Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

The impact of the endoplasmic reticulum protein-folding environment on cancer development

Key Points

  • Defective protein folding in the endoplasmic reticulum (ER) and unfolded protein response (UPR) activation are documented in many human cancer types, which is attributed to both intrinsic and extrinsic factors.

  • UPR activation is a vital step for oncogenic transformation, as UPR signalling molecules interact with well-established oncogene and tumour suppressor gene networks to modulate their function during cancer development.

  • Conditions of low nutrient supply (for example, glucose or oxygen deprivation), as well as excess nutrients (fatty acids, cholesterol and glucose) induce ER stress and UPR activation. UPR induction promotes cancer cell survival through induction of autophagy and adaptation to the stressful microenvironment.

  • ER stress and UPR activation possibly promote cancer development and progression through modulating inflammatory responses.

  • The UPR is indispensable in cells in the tumour microenvironment to either promote or inhibit cancer progression.

  • Targeting the UPR, through single or combination therapy, provides a promising therapeutic approach for many different cancers.

Abstract

The endoplasmic reticulum (ER) is an essential organelle in eukaryotic cells for the storage and regulated release of calcium and as the entrance to the secretory pathway. Protein misfolding in the ER causes accumulation of misfolded proteins (ER stress) and activation of the unfolded protein response (UPR), which has evolved to maintain a productive ER protein-folding environment. Both ER stress and UPR activation are documented in many different human cancers. In this Review, we summarize the impact of ER stress and UPR activation on every aspect of cancer and discuss outstanding questions for which answers will pave the way for therapeutics.

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: The unfolded protein response (UPR) signalling pathways.
Figure 2: Crosstalk between the unfolded protein response (UPR) components and oncogene or tumour suppressor gene networks in cancer cells.
Figure 3: The unfolded protein response (UPR) and inflammation.
Figure 4: The cancer-supporting role of the unfolded protein response (UPR).

Similar content being viewed by others

References

  1. Dorner, A. J., Bole, D. G. & Kaufman, R. J. The relationship of N-linked glycosylation and heavy chain-binding protein association with the secretion of glycoproteins. J. Cell Biol. 105, 2665–2674 (1987).

    CAS  PubMed  Google Scholar 

  2. Kozutsumi, Y., Segal, M., Normington, K., Gething, M. J. & Sambrook, J. The presence of malfolded proteins in the endoplasmic reticulum signals the induction of glucose-regulated proteins. Nature 332, 462–464 (1988).

    CAS  PubMed  Google Scholar 

  3. Dorner, A. J., Wasley, L. C. & Kaufman, R. J. Increased synthesis of secreted proteins induces expression of glucose-regulated proteins in butyrate-treated Chinese hamster ovary cells. J. Biol. Chem. 264, 20602–20607 (1989).

    CAS  PubMed  Google Scholar 

  4. Denoyelle, C. et al. Anti-oncogenic role of the endoplasmic reticulum differentially activated by mutations in the MAPK pathway. Nature Cell Biol. 8, 1053–1063 (2006). This study unraveled a potential tumour-suppressive function of the UPR through mediating oncogenic HRAS-driven cellular senescence in melanocytes independent of canonical senescence mediators such as p53 or p16 (also known as INK4A).

    CAS  PubMed  Google Scholar 

  5. Lee, D. Y. & Sugden, B. The LMP1 oncogene of EBV activates PERK and the unfolded protein response to drive its own synthesis. Blood 111, 2280–2289 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Hsiao, J. R. et al. Endoplasmic reticulum stress triggers XBP-1-mediated up-regulation of an EBV oncoprotein in nasopharyngeal carcinoma. Cancer Res. 69, 4461–4467 (2009).

    CAS  PubMed  Google Scholar 

  7. Hart, L. S. et al. ER stress-mediated autophagy promotes Myc-dependent transformation and tumor growth. J. Clin. Invest. 122, 4621–4634 (2012). This study showed that the PERK–eIF2α branch of the UPR can promote cancer formation and progression in MYC-dependent mouse lymphoma models by inducing cytoprotective autophagy.

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Ozcan, U. et al. Loss of the tuberous sclerosis complex tumor suppressors triggers the unfolded protein response to regulate insulin signaling and apoptosis. Mol. Cell 29, 541–551 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Kang, Y. J., Lu, M. K. & Guan, K. L. The TSC1 and TSC2 tumor suppressors are required for proper ER stress response and protect cells from ER stress-induced apoptosis. Cell Death Differ. 18, 133–144 (2011).

    CAS  PubMed  Google Scholar 

  10. Yeung, B. H. et al. Glucose-regulated protein 78 as a novel effector of BRCA1 for inhibiting stress-induced apoptosis. Oncogene 27, 6782–6789 (2008).

    CAS  PubMed  Google Scholar 

  11. Fang, M. et al. The ER UDPase ENTPD5 promotes protein N-glycosylation, the Warburg effect, and proliferation in the PTEN pathway. Cell 143, 711–724 (2010).

    CAS  PubMed  Google Scholar 

  12. Marada, S., Stewart, D. P., Bodeen, W. J., Han, Y. G. & Ogden, S. K. The unfolded protein response selectively targets active smoothened mutants. Mol. Cell. Biol. 33, 2375–2387 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Babour, A., Bicknell, A. A., Tourtellotte, J. & Niwa, M. A surveillance pathway monitors the fitness of the endoplasmic reticulum to control its inheritance. Cell 142, 256–269 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Kufe, D. W. Mucins in cancer: function, prognosis and therapy. Nature Rev. Cancer 9, 874–885 (2009).

    CAS  Google Scholar 

  15. Mahadevan, N. R. et al. Transmission of endoplasmic reticulum stress and pro-inflammation from tumor cells to myeloid cells. Proc. Natl Acad. Sci. USA 108, 6561–6566 (2011). This study demonstrated for the first time that ER stress can be transmitted from tumour cells to macrophages to elicit a pro-inflammatory response.

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Wang, Y. et al. The unfolded protein response induces the angiogenic switch in human tumor cells through the PERK/ATF4 pathway. Cancer Res. 72, 5396–5406 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Verfaillie, T. et al. PERK is required at the ER-mitochondrial contact sites to convey apoptosis after ROS-based ER stress. Cell Death Differ. 19, 1880–1891 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Marcu, M. G. et al. Heat shock protein 90 modulates the unfolded protein response by stabilizing IRE1α. Mol. Cell. Biol. 22, 8506–8513 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Bertolotti, A., Zhang, Y., Hendershot, L. M., Harding, H. P. & Ron, D. Dynamic interaction of BiP and ER stress transducers in the unfolded-protein response. Nature Cell Biol. 2, 326–332 (2000).

