sandhoff, K. Cancer: killing from the inside. Nature 502, 312–313 (2013).

    CAS  PubMed  Google Scholar 

  212. Petersen, N. H. et al. Transformation-associated changes in sphingolipid metabolism sensitize cells to lysosomal cell death induced by inhibitors of acid sphingomyelinase. Cancer Cell 24, 379–393 (2013). This paper describes lysosome function as a cancer drug target.

    CAS  PubMed  Google Scholar 

  213. Belotti, D. et al. Paclitaxel (Taxol(R)) inhibits motility of paclitaxel-resistant human ovarian carcinoma cells. Clin. Cancer Res. 2, 1725–1730 (1996).

    CAS  PubMed  Google Scholar 

  214. Bonezzi, K. et al. Inhibition of SIRT2 potentiates the anti-motility activity of taxanes: implications for antineoplastic combination therapies. Neoplasia 14, 846–854 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  215. Rotty, J. D., Wu, C. & Bear, J. E. New insights into the regulation and cellular functions of the ARP2/3 complex. Nature Rev. Mol. Cell Biol. 14, 7–12 (2013).

    CAS  Google Scholar 

  216. Fernando, H. S., Kynaston, H. G. & Jiang, W. G. WASP and WAVE proteins: vital intrinsic regulators of cell motility and their role in cancer (review). Int. J. Mol. Med. 23, 141–148 (2009).

    CAS  PubMed  Google Scholar 

  217. Teng, Y., Ngoka, L., Mei, Y., Lesoon, L. & Cowell, J. K. HSP90 and HSP70 proteins are essential for stabilization and activation of WASF3 metastasis-promoting protein. J. Biol. Chem. 287, 10051–10059 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  218. DeVita, V. T. Jr & Chu, E. A history of cancer chemotherapy. Cancer Res. 68, 8643–8653 (2008).

    CAS  PubMed  Google Scholar 

  219. Farber, S. & Diamond, L. K. Temporary remissions in acute leukemia in children produced by folic acid antagonist, 4-aminopteroyl-glutamic acid. N. Engl. J. Med. 238, 787–793 (1948).

    CAS  PubMed  Google Scholar 

  220. Li, M. C., Hertz, R. & Bergenstal, D. M. Therapy of choriocarcinoma and related trophoblastic tumors with folic acid and purine antagonists. N. Engl. J. Med. 259, 66–74 (1958).

    CAS  PubMed  Google Scholar 

  221. Johnson, I. S., Armstrong, J. G., Gorman, M. & Burnett, J. P. Jr. The vinca alkaloids: a new class of oncolytic agents. Cancer Res. 23, 1390–1427 (1963).

    CAS  PubMed  Google Scholar 

  222. Freireich, E. J., Karon, M. & Frei, E. I. Quadruple combination therapy (VAMP) for acute lymphocytic leukemia of childhood. Proc. Am. Assoc. Cancer Res. 5, 20 (1964).

    Google Scholar 

  223. Wani, M. C., Taylor, H. L., Wall, M. E., Coggon, P. & McPhail, A. T. Plant antitumor agents. VI. The isolation and structure of taxol, a novel antileukemic and antitumor agent from Taxus brevifolia. J. Am. Chem. Soc. 93, 2325–2327 (1971).

    CAS  PubMed  Google Scholar 

  224. Wall, M. E. Camptothecin and taxol: discovery to clinic. Med. Res. Rev. 18, 299–314 (1998).

    CAS  PubMed  Google Scholar 

  225. Druker, B. J. et al. Effects of a selective inhibitor of the Abl tyrosine kinase on the growth of Bcr-Abl positive cells. Nature Med. 2, 561–566 (1996).

    CAS  PubMed  Google Scholar 

  226. Roche-Lestienne, C. et al. Several types of mutations of the Abl gene can be found in chronic myeloid leukemia patients resistant to STI571, and they can pre-exist to the onset of treatment. Blood 100, 1014–1018 (2002).

    CAS  PubMed  Google Scholar 

  227. von Bubnoff, N., Schneller, F., Peschel, C. & Duyster, J. BCR-ABL gene mutations in relation to clinical resistance of Philadelphia-chromosome-positive leukaemia to STI571: a prospective study. Lancet 359, 487–491 (2002).

