1932

Abstract

The endoplasmic reticulum (ER) is the site of synthesis and folding of membrane and secretory proteins, which, collectively, represent a large fraction of the total protein output of a mammalian cell. Therefore, the protein flux through the ER must be carefully monitored for abnormalities, including the buildup of misfolded proteins. Mammalian cells have evolved an intricate set of signaling pathways from the ER to the cytosol and nucleus, to allow the cell to respond to the presence of misfolded proteins within the ER. These pathways, known collectively as the unfolded protein response, are important for normal cellular homeostasis and organismal development and may also play key roles in the pathogenesis of many diseases. This review provides background information on the unfolded protein response and discusses a selection of diseases whose pathogenesis involves ER stress.

Loading

Article metrics loading...

/content/journals/10.1146/annurev.pathmechdis.3.121806.151434
2008-02-28
2024-04-23
Loading full text...

Full text loading...

/deliver/fulltext/pm/3/1/annurev.pathmechdis.3.121806.151434.html?itemId=/content/journals/10.1146/annurev.pathmechdis.3.121806.151434&mimeType=html&fmt=ahah

Literature Cited

  1. Bernales S, Papa FR, Walter P. 1.  2006. Intracellular signaling by the unfolded protein response. Annu. Rev. Cell Dev. Biol. 22:487–508 [Google Scholar]
  2. Calfon M, Zeng H, Urano F, Till JH, Hubbard SR et al.2.  2002. IRE1 couples endoplasmic reticulum load to secretory capacity by processing the XBP-1 mRNA. Nature 415:92–96 [Google Scholar]
  3. Yoshida H, Matsui T, Yamamoto A, Okada T, Mori K. 3.  2001. XBP1 mRNA is induced by ATF6 and spliced by IRE1 in response to ER stress to produce a highly active transcription factor. Cell 107:881–91 [Google Scholar]
  4. Sidrauski C, Walter P. 3a.  1997. The transmembrane kinase Ire1p is a site-specific endonuclease that initiates mRNA splicing in the unfolded protein response. Cell 90:1031–39 [Google Scholar]
  5. Lee AH, Iwakoshi NN, Glimcher LH. 4.  2003. XBP-1 regulates a subset of endoplasmic reticulum resident chaperone genes in the unfolded protein response. Mol. Cell. Biol. 23:7448–59 [Google Scholar]
  6. Shaffer AL, Shapiro-Shelef M, Iwakoshi NN, Lee AH, Qian SB et al.5.  2004. XBP1, downstream of Blimp-1, expands the secretory apparatus and other organelles, and increases protein synthesis in plasma cell differentiation. Immunity 21:81–93 [Google Scholar]
  7. Urano F, Wang X, Bertolotti A, Zhang Y, Chung P et al.6.  2000. Coupling of stress in the ER to activation of JNK protein kinases by transmembrane protein kinase IRE1. Science 287:664–66 [Google Scholar]
  8. Nishitoh H, Matsuzawa A, Tobiume K, Saegusa K, Takeda K et al.7.  2002. ASK1 is essential for endoplasmic reticulum stress-induced neuronal cell death triggered by expanded polyglutamine repeats. Genes Dev. 16:1345–55 [Google Scholar]
  9. Barr RK, Bogoyevitch MA. 8.  2001. The c-Jun N-terminal protein kinase family of mitogen-activated protein kinases (JNK MAPKs). Int. J. Biochem. Cell Biol. 33:1047–63 [Google Scholar]
  10. Takeda K, Matsuzawa A, Nishitoh H, Ichijo H. 9.  2003. Roles of MAPKKK ASK1 in stress-induced cell death. Cell Struct. Funct. 28:23–29 [Google Scholar]
  11. Yoneda T, Imaizumi K, Oono K, Yui D, Gomi F et al.10.  2001. Activation of caspase-12, an endoplastic reticulum (ER) resident caspase, through tumor necrosis factor receptor-associated factor 2-dependent mechanism in response to the ER stress. J. Biol. Chem. 276:13935–40 [Google Scholar]
  12. Nakagawa T, Zhu H, Morishima N, Li E, Xu J et al.11.  2000. Caspase-12 mediates endoplasmic-reticulum-specific apoptosis and cytotoxicity by amyloid-β. Nature 403:98–103 [Google Scholar]
  13. Saleh M, Mathison JC, Wolinski MK, Bensinger SJ, Fitzgerald P et al.12.  2006. Enhanced bacterial clearance and sepsis resistance in caspase-12-deficient mice. Nature 440:1064–68 [Google Scholar]
  14. Xue Y, Daly A, Yngvadottir B, Liu M, Coop G et al.13.  2006. Spread of an inactive form of caspase-12 in humans is due to recent positive selection. Am. J. Hum. Genet. 78:659–70 [Google Scholar]
  15. Hitomi J, Katayama T, Eguchi Y, Kudo T, Taniguchi M et al.14.  2004. Involvement of caspase-4 in endoplasmic reticulum stress-induced apoptosis and Aβ-induced cell death. J. Cell Biol. 165:347–56 [Google Scholar]
