Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter September 29, 2011

ERp57/GRP58: A protein with multiple functions

  • Carlo Turano EMAIL logo , Elisa Gaucci , Caterina Grillo and Silvia Chichiarelli

Abstract

The protein ERp57/GRP58 is a stress-responsive protein and a component of the protein disulfide isomerase family. Its functions in the endoplasmic reticulum are well known, concerning mainly the proper folding and quality control of glycoproteins, and participation in the assembly of the major histocompatibility complex class 1. However, ERp57 is present in many other subcellular locations, where it is involved in a variety of functions, primarily suggested by its participation in complexes with other proteins and even with DNA. While in some instances these roles need to be confirmed by further studies, a great number of observations support the participation of ERp57 in signal transduction from the cell surface, in regulatory processes taking place in the nucleus, and in multimeric protein complexes involved in DNA repair.

[1] Turano, C., Coppari, S., Altieri, F. and Ferraro, A. Proteins of the PDI family: unpredicted non-ER locations and functions. J. Cell. Physiol. 193 (2002) 154–163. http://dx.doi.org/10.1002/jcp.1017210.1002/jcp.10172Search in Google Scholar PubMed

[2] Khanal, R.C. and Nemere, I. The ERp57/GRp58/1,25D3-MARRS receptor: multiple functional roles in diverse cell systems. Curr. Med. Chem. 14 (2007) 1087–1093. http://dx.doi.org/10.2174/09298670778036287110.2174/092986707780362871Search in Google Scholar PubMed

[3] Coe, H. and Michalak, M. ERp57, a multifunctional endoplasmic reticulum resident oxidoreductase. Int. J. Biochem. Cell Biol. 42 (2010) 796–799. http://dx.doi.org/10.1016/j.biocel.2010.01.00910.1016/j.biocel.2010.01.009Search in Google Scholar PubMed

[4] Bennett, C.F., Balcarek, J.M., Varrichio, A. and Crooke, S.T. Molecular cloning and complete amino-acid sequence of form-I phosphoinositidespecific phospholipase C. Nature 334 (1988) 268–270. http://dx.doi.org/10.1038/334268a010.1038/334268a0Search in Google Scholar PubMed

[5] Lee, A.S. The accumulation of three specific proteins related to glucoseregulated proteins in a temperature-sensitive hamster mutant cell line K12. J. Cell. Physiol. 106 (1981) 119–125. http://dx.doi.org/10.1002/jcp.104106011310.1002/jcp.1041060113Search in Google Scholar PubMed

[6] Ferrari, D.M., Söling, H.D. The protein disulphide-isomerase family: unravelling a string of folds. Biochem. J. 339 (1999) 1–10. http://dx.doi.org/10.1042/0264-6021:339000110.1042/0264-6021:3390001Search in Google Scholar

[7] Silvennoinen, L., Myllyharju, J., Ruoppolo, M., Orrù, S., Caterino, M., Kivirikko, K.I. and Koivunen, P. Identification and characterization of structural domains of human ERp57: association with calreticulin requires several domains. J. Biol. Chem. 279 (2004) 13607–13615. http://dx.doi.org/10.1074/jbc.M31305420010.1074/jbc.M313054200Search in Google Scholar PubMed

[8] Russell, S.J., Ruddock, L.W., Salo, K.E., Oliver, J.D., Roebuck, Q.P., Llewellyn, D.H., Roderick, H.L., Koivunen, P., Myllyharju, J. and High, S. The primary substrate binding site in the b’ domain of ERp57 is adapted for endoplasmic reticulum lectin association. J. Biol. Chem. 279 (2004) 18861–18869. http://dx.doi.org/10.1074/jbc.M40057520010.1074/jbc.M400575200Search in Google Scholar PubMed

[9] Kozlov, G., Maattanen, P., Schrag, J.D., Pollock, S., Cygler, M., Nagar, B., Thomas, D.Y. and Gehring, K. Crystal structure of the bb’ domains of the protein disulfideisomerase ERp57. Structure 14 (2006) 1331–1339. http://dx.doi.org/10.1016/j.str.2006.06.01910.1016/j.str.2006.06.019Search in Google Scholar PubMed

[10] Klappa, P., Ruddock, L.W., Darby, N.J. and Freedman, R.B. The b′ domain provides the principal peptide-binding site of protein disulfide isomerase but all domains contribute to binding of misfolded proteins. EMBO J. 17 (1998) 927–935. http://dx.doi.org/10.1093/emboj/17.4.92710.1093/emboj/17.4.927Search in Google Scholar PubMed PubMed Central

[11] Gaucci, E., Chichiarelli, S., Grillo, C., Del Vecchio, E., Eufemi, M. and Turano, C. The binding of antibiotics to ERp57/GRP58. J. Antibiot. (Tokyo) 61 (2008) 400–402. http://dx.doi.org/10.1038/ja.2008.5610.1038/ja.2008.56Search in Google Scholar