    CAS  PubMed  Google Scholar 

  20. Liu, C. Y., Schroder, M. & Kaufman, R. J. Ligand-independent dimerization activates the stress response kinases IRE1 and PERK in the lumen of the endoplasmic reticulum. J. Biol. Chem. 275, 24881–24885 (2000).

    CAS  PubMed  Google Scholar 

  21. Scheuner, D. et al. Translational control is required for the unfolded protein response and in vivo glucose homeostasis. Mol. Cell 7, 1165–1176 (2001).

    CAS  PubMed  Google Scholar 

  22. Das, S., Ghosh, R. & Maitra, U. Eukaryotic translation initiation factor 5 functions as a GTPase-activating protein. J. Biol. Chem. 276, 6720–6726 (2001).

    CAS  PubMed  Google Scholar 

  23. Jennings, M. D. & Pavitt, G. D. eIF5 is a dual function GAP and GDI for eukaryotic translational control. Small GTPases 1, 118–123 (2010).

    PubMed  PubMed Central  Google Scholar 

  24. Siekierka, J., Manne, V. & Ochoa, S. Mechanism of translational control by partial phosphorylation of the α subunit of eukaryotic initiation factor 2. Proc. Natl Acad. Sci. USA 81, 352–356 (1984).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Rowlands, A. G., Montine, K. S., Henshaw, E. C. & Panniers, R. Physiological stresses inhibit guanine-nucleotide-exchange factor in Ehrlich cells. Eur. J. Biochem. 175, 93–99 (1988).

    CAS  PubMed  Google Scholar 

  26. Harding, H. P. et al. Regulated translation initiation controls stress-induced gene expression in mammalian cells. Mol. Cell 6, 1099–1108 (2000).

    CAS  PubMed  Google Scholar 

  27. Cullinan, S. B. et al. Nrf2 is a direct PERK substrate and effector of PERK-dependent cell survival. Mol. Cell. Biol. 23, 7198–7209 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Zhang, W. et al. ER stress potentiates insulin resistance through PERK-mediated FOXO phosphorylation. Genes Dev. 27, 441–449 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Bobrovnikova-Marjon, E. et al. PERK utilizes intrinsic lipid kinase activity to generate phosphatidic acid, mediate Akt activation, and promote adipocyte differentiation. Mol. Cell. Biol. 32, 2268–2278 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Lu, P. D. et al. Cytoprotection by pre-emptive conditional phosphorylation of translation initiation factor 2. EMBO J. 23, 169–179 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Yaman, I. et al. The zipper model of translational control: a small upstream ORF is the switch that controls structural remodeling of an mRNA leader. Cell 113, 519–531 (2003).

    CAS  PubMed  Google Scholar 

  32. Palam, L. R., Baird, T. D. & Wek, R. C. Phosphorylation of eIF2 facilitates ribosomal bypass of an inhibitory upstream ORF to enhance CHOP translation. J. Biol. Chem. 286, 10939–10949 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Marciniak, S. J. et al. CHOP induces death by promoting protein synthesis and oxidation in the stressed endoplasmic reticulum. Genes Dev. 18, 3066–3077 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Song, B., Scheuner, D., Ron, D., Pennathur, S. & Kaufman, R. J. Chop deletion reduces oxidative stress, improves β cell function, and promotes cell survival in multiple mouse models of diabetes. J. Clin. Invest. 118, 3378–3389 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Chitnis, N. S. et al. miR-211 is a prosurvival microRNA that regulates chop expression in a PERK-dependent manner. Mol. Cell 48, 353–364 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Woo, C. W., Kutzler, L., Kimball, S. R. & Tabas, I. Toll-like receptor activation suppresses ER stress factor CHOP and translation inhibition through activation of eIF2B. Nature Cell Biol. 14, 192–200 (2012).

    CAS  PubMed  Google Scholar 

  37. Rutkowski, D. T. et al. Adaptation to ER stress is mediated by differential stabilities of pro-survival and pro-apoptotic mRNAs and proteins. PLoS Biol. 4, e374 (2006).

    PubMed  PubMed Central  Google Scholar 

  38. Lin, J. H. et al. IRE1 signaling affects cell fate during the unfolded protein response. Science 318, 944–949 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Scheuner, D. et al. Double-stranded RNA-dependent protein kinase phosphorylation of the α-subunit of eukaryotic translation initiation factor 2 mediates apoptosis. J. Biol. Chem. 281, 21458–21468 (2006).

    CAS  PubMed  Google Scholar 

  40. McCullough, K. D., Martindale, J. L., Klotz, L. O., Aw, T. Y. & Holbrook, N. J. Gadd153 sensitizes cells to endoplasmic reticulum stress by down-regulating Bcl2 and perturbing the cellular redox state. Mol. Cell. Biol. 21, 1249–1259 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Puthalakath, H. et al. ER stress triggers apoptosis by activating BH3-only protein Bim. Cell 129, 1337–1349 (2007).

    CAS  PubMed  Google Scholar 

  42. Szegezdi, E., Logue, S. E., Gorman, A. M. & Samali, A. Mediators of endoplasmic reticulum stress-induced apoptosis. EMBO Rep. 7, 880–885 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Galehdar, Z. et al. Neuronal apoptosis induced by endoplasmic reticulum stress is regulated by ATF4-CHOP-mediated induction of the Bcl-2 homology 3-only member PUMA. J. Neurosci. 30, 16938–16948 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Hsin, I. L. et al. Lipocalin 2, a new GADD153 target gene, as an apoptosis inducer of endoplasmic reticulum stress in lung cancer cells. Toxicol. Appl. Pharmacol. 263, 330–337 (2012).