  16. Tirasophon W, Lee K, Callaghan B, Welihinda A, Kaufman RJ. 15.  2000. The endoribonuclease activity of mammalian IRE1 autoregulates its mRNA and is required for the unfolded protein response. Genes Dev. 14:2725–36 [Google Scholar]
  17. Hollien J, Weissman JS. 16.  2006. Decay of endoplasmic reticulum-localized mRNAs during the unfolded protein response. Science 313:104–7 [Google Scholar]
  18. Harding HP, Zhang Y, Ron D. 17.  1999. Protein translation and folding are coupled by an endoplasmic-reticulum-resident kinase. Nature 397:271–74 [Google Scholar]
  19. Anderson P, Kedersha N. 18.  2002. Visibly stressed: the role of eIF2, TIA-1, and stress granules in protein translation. Cell Stress Chaperones 7:213–21 [Google Scholar]
  20. Jiang HY, Jiang L, Wek RC. 19.  2007. The eukaryotic initiation factor-2 kinase pathway facilitates differential GADD45 a expression in response to environmental stress. J. Biol. Chem. 282:3755–65 [Google Scholar]
  21. Wek RC, Jiang HY, Anthony TG. 20.  2006. Coping with stress: eIF2 kinases and translational control. Biochem. Soc. Trans. 34:7–11 [Google Scholar]
  22. Harding HP, Novoa I, Zhang Y, Zeng H, Wek R et al.21.  2000. Regulated translation initiation controls stress-induced gene expression in mammalian cells. Mol. Cell 6:1099–108 [Google Scholar]
  23. Novoa I, Zeng H, Harding HP, Ron D. 22.  2001. Feedback inhibition of the unfolded protein response by GADD34-mediated dephosphorylation of eIF2α. J. Cell Biol. 153:1011–22 [Google Scholar]
  24. Harding HP, Zhang Y, Bertolotti A, Zeng H, Ron D. 23.  2000. Perk is essential for translational regulation and cell survival during the unfolded protein response. Mol. Cell 5:897–904 [Google Scholar]
  25. Zinszner H, Kuroda M, Wang X, Batchvarova N, Lightfoot RT et al.24.  1998. CHOP is implicated in programmed cell death in response to impaired function of the endoplasmic reticulum. Genes Dev. 12:982–95 [Google Scholar]
  26. Marciniak SJ, Yun CY, Oyadomari S, Novoa I, Zhang Y et al.25.  2004. CHOP induces death by promoting protein synthesis and oxidation in the stressed endoplasmic reticulum. Genes Dev. 18:3066–77 [Google Scholar]
  27. Ohoka N, Yoshii S, Hattori T, Onozaki K, Hayashi H. 26.  2005. TRB3, a novel ER stress-inducible gene, is induced via ATF4-CHOP pathway and is involved in cell death. EMBO J. 24:1243–55 [Google Scholar]
  28. Zhang K, Shen X, Wu J, Sakaki K, Saunders T et al.27.  2006. Endoplasmic reticulum stress activates cleavage of CREBH to induce a systemic inflammatory response. Cell 124:587–99 [Google Scholar]
  29. Kondo S, Murakami T, Tatsumi K, Ogata M, Kanemoto S et al.28.  2005. OASIS, a CREB/ATF-family member, modulates UPR signalling in astrocytes. Nat. Cell Biol. 7:186–94 [Google Scholar]
  30. Kondo S, Saito A, Hino S, Murakami T, Ogata M et al.29.  2007. BBF2H7, a novel transmembrane bZIP transcription factor, is a new type of endoplasmic reticulum stress transducer. Mol. Cell. Biol. 27:1716–29 [Google Scholar]
  31. Stirling J, O'Hare P. 30.  2006. CREB4, a transmembrane bZip transcription factor and potential new substrate for regulation and cleavage by S1P. Mol. Biol. Cell 17:413–26 [Google Scholar]
  32. DenBoer LM, Hardy-Smith PW, Hogan MR, Cockram GP, Audas TE et al.31.  2005. Luman is capable of binding and activating transcription from the unfolded protein response element. Biochem. Biophys. Res. Commun. 331:113–19 [Google Scholar]
  33. Haze K, Yoshida H, Yanagi H, Yura T, Mori K. 32.  1999. Mammalian transcription factor ATF6 is synthesized as a transmembrane protein and activated by proteolysis in response to endoplasmic reticulum stress. Mol. Biol. Cell 10:3787–99 [Google Scholar]
  34. Ye J, Rawson RB, Komuro R, Chen X, Dave UP et al.33.  2000. ER stress induces cleavage of membrane-bound ATF6 by the same proteases that process SREBPs. Mol. Cell 6:1355–64 [Google Scholar]
  35. Yoshida H, Okada T, Haze K, Yanagi H, Yura T et al.34.  2000. ATF6 activated by proteolysis binds in the presence of NF-Y (CBF) directly to the cis-acting element responsible for the mammalian unfolded protein response. Mol. Cell. Biol. 20:6755–67 [Google Scholar]
  36. Nakanishi K, Sudo T, Morishima N. 35.  2005. Endoplasmic reticulum stress signaling transmitted by ATF6 mediates apoptosis during muscle development. J. Cell Biol. 169:555–60 [Google Scholar]