[12] Dick, T.P., Bangia, N., Peaper, D.R. and Cresswell, P. Disulfide bond isomerization and the assembly of MHC class I-peptide complexes. Immunity 16 (2002) 87–98. http://dx.doi.org/10.1016/S1074-7613(02)00263-710.1016/S1074-7613(02)00263-7Search in Google Scholar

[13] Grillo, C., D’Ambrosio, C., Scaloni, A., Maceroni, M., Merluzzi, S., Turano, C. and Altieri, F. Cooperative activity of Ref-1/APE and ERp57 in reductive activation of transcription factors. Free Radic. Biol. Med. 41 (2006) 1113–1123. http://dx.doi.org/10.1016/j.freeradbiomed.2006.06.01610.1016/j.freeradbiomed.2006.06.016Search in Google Scholar

[14] Donella-Deana, A., James, P., Staudenmann, W., Cesaro, L., Marin, O., Brunati, A.M., Ruzzene, M. and Pinna, L.A. Isolation from spleen of a 57-kDa protein substrate of the tyrosine kinase Lyn. Identification as a protein related to protein disulfide-isomerase and localisation of the phosphorylation sites. Eur. J. Biochem. 235 (1996) 18–25. http://dx.doi.org/10.1111/j.1432-1033.1996.00018.x10.1111/j.1432-1033.1996.00018.xSearch in Google Scholar

[15] Kita, K., Okumura, N., Takao, T., Watanabe, M., Matsubara, T., Nishimura, O. and Nagai, K. Evidence for phosphorylation of rat liver glucose-regulated protein 58, GRP58/ERp57/ER-60, induced by fasting and leptin. FEBS Lett. 580 (2006) 199–205. http://dx.doi.org/10.1016/j.febslet.2005.11.07410.1016/j.febslet.2005.11.074Search in Google Scholar

[16] Zhou, L., McKenzie, B.A., Eccleston, E.D. Jr, Srivastava, S.P., Chen, N., Erickson, R.R. and Holtzman, J.L. The covalent binding of [14C]acetaminophen to mouse hepatic microsomal proteins: the specific binding to calreticulin and the two forms of the thiol:protein disulfide oxidoreductases. Chem. Res. Toxicol. 9 (1996) 1176–1182. http://dx.doi.org/10.1021/tx960069d10.1021/tx960069dSearch in Google Scholar

[17] Martin, J.L., Pumford, N.R., LaRosa, A.C., Martin, B.M., Gonzaga, H.M., Beaven, M.A. and Pohl, L.R. A metabolite of halothane covalently binds to an endoplasmic reticulum protein that is highly homologous to phosphatidylinositol-specific phospholipase C-alpha but has no activity. Biochem. Biophys. Res. Commun. 178 (1991) 679–685. http://dx.doi.org/10.1016/0006-291X(91)90161-Y10.1016/0006-291X(91)90161-YSearch in Google Scholar

[18] Laragione, T., Gianazza, E., Tonelli, R., Bigini, P., Mennini, T., Casoni, F., Massignan, T., Bonetto, V. and Ghezzi, P. Regulation of redox-sensitive exofacial protein thiols in CHO cells. Biol. Chem. 387 (2006) 1371–1376. http://dx.doi.org/10.1515/BC.2006.17210.1515/BC.2006.172Search in Google Scholar

[19] van der Vlies, D., Pap, E.H., Post, J.A., Celis, J.E. and Wirtz, K.W. Endoplasmic reticulum resident proteins of normal human dermal fibroblasts are the major targets for oxidative stress induced by hydrogen peroxide. Biochem. J. 366 (2002) 825–830. Search in Google Scholar

[20] Grillo, C., D’Ambrosio, C., Consalvi, V., Chiaraluce, R., Scaloni, A., Maceroni, M., Eufemi, M. and Altieri, F. DNA-binding activity of the ERp57 C-terminal domain is related to a redox-dependent conformational change. J. Biol. Chem. 282 (2007) 10299–10310. http://dx.doi.org/10.1074/jbc.M70096620010.1074/jbc.M700966200Search in Google Scholar

[21] Freedman, R.B., Hirst, T.R. and Tuite, M.F. Protein disulphide isomerase: building bridges in protein folding. Trends Biochem. Sci. 19 (1994) 331–336. http://dx.doi.org/10.1016/0968-0004(94)90072-810.1016/0968-0004(94)90072-8Search in Google Scholar

[22] Okudo, H., Kito, M., Moriyama, T., Ogawa, T. and Urade, R. Transglutaminase activity of human ER-60. Biosci. Biotechnol. Biochem. 66 (2002) 1423–1426. http://dx.doi.org/10.1271/bbb.66.142310.1271/bbb.66.1423Search in Google Scholar

