Skip to main content Accessibility help
×
Hostname: page-component-7479d7b7d-rvbq7 Total loading time: 0 Render date: 2024-07-11T17:22:29.415Z Has data issue: false hasContentIssue false

8 - The cellular response to protein unfolding stress

from III - Protein folding and secretion

Published online by Cambridge University Press:  05 October 2013

M. Schröder
Affiliation:
School of Biological and Biomedical Sciences Science Laboratories South Road University of Durham Durham DH1 3LE UK
G. D. Robson
Affiliation:
University of Manchester
Pieter van West
Affiliation:
University of Aberdeen
Geoffrey Gadd
Affiliation:
University of Dundee
Get access

Summary

Introduction

Yeast and fungi are important for the production of industrial and therapeutic proteins (Tables 8.1 and 8.2). Product titres can reach several g/l, e.g. for human serum albumin expressed in Pichia pastoris (~7 g/l) (Kobayashi et al., 2000) or cellobiohydrolase expressed in Trichoderma reesei (~20 g/l) (Durand, Clanet & Tiraby, 1988; Nakari-Setälä & Penttilä, 1995). However, in most cases product titres are ~1000-fold lower (Archer, Jeenes & Mackenzie, 1994; Penttilä, 1998; Cereghino & Cregg, 2000). Identification and engineering of the bottleneck in these, usually heterologous, protein production processes will increase their cost efficiency and competitiveness.

The first bottlenecks that were identified were gene copy number (Clare et al., 1991; McGrew et al., 1997; Vassileva et al., 2001) and transcription efficiency (Outchkourov, Stiekema & Jongsma, 2002). Both were overcome by increasing gene dosage (Clare et al., 1991; Parekh, Forrester & Wittrup, 1995; McGrew et al., 1997; Vassileva et al., 2001) and optimization of codon usage (Gouka, Punt & van den Hondel, 1997a; Kraševec, van den Hondel & Komel, 2000; Moralejo et al., 2000; Outchkourov et al., 2002; Cardoza et al., 2003). Reports emerged stating that with increased gene copy number protein production decreased (Parekh et al., 1995). Recombinant protein was shown to remain associated with the cell (Hohenblum, Borth & Mattanovich, 2003) and localized to the endoplasmic reticulum (ER) (Kauffman et al., 2002). Thus, protein folding and posttranslational modification in the ER are a major bottleneck for protein secretion.

Type
Chapter
Information
Exploitation of Fungi , pp. 117 - 139
Publisher: Cambridge University Press
Print publication year: 2007

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Al-Sheikh, H., Watson, A. J., Lacey, G. A., Punt, P. J., MacKenzie, D. A., Jeenes, D. J., Pakula, T., Penttilä, M., Alcocer, M. J. & Archer, D. B. (2004). Endoplasmic reticulum stress leads to the selective transcriptional downregulation of the glucoamylase gene in Aspergillus niger. Molecular Microbiology, 53, 1731–42.CrossRefGoogle ScholarPubMed
Archer, D. B., Jeenes, D. J. & Mackenzie, D. A. (1994). Strategies for improving heterologous protein production from filamentous fungi. Antonie Van Leeuwenhoek, 65, 245–50.CrossRefGoogle ScholarPubMed
Bao, W. G. & Fukuhara, H. (2001). Secretion of human proteins from yeast: stimulation by duplication of polyubiquitin and protein disulfide isomerase genes in Kluyveromyces lactis. Gene, 272, 103–10.CrossRefGoogle ScholarPubMed
Bertolotti, A., Zhang, Y., Hendershot, L. M., Harding, H. P. & Ron, D. (2000). Dynamic interaction of BiP and ER stress transducers in the unfolded-protein response. Nature Cell Biology, 2, 326–32.CrossRefGoogle ScholarPubMed