    CAS  PubMed  Google Scholar 

  228. Baselga, J. et al. Phase II study of weekly intravenous recombinant humanized anti-p185HER2 monoclonal antibody in patients with HER2/neu-overexpressing metastatic breast cancer. J. Clin. Oncol. 14, 737–744 (1996).

    CAS  PubMed  Google Scholar 

  229. Davies, H. et al. Mutations of the BRAF gene in human cancer. Nature 417, 949–954 (2002).

    CAS  PubMed  Google Scholar 

  230. Yang, H. et al. RG7204 (PLX4032), a selective BRAFV600E inhibitor, displays potent antitumor activity in preclinical melanoma models. Cancer Res. 70, 5518–5527 (2010).

    CAS  PubMed  Google Scholar 

  231. Hoeflich, K. P. et al. Antitumor efficacy of the novel RAF inhibitor GDC-0879 is predicted by BRAFV600E mutational status and sustained extracellular signal-regulated kinase/mitogen-activated protein kinase pathway suppression. Cancer Res. 69, 3042–3051 (2009).

    CAS  PubMed  Google Scholar 

  232. Flaherty, K. T. et al. Inhibition of mutated, activated BRAF in metastatic melanoma. N. Engl. J. Med. 363, 809–819 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  233. Johannessen, C. M. et al. COT drives resistance to RAF inhibition through MAP kinase pathway reactivation. Nature 468, 968–972 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  234. Nazarian, R. et al. Melanomas acquire resistance to B-RAF(V600E) inhibition by RTK or N-RAS upregulation. Nature 468, 973–977 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  235. Yoshida, M., Kijima, M., Akita, M. & Beppu, T. Potent and specific inhibition of mammalian histone deacetylase both in vivo and in vitro by trichostatin A. J. Biol. Chem. 265, 17174–17179 (1990).

    CAS  PubMed  Google Scholar 

  236. Whitesell, L., Mimnaugh, E. G., De Costa, B., Myers, C. E. & Neckers, L. M. Inhibition of heat shock protein HSP90–pp60v-src heteroprotein complex formation by benzoquinone ansamycins: essential role for stress proteins in oncogenic transformation. Proc. Natl Acad. Sci. USA 91, 8324–8328 (1994).

    CAS  PubMed  PubMed Central  Google Scholar 

  237. Rock, K. L. et al. Inhibitors of the proteasome block the degradation of most cell proteins and the generation of peptides presented on MHC class I molecules. Cell 78, 761–771 (1994).

    CAS  PubMed  Google Scholar 

  238. Butler, L. M. et al. Suberoylanilide hydroxamic acid, an inhibitor of histone deacetylase, suppresses the growth of prostate cancer cells in vitro and in vivo. Cancer Res. 60, 5165–5170 (2000).

    CAS  PubMed  Google Scholar 

  239. Weisberg, E., Manley, P. W., Cowan-Jacob, S. W., Hochhaus, A. & Griffin, J. D. Second generation inhibitors of BCR-ABL for the treatment of imatinib-resistant chronic myeloid leukaemia. Nature Rev. Cancer 7, 345–356 (2007).

    CAS  Google Scholar 

  240. Baselga, J. & Swain, S. M. Novel anticancer targets: revisiting ERBB2 and discovering ERBB3. Nature Rev. Cancer 9, 463–475 (2009).

    CAS  Google Scholar 

  241. Stern, H. M. Improving treatment of HER2-positive cancers: opportunities and challenges. Sci. Transl. Med. 4, 127rv122 (2012).

    Google Scholar 

  242. Burris, H. A. Dual kinase inhibition in the treatment of breast cancer: initial experience with the EGFR/ErbB-2 inhibitor lapatinib. Oncol. 9 (Suppl. 3), 10–15 (2004).

    CAS  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  244. Poulikakos, P. I. et al. RAF inhibitor resistance is mediated by dimerization of aberrantly spliced BRAF(V600E). Nature 480, 387–390 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  245. Straussman, R. et al. Tumour micro-environment elicits innate resistance to RAF inhibitors through HGF secretion. Nature 487, 500–504 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  246. Villanueva, J. et al. Acquired resistance to BRAF inhibitors mediated by a RAF kinase switch in melanoma can be overcome by cotargeting MEK and IGF-1R/PI3K. Cancer Cell 18, 683–695 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  247. Metcalfe, C. & de Sauvage, F. J. Hedgehog fights back: mechanisms of acquired resistance against Smoothened antagonists. Cancer Res. 71, 5057–5061 (2011).