[23] Urade, R., Nasu, M., Moriyama, T., Wada, K. and Kito, M. Protein degradation by the phosphoinositide-specific phospholipase C-alpha family from rat liver endoplasmic reticulum. J. Biol. Chem. 267 (1992) 15152–15159. Search in Google Scholar

[24] Murthy, M.S. and Pande, S.V. A stress-regulated protein, GRP58, a member of thioredoxin superfamily, is a carnitine palmitoyltransferase isoenzyme. Biochem. J. 304 (1994) 31–34. Search in Google Scholar

[25] Oliver, J.D., van der Wal, F.J., Bulleid, N.J. and High, S. Interaction of the thiol-dependent reductase ERp57 with nascent glycoproteins. Science 275 (1997) 86–88. http://dx.doi.org/10.1126/science.275.5296.8610.1126/science.275.5296.86Search in Google Scholar

[26] Molinari, M. and Helenius, A. Glycoproteins form mixed disulphides with oxidoreductases during folding in living cells. Nature 402 (1999) 90–93. http://dx.doi.org/10.1038/4706210.1038/47062Search in Google Scholar

[27] Oliver, J.D., Roderick, H.L., Llewellyn, D.H. and High, S. ERp57 functions as a subunit of specific complexes formed with the ER lectins calreticulin and calnexin. Mol. Biol. Cell. 10 (1999) 2573–2582. Search in Google Scholar

[28] Jessop, C.E., Chakravarthi, S., Garbi, N., Hämmerling, G.J., Lovell, S. and Bulleid, N.J. ERp57 is essential for efficient folding of glycoproteins sharing common structural domains. EMBO J. 26 (2007) 28–40 http://dx.doi.org/10.1038/sj.emboj.760150510.1038/sj.emboj.7601505Search in Google Scholar

[29] Lindquist, J.A., Jensen, O.N., Mann, M. and Hämmerling, G.J. ER-60, a chaperone with thiol-dependent reductase activity involved in MHC class I assembly. EMBO J. 17 (1998) 2186–2195. http://dx.doi.org/10.1093/emboj/17.8.218610.1093/emboj/17.8.2186Search in Google Scholar

[30] Dick, T.P., Bangia, N., Peaper, D.R. and Cresswell, P. Disulfide bond isomerization and the assembly of MHC class I-peptide complexes. Immunity 16 (2002) 87–98. http://dx.doi.org/10.1016/S1074-7613(02)00263-710.1016/S1074-7613(02)00263-7Search in Google Scholar

[31] Dong, G., Wearsch, P.A., Peaper, D.R., Cresswell, P. and Reinisch, K.M. Insights into MHC class I peptide loading from the structure of the tapasin-ERp57 thioloxidoreductase heterodimer. Immunity 30 (2009) 21–32. http://dx.doi.org/10.1016/j.immuni.2008.10.01810.1016/j.immuni.2008.10.018Search in Google Scholar PubMed PubMed Central

[32] Zhang, Y., Kozlov, G., Pocanschi, C.L., Brockmeier, U., Ireland, B.S., Maattanen, P., Howe, C., Elliott, T., Gehring, K. and Williams, D.B. ERp57 does not require interactions with calnexin and calreticulin to promote assembly of class I histocompatibility molecules, and it enhances peptide loading independently of its redox activity. J. Biol. Chem. 284 (2009) 10160–10173. http://dx.doi.org/10.1074/jbc.M80835620010.1074/jbc.M808356200Search in Google Scholar PubMed PubMed Central

[33] Peaper, D.R. and Cresswell, P. The redox activity of ERp57 is not essential for its functions in MHC class I peptide loading. Proc. Natl. Acad. Sci. USA 105 (2008) 10477–10482. http://dx.doi.org/10.1073/pnas.080504410510.1073/pnas.0805044105Search in Google Scholar PubMed PubMed Central

[34] Garbi, N., Tanaka, S., Momburg, F., and Hammerling, G.J. Impaired assembly of the major histocompatibility complex class I peptide-loading complex in mice deficient in the oxidoreductase ERp57. Nat. Immunol. 7 (2006) 93–102. http://dx.doi.org/10.1038/ni128810.1038/ni1288Search in Google Scholar PubMed

[35] Li, Y. and Camacho, P. Ca2+-dependent redox modulation of SERCA 2b by ERpERp57. J. Cell Biol. 164 (2004) 35–46. http://dx.doi.org/10.1083/jcb.20030701010.1083/jcb.200307010Search in Google Scholar