    CAS  PubMed  Google Scholar 

  45. Ohoka, N., Yoshii, S., Hattori, T., Onozaki, K. & Hayashi, H. TRB3, a novel ER stress-inducible gene, is induced via ATF4-CHOP pathway and is involved in cell death. EMBO J. 24, 1243–1255 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Zou, W., Yue, P., Khuri, F. R. & Sun, S. Y. Coupling of endoplasmic reticulum stress to CDDO-Me-induced up-regulation of death receptor 5 via a CHOP-dependent mechanism involving JNK activation. Cancer Res. 68, 7484–7492 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Yamaguchi, H. & Wang, H. G. CHOP is involved in endoplasmic reticulum stress-induced apoptosis by enhancing DR5 expression in human carcinoma cells. J. Biol. Chem. 279, 45495–45502 (2004).

    CAS  PubMed  Google Scholar 

  48. Lu, M. et al. Cell death. Opposing unfolded-protein-response signals converge on death receptor 5 to control apoptosis. Science 345, 98–101 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Han, J. et al. ER-stress-induced transcriptional regulation increases protein synthesis leading to cell death. Nature Cell Biol. 15, 481–490 (2013). This study identified a novel mode of CHOP-mediated cell death through coupling with ATF4 that initiates robust de novo protein synthesis and oxidative stress.

    CAS  PubMed  Google Scholar 

  50. Malhotra, J. D. et al. Antioxidants reduce endoplasmic reticulum stress and improve protein secretion. Proc. Natl Acad. Sci. USA 105, 18525–18530 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Zinszner, H. et al. CHOP is implicated in programmed cell death in response to impaired function of the endoplasmic reticulum. Genes Dev. 12, 982–995 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Bobrovnikova-Marjon, E. et al. PERK promotes cancer cell proliferation and tumor growth by limiting oxidative DNA damage. Oncogene 29, 3881–3895 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Nagy, P., Varga, A., Pircs, K., Hegedus, K. & Juhasz, G. Myc-driven overgrowth requires unfolded protein response-mediated induction of autophagy and antioxidant responses in Drosophila melanogaster. PLoS Genet. 9, e1003664 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Talloczy, Z. et al. Regulation of starvation- and virus-induced autophagy by the eIF2α kinase signaling pathway. Proc. Natl Acad. Sci. USA 99, 190–195 (2002).

    CAS  PubMed  Google Scholar 

  55. B'Chir, W. et al. The eIF2α/ATF4 pathway is essential for stress-induced autophagy gene expression. Nucleic Acids Res. 41, 7683–7699 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Huber, A. L. et al. p58(IPK)-mediated attenuation of the proapoptotic PERK-CHOP pathway allows malignant progression upon low glucose. Mol. Cell 49, 1049–1059 (2013).

    CAS  PubMed  Google Scholar 

  57. Nakagawa, H. et al. ER stress combines with hypernutrition to cause TNF-dependent spontaneous HCC development. Cancer Cell (in the press). This study established a connection between high-fat-diet-induced liver ER stress and hepatocellular carcinoma tumorigenesis.

  58. Kan, Z. et al. Diverse somatic mutation patterns and pathway alterations in human cancers. Nature 466, 869–873 (2010).

    CAS  PubMed  Google Scholar 

  59. Gupta, S. et al. HSP72 protects cells from ER stress-induced apoptosis via enhancement of IRE1α-XBP1 signaling through a physical interaction. PLoS Biol. 8, e1000410 (2010).

    PubMed  PubMed Central  Google Scholar 

  60. Volmer, R., van der Ploeg, K. & Ron, D. Membrane lipid saturation activates endoplasmic reticulum unfolded protein response transducers through their transmembrane domains. Proc. Natl Acad. Sci. USA 110, 4628–4633 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Gardner, B. M. & Walter, P. Unfolded proteins are Ire1-activating ligands that directly induce the unfolded protein response. Science 333, 1891–1894 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Zhou, J. et al. The crystal structure of human IRE1 luminal domain reveals a conserved dimerization interface required for activation of the unfolded protein response. Proc. Natl Acad. Sci. USA 103, 14343–14348 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Cho, J. A. et al. The unfolded protein response element IRE1α senses bacterial proteins invading the ER to activate RIG-I and innate immune signaling. Cell Host Microbe 13, 558–569 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Lee, K. et al. IRE1-mediated unconventional mRNA splicing and S2P-mediated ATF6 cleavage merge to regulate XBP1 in signaling the unfolded protein response. Genes Dev. 16, 452–466 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Calfon, M. et al. IRE1 couples endoplasmic reticulum load to secretory capacity by processing the XBP-1 mRNA. Nature 415, 92–96 (2002).

    CAS  PubMed  Google Scholar 

  66. Yanagitani, K. et al. Cotranslational targeting of XBP1 protein to the membrane promotes cytoplasmic splicing of its own mRNA. Mol. Cell 34, 191–200 (2009).

    CAS  PubMed  Google Scholar 

  67. Shaffer, A. L. et al. XBP1, downstream of Blimp-1, expands the secretory apparatus and other organelles, and increases protein synthesis in plasma cell differentiation. Immunity 21, 81–93 (2004).

    CAS  PubMed  Google Scholar 

  68. Guo, F. J. et al. XBP1S protects cells from ER stress-induced apoptosis through Erk1/2 signaling pathway involving CHOP. Histochem. Cell Biol. 138, 447–460 (2012).

    CAS  PubMed  Google Scholar 

  69. Cross, B. C. et al. The molecular basis for selective inhibition of unconventional mRNA splicing by an IRE1-binding small molecule. Proc. Natl Acad. Sci. USA 109, E869–878 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Han, D. et al. IRE1α kinase activation modes control alternate endoribonuclease outputs to determine divergent cell fates. Cell 138, 562–575 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Li, H., Korennykh, A. V., Behrman, S. L. & Walter, P. Mammalian endoplasmic reticulum stress sensor IRE1 signals by dynamic clustering. Proc. Natl Acad. Sci. USA 107, 16113–16118 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Ishiwata-Kimata, Y., Promlek, T., Kohno, K. & Kimata, Y. BiP-bound and nonclustered mode of Ire1 evokes a weak but sustained unfolded protein response. Genes Cells 18, 288–301 (2013).

    CAS  PubMed  Google Scholar 

  73. Qiu, Y. et al. A crucial role for RACK1 in the regulation of glucose-stimulated IRE1α activation in pancreatic β cells. Sci.Signal. 3, ra7 (2010).