  37. Gotoh T, Oyadomari S, Mori K, Mori M. 36.  2002. Nitric oxide-induced apoptosis in RAW 264.7 macrophages is mediated by endoplasmic reticulum stress pathway involving ATF6 and CHOP. J. Biol. Chem. 277:12343–50 [Google Scholar]
  38. Oakes SA, Lin SS, Bassik MC. 37.  2006. The control of endoplasmic reticulum-initiated apoptosis by the BCL-2 family of proteins. Curr. Mol. Med. 6:99–109 [Google Scholar]
  39. Wei MC, Zong WX, Cheng EH, Lindsten T, Panoutsakopoulou V et al.38.  2001. Proapoptotic BAX and BAK: a requisite gateway to mitochondrial dysfunction and death. Science 292:727–30 [Google Scholar]
  40. Nutt LK, Pataer A, Pahler J, Fang B, Roth J et al.39.  2002. Bax and Bak promote apoptosis by modulating endoplasmic reticular and mitochondrial Ca2+ stores. J. Biol. Chem. 277:9219–25 [Google Scholar]
  41. Zong WX, Li C, Hatzivassiliou G, Lindsten T, Yu QC et al.40.  2003. Bax and Bak can localize to the endoplasmic reticulum to initiate apoptosis. J. Cell Biol. 162:59–69 [Google Scholar]
  42. Oakes SA, Scorrano L, Opferman JT, Bassik MC, Nishino M et al.41.  2005. Proapoptotic BAX and BAK regulate the type 1 inositol trisphosphate receptor and calcium leak from the endoplasmic reticulum. Proc. Natl. Acad. Sci. USA 102:105–10 [Google Scholar]
  43. Hetz C, Bernasconi P, Fisher J, Lee AH, Bassik MC et al.42.  2006. Proapoptotic BAX and BAK modulate the unfolded protein response by a direct interaction with IRE1α. Science 312:572–76 [Google Scholar]
  44. Rutkowski DT, Arnold SM, Miller CN, Wu J, Li J et al.43.  2006. Adaptation to ER stress is mediated by differential stabilities of prosurvival and proapoptotic mRNAs and proteins. PLoS Biol. 4:e374 [Google Scholar]
  45. Lin JH, Li H, Yasumura D, Cohen HR, Zhang C et al.43a.  2007. IRE1 signaling affects cell fate during the unfolded protein response. Science. In press [Google Scholar]
  46. Delepine M, Nicolino M, Barrett T, Golamaully M, Lathrop GM, Julier C. 44.  2000. EIF2AK3, encoding translation initiation factor 2-α kinase 3, is mutated in patients with Wolcott-Rallison syndrome. Nat. Genet. 25:406–9 [Google Scholar]
  47. Harding HP, Zeng H, Zhang Y, Jungries R, Chung P et al.45.  2001. Diabetes mellitus and exocrine pancreatic dysfunction in perk−/− mice reveals a role for translational control in secretory cell survival. Mol. Cell. 7:1153–63 [Google Scholar]
  48. Scheuner D, Song B, McEwen E, Liu C, Laybutt R et al.46.  2001. Translational control is required for the unfolded protein response and in vivo glucose homeostasis. Mol. Cell 7:1165–76 [Google Scholar]
  49. Zhang W, Feng D, Li Y, Iida K, McGrath B et al.47.  2006. PERK EIF2AK3 control of pancreatic β cell differentiation and proliferation is required for postnatal glucose homeostasis. Cell Metab. 4:491–97 [Google Scholar]
  50. Lee AH, Chu GC, Iwakoshi NN, Glimcher LH. 48.  2005. XBP-1 is required for biogenesis of cellular secretory machinery of exocrine glands. EMBO J. 24:4368–80 [Google Scholar]
  51. Inoue H, Tanizawa Y, Wasson J, Behn P, Kalidas K et al.49.  1998. A gene encoding a transmembrane protein is mutated in patients with diabetes mellitus and optic atrophy (Wolfram syndrome). Nat. Genet. 20:143–48 [Google Scholar]
  52. Strom TM, Hortnagel K, Hofmann S, Gekeler F, Scharfe C et al.50.  1998. Diabetes insipidus, diabetes mellitus, optic atrophy and deafness (DIDMOAD) caused by mutations in a novel gene (wolframin) coding for a predicted transmembrane protein. Hum. Mol. Genet. 7:2021–28 [Google Scholar]
  53. Fonseca SG, Fukuma M, Lipson KL, Nguyen LX, Allen JR et al.51.  2005. WFS1 is a novel component of the unfolded protein response and maintains homeostasis of the endoplasmic reticulum in pancreatic β-cells. J. Biol. Chem. 280:39609–15 [Google Scholar]
  54. Ueda K, Kawano J, Takeda K, Yujiri T, Tanabe K et al.52.  2005. Endoplasmic reticulum stress induces Wfs1 gene expression in pancreatic β-cells via transcriptional activation. Eur. J. Endocrinol. 153:167–76 [Google Scholar]
  55. Ishihara H, Takeda S, Tamura A, Takahashi R, Yamaguchi S et al.53.  2004. Disruption of the WFS1 gene in mice causes progressive β-cell loss and impaired stimulus-secretion coupling in insulin secretion. Hum. Mol. Genet. 13:1159–70 [Google Scholar]