[36] Schelhaas, M., Malmström, J., Pelkmans, L., Haugstetter, J., Ellgaard, L., Grünewald, K. and Helenius, A. Simian Virus 40 depends on ER protein folding and quality control factors for entry into host cells. Cell 131 (2007) 516–529. http://dx.doi.org/10.1016/j.cell.2007.09.03810.1016/j.cell.2007.09.038Search in Google Scholar

[37] Desjardins, M. ER-mediated phagocytosis: a new membrane for new functions. Nat. Rev. Immunol. 3 (2003) 280–291. http://dx.doi.org/10.1038/nri105310.1038/nri1053Search in Google Scholar

[38] Frickel, E.M., Riek, R., Jelesarov, I., Helenius, A., Wuthrich, K. and Ellgaard, L. TROSY-NMR reveals interaction between ERp57 and the tip of the calreticulin P-domain. Proc. Natl. Acad. Sci. USA 99 (2002) 1954–1959. http://dx.doi.org/10.1073/pnas.04269909910.1073/pnas.042699099Search in Google Scholar

[39] Hirano, N., Shibasaki, F., Sakai, R., Tanaka, T., Nishida, J., Yazaki, Y., Takenawa, T. and Hirai, H. Molecular cloning of the human glucoseregulated protein ERp57/GRP58, a thiol-dependent reductase. Identification of its secretory form and inducible expression by the oncogenic transformation. Eur. J. Biochem. 234 (1995) 336–342. http://dx.doi.org/10.1111/j.1432-1033.1995.336_c.x10.1111/j.1432-1033.1995.336_c.xSearch in Google Scholar

[40] Johnson, S., Michalak, M., Opas, M. and Eggleton, P. The ins and outs of calreticulin: from the ER lumen to the extracellular space. Trends Cell Biol. 11 (2001) 122–129. http://dx.doi.org/10.1016/S0962-8924(01)01926-210.1016/S0962-8924(01)01926-2Search in Google Scholar

[41] Afshar, N., Black, B.E. and Paschal, B.M. Retrotranslocation of the chaperone calreticulin from the endoplasmic reticulum lumen to the cytosol. Mol. Cell. Biol. 25 (2005) 8844–8853. http://dx.doi.org/10.1128/MCB.25.20.8844-8853.200510.1128/MCB.25.20.8844-8853.2005Search in Google Scholar PubMed PubMed Central

[42] Ellerman, D.A., Myles, D.G. And Primakoff P. A role for sperm surface protein disulfide isomerase activity in gamete fusion: evidence for the participation of ERp57. Dev. Cell. 10 (2006) 831–837. http://dx.doi.org/10.1016/j.devcel.2006.03.01110.1016/j.devcel.2006.03.011Search in Google Scholar PubMed

[43] Nemere, I., Farach-Carson, M.C., Rohe, B., Sterling, T.M., Norman, A.W., Boyan, B.D. and Safford, S.E. Ribozyme knockdown functionally links a 1,25(OH)2D3 membrane binding protein (1,25D3-MARRS) and phosphate uptake in intestinal cells. Proc. Natl. Acad. Sci. USA 101 (2004) 7392–7397. http://dx.doi.org/10.1073/pnas.040220710110.1073/pnas.0402207101Search in Google Scholar PubMed PubMed Central

[44] Boyan, B.D., Wong, K.L., Fang, M. and Schwartz, Z. 1alpha,25(OH)2D3 is an autocrine regulator of extracellular matrix turnover and growth factor release via ERp60 activated matrix vesicle metalloproteinases. J. Steroid. Biochem. Mol. Biol. 103 (2007) 467–472. http://dx.doi.org/10.1016/j.jsbmb.2006.11.00310.1016/j.jsbmb.2006.11.003Search in Google Scholar PubMed PubMed Central

[45] Chen, J., Olivares-Navarrete, R., Wang, Y., Herman, T.R., Boyan, B.D. and Schwartz, Z. Protein-disulfide isomerase-associated 3 (Pdia3) mediates the membrane response to 1,25-dihydroxyvitamin D3 in osteoblasts. J. Biol. Chem. 285 (2010) 37041–37050. http://dx.doi.org/10.1074/jbc.M110.15711510.1074/jbc.M110.157115Search in Google Scholar PubMed PubMed Central

[46] Tunsophon, S. and Nemere, I. Protein kinase C isotypes in signal transduction for the 1,25D3-MARRS receptor (ERp57/PDIA3) in steroid hormone-stimulated phosphate uptake. Steroids 75 (2010) 307–313. http://dx.doi.org/10.1016/j.steroids.2010.01.00410.1016/j.steroids.2010.01.004Search in Google Scholar PubMed

