Derlin-2 and Derlin-3 are regulated by the mammalian unfolded protein response and are required for ER-associated degradation

Yukako Oda, Tetsuya Okada, Hiderou Yoshida, Randal J Kaufman, Kazuhiro Nagata, Kazutoshi Mori, Yukako Oda, Tetsuya Okada, Hiderou Yoshida, Randal J Kaufman, Kazuhiro Nagata, Kazutoshi Mori

Abstract

Proteins that are unfolded or misfolded in the endoplasmic reticulum (ER) must be refolded or degraded to maintain the homeostasis of the ER. Components of both productive folding and ER-associated degradation (ERAD) mechanisms are known to be up-regulated by the unfolded protein response (UPR). We describe two novel components of mammalian ERAD, Derlin-2 and -3, which show weak homology to Der1p, a transmembrane protein involved in yeast ERAD. Both Derlin-2 and -3 are up-regulated by the UPR, and at least Derlin-2 is a target of the IRE1 branch of the response, which is known to up-regulate ER degradation enhancing alpha-mannosidase-like protein (EDEM) and EDEM2, receptor-like molecules for misfolded glycoprotein. Overexpression of Derlin-2 or -3 accelerated degradation of misfolded glycoprotein, whereas their knockdown blocked degradation. Derlin-2 and -3 are associated with EDEM and p97, a cytosolic ATPase responsible for extraction of ERAD substrates. These findings indicate that Derlin-2 and -3 provide the missing link between EDEM and p97 in the process of degrading misfolded glycoproteins.