  56. Riggs AC, Bernal-Mizrachi E, Ohsugi M, Wasson J, Fatrai S et al.54.  2005. Mice conditionally lacking the Wolfram gene in pancreatic islet β cells exhibit diabetes as a result of enhanced endoplasmic reticulum stress and apoptosis. Diabetologia 48:2313–21 [Google Scholar]
  57. Wang J, Takeuchi T, Tanaka S, Kubo SK, Kayo T et al.55.  1999. A mutation in the insulin 2 gene induces diabetes with severe pancreatic β-cell dysfunction in the Mody mouse. J. Clin. Invest. 103:27–37 [Google Scholar]
  58. Leroux L, Desbois P, Lamotte L, Duvillie B, Cordonnier N et al.56.  2001. Compensatory responses in mice carrying a null mutation for Ins1 or Ins2. Diabetes 50:S150–53 [Google Scholar]
  59. Oyadomari S, Koizumi A, Takeda K, Gotoh T, Akira S et al.57.  2002. Targeted disruption of the Chop gene delays endoplasmic reticulum stress–mediated diabetes. J. Clin. Invest. 109:525–32 [Google Scholar]
  60. Cardozo AK, Ortis F, Storling J, Feng YM, Rasschaert J et al.58.  2005. Cytokines downregulate the sarcoendoplasmic reticulum pump Ca2+ ATPase 2b and deplete endoplasmic reticulum Ca2+, leading to induction of endoplasmic reticulum stress in pancreatic β-cells. Diabetes 54:452–61 [Google Scholar]
  61. Oyadomari S, Takeda K, Takiguchi M, Gotoh T, Matsumoto M et al.59.  2001. Nitric oxide-induced apoptosis in pancreatic β cells is mediated by the endoplasmic reticulum stress pathway. Proc. Natl. Acad. Sci. USA 98:10845–50 [Google Scholar]
  62. Laybutt DR, Preston AM, Akerfeldt MC, Kench JG, Busch AK et al.60.  2007. Endoplasmic reticulum stress contributes to β cell apoptosis in type 2 diabetes. Diabetologia 1:1 [Google Scholar]
  63. Lipson KL, Fonseca SG, Ishigaki S, Nguyen LX, Foss E et al.61.  2006. Regulation of insulin biosynthesis in pancreatic β cells by an endoplasmic reticulum-resident protein kinase IRE1. Cell Metab. 4:245–54 [Google Scholar]
  64. Cnop M, Ladriere L, Hekerman P, Ortis F, Cardozo AK et al.62.  2007. Selective inhibition of eukaryotic translation initiation factor 2α dephosphorylation potentiates fatty acid-induced endoplasmic reticulum stress and causes pancreatic β-cell dysfunction and apoptosis. J. Biol. Chem. 282:3989–97 [Google Scholar]
  65. Scheuner D, Vander Mierde D, Song B, Flamez D, Creemers JW et al.63.  2005. Control of mRNA translation preserves endoplasmic reticulum function in β cells and maintains glucose homeostasis. Nat. Med. 11:757–64 [Google Scholar]
  66. Özcan U, Cao Q, Yilmaz E, Lee AH, Iwakoshi NN et al.64.  2004. Endoplasmic reticulum stress links obesity, insulin action, and type 2 diabetes. Science 306:457–61 [Google Scholar]
  67. Du K, Herzig S, Kulkarni RN, Montminy M. 65.  2003. TRB3: a tribbles homolog that inhibits Akt/PKB activation by insulin in liver. Science 300:1574–77 [Google Scholar]
  68. Özcan U, Yilmaz E, Özcan L, Furuhashi M, Vaillancourt E et al.66.  2006. Chemical chaperones reduce ER stress and restore glucose homeostasis in a mouse model of type 2 diabetes. Science 313:1137–40 [Google Scholar]
  69. Nakatani Y, Kaneto H, Kawamori D, Yoshiuchi K, Hatazaki M et al.67.  2005. Involvement of endoplasmic reticulum stress in insulin resistance and diabetes. J. Biol. Chem. 280:847–51 [Google Scholar]
  70. Thameem F, Farook VS, Bogardus C, Prochazka M. 67a.  2006. Association of amino acid variants in the activating transcription factor 6 gene (ATF6) on 1q21-q23 with type 2 diabetes in Pima Indians. Diabetes 55:839–42 [Google Scholar]
  71. Meex SJR, van Greevenbroek MMJ, Ayoubi TA, Vlietinck R, van Vliet-Ostaptchouk JV et al.67b.  2007. Activating transcription factor 6 polymorphisms and haplotypes are associated with impaired glucose homeostasis and type 2 diabetes in Dutch Caucasians. J. Clin. Endocrinol. Metabol. 92:2720–25 [Google Scholar]
  72. Kharroubi I, Ladriere L, Cardozo AK, Dogusan Z, Cnop M, Eizirik DL. 68.  2004. Free fatty acids and cytokines induce pancreatic β-cell apoptosis by different mechanisms: role of nuclear factor-κB and endoplasmic reticulum stress. Endocrinology 145:5087–96 [Google Scholar]
  73. Harding HP, Zhang Y, Zeng H, Novoa I, Lu PD et al.69.  2003. An integrated stress response regulates amino acid metabolism and resistance to oxidative stress. Mol. Cell 11:619–33 [Google Scholar]
  74. Yusta B, Baggio LL, Estall JL, Koehler JA, Holland DP et al.70.  2006. GLP-1 receptor activation improves β cell function and survival following induction of endoplasmic reticulum stress. Cell Metab. 4:391–406 [Google Scholar]