[47] Nemere, I., Garbi, N., Hämmerling, G.J. and Khanal, R.C. Intestinal cell calcium uptake and the targeted knockout of the 1,25D3-MARRS (membrane-associated, rapid response steroid-binding) receptor/PDIA3/Erp57. J. Biol. Chem. 285 (2010) 31859–31866. http://dx.doi.org/10.1074/jbc.M110.11695410.1074/jbc.M110.116954Search in Google Scholar PubMed PubMed Central

[48] Richard, C.L., Farach-Carson, M.C., Rohe, B., Nemere, I. and Meckling, K.A. Involvement of 1,25D3-MARRS (membrane associated, rapid response steroid-binding), a novel vitamin D receptor, in growth inhibition of breast cancer cells. Exp. Cell Res. 316 (2010) 695–703. http://dx.doi.org/10.1016/j.yexcr.2009.12.01510.1016/j.yexcr.2009.12.015Search in Google Scholar PubMed

[49] Wu, W., Beilhartz, G., Roy, Y., Richard, C.L., Curtin, M., Brown, L., Cadieux, D., Coppolino, M., Farach-Carson, M.C., Nemere, I. and Meckling, K.A. Nuclear translocation of the 1,25D3-MARRS (membrane associated rapid response to steroids) receptor protein and NFkappaB in differentiating NB4 leukemia cells. Exp. Cell Res. 316 (2010) 1101–1108. http://dx.doi.org/10.1016/j.yexcr.2010.01.01010.1016/j.yexcr.2010.01.010Search in Google Scholar PubMed

[50] Mah, S.J., Ades, A.M., Mir, R., Siemens, I.R., Williamson, J.R. and Fluharty, S.J. Association of solubilized angiotensin II receptors with phospholipase C-alpha in murine neuroblastoma NIE-115 cells. Mol. Pharmacol. 42 (1992) 217–226. Search in Google Scholar

[51] Aiyar, N., Bennett, C.F., Nambi, P., Valinski, W., Angioli, M., Minnich, M. and Crooke, S.T. Solubilization of rat liver vasopressin receptors as a complex with a guanine-nucleotide-binding protein and phosphoinositidespecific phospholipase C. Biochem. J. 261 (1989) 63–70. Search in Google Scholar

[52] Altieri, F., Maras, B., Eufemi, M., Ferraro, A. and Turano, C. Purification of a 57kDa nuclear matrix protein associated with thiol:protein-disulfide oxidoreductase and phospholipase C activities. Biochem. Biophys. Res. Commun. 194 (1993) 992–1000. http://dx.doi.org/10.1006/bbrc.1993.191910.1006/bbrc.1993.1919Search in Google Scholar PubMed

[53] Srivastava, S.P., Fuchs, J.A. and Holtzman, J.L. The reported cDNA sequence for phospholipase C alpha encodes protein disulfide isomerase, isozyme Q-2 and not phospholipase-C. Biochem. Biophys. Res. Commun. 193 (1993) 971–978. http://dx.doi.org/10.1006/bbrc.1993.172010.1006/bbrc.1993.1720Search in Google Scholar PubMed

[54] Tokutomi, Y., Araki, N., Kataoka, K., Yamamoto, E. and Kim-Mitsuyama, S. Oxidation of Prx2 and phosphorylation of GRP58 by angiotensin II in human coronary smooth muscle cells identified by 2D-DIGE analysis. Biochem. Biophys. Res. Commun. 364 (2007) 822–830. http://dx.doi.org/10.1016/j.bbrc.2007.10.09510.1016/j.bbrc.2007.10.095Search in Google Scholar PubMed

[55] Zhu, L., Santos, N.C. and Kim, K.H. Disulfide isomerase glucose-regulated protein 58 is required for the nuclear localization and degradation of retinoic acid receptor alpha. Reproduction 139 (2010) 717–731. http://dx.doi.org/10.1530/REP-09-052710.1530/REP-09-0527Search in Google Scholar PubMed

[56] Ndubuisi, M.I., Guo, G.G., Fried, V.A., Etlinger, J.D. and Sehgal, P.B. Cellular physiology of STAT3: Where’s the cytoplasmic monomer? J. Biol. Chem. 274 (1999) 25499–25509. http://dx.doi.org/10.1074/jbc.274.36.2549910.1074/jbc.274.36.25499Search in Google Scholar PubMed

[57] Sehgal, P.B., Guo, G.G., Shah, M., Kumar, V. and Patel, K. Cytokine signaling: STATS in plasma membrane rafts. J. Biol. Chem. 277 (2002) 12067–12074. http://dx.doi.org/10.1074/jbc.M20001820010.1074/jbc.M200018200Search in Google Scholar PubMed