Figures

Figure 1.
Figure 1.
Identification of Derlin-2 and -3. (A) The levels of CGI-101, BiP, and β-actin mRNA are determined by microarray analysis in HeLa cells treated with or without 2 μg/ml tunicamycin for 8 h. Fold induction was determined, with the means from six independent experiments presented with SDs (error bars). CGI-101 is identical to Derlin-2. (B) Hydropathy plots of human Derlin-2 (CGI-101), -3 (FLJ43842), and -1 are shown. Hydrophobicity and hydrophilicity (expressed by positive and negative numbers, respectively) of the amino acid sequences of three human Derlins were obtained according to the method of Kyte and Doolittle (1982). Black bars mark hydrophobic regions that span the membrane. (C) Amino acid sequence alignment of human Derlin-2, Derlin-3 tv1, Derlin-3 tv2, Derlin-1, and yeast Der1p is shown. Identical amino acids are indicated by white letters in black boxes. Two transcriptional variants for Derlin-3 are deposited in the data bank. Derlin-3 tv1 lacks the 30 COOH-terminal amino acids present in Derlin-3 tv2.
Figure 2.
Figure 2.
Structure and tissue distribution of Derlin-2 and -3 mRNA. (A) Schematic structures of transcripts deposited in the data bank for human (available from GenBank/EMBL/DDBJ under accession no. NM_024295) and mouse (NM_024207) Derlin-1, human (NM_016041) and mouse (NM_033562) Derlin-2, and human (NM_001002862) and mouse (NM_024440) Derlin-3 are shown. Black underlines indicate the respective region of cDNA probe used for Northern blot analysis. (B) A nylon membrane onto which ∼2 μg each of poly A+ RNA prepared from eight human tissues was blotted after separation through gel electrophoresis (Human MTN blot) was hybridized with a DIG-labeled cDNA probe specific to human Derlin-1, Derlin-2, Derlin-3, or β-actin. Migration positions of molecular weight markers are indicated on the left of each panel. (C) A second nylon membrane onto which ∼1 μg each of poly A+ RNA prepared from 12 human tissues was blotted after separation through gel electrophoresis (Human 12-lane MTN blot) was hybridized as in B.
Figure 3.
Figure 3.
Involvement of the IRE1–XBP1 pathway in the induction of Derlin-1 and -2 in response to ER stress. (A) IRE1α+/+, IRE1α−/−, XBP1+/+, and XBP1−/− MEFs were treated with 10 μg/ml tunicamycin (Tm) for the indicated periods. Total RNAs were isolated and analyzed by Northern blot hybridization using a DIG-labeled cDNA probe specific to mouse Derlin-2, Derlin-1, EDEM, BiP, or GAPDH. Closed and open arrowheads indicate the migration positions of 28S ribosomal RNA (4.7 kb) and 18S ribosomal RNA (1.9 kb), respectively. (B) 293T or HeLa cells were treated with 2 μg/ml tunicamycin (Tm) for the indicated periods. Total RNAs were analyzed as in A using a DIG-labeled cDNA probe specific to human Derlin-2, Derlin-3, EDEM, BiP, or GAPDH.
Figure 4.
Figure 4.
Characterization of Derlin-2 and -3. (A) HeLa cells were trans-fected with plasmid to express each Derlin tagged with the c-myc epitope at the respective NH2 terminus. Transfected cells were fixed and stained with anti-myc and anti-Sec61β antibodies. (B) HEK293 cells were trans-fected with plasmid to express each Derlin tagged with the c-myc epitope at the respective COOH terminus. Postnuclear supernatant of transfected cells was incubated with increasing amounts of trypsin (0, 4, 8, and 16 μg for lane 1, 2, 3, and 4, respectively) for 15 min at 4°C. Immunoblotting analysis of the samples was performed using anti-myc, anti–NH2 terminus of calnexin (CNX[N]), anti–COOH terminus of calnexin (CNX[C]), and anti-calreticulin (CRT) antibodies. (C) HEK293 cells were transfected with plasmid to express Derlin-2 (a) or Derlin-3 tv1 or tv2 (b) tagged with the c-myc epitope at either the NH2 (N) or COOH (C) terminus. 36 h later, transfected cells were pulse labeled with 35S-methionine and cysteine for 15 min and then chased for the indicated periods. Cells were lysed with buffer containing 1% NP-40 and subjected to immunoprecipitation analysis using anti-myc antibody.
Figure 5.
Figure 5.
Effects of overexpression of Derlin-2 and -3 on degradation of NHK and NHK(QQQ). (A and B) HEK293 cells were mock-transfected or transfected with plasmid to express Derlin-2 or Derlin-3 tv2 tagged with the c-myc epitope at the respective NH2 terminus together with plasmid to express NHK (A) or NHK(QQQ) (B). Transfected cells were pulse-chased and subjected to immunoprecipitation analysis using anti–α1-PI antibody as in Fig. 4 C. Migration positions of NHK, NHK(QQQ), Derlin-2, and Derlin-3 are indicated. The radioactivity of each NHK band was determined and expressed as relative to the summation of radioactivity of nine NHK bands obtained in each experiment. The relative radioactivity of each band was normalized with the value at chase period 0 h. The means from three independent experiments with SDs (error bars) are plotted against the chase period (bottom).
Figure 6.
Figure 6.