Blond-Elguindi, S., Cwirla, S. E., Dower, W. J., Lipshutz, R. J., Sprang, S. R., Sambrook, J. F. & Gething, M.-J. (1993). Affinity panning of a library of peptides displayed on bacteriophages reveals the binding specificity of BiP. Cell, 75, 717–28.CrossRefGoogle Scholar
Bowdish, K. S., Yuan, H. E. & Mitchell, A. P. (1995). Positive control of yeast meiotic genes by the negative regulator UME6. Molecular and Cellular Biology, 15, 2955–61.CrossRefGoogle ScholarPubMed
Brickner, J. H. & Walter, P. (2004). Gene recruitment of the activated INO1 locus to the nuclear membrane. Public Library of Science Biology, 2, 1–11.Google ScholarPubMed
Broekhuijsen, M. P., Mattern, I. E., Contreras, R., Kinghorn, J. R. & Hondel, C. A. M. J. J. (1993). Secretion of heterologous proteins by Aspergillus niger: production of active human interleukin-6 in a protease-deficient mutant by KEX2-like processing of a glucoamylase-hIL6 fusion protein. Journal of Biotechnology, 31, 135–45.CrossRefGoogle Scholar
Cardoza, R. E., Gutierrez, S., Ortega, N., Colina, A., Casqueiro, J. & Martin, J. F. (2003). Expression of a synthetic copy of the bovine chymosin gene in Aspergillus awamori from constitutive and pH-regulated promoters and secretion using two different pre-pro sequences. Biotechnology and Bioengineering, 83, 249–59.CrossRefGoogle ScholarPubMed
Carman, G. M. & Henry, S. A. (1989). Phospholipid biosynthesis in yeast. Annual Review of Biochemistry, 58, 635–69.CrossRefGoogle Scholar
Cereghino, J. L. & Cregg, J. M. (2000). Heterologous protein expression in the methylotrophic yeast Pichia pastoris. FEMS Microbiology Reviews, 24, 45–66.CrossRefGoogle ScholarPubMed
Chang, H. J., Jones, E. W. & Henry, S. A. (2002). Role of the unfolded protein response pathway in regulation of INO1 and in the sec14 bypass mechanism in Saccharomyces cerevisiae. Genetics, 162, 29–43.Google ScholarPubMed
Chapman, R. E. & Walter, P. (1997). Translational attenuation mediated by an mRNA intron. Current Biology, 7, 850–9.CrossRefGoogle ScholarPubMed
Chen, Y., Pioli, D. & Piper, P. W. (1994). Overexpression of the gene for polyubiquitin in yeast confers increased secretion of a human leucocyte protease inhibitor. Biotechnology (NY), 12, 819–23.Google ScholarPubMed
Clare, J. J., Romanos, M. A., Rayment, F. B., Rowedder, J. E., Smith, M. A., Payne, M. M., Sreekrishna, K. & Henwood, C. A. (1991). Production of mouse epidermal growth factor in yeast: high-level secretion using Pichia pastoris strains containing multiple gene copies. Gene, 105, 205–12.CrossRefGoogle ScholarPubMed
Clark, M. W. & Abelson, J. (1987). The subnuclear localization of tRNA ligase in yeast. Journal of Cell Biology, 105, 1515–26.CrossRefGoogle Scholar
Conesa, A., Jeenes, D., Archer, D. B., Hondel, C. A. M. J. J. & Punt, P. J. (2002). Calnexin overexpression increases manganese peroxidase production in Aspergillus niger. Applied and Environmental Microbiology, 68, 846–51.CrossRefGoogle ScholarPubMed
Cox, J. S., Chapman, R. E. & Walter, P. (1997). The unfolded protein response coordinates the production of endoplasmic reticulum protein and endoplasmic reticulum membrane. Molecular Biology of the Cell, 8, 1805–14.CrossRefGoogle ScholarPubMed
Dobson, C. M., Sali, A. & Karplus, M. (1998). Protein folding: a perspective from theory and experiment. Angewandte Chemie International Edition in English, 37, 868–93.3.0.CO;2-H>CrossRefGoogle ScholarPubMed
Dunn, A., Luz, J. M., Natalia, D., Gamble, J. A., Freedman, R. B. & Tuite, M. F. (1995). Protein disulphide isomerase (PDI) is required for the secretion of a native disulphide-bonded protein from Saccharomyces cerevisiae. Biochemical Society Transactions, 23, 78S.CrossRefGoogle ScholarPubMed