    PubMed  PubMed Central  Google Scholar 

  74. Qiu, Q. et al. Toll-like receptor-mediated IRE1α activation as a therapeutic target for inflammatory arthritis. EMBO J. 32, 2477–2490 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Eletto, D., Eletto, D., Dersh, D., Gidalevitz, T. & Argon, Y. Protein disulfide isomerase A6 controls the decay of IRE1α signaling via disulfide-dependent association. Mol. Cell 53, 562–576 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  76. Groenendyk, J. et al. Interplay between PDIA6 and miR-322 controls adaptive response to disrupted endoplasmic reticulum calcium homeostasis. Sci.Signal. 7, ra54 (2014).

    PubMed  PubMed Central  Google Scholar 

  77. Yoshida, H., Uemura, A. & Mori, K. pXBP1(U), a negative regulator of the unfolded protein response activator pXBP1(S), targets ATF6 but not ATF4 in proteasome-mediated degradation. Cell Struct. Funct. 34, 1–10 (2009).

    CAS  PubMed  Google Scholar 

  78. Mishiba, K. I. et al. Defects in IRE1 enhance cell death and fail to degrade mRNAs encoding secretory pathway proteins in the Arabidopsis unfolded protein response. Proc. Natl Acad. Sci. USA 110, 5713–5718 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Tirasophon, W., Lee, K., Callaghan, B., Welihinda, A. & Kaufman, R. J. The endoribonuclease activity of mammalian IRE1 autoregulates its mRNA and is required for the unfolded protein response. Genes Dev. 14, 2725–2736 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  80. Hollien, J. & Weissman, J. S. Decay of endoplasmic reticulum-localized mRNAs during the unfolded protein response. Science 313, 104–107 (2006).

    CAS  PubMed  Google Scholar 

  81. Ghosh, R. et al. Allosteric Inhibition of the IRE1α RNase Preserves Cell Viability and Function during Endoplasmic Reticulum Stress. Cell 158, 534–548 (2014). This study identified somatic IRE1A mutations in human cancer patients and characterized the physiological role of RIDD in regulation of cell apoptosis using genetic and pharmaceutical approaches.

    CAS  PubMed  PubMed Central  Google Scholar 

  82. Upton, J. P. et al. IRE1α cleaves select microRNAs during ER stress to derepress translation of proapoptotic Caspase-2. Science 338, 818–822 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Sandow, J. J. et al. ER stress does not cause upregulation and activation of caspase-2 to initiate apoptosis. Cell Death Differ. 21, 475–480 (2014).

    CAS  PubMed  Google Scholar 

  84. Urano, F. et al. Coupling of stress in the ER to activation of JNK protein kinases by transmembrane protein kinase IRE1. Science 287, 664–666 (2000).

    CAS  PubMed  Google Scholar 

  85. Reimold, A. M. et al. Plasma cell differentiation requires the transcription factor XBP-1. Nature 412, 300–307 (2001).

    CAS  PubMed  Google Scholar 

  86. Iwakoshi, N. N. et al. Plasma cell differentiation and the unfolded protein response intersect at the transcription factor XBP-1. Nature Immunol. 4, 321–329 (2003).

    CAS  Google Scholar 

  87. Zhang, K. et al. The unfolded protein response sensor IRE1α is required at 2 distinct steps in B cell lymphopoiesis. J. Clin. Invest. 115, 268–281 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Carrasco, D. R. et al. The differentiation and stress response factor XBP-1 drives multiple myeloma pathogenesis. Cancer Cell 11, 349–360 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Hong, S. Y. & Hagen, T. Multiple myeloma Leu167Ile (c.499C>A) mutation prevents XBP1 mRNA splicing. Br. J. Haematol. 161, 898–901 (2013).

    CAS  PubMed  Google Scholar 

  90. Leung-Hagesteijn, C. et al. Xbp1s-negative tumor B cells and pre-plasmablasts mediate therapeutic proteasome inhibitor resistance in multiple myeloma. Cancer Cell 24, 289–304 (2013). This study demonstrated that human multiple myeloma tumour lines with loss-of-function mutations in either IRE1α or XBP1s, displaying pre-plasmablast characteristics, are resistant to proteasome inhibitor treatment.

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Greenman, C. et al. Patterns of somatic mutation in human cancer genomes. Nature 446, 153–158 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Niederreiter, L. et al. ER stress transcription factor Xbp1 suppresses intestinal tumorigenesis and directs intestinal stem cells. J. Exp. Med. 210, 2041–2056 (2013). This study identified XBP1 in the intestinal epithelium as a tumour suppressor in both colitis-associated and APC mutation-driven mouse tumour models.

    CAS  PubMed  PubMed Central  Google Scholar 

  93. Chen, X. et al. XBP1 promotes triple-negative breast cancer by controlling the HIF1α pathway. Nature 508, 103–107 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Ye, J. et al. ER stress induces cleavage of membrane-bound ATF6 by the same proteases that process SREBPs. Mol. Cell 6, 1355–1364 (2000).

    CAS  PubMed  Google Scholar 

  95. Yoshida, H., Matsui, T., Yamamoto, A., Okada, T. & Mori, K. XBP1 mRNA is induced by ATF6 and spliced by IRE1 in response to ER stress to produce a highly active transcription factor. Cell 107, 881–891 (2001).

    CAS  PubMed  Google Scholar 

  96. Wu, J. et al. ATF6α optimizes long-term endoplasmic reticulum function to protect cells from chronic stress. Dev. Cell 13, 351–364 (2007).

    CAS  PubMed  Google Scholar 

  97. Guan, D. et al. N-glycosylation of ATF6β is essential for its proteolytic cleavage and transcriptional repressor function to ATF6α. J. Cell Biochem. 108, 825–831 (2009).

    CAS  PubMed  Google Scholar 

  98. Teske, B. F. et al. The eIF2 kinase PERK and the integrated stress response facilitate activation of ATF6 during endoplasmic reticulum stress. Mol. Biol. Cell 22, 4390–4405 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  99. Arai, M. et al. Transformation-associated gene regulation by ATF6α during hepatocarcinogenesis. FEBS Lett. 580, 184–190 (2006).

    CAS  PubMed  Google Scholar 

  100. Wu, X. et al. A missense polymorphism in ATF6 gene is associated with susceptibility to hepatocellular carcinoma probably by altering ATF6 level. Int. J. Cancer 135, 61–68 (2014).