Effects of knockdown of Derlin-2 and -3 on degradation of NHK. (A) HEK293 cells were untransfected or transfected with the shRNA vector pSUPER alone or pSUPER carrying a sequence corresponding to a part of Derlin-2 or -3. Total RNAs were isolated 64 h after transfection and analyzed by Northern blot hybridization using a DIG-labeled cDNA probe specific to human Derlin-1, Derlin-2, Derlin-3, or GAPDH. (B) HEK293 cells were transfected with pSUPER alone or pSUPER carrying a sequence corresponding to a part of Derlin-2, Derlin-3, or both together with plasmid to express NHK. Pulse-chase and subsequent immunoprecipitation were performed as in Fig. 5 A 64 h after transfection. The results of three independent experiments are shown (left). Normalized radioactivity of each NHK band was determined and is presented as in Fig. 5 A (right). Error bars depict means ± SD. (C) HEK293 cells were transfected with (+) or without (−) plasmid to express Derlin-1 or -2 tagged with the c-myc epitope at the respective NH2 terminus with (+) or without (−) plasmid to express Derlin-3 tv2 tagged with the HA epitope at the NH2 terminus. 36 h later, transfected cells were labeled with 35S-methionine and cysteine for 2 h, lysed with buffer containing 1% NP-40, and subjected to immunoprecipitation analysis using anti-myc or anti-HA antibody as indicated. Migration positions of Derlin-1, Derlin-2, and Derlin-3 tv2 are marked. (D) HEK293 cells were transfected with pSUPER alone or pSUPER derivatives as in B together with plasmid to express the wild-type (wt) α1-PI. Pulse-chase and subsequent immunoprecipitation from cell lysate as well as from media were performed as in B. Migration positions of high-mannose and complex wild-type α1-PI are indicated.
Figure 7.
Figure 7.
Association of Derlin-2 and -3 with p97, EDEM, and NHK. (A) HEK293 cells were mock-transfected or transfected with plasmid to express each Derlin tagged with the c-myc epitope at the respective COOH terminus alone (a) or with plasmid to express FLAG-tagged p97 (b). HEK293 cells were also transfected with plasmid to express each Derlin tagged with the c-myc epitope at the respective NH2 terminus (c). 24 h later, transfected cells were labeled with 35S-methionine and cysteine for 1 h, lysed with buffer containing 1% NP-40, and subjected to immunoprecipitation analysis using anti-myc or anti-p97 antibody as indicated. Migration positions of endogenous and FLAG-tagged p97 are marked. The asterisk shows a nonspecific band. The short open arrowhead indicates p97 coimmunoprecipitated with Derlins, whereas long open arrowheads indicate Derlins coimmunoprecipitated with p97. (B, a) HEK293 cells were mock-transfected or transfected with plasmid to express each Derlin tagged with the c-myc epitope at the respective NH2 terminus together with plasmid to express HA-tagged EDEM. 24 h later, transfected cells were labeled with 35S-methionine and cysteine for 2 h, lysed, and subjected to immunoprecipitation analysis using anti-myc or anti-HA antibody as indicated. The migration position of EDEM is marked. The asterisk shows a nonspecific band. The short open arrowhead indicates EDEM coimmunoprecipitated with Derlins, whereas long open arrowheads indicate Derlins coimmunoprecipitated with EDEM. (b and c) HEK293 cells were transfected with (+) or without (−) plasmid to express HA-tagged EDEM with or without plasmid to express Derlin-1 or -2 tagged with the c-myc epitope at the respective NH2 terminus. Transfected cells pulse labeled and then lysed as in a were subjected to immunoprecipitation analysis using anti-HA or anti-p97 antibody as indicated. (c) The amount of Derlin-1 expression plasmid to transfect HEK293 cells was twice that of Derlin-2 expression plasmid. Migration positions of p97, EDEM, and Derlins coimmunoprecipitated with EDEM or p97 are indicated. (C) HEK293 cells were transfected with (+) plasmid to express NHK and with or without plasmid to express Derlin-1 or -2 tagged with the c-myc epitope at the respective NH2 terminus. Transfected cells were pulse labeled and then lysed as in panel a. Equal amounts of cell lysate were subjected to immunoprecipitation analysis using anti–α1-PI, anti-p97, or anti-myc antibody as indicated. Migration positions of p97, NHK, and Derlins are indicated.
Figure 8.
Figure 8.
Effects of knockdown of Derlin-1 on degradation of NHK. (A) HEK293 cells were untransfected or transfected with pSUPER alone or pSUPER carrying a sequence corresponding to a part of Derlin-1, -2, or -3. 64 h later, transfected cells were subjected to immunoblotting analysis using antibody against Derlin-1, Derlin-2, or GAPDH. (B) HEK293 cells were transfected with pSUPER alone or pSUPER carrying a sequence corresponding to a part of Derlin-1 or -3 together with plasmid to express NHK. 64 h later, transfected cells were pulse-chased, lysed, and subjected to immunoprecipitation as in Fig. 6 B (left). The results of three independent experiments are shown. Normalized radioactivity of each NHK band was also determined and is presented as in Fig. 5 A (right). Error bars depict mean ± SD.