Durand, H., Clanet, M. & Tiraby, G. (1988). Genetic improvement of Trichoderma reesei for large-scale cellulase production. Enzyme and Microbial Technology, 10, 341–6.CrossRefGoogle Scholar
Ellgaard, L. & Helenius, A. (2003). Quality control in the endoplasmic reticulum. Nature Reviews Molecular and Cellular Biology, 4, 181–91.CrossRefGoogle ScholarPubMed
Flynn, G. C., Pohl, J., Flocco, M. T. & Rothman, J. E. (1991). Peptide-binding specificity of the molecular chaperone BiP. Nature, 353, 726–30.CrossRefGoogle ScholarPubMed
Friedlander, R., Jarosch, E., Urban, J., Volkwein, C. & Sommer, T. (2000). A regulatory link between ER-associated protein degradation and the unfolded-protein response. Nature Cell Biology, 2, 379–84.CrossRefGoogle ScholarPubMed
Fülöp, V. & Jones, D. T. (1999). b propellers: structural rigidity and functional diversity. Current Opinion in Structural Biology, 9, 715–21.CrossRefGoogle Scholar
Gimeno, C. J., Ljungdahl, P. O., Styles, C. A. & Fink, G. R. (1992). Unipolar cell divisions in the yeast S. cerevisiae lead to filamentous growth: regulation by starvation and RAS. Cell, 68, 1077–90.CrossRefGoogle ScholarPubMed
Goldmark, J. P., Fazzio, T. G., Estep, P. W., Church, G. M. & Tsukiyama, T. (2000). The Isw2 chromatin remodeling complex represses early meiotic genes upon recruitment by Ume6p. Cell, 103, 423–33.CrossRefGoogle ScholarPubMed
Gonzalez, T. N., Sidrauski, C., Dörfler, S. & Walter, P. (1999). Mechanism of non-spliceosomal mRNA splicing in the unfolded protein response pathway. EMBO Journal, 18, 3119–32.CrossRefGoogle ScholarPubMed
Gouka, R. J., Punt, P. J. & Hondel, C. A. M. J. J. (1997a). Glucoamylase gene fusions alleviate limitations for protein production in Aspergillus awamori at the transcriptional and (post) translational levels. Applied and Environmental Microbiology, 63, 488–97.Google Scholar
Gouka, R. J., Punt, P. J. & Hondel, C. A. M. J. J. (1997b). Efficient production of secreted proteins by Aspergillus: progress, limitations and prospects. Applied Microbiology and Biotechnology, 47, 1–11.CrossRefGoogle Scholar
Harmsen, M. M., Bruyne, M. I., Raue, H. A. & Maat, J. (1996). Overexpression of binding protein and disruption of the PMR1 gene synergistically stimulate secretion of bovine prochymosin but not plant thaumatin in yeast. Applied Microbiology and Biotechnology, 46, 365–70.CrossRefGoogle ScholarPubMed
Hayano, T., Hirose, M. & Kikuchi, M. (1995). Protein disulfide isomerase mutant lacking its isomerase activity accelerates protein folding in the cell. FEBS Letters, 377, 505–11.Google Scholar
Haynes, C. M., Titus, E. A. & Cooper, A. A. (2004). Degradation of misfolded proteins prevents ER-derived oxidative stress and cell death. Molecular Cell, 15, 767–76.CrossRefGoogle ScholarPubMed
Hessing, J. G. M., Rotterdam, C., Verbakel, J. M., Roza, M., Maat, J., Gorcom, R. F. M. & Hondel, C. A. M. J. J. (1994). Isolation and characterization of a 1,4-β-endoxylanase gene of A. awamori. Current Genetics, 26, 228–32.CrossRefGoogle ScholarPubMed
Hohenblum, H., Borth, N. & Mattanovich, D. (2003). Assessing viability and cell-associated product of recombinant protein producing Pichia pastoris with flow cytometry. Journal of Biotechnology, 102, 281–90.CrossRefGoogle ScholarPubMed
Hohenblum, H., Gasser, B., Maurer, M., Borth, N. & Mattanovich, D. (2004). Effects of gene dosage, promoters, and substrates on unfolded protein stress of recombinant Pichia pastoris. Biotechnology and Bioengineering, 85, 367–75.CrossRefGoogle ScholarPubMed