  75. Lu PD, Jousse C, Marciniak SJ, Zhang Y, Novoa I et al.71.  2004. Cytoprotection by pre-emptive conditional phosphorylation of translation initiation factor 2. EMBO J. 23:169–79 [Google Scholar]
  76. Baltzis D, Qu LK, Papadopoulou S, Blais JD, Bell JC et al.72.  2004. Resistance to vesicular stomatitis virus infection requires a functional cross talk between the eukaryotic translation initiation factor 2α kinases PERK and PKR. J. Virol. 78:12747–61 [Google Scholar]
  77. Perkins DJ, Barber GN. 73.  2004. Defects in translational regulation mediated by the α subunit of eukaryotic initiation factor 2 inhibit antiviral activity and facilitate the malignant transformation of human fibroblasts. Mol. Cell. Biol. 24:2025–40 [Google Scholar]
  78. Chou J, Roizman B. 74.  1994. Herpes simplex virus 1 γ(1)34.5 gene function, which blocks the host response to infection, maps in the homologous domain of the genes expressed during growth arrest and DNA damage. Proc. Natl. Acad. Sci. USA 91:5247–51 [Google Scholar]
  79. Whitley RJ, Kern ER, Chatterjee S, Chou J, Roizman B. 75.  1993. Replication, establishment of latency, and induced reactivation of herpes simplex virus γ 1 34.5 deletion mutants in rodent models. J. Clin. Invest. 91:2837–43 [Google Scholar]
  80. Boyce M, Bryant KF, Jousse C, Long K, Harding HP et al.76.  2005. A selective inhibitor of eIF2α dephosphorylation protects cells from ER stress. Science 307:935–39 [Google Scholar]
  81. Netherton CL, Parsley JC, Wileman T. 77.  2004. African swine fever virus inhibits induction of the stress-induced proapoptotic transcription factor CHOP/GADD153. J. Virol. 78:10825–28 [Google Scholar]
  82. Rivera J, Abrams C, Hernaez B, Alcazar A, Escribano JM et al.78.  2007. The MyD116 African swine fever virus homologue interacts with the catalytic subunit of protein phosphatase 1 and activates its phosphatase activity. J. Virol. 81:2923–29 [Google Scholar]
  83. Pavio N, Romano PR, Graczyk TM, Feinstone SM, Taylor DR. 79.  2003. Protein synthesis and endoplasmic reticulum stress can be modulated by the hepatitis C virus envelope protein E2 through the eukaryotic initiation factor 2α kinase PERK. J. Virol. 77:3578–85 [Google Scholar]
  84. Mulvey M, Arias C, Mohr I. 80.  2007. Maintenance of ER homeostasis in HSV-1 infected cells through the association of a viral glycoprotein with PERK, a cellular ER-stress sensor. J. Virol. 17:17 [Google Scholar]
  85. Liberman E, Fong YL, Selby MJ, Choo QL, Cousens L et al.81.  1999. Activation of the grp78 and grp94 promoters by hepatitis C virus E2 envelope protein. J. Virol. 73:3718–22 [Google Scholar]
  86. Isler JA, Skalet AH, Alwine JC. 82.  2005. Human cytomegalovirus infection activates and regulates the unfolded protein response. J. Virol. 79:6890–99 [Google Scholar]
  87. Tardif KD, Mori K, Kaufman RJ, Siddiqui A. 83.  2004. Hepatitis C virus suppresses the IRE1-XBP1 pathway of the unfolded protein response. J. Biol. Chem. 279:17158–64 [Google Scholar]
  88. Yoshida H, Matsui T, Hosokawa N, Kaufman RJ, Nagata K et al.84.  2003. A time-dependent phase shift in the mammalian unfolded protein response. Dev. Cell 4:265–71 [Google Scholar]
  89. Tirosh B, Iwakoshi NN, Lilley BN, Lee AH, Glimcher LH et al.85.  2005. Human cytomegalovirus protein US11 provokes an unfolded protein response that may facilitate the degradation of class I major histocompatibility complex products. J. Virol. 79:2768–79 [Google Scholar]
  90. Smith JA, Schmechel SC, Raghavan A, Abelson M, Reilly C et al.86.  2006. Reovirus induces and benefits from an integrated cellular stress response. J. Virol. 80:2019–33 [Google Scholar]
  91. Huang ZM, Tan T, Yoshida H, Mori K, Ma Y et al.87.  2005. Activation of hepatitis B virus S promoter by a cell type-restricted IRE1-dependent pathway induced by endoplasmic reticulum stress. Mol. Cell. Biol. 25:7522–33 [Google Scholar]
  92. Liu N, Scofield VL, Qiang W, Yan M, Kuang X et al.88.  2006. Interaction between endoplasmic reticulum stress and caspase 8 activation in retrovirus MoMuLV-ts1-infected astrocytes. Virology 348:398–405 [Google Scholar]
  93. Dimcheff DE, Askovic S, Baker AH, Johnson-Fowler C, Portis JL. 89.  2003. Endoplasmic reticulum stress is a determinant of retrovirus-induced spongiform neurodegeneration. J. Virol. 77:12617–29 [Google Scholar]