    CAS  PubMed  Google Scholar 

  248. Von Hoff, D. D. et al. Inhibition of the hedgehog pathway in advanced basal-cell carcinoma. N. Engl. J. Med. 361, 1164–1172 (2009).

    CAS  PubMed  Google Scholar 

  249. Rudin, C. M. et al. Treatment of medulloblastoma with hedgehog pathway inhibitor GDC-0449. N. Engl. J. Med. 361, 1173–1178 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  250. Verstovsek, S. et al. A double-blind, placebo-controlled trial of ruxolitinib for myelofibrosis. N. Engl. J. Med. 366, 799–807 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  251. Mesa, R. A., Yasothan, U. & Kirkpatrick, P. Ruxolitinib. Nature Rev. Drug Discov. 11, 103–104 (2012).

    CAS  Google Scholar 

  252. Deshpande, A. et al. Kinase domain mutations confer resistance to novel inhibitors targeting JAK2V617F in myeloproliferative neoplasms. Leukemia 26, 708–715 (2012).

    CAS  PubMed  Google Scholar 

  253. Engelman, J. A. et al. MET amplification leads to gefitinib resistance in lung cancer by activating ERBB3 signaling. Science 316, 1039–1043 (2007).

    CAS  PubMed  Google Scholar 

  254. Garofalo, M. et al. EGFR and MET receptor tyrosine kinase-altered microRNA expression induces tumorigenesis and gefitinib resistance in lung cancers. Nature Med. 18, 74–82 (2012).

    CAS  Google Scholar 

  255. Kobayashi, S. et al. EGFR mutation and resistance of non-small-cell lung cancer to gefitinib. N. Engl. J. Med. 352, 786–792 (2005).

    CAS  PubMed  Google Scholar 

  256. Bean, J. et al. MET amplification occurs with or without T790M mutations in EGFR mutant lung tumors with acquired resistance to gefitinib or erlotinib. Proc. Natl Acad. Sci. USA 104, 20932–20937 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  257. Wang, W. L. et al. Mechanisms of resistance to imatinib and sunitinib in gastrointestinal stromal tumor. Cancer Chemother. Pharmacol. 67 (Suppl. 1), 15–24 (2011).

    CAS  Google Scholar 

  258. Gajiwala, K. S. et al. KIT kinase mutants show unique mechanisms of drug resistance to imatinib and sunitinib in gastrointestinal stromal tumor patients. Proc. Natl Acad. Sci. USA 106, 1542–1547 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  259. Villanueva, A. & Llovet, J. M. Second-line therapies in hepatocellular carcinoma: emergence of resistance to sorafenib. Clin. Cancer Res. 18, 1824–1826 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  260. Chen, K. F. et al. Activation of phosphatidylinositol 3-kinase/Akt signaling pathway mediates acquired resistance to sorafenib in hepatocellular carcinoma cells. J. Pharmacol. Exp. Ther. 337, 155–161 (2011).

    CAS  PubMed  Google Scholar 

  261. Motzer, R. J. et al. Pazopanib versus sunitinib in metastatic renal-cell carcinoma. N. Engl. J. Med. 369, 722–731 (2013).

    CAS  PubMed  Google Scholar 

  262. Santoro, A. et al. Tivantinib for second-line treatment of advanced hepatocellular carcinoma: a randomised, placebo-controlled phase 2 study. Lancet Oncol. 14, 55–63 (2013).

    CAS  PubMed  Google Scholar 

  263. Baselga, J. et al. Phase II study of weekly intravenous trastuzumab (Herceptin) in patients with HER2/ neu-overexpressing metastatic breast cancer. Seminars Oncol. 26, 78–83 (1999).

    CAS  Google Scholar 

  264. Hudis, C. A. Trastuzumab — mechanism of action and use in clinical practice. N. Engl. J. Med. 357, 39–51 (2007).

    CAS  PubMed  Google Scholar 

  265. Katz, J., Janik, J. E. & Younes, A. Brentuximab vedotin (SGN-35). Clin. Cancer Res. 17, 6428–6436 (2011).

    CAS  PubMed  Google Scholar 

  266. Smith, M. R. Rituximab (monoclonal anti-CD20 antibody): mechanisms of action and resistance. Oncogene 22, 7359–7368 (2003).

    CAS  PubMed  Google Scholar 

  267. Bardelli, A. & Siena, S. Molecular mechanisms of resistance to cetuximab and panitumumab in colorectal cancer. J. Clin. Oncol. 28, 1254–1261 (2010).