References

    1. Braun, S., K. Matuschewski, M. Rape, S. Thoms, and S. Jentsch. 2002. Role of the ubiquitin-selective CDC48(UFD1/NPL4)chaperone (segregase) in ERAD of OLE1 and other substrates. EMBO J. 21:615–621.
    1. Brummelkamp, T.R., R. Bernards, and R. Agami. 2002. A system for stable expression of short interfering RNAs in mammalian cells. Science. 296:550–553.
    1. Dai, R.M., and C.C. Li. 2001. Valosin-containing protein is a multi-ubiquitin chain-targeting factor required in ubiquitin-proteasome degradation. Nat. Cell Biol. 3:740–744.
    1. Gething, M.J., and J. Sambrook. 1992. Protein folding in the cell. Nature. 355:33–45.
    1. Harding, H.P., M. Calfon, F. Urano, I. Novoa, and D. Ron. 2002. Transcriptional and translational control in the mammalian unfolded protein response. Annu. Rev. Cell Dev. Biol. 18:575–599.
    1. Hartmann, E., T. Sommer, S. Prehn, D. Gorlich, S. Jentsch, and T.A. Rapoport. 1994. Evolutionary conservation of components of the protein translocation complex. Nature. 367:654–657.
    1. Helenius, A., T. Marquardt, and I. Braakman. 1992. The endoplasmic reticulum as a protein folding compartment. Trends Cell Biol. 2:227–231.
    1. Hiller, M.M., A. Finger, M. Schweiger, and D.H. Wolf. 1996. ER degradation of a misfolded luminal protein by the cytosolic ubiquitin-proteasome pathway. Science. 273:1725–1728.
    1. Hosokawa, N., I. Wada, K. Hasegawa, T. Yorihuzi, L.O. Tremblay, A. Herscovics, and K. Nagata. 2001. A novel ER α-mannosidase-like protein accelerates ER-associated degradation. EMBO Rep. 2:415–422.
    1. Jakob, C.A., P. Burda, J. Roth, and M. Aebi. 1998. Degradation of misfolded endoplasmic reticulum glycoproteins in Saccharomyces cerevisiae is determined by a specific oligosaccharide structure. J. Cell Biol. 142:1223–1233.
    1. Kalies, K.U., S. Allan, T. Sergeyenko, H. Kroger, and K. Romisch. 2005. The protein translocation channel binds proteasomes to the endoplasmic reticulum membrane. EMBO J. 24:2284–2293.
    1. Kaneko, M., and Y. Nomura. 2003. ER signaling in unfolded protein response. Life Sci. 74:199–205.
    1. Kaneko, M., M. Ishiguro, Y. Niinuma, M. Uesugi, and Y. Nomura. 2002. Human HRD1 protects against ER stress-induced apoptosis through ER-associated degradation. FEBS Lett. 532:147–152.
    1. Kaufman, R.J. 1999. Stress signaling from the lumen of the endoplasmic reticulum: coordination of gene transcriptional and translational controls. Genes Dev. 13:1211–1233.
    1. Kaufman, R.J., D. Scheuner, M. Schroder, X. Shen, K. Lee, C.Y. Liu, and S.M. Arnold. 2002. The unfolded protein response in nutrient sensing and differentiation. Nat. Rev. Mol. Cell Biol. 3:411–421.
    1. Knop, M., A. Finger, T. Braun, K. Hellmuth, and D.H. Wolf. 1996. Der1, a novel protein specifically required for endoplasmic reticulum degradation in yeast. EMBO J. 15:753–763.
    1. Kobayashi, T., K. Tanaka, K. Inoue, and A. Kakizuka. 2002. Functional ATPase activity of p97/valosin-containing protein (VCP) is required for the quality control of endoplasmic reticulum in neuronally differentiated mammalian PC12 cells. J. Biol. Chem. 277:47358–47365.
    1. Kopito, R.R. 1997. ER quality control: the cytoplasmic connection. Cell. 88:427–430.
    1. Kyte, J., and R.F. Doolittle. 1982. A simple method for displaying the hydropathic character of a protein. J. Mol. Biol. 157:105–132.
    1. Lee, A.H., N.N. Iwakoshi, and L.H. Glimcher. 2003. XBP-1 regulates a subset of endoplasmic reticulum resident chaperone genes in the unfolded protein response. Mol. Cell. Biol. 23:7448–7459.
    1. Lee, K., W. Tirasophon, X. Shen, M. Michalak, R. Prywes, T. Okada, H. Yoshida, K. Mori, and R.J. Kaufman. 2002. 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.
    1. Lilley, B.N., and H.L. Ploegh. 2004. A membrane protein required for dislocation of misfolded proteins from the ER. Nature. 429:834–840.
    1. Lilley, B.N., and H.L. Ploegh. 2005. Multiprotein complexes that link dislocation, ubiquitination, and extraction of misfolded proteins from the endoplasmic reticulum membrane. Proc. Natl. Acad. Sci. USA. 102:14296–14301.