  94. Dimcheff DE, Faasse MA, McAtee FJ, Portis JL. 90.  2004. Endoplasmic reticulum (ER) stress induced by a neurovirulent mouse retrovirus is associated with prolonged BiP binding and retention of a viral protein in the ER. J. Biol. Chem. 279:33782–90 [Google Scholar]
  95. Williams BL, Lipkin WI. 91.  2006. Endoplasmic reticulum stress and neurodegeneration in rats neonatally infected with borna disease virus. J. Virol. 80:8613–26 [Google Scholar]
  96. Jordan R, Wang L, Graczyk TM, Block TM, Romano PR. 92.  2002. Replication of a cytopathic strain of bovine viral diarrhea virus activates PERK and induces endoplasmic reticulum stress-mediated apoptosis of MDBK cells. J. Virol. 76:9588–99 [Google Scholar]
  97. Su HL, Liao CL, Lin YL. 93.  2002. Japanese encephalitis virus infection initiates endoplasmic reticulum stress and an unfolded protein response. J. Virol. 76:4162–71 [Google Scholar]
  98. Davies SE, Portmann BC, O'Grady JG, Aldis PM, Chaggar K et al.94.  1991. Hepatic histological findings after transplantation for chronic hepatitis B virus infection, including a unique pattern of fibrosing cholestatic hepatitis. Hepatology 13:150–57 [Google Scholar]
  99. Meuleman P, Libbrecht L, Wieland S, De Vos R, Habib N et al.95.  2006. Immune suppression uncovers endogenous cytopathic effects of the hepatitis B virus. J. Virol. 80:2797–807 [Google Scholar]
  100. Lau JY, Bain VG, Davies SE, O'Grady JG, Alberti A et al.96.  1992. High-level expression of hepatitis B viral antigens in fibrosing cholestatic hepatitis. Gastroenterology 102:956–62 [Google Scholar]
  101. Foo NC, Ahn BY, Ma X, Hyun W, Yen TS. 97.  2002. Cellular vacuolization and apoptosis induced by hepatitis B virus large surface protein. Hepatology 36:1400–7 [Google Scholar]
  102. Xu Z, Jensen G, Yen TS. 98.  1997. Activation of hepatitis B virus S promoter by the viral large surface protein via induction of stress in the endoplasmic reticulum. J. Virol. 71:7387–92 [Google Scholar]
  103. Berson EL. 99.  1993. Retinitis pigmentosa. The Friedenwald Lecture.. Invest. Ophthalmol. Vis. Sci. 34:1659–76 [Google Scholar]
  104. Sohocki MM, Daiger SP, Bowne SJ, Rodriquez JA, Northrup H et al.100.  2001. Prevalence of mutations causing retinitis pigmentosa and other inherited retinopathies. Hum. Mutat. 17:42–51 [Google Scholar]
  105. Liu X, Garriga P, Khorana HG. 101.  1996. Structure and function in rhodopsin: correct folding and misfolding in two point mutants in the intradiscal domain of rhodopsin identified in retinitis pigmentosa. Proc. Natl. Acad. Sci. USA 93:4554–59 [Google Scholar]
  106. Kaushal S, Khorana HG. 102.  1994. Structure and function in rhodopsin. 7. Point mutations associated with autosomal dominant retinitis pigmentosa. Biochemistry 33:6121–28 [Google Scholar]
  107. Sung CH, Schneider BG, Agarwal N, Papermaster DS, Nathans J. 103.  1991. Functional heterogeneity of mutant rhodopsins responsible for autosomal dominant retinitis pigmentosa. Proc. Natl. Acad. Sci. USA 88:8840–44 [Google Scholar]
  108. Ryoo HD, Domingos PM, Kang MJ, Steller H. 104.  2007. Unfolded protein response in a Drosophila model for retinal degeneration. EMBO J. 26:242–52 [Google Scholar]
  109. 105. Natl. Inst. Aging Reagan Inst. Work. Group Diagn. Criteria Neuropathol. Assess. Alzheimer's Dis. 1997. Consensus recommendations for the postmortem diagnosis of Alzheimer's disease. Neurobiol. Aging 18:S1–2 [Google Scholar]
  110. Unterberger U, Hoftberger R, Gelpi E, Flicker H, Budka H et al.106.  2006. Endoplasmic reticulum stress features are prominent in Alzheimer disease but not in prion diseases in vivo. J. Neuropathol. Exp. Neurol. 65:348–57 [Google Scholar]
  111. Hoozemans JJ, Veerhuis R, Van Haastert ES, Rozemuller JM, Baas F et al.107.  2005. The unfolded protein response is activated in Alzheimer's disease. Acta Neuropathol. 110:165–72 [Google Scholar]
  112. Selkoe DJ. 108.  2001. Alzheimer's disease: genes, proteins, and therapy. Physiol. Rev. 81:741–66 [Google Scholar]
  113. Schenk D, Barbour R, Dunn W, Gordon G, Grajeda H et al.109.  1999. Immunization with amyloid-β attenuates Alzheimer-disease-like pathology in the PDAPP mouse. Nature 400:173–77 [Google Scholar]
  114. Ferreiro E, Resende R, Costa R, Oliveira CR, Pereira CM. 110.  2006. An endoplasmic-reticulum-specific apoptotic pathway is involved in prion and amyloid-β peptides neurotoxicity. Neurobiol. Dis. 23:669–78 [Google Scholar]