    CAS  PubMed  Google Scholar 

  268. Montagut, C. et al. Identification of a mutation in the extracellular domain of the epidermal growth factor receptor conferring cetuximab resistance in colorectal cancer. Nature Med. 18, 221–223 (2012).

    CAS  PubMed  Google Scholar 

  269. Misale, S. et al. Emergence of KRAS mutations and acquired resistance to anti-EGFR therapy in colorectal cancer. Nature 486, 532–536 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  270. Loges, S., Schmidt, T. & Carmeliet, P. Mechanisms of resistance to anti-angiogenic therapy and development of third-generation anti-angiogenic drug candidates. Genes Cancer 1, 12–25 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  271. Holmgaard, R. B., Zamarin, D., Munn, D. H., Wolchok, J. D. & Allison, J. P. Indoleamine 2,3-dioxygenase is a critical resistance mechanism in antitumor T cell immunotherapy targeting CTLA-4. J. Exp. Med. 210, 1389–1402 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  272. Janne, P. A., Gray, N. & Settleman, J. Factors underlying sensitivity of cancers to small-molecule kinase inhibitors. Nature Rev. Drug Discov. 8, 709–723 (2009).

    CAS  Google Scholar 

  273. Lopez-Contreras, A. J., Gutierrez-Martinez, P., Specks, J., Rodrigo-Perez, S. & Fernandez-Capetillo, O. An extra allele of Chk1 limits oncogene-induced replicative stress and promotes transformation. J. Exp. Med. 209, 455–461 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  274. Murga, M. et al. Exploiting oncogene-induced replicative stress for the selective killing of Myc-driven tumors. Nature Struct. Mol. Biol. 18, 1331–1335 (2011).

    CAS  Google Scholar 

  275. Malumbres, M. Oncogene-induced mitotic stress: 53 and pRb get mad too. Cancer Cell 19, 691–692 (2011).

    CAS  PubMed  Google Scholar 

  276. Kops, G. J., Weaver, B. A. & Cleveland, D. W. On the road to cancer: aneuploidy and the mitotic checkpoint. Nature Rev. Cancer 5, 773–785 (2005).

    CAS  Google Scholar 

  277. Kaye, S. B. et al. Phase II, open-label, randomized, multicenter study comparing the efficacy and safety of olaparib, a poly (ADP-ribose) polymerase inhibitor, and pegylated liposomal doxorubicin in patients with BRCA1 or BRCA2 mutations and recurrent ovarian cancer. J. Clin. Oncol. 30, 372–379 (2012).

    CAS  PubMed  Google Scholar 

  278. Plummer, R. et al. A phase II study of the potent PARP inhibitor, rucaparib (PF-01367338, AG014699), with temozolomide in patients with metastatic melanoma demonstrating evidence of chemopotentiation. Cancer Chemother. Pharmacol. 71, 1191–1199 (2013).

    CAS  PubMed  Google Scholar 

  279. Kummar, S. et al. A phase I study of veliparib in combination with metronomic cyclophosphamide in adults with refractory solid tumors and lymphomas. Clin. Cancer Res. 18, 1726–1734 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  280. Sulli, G., Di Micco, R. & d'Adda di Fagagna, F. Crosstalk between chromatin state and DNA damage response in cellular senescence and cancer. Nature Rev. Cancer 12, 709–720 (2012).

    CAS  Google Scholar 

  281. Olsen, E. A. et al. Phase IIb multicenter trial of vorinostat in patients with persistent, progressive, or treatment refractory cutaneous T-cell lymphoma. J. Clin. Oncol. 25, 3109–3115 (2007).

    CAS  PubMed  Google Scholar 

  282. Duvic, M. et al. Phase 2 trial of oral vorinostat (suberoylanilide hydroxamic acid, SAHA) for refractory cutaneous T-cell lymphoma (CTCL). Blood 109, 31–39 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  283. Neri, P., Bahlis, N. J. & Lonial, S. Panobinostat for the treatment of multiple myeloma. Expert Opin. Investigat. Drugs 21, 733–747 (2012).

    CAS  Google Scholar 

  284. Kaminskas, E., Farrell, A. T., Wang, Y. C., Sridhara, R. & Pazdur, R. FDA drug approval summary: azacitidine (5-azacytidine, Vidaza) for injectable suspension. Oncologist 10, 176–182 (2005).