    CAS  PubMed  Google Scholar 

  101. Arap, M. A. et al. Cell surface expression of the stress response chaperone GRP78 enables tumor targeting by circulating ligands. Cancer Cell 6, 275–284 (2004).

    CAS  PubMed  Google Scholar 

  102. Lee, A. S. GRP78 induction in cancer: therapeutic and prognostic implications. Cancer Res. 67, 3496–3499 (2007).

    CAS  PubMed  Google Scholar 

  103. Rzymski, T., Milani, M., Singleton, D. C. & Harris, A. L. Role of ATF4 in regulation of autophagy and resistance to drugs and hypoxia. Cell Cycle 8, 3838–3847 (2009).

    CAS  PubMed  Google Scholar 

  104. Rouschop, K. M. et al. PERK/eIF2α signaling protects therapy resistant hypoxic cells through induction of glutathione synthesis and protection against ROS. Proc. Natl Acad. Sci. USA 110, 4622–4627 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  105. Schewe, D. M. & Aguirre-Ghiso, J. A. ATF6α-Rheb-mTOR signaling promotes survival of dormant tumor cells in vivo. Proc. Natl Acad. Sci. USA 105, 10519–10524 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  106. Bruchmann, A. et al. Bcl-2 associated athanogene 5 (Bag5) is overexpressed in prostate cancer and inhibits ER-stress induced apoptosis. BMC Cancer 13, 96 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Tay, K. H. et al. Sustained IRE1 and ATF6 signaling is important for survival of melanoma cells undergoing ER stress. Cell Signal 26, 287–294 (2014).

    CAS  PubMed  Google Scholar 

  108. Reddy, R. K. et al. Endoplasmic reticulum chaperone protein GRP78 protects cells from apoptosis induced by topoisomerase inhibitors: role of ATP binding site in suppression of caspase-7 activation. J. Biol. Chem. 278, 20915–20924 (2003).

    CAS  PubMed  Google Scholar 

  109. Nakajima, S. et al. Selective abrogation of BiP/GRP78 blunts activation of NF-κB through the ATF6 branch of the UPR: involvement of C/EBPβ and mTOR-dependent dephosphorylation of Akt. Mol. Cell. Biol. 31, 1710–1718 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Du, K., Herzig, S., Kulkarni, R. N. & Montminy, M. TRB3: a tribbles homolog that inhibits Akt/PKB activation by insulin in liver. Science 300, 1574–1577 (2003).

    CAS  PubMed  Google Scholar 

  111. Hamanaka, R. B., Bennett, B. S., Cullinan, S. B. & Diehl, J. A. PERK and GCN2 contribute to eIF2α phosphorylation and cell cycle arrest after activation of the unfolded protein response pathway. Mol. Biol. Cell 16, 5493–5501 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Dorner, A. J., Wasley, L. C. & Kaufman, R. J. Protein dissociation from GRP78 and secretion are blocked by depletion of cellular ATP levels. Proc. Natl Acad. Sci. USA 87, 7429–7432 (1990).

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Braakman, I., Helenius, J. & Helenius, A. Role of ATP and disulphide bonds during protein folding in the endoplasmic reticulum. Nature 356, 260–262 (1992).

    CAS  PubMed  Google Scholar 

  114. Saito, S. et al. Chemical genomics identifies the unfolded protein response as a target for selective cancer cell killing during glucose deprivation. Cancer Res. 69, 4225–4234 (2009).

    CAS  PubMed  Google Scholar 

  115. Fu, S. et al. Aberrant lipid metabolism disrupts calcium homeostasis causing liver endoplasmic reticulum stress in obesity. Nature 473, 528–531 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Ozcan, U. et al. Endoplasmic reticulum stress links obesity, insulin action, and type 2 diabetes. Science 306, 457–461 (2004).

    PubMed  Google Scholar 

  117. Taubes, G. Cancer research. Unraveling the obesity-cancer connection. Science 335, 28–32 (2012).

    CAS  PubMed  Google Scholar 

  118. Park, E. J. et al. Dietary and genetic obesity promote liver inflammation and tumorigenesis by enhancing IL-6 and TNF expression. Cell 140, 197–208 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  119. Suh, D. H., Kim, M. K., Kim, H. S., Chung, H. H. & Song, Y. S. Unfolded protein response to autophagy as a promising druggable target for anticancer therapy. Ann. NY Acad. Sci. 1271, 20–32 (2012).

    CAS  PubMed  Google Scholar 

  120. Kouroku, Y. et al. ER stress (PERK/eIF2α phosphorylation) mediates the polyglutamine-induced LC3 conversion, an essential step for autophagy formation. Cell Death Differ. 14, 230–239 (2007).

    CAS  PubMed  Google Scholar 

  121. Ogata, M. et al. Autophagy is activated for cell survival after endoplasmic reticulum stress. Mol. Cell. Biol. 26, 9220–9231 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Hetz, C. et al. XBP-1 deficiency in the nervous system protects against amyotrophic lateral sclerosis by increasing autophagy. Genes Dev. 23, 2294–2306 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  123. Gade, P. et al. An IFN-γ-stimulated ATF6-C/EBP-β-signaling pathway critical for the expression of death associated protein kinase 1 and induction of autophagy. Proc. Natl Acad. Sci. USA 109, 10316–10321 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  124. Sakaki, K., Wu, J. & Kaufman, R. J. Protein kinase Cθ is required for autophagy in response to stress in the endoplasmic reticulum. J. Biol. Chem. 283, 15370–15380 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  125. Hoyer-Hansen, M. et al. Control of macroautophagy by calcium, calmodulin-dependent kinase kinase-β, and Bcl-2. Mol. Cell 25, 193–205 (2007).

    PubMed  Google Scholar 

  126. Qin, L., Wang, Z., Tao, L. & Wang, Y. ER stress negatively regulates AKT/TSC/mTOR pathway to enhance autophagy. Autophagy 6, 239–247 (2010).

    CAS  PubMed  Google Scholar 

  127. Rouschop, K. M. et al. The unfolded protein response protects human tumor cells during hypoxia through regulation of the autophagy genes MAP1LC3B and ATG5. J. Clin. Invest. 120, 127–141 (2010).