[58] Guo, G.G., Patel, K., Kumar, V., Shah, M., Fried, V.A., Etlinger, J.D. and Sehgal, P.B. Association of the chaperone glucose-regulated protein 58 (GRP58/ER-60/ERp57) with Stat3 in cytosol and plasma membrane complexes. J. Interferon Cytokine Res. 22 (2002) 555–563. http://dx.doi.org/10.1089/1079990025298203410.1089/10799900252982034Search in Google Scholar PubMed

[59] Eufemi, M., Coppari, S., Altieri, F., Grillo, C., Ferraro, A. and Turano, C. ERp57 is present in STAT3-DNA complexes. Biochem. Biophys. Res. Commun. 323 (2004) 1306–1312. http://dx.doi.org/10.1016/j.bbrc.2004.09.00910.1016/j.bbrc.2004.09.009Search in Google Scholar PubMed

[60] Chichiarelli, S., Gaucci, E., Ferraro, A., Grillo, C., Altieri, F., Cocchiola, R., Arcangeli, V., Turano, C. and Eufemi, M. Role of ERp57 in the signaling and transcriptional activity of STAT3 in a melanoma cell line. Arch. Biochem. Biophys. 494 (2010) 178–183. http://dx.doi.org/10.1016/j.abb.2009.12.00410.1016/j.abb.2009.12.004Search in Google Scholar PubMed

[61] Wyse, B., Ali, N. and Ellison, D.H. Interaction with grp58 increases activity of the thiazide-sensitive Na-Cl cotransporter. Am. J. Physiol. Renal Physiol. 282 (2002) F424–430. 10.1152/ajprenal.0028.2001Search in Google Scholar PubMed

[62] Panaretakis, T., Joza, N., Modjtahedi, N., Tesniere, A., Vitale, I., Durchschlag, M., Fimia, G.M., Kepp, O., Piacentini, M., Froehlich, K.U., van Endert, P., Zitvogel, L., Madeo, F. and Kroemer, G. The co-translocation of ERp57 and calreticulin determines the immunogenicity of cell death. Cell Death Differ. 15 (2008) 1499–1509. http://dx.doi.org/10.1038/cdd.2008.6710.1038/cdd.2008.67Search in Google Scholar PubMed

[63] Obeid, M. ERP57 membrane translocation dictates the immunogenicity of tumor cell death by controlling the membrane translocation of calreticulin. J. Immunol. 181 (2008) 2533–2543. Search in Google Scholar

[64] Ramírez-Rangel, I., Bracho-Valdés, I., Vázquez-MacÍas, A., Carretero-Ortega, J., Reyes-Cruz, G. and Vázquez-Prado, J. Regulation of mTORC1 complex assembly and signaling by GRp58/ERp57. Mol. Cell. Biol. 31 (2011) 1657–1671. http://dx.doi.org/10.1128/MCB.00824-1010.1128/MCB.00824-10Search in Google Scholar PubMed PubMed Central

[65] Sarbassov, D.D. and Sabatini, D.M. Redox regulation of the nutrientsensitive raptor-mTOR pathway and complex. J. Biol. Chem. 280 (2005) 39505–39509. http://dx.doi.org/10.1074/jbc.M50609620010.1074/jbc.M506096200Search in Google Scholar PubMed

[66] Scherz-Shouval, R., Shvets, E., Fass, E., Shorer, H., Gil, L. and Elazar, Z. Reactive oxygen species are essential for autophagy and specifically regulate the activity of Atg4. EMBO J. 26 (2007) 1749–1760. http://dx.doi.org/10.1038/sj.emboj.760162310.1038/sj.emboj.7601623Search in Google Scholar PubMed PubMed Central

[67] Ohtani, H., Wakui, H., Ishino, T., Komatsuda, A. and Miura, A.B. An isoform of protein disulfide isomerase is expressed in the developing acrosome of spermatids during rat spermiogenesis and is transported into the nucleus of mature spermatids and epididymal spermatozoa. Histochemistry 100 (1993) 423–429. http://dx.doi.org/10.1007/BF0026782210.1007/BF00267822Search in Google Scholar PubMed

[68] Coppari, S., Altieri, F., Ferraro, A., Chichiarelli, S., Eufemi, M. and Turano, C. Nuclear localization and DNA interaction of protein disulfide isomerase ERp57 in mammalian cells. J. Cell. Biochem. 85 (2002) 325–333. http://dx.doi.org/10.1002/jcb.1013710.1002/jcb.10137Search in Google Scholar

[69] Krynetski, E.Y., Krynetskaia, N.F., Bianchi, M.E. and Evans, W.E. A nuclear protein complex containing high mobility group proteins B1 and B2, heat shock cognate protein 70, ERp60, and glyceraldehyde-3-phosphate dehydrogenase is involved in the cytotoxic response to DNA modified by incorporation of anticancer nucleoside analogues. Cancer Res. 63 (2003) 100–106. Search in Google Scholar