    CAS  PubMed  Google Scholar 

  285. Lubbert, M. et al. Low-dose decitabine versus best supportive care in elderly patients with intermediate- or high-risk myelodysplastic syndrome (MDS) ineligible for intensive chemotherapy: final results of the randomized phase III study of the European Organisation for Research and Treatment of Cancer Leukemia Group and the German MDS Study Group. J. Clin. Oncol. 29, 1987–1996 (2011).

    PubMed  Google Scholar 

  286. Kantarjian, H. M. et al. Multicenter, randomized, open-label, phase III trial of decitabine versus patient choice, with physician advice, of either supportive care or low-dose cytarabine for the treatment of older patients with newly diagnosed acute myeloid leukemia. J. Clin. Oncol. 30, 2670–2677 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  287. Morimoto, R. I. Proteotoxic stress and inducible chaperone networks in neurodegenerative disease and aging. Genes Dev. 22, 1427–1438 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  288. Socinski, M. A. et al. A multicenter Phase II study of ganetespib monotherapy in patients with genotypically defined advanced non-small cell lung cancer. Clin. Cancer Res. 19, 3068–3077 (2013).

    CAS  PubMed  Google Scholar 

  289. 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 

  290. Hrebackova, J., Hrabeta, J. & Eckschlager, T. Valproic acid in the complex therapy of malignant tumors. Curr. Drug Targets 11, 361–379 (2010).

    CAS  PubMed  Google Scholar 

  291. Duenas-Gonzalez, A. et al. Valproic acid as epigenetic cancer drug: preclinical, clinical and transcriptional effects on solid tumors. Cancer Treatment Rev. 34, 206–222 (2008).

    CAS  Google Scholar 

  292. Jagannath, S. et al. An open-label single-arm pilot phase II study (PX-171-003-A0) of low-dose, single-agent carfilzomib in patients with relapsed and refractory multiple myeloma. Clin. Lymphoma Myeloma Leuk. 12, 310–318 (2012).

    CAS  PubMed  Google Scholar 

  293. Siegel, D. S. et al. A phase 2 study of single-agent carfilzomib (PX-171-003-A1) in patients with relapsed and refractory multiple myeloma. Blood 120, 2817–2825 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  294. Boulon, S., Westman, B. J., Hutten, S., Boisvert, F. M. & Lamond, A. I. The nucleolus under stress. Mol. Cell 40, 216–227 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  295. Derheimer, F. A. et al. RPA and ATR link transcriptional stress to p53. Proc. Natl Acad. Sci. USA 104, 12778–12783 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  296. Gilkes, D. M., Chen, L. & Chen, J. MDMX regulation of p53 response to ribosomal stress. EMBO J. 25, 5614–5625 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  297. Schonthal, A. H. Pharmacological targeting of endoplasmic reticulum stress signaling in cancer. Biochem. Pharmacol. 85, 653–666 (2013).

    CAS  PubMed  Google Scholar 

  298. Xu, C., Bailly-Maitre, B. & Reed, J. C. Endoplasmic reticulum stress: cell life and death decisions. J. Clin. Invest. 115, 2656–2664 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  299. Thastrup, O., Cullen, P. J., Drobak, B. K., Hanley, M. R. & Dawson, A. P. Thapsigargin, a tumor promoter, discharges intracellular Ca2+ stores by specific inhibition of the endoplasmic reticulum Ca2+-ATPase. Proc. Natl Acad. Sci. USA 87, 2466–2470 (1990).

    CAS  PubMed  PubMed Central  Google Scholar 

  300. Ganley, I. G., Wong, P. M., Gammoh, N. & Jiang, X. Distinct autophagosomal-lysosomal fusion mechanism revealed by thapsigargin-induced autophagy arrest. Mol. Cell 42, 731–743 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  301. Ohuchi, K. et al. Stimulation of arachidonic acid metabolism in rat peritoneal macrophages by thapsigargin, a non-(12-O-tetradecanoylphorbol-13- acetate) (TPA)-type tumor promoter. J. Cancer Res. Clin. Oncol. 113, 319–324 (1987).

    CAS  PubMed  Google Scholar 

  302. Wellen, K. E. & Thompson, C. B. Cellular metabolic stress: considering how cells respond to nutrient excess. Mol. Cell 40, 323–332 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  303. Benz, C. C. & Yau, C. Ageing, oxidative stress and cancer: paradigms in parallax. Nature Rev. Cancer 8, 875–879 (2008).