Kadosh, D. & Struhl, K. (1997). Repression by Ume6 involves recruitment of a complex containing Sin3 corepressor and Rpd3 histone deacetylase to target promoters. Cell, 89, 365–71.CrossRefGoogle ScholarPubMed
Kadosh, D. & Struhl, K. (1998). Histone deacetylase activity of Rpd3 is important for transcriptional repression in vivo. Genes and Development, 12, 797–805.CrossRefGoogle ScholarPubMed
Kauffman, K. J., Pridgen, E. M., Doyle, F. J., 3rd, Dhurjati, P. S. & Robinson, A. S. (2002). Decreased protein expression and intermittent recoveries in BiP levels result from cellular stress during heterologous protein expression in Saccharomyces cerevisiae. Biotechnology Progress, 18, 942–50.CrossRefGoogle ScholarPubMed
Kawahara, T., Yanagi, H., Yura, T. & Mori, K. (1997). Endoplasmic reticulum stress-induced mRNA splicing permits synthesis of transcription factor Hac1p/Ern4p that activates the unfolded protein response. Molecular Biology of the Cell, 8, 1845–62.CrossRefGoogle ScholarPubMed
Kobayashi, K., Kuwae, S., Ohya, T., Ohda, T., Ohyama, M. & Tomomitsu, K. (2000). High level secretion of recombinant human serum albumin by fed-batch fermentation of the methylotrophic yeast, Pichia pastoris, based on optimal methanol feeding strategy. Journal of Bioscience and Bioengineering, 90, 280–8.CrossRefGoogle ScholarPubMed
Krasevec, N., Hondel, C. A. M. J. J. & Komel, R. (2000). Expression of human lymphotoxin alpha in Aspergillus niger. Pflugers Arch, 440, R83–5.CrossRefGoogle ScholarPubMed
Kron, S. J., Styles, C. A. & Fink, G. R. (1994). Symmetric cell division in pseudohyphae of the yeast Saccharomyces cerevisiae. Molecular Biology of the Cell, 5, 1003–22.CrossRefGoogle ScholarPubMed
Kupiec, M., Byers, B., Esposito, R. E. & Mitchell, A. P. (1997). Meiosis and sporulation in Saccharomyces cerevisiae. In The Molecular and Cellular Biology of the Yeast Saccharomyces, Vol. 3, ed. Pringle, J. R., Broach, J. R. & Jones, E. W.. Plainview, NY: Cold Spring Harbor Laboratory Press, pp. 889–1036.Google Scholar
Liu, C. Y., Schröder, M. & Kaufman, R. J. (2000). Ligand-independent dimerization activates the stress response kinases IRE1 and PERK in the lumen of the endoplasmic reticulum. Journal of Biological Chemistry, 275, 24881–5.CrossRefGoogle ScholarPubMed
Loewen, C. J., Gaspar, M. L., Jesch, S. A., Delon, C., Ktistakis, N. T., Henry, S. A. & Levine, T. P. (2004). Phospholipid metabolism regulated by a transcription factor sensing phosphatidic acid. Science, 304, 1644–7.CrossRefGoogle ScholarPubMed
Madeo, F., Fröhlich, E. & Fröhlich, K.-U. (1997). A yeast mutant showing diagnostic markers of early and late apoptosis. Journal of Cell Biology, 139, 729–34.CrossRefGoogle ScholarPubMed
Madeo, F., Fröhlich, E., Ligr, M., Grey, M., Sigrist, S. J., Wolf, D. H. & Fröhlich, K.-U. (1999). Oxygen stress: a regulator of apoptosis in yeast. Journal of Cell Biology, 145, 757–67.CrossRefGoogle Scholar
McGrew, J. T., Leiske, D., Dell, B., Klinke, R., Krasts, D., Wee, S. F., Abbott, N., Armitage, R. & Harrington, K. (1997). Expression of trimeric CD40 ligand in Pichia pastoris: use of a rapid method to detect high-level expressing transformants. Gene, 187, 193–200.CrossRefGoogle ScholarPubMed
Mitchell, A. P. (1994). Control of meiotic gene expression in Saccharomyces cerevisiae. Microbiological Reviews, 58, 56–70.Google ScholarPubMed
Molinari, M., Calanca, V., Galli, C., Lucca, P. & Paganetti, P. (2003). Role of EDEM in the release of misfolded glycoproteins from the calnexin cycle. Science, 299, 1397–400.CrossRefGoogle ScholarPubMed