    1. Mast, S.W., K. Diekman, K. Karaveg, A. Davis, R.N. Sifers, and K.W. Moremen. 2004. Human EDEM2, a novel homolog of family 47 glycosidases, is involved in ER-associated degradation of glycoproteins. Glycobiology. 15:421–436.
    1. Meyer, H.H., Y. Wang, and G. Warren. 2002. Direct binding of ubiquitin conjugates by the mammalian p97 adaptor complexes, p47 and Ufd1-Npl4. EMBO J. 21:5645–5652.
    1. Molinari, M., V. Calanca, C. Galli, P. Lucca, and P. Paganetti. 2003. Role of EDEM in the release of misfolded glycoproteins from the calnexin cycle. Science. 299:1397–1400.
    1. Mori, K. 2000. Tripartite management of unfolded proteins in the endoplasmic reticulum. Cell. 101:451–454.
    1. Mori, K. 2003. Frame switch splicing and regulated intramembrane proteolysis: key words to understand the unfolded protein response. Traffic. 4:519–528.
    1. Oda, Y., N. Hosokawa, I. Wada, and K. Nagata. 2003. EDEM as an acceptor of terminally misfolded glycoproteins released from calnexin. Science. 299:1394–1397.
    1. Okada, T., H. Yoshida, R. Akazawa, M. Negishi, and K. Mori. 2002. Distinct roles of activating transcription factor 6 (ATF6) and double-stranded RNA-activated protein kinase-like endoplasmic reticulum kinase (PERK) in transcription during the mammalian unfolded protein response. Biochem. J. 366:585–594.
    1. Olivari, S., C. Galli, H. Alanen, L. Ruddock, and M. Molinari. 2005. A novel stress-induced EDEM variant regulating endoplasmic reticulum-associated glycoprotein degradation. J. Biol. Chem. 280:2424–2428.
    1. Patil, C., and P. Walter. 2001. Intracellular signaling from the endoplasmic reticulum to the nucleus: the unfolded protein response in yeast and mammals. Curr. Opin. Cell Biol. 13:349–356.
    1. Ron, D. 2002. Translational control in the endoplasmic reticulum stress response. J. Clin. Invest. 110:1383–1388.
    1. Sambrook, J., E.F. Fritsch, and T. Maniatis. 1989. Molecular Cloning: A Laboratory Manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York.
    1. Schroder, M., and R.J. Kaufman. 2005. The mammalian unfolded protein response. Annu. Rev. Biochem. 74:739–789.
    1. Sifers, R.N., S. Brashears-Macatee, V.J. Kidd, H. Muensch, and S.L. Woo. 1988. A frameshift mutation results in a truncated α1-antitrypsin that is retained within the rough endoplasmic reticulum. J. Biol. Chem. 263:7330–7335.
    1. Tsai, B., Y. Ye, and T.A. Rapoport. 2002. Retro-translocation of proteins from the endoplasmic reticulum into the cytosol. Nat. Rev. Mol. Cell Biol. 3:246–255.
    1. Wiertz, E.J., T.R. Jones, L. Sun, M. Bogyo, H.J. Geuze, and H.L. Ploegh. 1996. a. The human cytomegalovirus US11 gene product dislocates MHC class I heavy chains from the endoplasmic reticulum to the cytosol. Cell. 84:769–779.
    1. Wiertz, E.J., D. Tortorella, M. Bogyo, J. Yu, W. Mothes, T.R. Jones, T.A. Rapoport, and H.L. Ploegh. 1996. b. Sec61-mediated transfer of a membrane protein from the endoplasmic reticulum to the proteasome for destruction. Nature. 384:432–438.
    1. Wilhovsky, S., R. Gardner, and R. Hampton. 2000. HRD gene dependence of endoplasmic reticulum-associated degradation. Mol. Biol. Cell. 11:1697–1708.
    1. Ye, Y., H.H. Meyer, and T.A. Rapoport. 2001. The AAA ATPase Cdc48/p97 and its partners transport proteins from the ER into the cytosol. Nature. 414:652–656.
    1. Ye, Y., Y. Shibata, C. Yun, D. Ron, and T.A. Rapoport. 2004. A membrane protein complex mediates retro-translocation from the ER lumen into the cytosol. Nature. 429:841–847.
    1. Ye, Y., Y. Shibata, M. Kikkert, S. van Voorden, E. Wiertz, and T.A. Rapoport. 2005. Recruitment of the p97 ATPase and ubiquitin ligases to the site of retrotranslocation at the endoplasmic reticulum membrane. Proc. Natl. Acad. Sci. USA. 102:14132–14138.
    1. Yoshida, H., T. Matsui, A. Yamamoto, T. Okada, and K. Mori. 2001. XBP1 mRNA is induced by ATF6 and spliced by IRE1 in response to ER stress to produce a highly active transcription factor. Cell. 107:881–891.
    1. Yoshida, H., T. Matsui, N. Hosokawa, R.J. Kaufman, K. Nagata, and K. Mori. 2003. A time-dependent phase shift in the mammalian unfolded protein response. Dev. Cell. 4:265–271.

Source: PubMed

3
Subskrybuj