  115. Kadowaki H, Nishitoh H, Urano F, Sadamitsu C, Matsuzawa A et al.111.  2005. Amyloid β induces neuronal cell death through ROS-mediated ASK1 activation. Cell Death Differ. 12:19–24 [Google Scholar]
  116. Kogel D, Schomburg R, Schurmann T, Reimertz C, Konig HG et al.112.  2003. The amyloid precursor protein protects PC12 cells against endoplasmic reticulum stress-induced apoptosis. J. Neurochem. 87:248–56 [Google Scholar]
  117. Esposito L, Gan L, Yu GQ, Essrich C, Mucke L. 113.  2004. Intracellularly generated amyloid-β peptide counteracts the antiapoptotic function of its precursor protein and primes proapoptotic pathways for activation by other insults in neuroblastoma cells. J. Neurochem. 91:1260–74 [Google Scholar]
  118. Koo EH, Kopan R. 114.  2004. Potential role of presenilin-regulated signaling pathways in sporadic neurodegeneration. Nat. Med. 10:Suppl.S26–33 [Google Scholar]
  119. Tu H, Nelson O, Bezprozvanny A, Wang Z, Lee SF et al.115.  2006. Presenilins form ER Ca2+ leak channels, a function disrupted by familial Alzheimer's disease-linked mutations. Cell 126:981–93 [Google Scholar]
  120. Tandon A, Fraser P. 116.  2002. The presenilins. Genome Biol. 3:3014–20 [Google Scholar]
  121. Leissring MA, Akbari Y, Fanger CM, Cahalan MD, Mattson MP et al.117.  2000. Capacitative calcium entry deficits and elevated luminal calcium content in mutant presenilin-1 knockin mice. J. Cell Biol. 149:793–98 [Google Scholar]
  122. Yoo AS, Cheng I, Chung S, Grenfell TZ, Lee H et al.118.  2000. Presenilin-mediated modulation of capacitative calcium entry. Neuron 27:561–72 [Google Scholar]
  123. Lytton J, Westlin M, Hanley MR. 119.  1991. Thapsigargin inhibits the sarcoplasmic or endoplasmic reticulum Ca-ATPase family of calcium pumps. J. Biol. Chem. 266:17067–71 [Google Scholar]
  124. Bailly-Maitre B, Fondevila C, Kaldas F, Droin N, Luciano F et al.120.  2006. Cytoprotective gene bi-1 is required for intrinsic protection from endoplasmic reticulum stress and ischemia-reperfusion injury. Proc. Natl. Acad. Sci. USA 103:2809–14 [Google Scholar]
  125. Koumenis C, Naczki C, Koritzinsky M, Rastani S, Diehl A et al.121.  2002. 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. 22:7405–16 [Google Scholar]
  126. Bi M, Naczki C, Koritzinsky M, Fels D, Blais J et al.122.  2005. ER stress-regulated translation increases tolerance to extreme hypoxia and promotes tumor growth. EMBO J. 24:3470–81 [Google Scholar]
  127. Koumenis C. 123.  2006. ER stress, hypoxia tolerance and tumor progression. Curr. Mol. Med. 6:55–69 [Google Scholar]
  128. Blais JD, Addison CL, Edge R, Falls T, Zhao H et al.124.  2006. Perk-dependent translational regulation promotes tumor cell adaptation and angiogenesis in response to hypoxic stress. Mol. Cell. Biol. 26:9517–32 [Google Scholar]
  129. Arnal M, Solary E, Brunet-Lecomte P, Lizard-Nacol S. 125.  1999. Expression of the gadd153 gene in normal and tumor breast tissues by a sensitive RT-PCR method. Int. J. Mol. Med. 4:545–48 [Google Scholar]
  130. Forus A, Florenes VA, Maelandsmo GM, Fodstad O, Myklebost O. 126.  1994. The protooncogene CHOP/GADD153, involved in growth arrest and DNA damage response, is amplified in a subset of human sarcomas. Cancer Genet. Cytogenet. 78:165–71 [Google Scholar]
  131. Southwood CM, Garbern J, Jiang W, Gow A. 127.  2002. The unfolded protein response modulates disease severity in Pelizaeus-Merzbacher disease. Neuron 36:585–96 [Google Scholar]
  132. Mayerhofer T, Kodym R. 128.  2003. Gadd153 restores resistance to radiation-induced apoptosis after thiol depletion. Biochem. Biophys. Res. Commun. 310:115–20 [Google Scholar]
  133. Romero-Ramirez L, Cao H, Nelson D, Hammond E, Lee AH et al.129.  2004. XBP1 is essential for survival under hypoxic conditions and is required for tumor growth. Cancer Res. 64:5943–47 [Google Scholar]
  134. Jamora C, Dennert G, Lee AS. 130.  1996. Inhibition of tumor progression by suppression of stress protein GRP78/BiP induction in fibrosarcoma B/C10ME. Proc. Natl. Acad. Sci. USA 93:7690–94 [Google Scholar]
  135. Pluquet O, Qu LK, Baltzis D, Koromilas AE. 131.  2005. Endoplasmic reticulum stress accelerates p53 degradation by the cooperative actions of Hdm2 and glycogen synthase kinase 3β. Mol. Cell. Biol. 25:9392–405 [Google Scholar]