Moralejo, F. J., Cardoza, R. E., Gutierrez, S., Sisniega, H., Faus, I. & Martin, J. F. (2000). Overexpression and lack of degradation of thaumatin in an aspergillopepsin A-defective mutant of Aspergillus awamori containing an insertion in the pepA gene. Applied Microbiology and Biotechnology, 54, 772–7.CrossRefGoogle Scholar
Moralejo, F. J., Watson, A. J., Jeenes, D. J., Archer, D. B. & Martín, J. F. (2001). A defined level of protein disulfide isomerase expression is required for optimal secretion of thaumatin by Aspegillus awamori. Molecular Genetics and Genomics, 266, 246–53.Google ScholarPubMed
Mori, K., Ogawa, N., Kawahara, T., Yanagi, H. & Yura, T. (2000). mRNA splicing-mediated C-terminal replacement of transcription factor Hac1p is required for efficient activation of the unfolded protein response. Proceedings of the National Academy of Sciences of the United States of America, 97, 4660–5.CrossRefGoogle ScholarPubMed
Mulder, H. J., Saloheimo, M., Penttilä, M. & Madrid, S. M. (2004). The transcription factor HACA mediates the unfolded protein response in Aspergillus niger, and up-regulates its own transcription. Molecular Genetics and Genomics, 271, 130–40.CrossRefGoogle ScholarPubMed
Nakai, K. & Horton, P. (1999). PSORT: a program for detecting sorting signals in proteins and predicting their subcellular localization. Trends in Biochemical Sciences, 24, 34–6.CrossRefGoogle ScholarPubMed
Nakari-Setälä, T. & Penttilä, M. (1995). Production of Trichoderma reesei cellulases on glucose-containing media. Applied and Environmental Microbiology, 61, 3650–5.Google ScholarPubMed
Ngiam, C., Jeenes, D. J., Punt, P. J., Hondel, C. A. M. J. J. & Archer, D. B. (2000). Characterization of a foldase, protein disulfide isomerase A, in the protein secretory pathway of Aspergillus niger. Applied and Environmental Microbiology, 66, 775–82.CrossRefGoogle ScholarPubMed
Nikawa, J., Akiyoshi, M., Hirata, S. & Fukuda, T. (1996). Saccharomyces cerevisiae IRE2/HAC1 is involved in IRE1-mediated KAR2 expression. Nucleic Acids Research, 24, 4222–6.CrossRefGoogle ScholarPubMed
Nikawa, J. & Yamashita, S. (1992). IRE1 encodes a putative protein kinase containing a membrane-spanning domain and is required for inositol phototrophy in Saccharomyces cerevisiae. Molecular Microbiology, 6, 1441–6.CrossRefGoogle ScholarPubMed
Niwa, M., Patil, C. K., DeRisi, J. & Walter, P. (2005). Genome-scale approaches for discovering novel nonconventional splicing substrates of the Ire1 nuclease. Genome Biology, 6, R3.CrossRefGoogle ScholarPubMed
Oda, Y., Hosokawa, N., Wada, I. & Nagata, K. (2003). EDEM as an acceptor of terminally misfolded glycoproteins released from calnexin. Science, 299, 1394–7.CrossRefGoogle ScholarPubMed
Ogawa, N. & Mori, K. (2004). Autoregulation of the HAC1 gene is required for sustained activation of the yeast unfolded protein response. Genes to Cells, 9, 95–104.CrossRefGoogle ScholarPubMed
Okushima, Y., Koizumi, N., Yamaguchi, Y., Kimata, Y., Kohno, K. & Sano, H. (2002). Isolation and characterization of a putative transducer of endoplasmic reticulum stress in Oryza sativa. Plant Cellular Physiology, 43, 532–9.CrossRefGoogle ScholarPubMed
Outchkourov, N. S., Stiekema, W. J. & Jongsma, M. A. (2002). Optimization of the expression of equistatin in Pichia pastoris. Protein Expression and Purification, 24, 18–24.CrossRefGoogle ScholarPubMed
Pakula, T., Saloheimo, M., Uusitalo, J., Huuskonen, A., Watson, A., Jeenes, D., Archer, D. & Pentilla, M. (2004). Method for production of secreted proteins in fungi. Sughrue Mion, PLLC: Canada. CA 2,438,356 (A1), 1–58.