[70] Krynetskaia, N.F., Phadke, M.S., Jadhav, S.H. and Krynetskiy, E.Y. Chromatin-associated proteins HMGB1/2 and PDIA3 trigger cellular response to chemotherapy-induced DNA damage. Mol. Cancer Ther. 8 (2009) 864–872. http://dx.doi.org/10.1158/1535-7163.MCT-08-069510.1158/1535-7163.MCT-08-0695Search in Google Scholar

[71] Cicchillitti, L., Di Michele, M., Urbani, A., Ferlini, C., Donat, M.B., Scambia, G. and Rotilio, D. Comparative proteomic analysis of paclitaxel sensitive A2780 epithelial ovarian cancer cell line and its resistant counterpart A2780TC1 by 2D-DIGE: the role of ERp57. J. Proteome Res. 8 (2009) 1902–1912. http://dx.doi.org/10.1021/pr800856b10.1021/pr800856bSearch in Google Scholar

[72] Cicchillitti, L., Della Corte, A., Di Michele, M., Donati, M.B., Rotilio, D. and Scambia, G. Characterisation of a multimeric protein complex associated with ERp57 within the nucleus in paclitaxel-sensitive and -resistant epithelial ovarian cancer cells: the involvement of specific conformational states of beta-actin. Int. J. Oncol. 37 (2010) 445–454. http://dx.doi.org/10.3892/ijo_0000069310.3892/ijo_00000693Search in Google Scholar

[73] Ferraro, A., Altieri, F., Coppari, S., Eufemi, M., Chichiarelli, S. and Turano, C. Binding of the protein disulfide isomerase isoform ERp60 to the nuclear matrix-associated regions of DNA. J. Cell. Biochem. 72 (1999) 528–539. http://dx.doi.org/10.1002/(SICI)1097-4644(19990315)72:4<528::AID-JCB8>3.0.CO;2-V10.1002/(SICI)1097-4644(19990315)72:4<528::AID-JCB8>3.0.CO;2-VSearch in Google Scholar

[74] Grillo, C., Coppari, S., Turano, C. and Altieri, F. The DNA-binding activity of protein disulfide isomerase ERp57 is associated with the a(′) domain. Biochem. Biophys. Res. Commun. 295 (2002) 67–73. http://dx.doi.org/10.1016/S0006-291X(02)00634-410.1016/S0006-291X(02)00634-4Search in Google Scholar

[75] Chichiarelli, S., Ferraro, A., Altieri, F., Eufemi, M., Coppari, S., Grillo, C., Arcangeli, V. and Turano, C. The stress protein ERp57/GRP58 binds specific DNA sequences in HeLa cells. J. Cell. Physiol. 210 (2007) 343–351. http://dx.doi.org/10.1002/jcp.2082410.1002/jcp.20824Search in Google Scholar

[76] Schultz-Norton, J.R., McDonald, W.H., Yates, J.R. and Nardulli, A.M. Protein disulfide isomerase serves as a molecular chaperone to maintain estrogen receptor alpha structure and function. Mol. Endocrinol. 20 (2006) 1982–1995. http://dx.doi.org/10.1210/me.2006-000610.1210/me.2006-0006Search in Google Scholar

[77] Coe, H., Jung, J., Groenendyk, J., Prins, D. and Michalak, M. ERp57 modulates STAT3 signaling from the lumen of the endoplasmic reticulum. J. Biol. Chem. 285 (2010) 6725–6738. http://dx.doi.org/10.1074/jbc.M109.05401510.1074/jbc.M109.054015Search in Google Scholar

[78] Sehgal, P.B. Plasma membrane rafts and chaperones in cytokine/STAT signaling. Acta Biochim. Pol. 50 (2003) 583–594. Search in Google Scholar

[79] Markus, M. and Benezra, R. Two isoforms of protein disulfide isomerase alter the dimerization status of E2A proteins by a redox mechanism. J. Biol. Chem. 274 (1999) 1040–1049. http://dx.doi.org/10.1074/jbc.274.2.104010.1074/jbc.274.2.1040Search in Google Scholar

[80] Ozaki, T., Yamashita, T. and Ishiguro, S. ERp57-associated mitochondrial micro-calpain truncates apoptosis-inducing factor. Biochim. Biophys. Acta 1783 (2008) 1955–1963. http://dx.doi.org/10.1016/j.bbamcr.2008.05.01110.1016/j.bbamcr.2008.05.011Search in Google Scholar PubMed

[81] Murray, J.I., Whitfield, M.L., Trinklein, N.D., Myers, R.M., Brown, P.O. and Botstein, D. Diverse and specific gene expression responses to stresses in cultured human cells. Mol. Biol. Cell 15 (2004) 2361–2374. http://dx.doi.org/10.1091/mbc.E03-11-079910.1091/mbc.e03-11-0799Search in Google Scholar PubMed PubMed Central