  136. Reimold AM, Iwakoshi NN, Manis J, Vallabhajosyula P, Szomolanyi-Tsuda E et al.132.  2001. Plasma cell differentiation requires the transcription factor XBP-1. Nature 412:300–7 [Google Scholar]
  137. Zhang K, Wong HN, Song B, Miller CN, Scheuner D et al.133.  2005. The unfolded protein response sensor IRE1α is required at 2 distinct steps in B cell lymphopoiesis. J. Clin. Invest. 115:268–81 [Google Scholar]
  138. Carrasco DR, Sukhdeo K, Protopopova M, Sinha R, Enos M et al.133a.  2007. The differentiation and stress response factor XBP-1 drives multiple myeloma pathogenesis. Cancer Cell 11:349–60 [Google Scholar]
  139. Richardson PG, Sonneveld P, Schuster MW, Irwin D, Stadtmauer EA et al.134.  2005. Bortezomib or high-dose dexamethasone for relapsed multiple myeloma. N. Engl. J. Med. 352:2487–98 [Google Scholar]
  140. Teicher BA, Ara G, Herbst R, Palombella VJ, Adams J. 135.  1999. The proteasome inhibitor PS-341 in cancer therapy. Clin. Cancer Res. 5:2638–45 [Google Scholar]
  141. Lee AH, Iwakoshi NN, Anderson KC, Glimcher LH. 136.  2003. Proteasome inhibitors disrupt the unfolded protein response in myeloma cells. Proc. Natl. Acad. Sci. USA 100:9946–51 [Google Scholar]
  142. Obeng EA, Carlson LM, Gutman DM, Harrington WJ Jr, Lee KP et al.137.  2006. Proteasome inhibitors induce a terminal unfolded protein response in multiple myeloma cells. Blood 107:4907–16 [Google Scholar]
  143. Jiang HY, Wek RC. 138.  2005. Phosphorylation of the α-subunit of the eukaryotic initiation factor-2 (eIF2α) reduces protein synthesis and enhances apoptosis in response to proteasome inhibition. J. Biol. Chem. 280:14189–202 [Google Scholar]
  144. Nawrocki ST, Carew JS, Pino MS, Highshaw RA, Dunner K Jr et al.139.  2005. Bortezomib sensitizes pancreatic cancer cells to endoplasmic reticulum stress-mediated apoptosis. Cancer Res. 65:11658–66 [Google Scholar]
  145. Gray MD, Mann M, Nitiss JL, Hendershot LM. 140.  2005. Activation of the UPR is necessary and sufficient for reducing topoisomerase IIα protein levels and decreasing sensitivity to topoisomerase targeted drugs. Mol. Pharmacol. 68:1699–707 [Google Scholar]
  146. Reddy RK, Mao C, Baumeister P, Austin RC, Kaufman RJ et al.141.  2003. 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. 278:20915–24 [Google Scholar]
  147. Carew JS, Nawrocki ST, Krupnik YV, Dunner K Jr, McConkey DJ et al.142.  2006. Targeting endoplasmic reticulum protein transport: a novel strategy to kill malignant B cells and overcome fludarabine resistance in CLL. Blood 107:222–31 [Google Scholar]
  148. Fabian Z, Csatary CM, Szeberenyi J, Csatary LK. 143.  2007. p53-Independent endoplasmic reticulum stress-mediated cytotoxicity of a Newcastle disease virus strain in tumor cell lines. J. Virol. 81:2817–30 [Google Scholar]
  149. Shen J, Hughes C, Chao C, Cai J, Bartels C et al.144.  1987. Coinduction of glucose-regulated proteins and doxorubicin resistance in Chinese hamster cells. Proc. Natl. Acad. Sci. USA 84:3278–82 [Google Scholar]
  150. Tomida A, Yun J, Tsuruo T. 145.  1996. Glucose-regulated stresses induce resistance to camptothecin in human cancer cells. Int. J. Cancer 68:391–96 [Google Scholar]
  151. Chatterjee S, Hirota H, Belfi CA, Berger SJ, Berger NA. 146.  1997. Hypersensitivity to DNA cross-linking agents associated with up-regulation of glucose-regulated stress protein GRP78. Cancer Res. 57:5112–16 [Google Scholar]
  152. Mandic A, Hansson J, Linder S, Shoshan MC. 147.  2003. Cisplatin induces endoplasmic reticulum stress and nucleus-independent apoptotic signaling. J. Biol. Chem. 278:9100–6 [Google Scholar]
  153. Dong D, Ko B, Baumeister P, Swenson S, Costa F et al.148.  2005. Vascular targeting and antiangiogenesis agents induce drug resistance effector GRP78 within the tumor microenvironment. Cancer Res. 65:5785–91 [Google Scholar]
  154. Kerkela R, Grazette L, Yacobi R, Iliescu C, Patten R et al.149.  2006. Cardiotoxicity of the cancer therapeutic agent imatinib mesylate. Nat. Med. 12:908–16 [Google Scholar]
  155. Oda Y, Okada T, Yoshida H, Kaufman RJ, Nagata K et al.150.  2006. Derlin-2 and Derlin-3 are regulated by the mammalian unfolded protein response and are required for ER-associated degradation. J. Cell Biol. 172:383–93 [Google Scholar]
/content/journals/10.1146/annurev.pathmechdis.3.121806.151434
Loading
/content/journals/10.1146/annurev.pathmechdis.3.121806.151434
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error