    CAS  PubMed  Google Scholar 

  128. Baldwin, A. C., Green, C. D., Olson, L. K., Moxley, M. A. & Corbett, J. A. A role for aberrant protein palmitoylation in FFA-induced ER stress and β-cell death. Am. J. Physiol. Endocrinol. Metab. 302, E1390–E1398 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  129. Kim, J., Park, Y. J., Jang, Y. & Kwon, Y. H. AMPK activation inhibits apoptosis and tau hyperphosphorylation mediated by palmitate in SH-SY5Y cells. Brain Res. 1418, 42–51 (2011).

    CAS  PubMed  Google Scholar 

  130. Oh, Y. S. et al. Exendin-4 inhibits glucolipotoxic ER stress in pancreatic β cells via regulation of SREBP1c and C/EBPβ transcription factors. J. Endocrinol. 216, 343–352 (2013).

    CAS  PubMed  Google Scholar 

  131. Colgan, S. M., Tang, D., Werstuck, G. H. & Austin, R. C. Endoplasmic reticulum stress causes the activation of sterol regulatory element binding protein-2. Int. J. Biochem. Cell Biol. 39, 1843–1851 (2007).

    CAS  PubMed  Google Scholar 

  132. Lee, J. S., Mendez, R., Heng, H. H., Yang, Z. Q. & Zhang, K. Pharmacological ER stress promotes hepatic lipogenesis and lipid droplet formation. Am. J. Transl. Res. 4, 102–113 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Williams, K. J. et al. An essential requirement for the SCAP/SREBP signaling axis to protect cancer cells from lipotoxicity. Cancer Res. 73, 2850–2862 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  134. Fuchs, C. D. et al. Absence of adipose triglyceride lipase protects from hepatic endoplasmic reticulum stress in mice. Hepatology 56, 270–280 (2012).

    CAS  PubMed  Google Scholar 

  135. Li, Y. et al. Free cholesterol-loaded macrophages are an abundant source of tumor necrosis factor-α and interleukin-6: model of NF-κB- and map kinase-dependent inflammation in advanced atherosclerosis. J. Biol. Chem. 280, 21763–21772 (2005).

    CAS  PubMed  Google Scholar 

  136. Bensellam, M., Laybutt, D. R. & Jonas, J. C. The molecular mechanisms of pancreatic β-cell glucotoxicity: recent findings and future research directions. Mol. Cell Endocrinol. 364, 1–27 (2012).

    CAS  PubMed  Google Scholar 

  137. Malhotra, J. D. & Kaufman, R. J. Endoplasmic reticulum stress and oxidative stress: a vicious cycle or a double-edged sword? Antioxid. Redox Signal. 9, 2277–2293 (2007).

    CAS  PubMed  Google Scholar 

  138. Back, S. H. et al. Translation attenuation through eIF2α phosphorylation prevents oxidative stress and maintains the differentiated state in β cells. Cell. Metab. 10, 13–26 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  139. Grivennikov, S. I., Greten, F. R. & Karin, M. Immunity, inflammation, and cancer. Cell 140, 883–899 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  140. Zhang, K. & Kaufman, R. J. From endoplasmic-reticulum stress to the inflammatory response. Nature 454, 455–462 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  141. Adolph, T. E. et al. Paneth cells as a site of origin for intestinal inflammation. Nature 503, 272–276 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  142. Zhang, K. et al. Endoplasmic reticulum stress activates cleavage of CREBH to induce a systemic inflammatory response. Cell 124, 587–599 (2006). This study showed for the first time that pro-inflammatory cytokines and lipopolysaccharide activate the UPR and induce cleavage of CREBH in the liver, providing a link by which ER stress initiates an acute inflammatory response.

    CAS  PubMed  Google Scholar 

  143. Oslowski, C. M. et al. Thioredoxin-interacting protein mediates ER stress-induced β cell death through initiation of the inflammasome. Cell. Metabolism 16, 265–273 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  144. Lerner, A. G. et al. IRE1α induces thioredoxin-interacting protein to activate the NLRP3 inflammasome and promote programmed cell death under irremediable ER stress. Cell. Metab. 16, 250–264 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  145. Menu, P. et al. ER stress activates the NLRP3 inflammasome via an UPR-independent pathway. Cell Death Dis. 3, e261 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  146. Condamine, T. et al. ER stress regulates myeloid-derived suppressor cell fate through TRAIL-R-mediated apoptosis. J. Clin. Invest. 124, 2626–2639 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  147. Mahadevan, N. R. et al. Cell-extrinsic effects of tumor ER stress imprint myeloid dendritic cells and impair CD8+ T cell priming. PLoS ONE 7, e51845 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  148. Oh, J. et al. Endoplasmic reticulum stress controls M2 macrophage differentiation and foam cell formation. J. Biol. Chem. 287, 11629–11641 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  149. Goodall, J. C. et al. Endoplasmic reticulum stress-induced transcription factor, CHOP, is crucial for dendritic cell IL-23 expression. Proc. Natl Acad. Sci. USA 107, 17698–17703 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  150. Wang, L. et al. IL-17 can promote tumor growth through an IL-6-Stat3 signaling pathway. J. Exp. Med. 206, 1457–1464 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  151. Granados, D. P. et al. ER stress affects processing of MHC class I-associated peptides. BMC Immunol. 10, 10 (2009).

    PubMed  PubMed Central  Google Scholar 

  152. Gao, B. et al. Assembly and antigen-presenting function of MHC class I molecules in cells lacking the ER chaperone calreticulin. Immunity 16, 99–109 (2002).

    CAS  PubMed  Google Scholar 

  153. Kropp, L. E., Garg, M. & Binder, R. J. Ovalbumin-derived precursor peptides are transferred sequentially from gp96 and calreticulin to MHC class I in the endoplasmic reticulum. J. Immunol. 184, 5619–5627 (2010).

    CAS  PubMed  Google Scholar 

  154. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell 144, 646–674 (2011).

    CAS  PubMed  Google Scholar 

  155. Senovilla, L. et al. An immunosurveillance mechanism controls cancer cell ploidy. Science 337, 1678–1684 (2012). This study revealed for the first time that ER stress, by inducing activation of the UPR and calreticulin surface exposure, renders cancer cells immunogenic and therefore constitutes an immunosurveillance system against cancer development.

    CAS  PubMed  Google Scholar 

  156. Pereira, E. R., Liao, N., Neale, G. A. & Hendershot, L. M. Transcriptional and post-transcriptional regulation of proangiogenic factors by the unfolded protein response. PLoS ONE 5, e12521 (2010).