[82] Rohe, B., Safford, S.E., Nemere, I., Farach-Carson, M.C. Regulation of expression of 1,25D3-MARRS/ERp57/PDIA3 in rat IEC-6 cells by TGF beta and 1,25(OH)2D3. Steroids 72 (2007) 144–150. http://dx.doi.org/10.1016/j.steroids.2006.11.01310.1016/j.steroids.2006.11.013Search in Google Scholar PubMed

[83] Corazzari, M., Lovat, P.E., Armstrong, J.L., Fimia, G.M., Hill, D.S., Birch-Machin, M., Redfern, C.P. and Piacentini, M. Targeting homeostatic mechanisms of endoplasmic reticulum stress to increase susceptibility of cancer cells to fenretinide-induced apoptosis: the role of stress proteins ERdj5 and ERp57. Br. J. Cancer 96 (2007) 1062–1071. http://dx.doi.org/10.1038/sj.bjc.660367210.1038/sj.bjc.6603672Search in Google Scholar PubMed PubMed Central

[84] Lovat, P.E., Corazzari, M., Armstrong, J.L., Martin, S., Pagliarini, V., Hill, D., Brown, A.M., Piacentini, M., Birch-Machin, M.A. and Redfern, C.P. Increasing melanoma cell death using inhibitors of protein disulfide isomerases to abrogate survival responses to endoplasmic reticulum stress. Cancer Res. 68 (2008) 5363–5369. http://dx.doi.org/10.1158/0008-5472.CAN-08-003510.1158/0008-5472.CAN-08-0035Search in Google Scholar PubMed PubMed Central

[85] Hetz, C., Russelakis-Carneiro, M., Wälchli, S., Carboni, S., Vial-Knecht, E., Maundrell, K., Castilla, J. and Soto, C. The disulfide isomerase Grp58 is a protective factor against prion neurotoxicity. J. Neurosci. 25 (2005) 2793–2802. http://dx.doi.org/10.1523/JNEUROSCI.4090-04.200510.1523/JNEUROSCI.4090-04.2005Search in Google Scholar PubMed PubMed Central

[86] Erickson, R.R., Dunning, L.M., Olson, D.A., Cohen, S.J., Davis, A.T., Wood, W.G., Kratzke, R.A. and Holtzman, J.L. In cerebrospinal fluid ER chaperones ERp57 and calreticulin bind beta-amyloid. Biochem. Biophys. Res. Commun. 332 (2005) 50–57. http://dx.doi.org/10.1016/j.bbrc.2005.04.09010.1016/j.bbrc.2005.04.090Search in Google Scholar PubMed

[87] Xu, D., Perez, R.E., Rezaiekhaligh, M.H., Bourdi, M. and Truog, W.E. Knockdown of ERp57 increases BiP/GRP78 induction and protects against hyperoxia and tunicamycin-induced apoptosis. Am. J. Physiol. Lung Cell. Mol. Physiol. 297 (2009) L44–51. http://dx.doi.org/10.1152/ajplung.90626.200810.1152/ajplung.90626.2008Search in Google Scholar PubMed

[88] Dukes, A.A., Van Laar, V.S., Cascio, M. and Hastings, T.G. Changes in endoplasmic reticulum stress proteins and aldolase A in cells exposed to dopamine. J. Neurochem. 106 (2008) 333–346. http://dx.doi.org/10.1111/j.1471-4159.2008.05392.x10.1111/j.1471-4159.2008.05392.xSearch in Google Scholar PubMed PubMed Central

[89] Kim-Han, J.S. and O’Malley, K.L. Cell stress induced by the parkinsonian mimetic, 6-hydroxydopamine, is concurrent with oxidation of the chaperone, ERp57, and aggresome formation. Antioxid. Redox Signal. 9 (2007) 2255–2264. http://dx.doi.org/10.1089/ars.2007.179110.1089/ars.2007.1791Search in Google Scholar PubMed

[90] Akazawa, Y.O., Saito, Y., Nishio, K., Horie, M., Kinumi, T., Masuo, Y., Yoshida, Y., Ashida, H. and Niki, E. Proteomic characterization of the striatum and midbrain treated with 6-hydroxydopamine: alteration of 58-kDa glucose-regulated protein and C/EBP homologous protein. Free Radic. Res. 44 (2010) 410–421. http://dx.doi.org/10.3109/1071576090353634910.3109/10715760903536349Search in Google Scholar PubMed

Published Online: 2011-9-29
Published in Print: 2011-12-1

© 2011 University of Wrocław, Poland

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 3.0 License.

Downloaded on 19.3.2024 from https://www.degruyter.com/document/doi/10.2478/s11658-011-0022-z/html
Scroll to top button