    CAS  Google Scholar 

  304. Diaz, L. A. Jr. et al. The molecular evolution of acquired resistance to targeted EGFR blockade in colorectal cancers. Nature 486, 537–540 (2012). References 268 and 304 provide a mechanistic explanation for the resistance to EGFR blockade in colorectal cancer.

    CAS  PubMed  PubMed Central  Google Scholar 

  305. Wilson, T. R. et al. Widespread potential for growth-factor-driven resistance to anticancer kinase inhibitors. Nature 487, 505–509 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We are indebted to H. Bastians, Y. Begus-Nahrmann, A. P. Kudelka, N. Malek, A. Stolz and G. Wulf for critically reading the manuscript. We apologize to all colleagues whose important work could not be cited owing to space limitations. Our work was funded by the German Research Foundation (DFG), the German Cancer Aid, the German José Carreras Leukemia Foundation, the Wilhelm Sander foundation and the European Union (the GANNET53 project).

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Matthias Dobbelstein or Ute Moll.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

PowerPoint slides

Glossary

Antibody–drug conjugates

Therapeutics consisting of a tumour surface-targeting antibody (a whole monoclonal antibody or a single-chain variable fragment) linked via a chemical linker to a cytotoxic molecule with anticancer activity.

Biomarkers

Parameters that can be measured in patients — for example, to indicate the expression levels of a gene in tumour cells. In cancer treatment, biomarkers serve to indicate the likelihood of cancer progression, death, remission or a cure (these are known as prognostic biomarkers), or to assess how well a particular therapeutic regimen will work for a given patient (these are known as predictive biomarkers).

Hallmarks of cancer

The general phenotypic traits of cancers. The original article by Hanahan and Weinberg defines six such hallmarks: self-sufficiency in growth signals; insensitivity to anti-growth signals; evasion of apoptosis; limitless reproductive potential; sustained angiogenesis; and tissue invasion and metastasis. This list has since been expanded.

Kinobead technologies

The use of beads that are linked to kinase inhibitor molecules. These molecules bind to a subset of kinases from a cell lysate, allowing their subsequent identification and quantification by mass spectrometry.

Monoclonal antibodies

Antibodies generated by a clonal B cell population (traditionally fusion hybrids of individual B cells with a myeloma cell). For cancer therapy, monoclonal antibodies are used to target tumour-associated molecules on the surface of cancer and/or stromal cells or secreted factors.

Non-oncogene addiction

The phenomenon whereby cancer cells critically depend on a particular adaptive mechanism to a greater extent than normal cells, although the corresponding genes do not fulfil the criteria of oncogenes; this mechanism is typically composed of a complex multicomponent machinery.

Oncogene addiction

The phenomenon whereby cancer cells depend on the continuous hyperactivation of one or more oncogenes and their products for their survival.

Philadelphia chromosome

A reciprocal translocation of the chromosomes 9 and 22, designated t(9;22)(q34;q11), found in more than 90% of chronic myeloid leukaemias.

REMARK guidelines

Guidelines entitled 'Reporting recommendations for tumour marker prognostic studies', which recommend standards for the evaluation of, and reporting on, biomarkers in clinical studies on cancer, including “study design, preplanned hypotheses, patient and specimen characteristics, assay methods, and statistical analysis methods”.

Small molecules

Organic compounds with a low molecular mass (the usual upper limit is 500–1,000 Da) that bind to specific macromolecules (such as proteins), thus altering their activity or function. Traditionally, small molecules with anticancer effects inhibit enzymatic activities, but more recently protein–protein interactions have also been successfully targeted.

Synthetic lethality

Originally defined as a combination of two genes that, when individually mutated, can be tolerated by an organism, but when mutated simultaneously, result in death. This concept was extended to cancer therapy; if a gene is synthetically lethal to a cancer-associated mutation of another gene, targeting (that is, inhibiting) the product of the first gene using drugs will also result in synthetic lethality.

Therapeutic window

The range of a drug dose that both confers the desired effect (such as a reduction in tumour growth) and can also be tolerated by the patient without causing prohibitively severe side effects as a result of damage to normal tissues.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Dobbelstein, M., Moll, U. Targeting tumour-supportive cellular machineries in anticancer drug development. Nat Rev Drug Discov 13, 179–196 (2014). https://doi.org/10.1038/nrd4201

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrd4201

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