    PubMed  PubMed Central  Google Scholar 

  157. Hu, D. et al. Overexpressed Derlin-1 inhibits ER expansion in the endothelial cells derived from human hepatic cavernous hemangioma. J. Biochem. Mol. Biol. 39, 677–685 (2006).

    CAS  PubMed  Google Scholar 

  158. Karali, E. et al. VEGF signals through ATF6 and PERK to promote endothelial cell survival and angiogenesis in the absence of ER stress. Mol. Cell 54, 559–572 (2014). This study indicated that VEGF stimulates the UPR in endothelial cells independent of ER stress, which contributes to VEGF-induced endothelial cell survival and angiogenesis.

    CAS  PubMed  Google Scholar 

  159. Salminen, A., Kauppinen, A., Hyttinen, J. M., Toropainen, E. & Kaarniranta, K. Endoplasmic reticulum stress in age-related macular degeneration: trigger for neovascularization. Mol. Med. 16, 535–542 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  160. Zeng, L. et al. Vascular endothelial cell growth-activated XBP1 splicing in endothelial cells is crucial for angiogenesis. Circulation 127, 1712–1722 (2013).

    CAS  PubMed  Google Scholar 

  161. Dong, D. et al. Critical role of the stress chaperone GRP78/BiP in tumor proliferation, survival, and tumor angiogenesis in transgene-induced mammary tumor development. Cancer Res. 68, 498–505 (2008).

    CAS  PubMed  Google Scholar 

  162. Virrey, J. J. et al. Stress chaperone GRP78/BiP confers chemoresistance to tumor-associated endothelial cells. Mol. Cancer Res. 6, 1268–1275 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  163. Obeng, E. A. et al. Proteasome inhibitors induce a terminal unfolded protein response in multiple myeloma cells. Blood 107, 4907–4916 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  164. Ling, S. C. et al. Response of myeloma to the proteasome inhibitor bortezomib is correlated with the unfolded protein response regulator XBP-1. Haematologica 97, 64–72 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  165. Atkins, C. et al. Characterization of a novel PERK kinase inhibitor with antitumor and antiangiogenic activity. Cancer Res. 73, 1993–2002 (2013).

    CAS  PubMed  Google Scholar 

  166. Krishnamoorthy, J. et al. Evidence for eIF2α phosphorylation-independent effects of GSK2656157, a novel catalytic inhibitor of PERK with clinical implications. Cell Cycle 13, 801–806 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  167. Tang, C. et al. Inhibition of ER stress-associated IRE-1/XBP-1 pathway reduces leukemic cell survival. J. Clin. Invest. 124, 2585–2598 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  168. Kern, J. et al. GRP-78 secreted by tumor cells blocks the antiangiogenic activity of bortezomib. Blood 114, 3960–3967 (2009).

    CAS  PubMed  Google Scholar 

  169. Ma, X. H. et al. Targeting ER stress-induced autophagy overcomes BRAF inhibitor resistance in melanoma. J. Clin. Invest. 124, 1406–1417 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  170. Mahoney, D. J. et al. Virus-tumor interactome screen reveals ER stress response can reprogram resistant cancers for oncolytic virus-triggered caspase-2 cell death. Cancer Cell 20, 443–456 (2011).

    CAS  PubMed  Google Scholar 

  171. Dorr, J. R. et al. Synthetic lethal metabolic targeting of cellular senescence in cancer therapy. Nature 501, 421–425 (2013).

    PubMed  Google Scholar 

  172. Blais, J. D. et al. Perk-dependent translational regulation promotes tumor cell adaptation and angiogenesis in response to hypoxic stress. Mol. Cell. Biol. 26, 9517–9532 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  173. Verras, M., Papandreou, I., Lim, A. L. & Denko, N. C. Tumor hypoxia blocks Wnt processing and secretion through the induction of endoplasmic reticulum stress. Mol. Cell. Biol. 28, 7212–7224 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  174. Koumenis, C. et al. Regulation of protein synthesis by hypoxia via activation of the endoplasmic reticulum kinase PERK and phosphorylation of the translation initiation factor eIF2α. Mol. Cell. Biol. 22, 7405–7416 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  175. Koritzinsky, M. et al. Two phases of disulfide bond formation have differing requirements for oxygen. J. Cell Biol. 203, 615–627 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  176. Koditz, J. et al. Oxygen-dependent ATF-4 stability is mediated by the PHD3 oxygen sensor. Blood 110, 3610–3617 (2007).

    PubMed  Google Scholar 

  177. Scortegagna, M. et al. Fine tuning of the UPR by the ubiquitin ligases Siah1/2. PLoS Genet. 10, e1004348 (2014).

    PubMed  PubMed Central  Google Scholar 

  178. Pereira, E. R., Frudd, K., Awad, W. & Hendershot, L. M. Endoplasmic reticulum (ER) stress and hypoxia response pathways interact to potentiate hypoxia-inducible factor 1 (HIF-1) transcriptional activity on targets like vascular endothelial growth factor (VEGF). J. Biol. Chem. 289, 3352–3364 (2014).

    CAS  PubMed  Google Scholar 

  179. Bi, M. et al. ER stress-regulated translation increases tolerance to extreme hypoxia and promotes tumor growth. EMBO J. 24, 3470–3481 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  180. Moore, C. E., Omikorede, O., Gomez, E., Willars, G. B. & Herbert, T. P. PERK activation at low glucose concentration is mediated by SERCA pump inhibition and confers preemptive cytoprotection to pancreatic β-cells. Mol. Endocrinol. 25, 315–326 (2011).

    CAS  PubMed  Google Scholar 

  181. Novoa, I., Zeng, H., Harding, H. P. & Ron, D. Feedback inhibition of the unfolded protein response by GADD34-mediated dephosphorylation of eIF2α. J. Cell Biol. 153, 1011–1022 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  182. Marchal, J. A. et al. The impact of PKR activation: from neurodegeneration to cancer. FASEB J. 28, 1965–1974 (2014).

    CAS  PubMed  Google Scholar 

  183. Ye, J. et al. The GCN2-ATF4 pathway is critical for tumour cell survival and proliferation in response to nutrient deprivation. EMBO J. 29, 2082–2096 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  184. Wang, Y. et al. Amino acid deprivation promotes tumor angiogenesis through the GCN2/ATF4 pathway. Neoplasia 15, 989–997 (2013).

    PubMed  PubMed Central  Google Scholar 

  185. Hwang, S. Y., Kim, M. K. & Kim, J. C. Cloning of hHRI, human heme-regulated eukaryotic initiation factor 2α kinase: down-regulated in epithelial ovarian cancers. Mol. Cells 10, 584–591 (2000).

    CAS  PubMed  Google Scholar 

  186. Martinon, F., Chen, X., Lee, A. H. & Glimcher, L. H. TLR activation of the transcription factor XBP1 regulates innate immune responses in macrophages. Nature Immunol. 11, 411–418 (2010).

    CAS  Google Scholar 

  187. Chen, R., Alvero, A. B., Silasi, D. A. & Mor, G. Inflammation, cancer and chemoresistance: taking advantage of the toll-like receptor signaling pathway. Am. J. Reprod. Immunol. 57, 93–107 (2007).

    CAS  PubMed  Google Scholar 

  188. Feng, Y. et al. Epithelial-to-mesenchymal transition activates PERK-eIF2a and sensitizes cells to endoplasmic reticulum stress. Cancer Discov. 4, 702–715 (2014).

    CAS  PubMed  Google Scholar 

  189. Avivar-Valderas, A. et al. Regulation of autophagy during ECM detachment is linked to a selective inhibition of mTORC1 by PERK. Oncogene 32, 4932–4940 (2013).

    CAS  PubMed  Google Scholar 

  190. Mounir, Z. et al. Tumor suppression by PTEN requires the activation of the PKR-eIF2α phosphorylation pathway. Sci. Signal. 2, ra85 (2009).

    PubMed  PubMed Central  Google Scholar 

  191. Yang, G. et al. PTEN deficiency causes dyschondroplasia in mice by enhanced hypoxia-inducible factor 1α signaling and endoplasmic reticulum stress. Development 135, 3587–3597 (2008).

    CAS  PubMed  Google Scholar 

  192. Horak, P. et al. TUSC3 loss alters the ER stress response and accelerates prostate cancer growth in vivo. Sci. Rep. 4, 3739 (2014).

    PubMed  PubMed Central  Google Scholar 

  193. Rzymski, T. et al. The unfolded protein response controls induction and activation of ADAM17/TACE by severe hypoxia and ER stress. Oncogene 31, 3621–3634 (2012).

    CAS  PubMed  Google Scholar 

  194. Qu, L. et al. Endoplasmic reticulum stress induces p53 cytoplasmic localization and prevents p53-dependent apoptosis by a pathway involving glycogen synthase kinase-3β. Genes Dev. 18, 261–277 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  195. Pluquet, O., Qu, L. K., Baltzis, D. & Koromilas, A. E. Endoplasmic reticulum stress accelerates p53 degradation by the cooperative actions of Hdm2 and glycogen synthase kinase 3β. Mol. Cell. Biol. 25, 9392–9405 (2005). References 194 and 195 unveiled a deleterious role of ER stress in the regulation of the p53 tumour suppressor protein.

    CAS  PubMed  PubMed Central  Google Scholar 

  196. Hu, P., Han, Z., Couvillon, A. D., Kaufman, R. J. & Exton, J. H. Autocrine tumor necrosis factor α links endoplasmic reticulum stress to the membrane death receptor pathway through IRE1α-mediated NF-κB activation and down-regulation of TRAF2 expression. Mol. Cell. Biol. 26, 3071–3084 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  197. Eferl, R. & Wagner, E. F. AP-1: a double-edged sword in tumorigenesis. Nature Rev. Cancer 3, 859–868 (2003).

    CAS  Google Scholar 

  198. Tam, A. B., Mercado, E. L., Hoffmann, A. & Niwa, M. ER stress activates NF-κB by integrating functions of basal IKK activity, IRE1 and PERK. PLoS ONE 7, e45078 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  199. Zhang, Y. et al. Peroxynitrite-induced neuronal apoptosis is mediated by intracellular zinc release and 12-lipoxygenase activation. J. Neurosci. 24, 10616–10627 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  200. Yamazaki, H. et al. Activation of the Akt-NF-κB pathway by subtilase cytotoxin through the ATF6 branch of the unfolded protein response. J. Immunol. 183, 1480–1487 (2009).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors apologize to those whose work could not be cited owing to length restraints. R.J.K. is supported by US National Institutes of Health (NIH) grants DK042394, DK088227 and HL052173.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Randal J. Kaufman.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

PowerPoint slides

Supplementary information

Supplementary information S1 (table)

Evidence of ER stress and UPR activation in various human cancer types (PDF 146 kb)

Supplementary information S2 (table)

Strategies to target the UPR components for cancer treatment# (PDF 199 kb)

Glossary

ER-associated degradation

(ERAD). A process by which misfolded proteins in the endoplasmic reticulum (ER) are targeted by retrotranslocation and ubiquitylation for subsequent degradation by the proteasome.

Mitochondria-associated ER membranes

(MAMs). A specialized endoplasmic reticulum (ER) membrane is directly juxtaposed to the mitochondrion to coordinate efficient communication between these two organelles.

Polysome

A cluster of ribosomes translating a single mRNA molecule.

Regulated IRE1-dependent decay

(RIDD). A process in which activated inositol-requiring protein 1 (IRE1) induces cleavage and degradation of microRNAs and of mRNAs encoding membrane and secreted proteins.

Regulated intramembrane proteolysis

A process in which endoplasmic reticulum (ER) transmembrane transcription factors are cleaved within the plane of the membrane to release cytosolic fragments that enter the nucleus to regulate gene transcription.

Inflammasome

A large intracellular multiprotein oligomeric complex that is activated by pattern recognition receptors to initiate an innate immune response by maturation of the inflammatory cytokines interleukin-1 (IL-1) and IL-18.

Type M2 macrophage

A subset of activated macrophages that are involved in immunosuppression and tissue repair.

MHC class I pathway

(Major histocompatability complex class I pathway). A pathway by which cells present peptides from cytosolic proteins to T cells.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, M., Kaufman, R. The impact of the endoplasmic reticulum protein-folding environment on cancer development. Nat Rev Cancer 14, 581–597 (2014). https://doi.org/10.1038/nrc3800

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrc3800

This article is cited by

Search

Quick links

Nature Briefing: Cancer

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

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