Pakula, T. M., Laxell, M., Huuskonen, A., Uusitalo, J., Saloheimo, M. & Penttilä, M. (2003). The effects of drugs inhibiting protein secretion in the filamentous fungus Trichoderma reesei. Evidence for down-regulation of genes that encode secreted proteins in the stressed cells. Journal of Biological Chemistry, 278, 45011–20.CrossRefGoogle ScholarPubMed
Palmiter, R. D. (1975). Quantitation of parameters that determine the rate of ovalbumin synthesis. Cell, 4, 189–97.CrossRefGoogle ScholarPubMed
Paoli, M. (2001). Protein folds propelled by diversity. Progress in Biophysics and Molecular Biology, 76, 103–30.CrossRefGoogle ScholarPubMed
Papa, F. R., Zhang, C., Shokat, K. & Walter, P. (2003). Bypassing a kinase activity with an ATP-competitive drug. Science, 302, 1533–7.CrossRefGoogle ScholarPubMed
Parekh, R., Forrester, K. & Wittrup, D. (1995). Multicopy overexpression of bovine pancreatic trypsin inhibitor saturates the protein folding and secretory capacity of Saccharomyces cerevisiae. Protein Expression and Purification, 6, 537–45.CrossRefGoogle ScholarPubMed
Patil, C. K., Li, H. & Walter, P. (2004). Gcn4p and novel upstream activating sequences regulate targets of the unfolded protein response. Public Library of Science Biology, 2, 1208–23.Google ScholarPubMed
Penttilä, M. (1998). Heterologous protein production in Trichoderma. In Trichoderma And Gliocladium, Vol. 2, ed. , H. G. E. & Kubicek, C. P.. London: Taylor & Francis, pp. 365–82.Google Scholar
Ponting, C. P. (2000). Proteins of the endoplasmic-reticulum-associated degradation pathway: domain detection and function prediction. Biochemical Journal, 351, 527–35.CrossRefGoogle ScholarPubMed
Punt, P. J., Gemeren, I. A., Drint-Kuijvenhoven, J., Hessing, J. G. M., Muijlwijk-Harteveld, G. M., Beijersbergen, A., Verrips, C. T. & Hondel, C. A. M. J. J. (1998). Analysis of the role of the gene bipA, encoding the major endoplasmic reticulum chaperone protein in the secretion of homologous and heterologous proteins in black Aspergilli. Applied Microbiology and Biotechnology, 50, 447–54.CrossRefGoogle ScholarPubMed
Punt, P. J., Veldhuisen, G. & Hondel, C. A. M. J. J. (1994). Protein targeting and secretion in filamentous fungi. A progress report. Antonie Van Leeuwenhoek, 65, 211–16.CrossRefGoogle ScholarPubMed
Robinson, A. S., Bockhaus, J. A., Voegler, A. C. & Wittrup, K. D. (1996). Reduction of BiP levels decreases heterologous protein secretion in Saccharomyces cerevisiae. Journal of Biological Chemistry, 271, 10017–22.CrossRefGoogle ScholarPubMed
Robinson, A. S., Hines, V. & Wittrup, K. D. (1994). Protein disulfide isomerase overexpression increases secretion of foreign proteins in Saccharomyces cerevisiae. Biotechnology (NY), 12, 381–4.CrossRefGoogle ScholarPubMed
Rubin-Bejerano, I., Mandel, S., Robzyk, K. & Kassir, Y. (1996). Induction of meiosis in Saccharomyces cerevisiae depends on conversion of the transcriptional represssor Ume6 to a positive regulator by its regulated association with the transcriptional activator Ime1. Molecular and Cellular Biology, 16, 2518–26.CrossRefGoogle ScholarPubMed
Rüegsegger, U., Leber, J. H. & Walter, P. (2001). Block of HAC1 mRNA translation by long-range base pairing is released by cytoplasmic splicing upon induction of the unfolded protein response. Cell, 107, 103–14.CrossRefGoogle ScholarPubMed
Saloheimo, M., Lund, M. & Penttilä, M. E. (1999). The protein disulphide isomerase gene of the fungus Trichoderma reesei is induced by endoplasmic reticulum stress and regulated by the carbon source. Molecular and General Genetics, 262, 35–45.CrossRefGoogle ScholarPubMed
Saloheimo, M., Valkonen, M. & Penttilä, M. (2003). Activation mechanisms of the HAC1-mediated unfolded protein response in filamentous fungi. Molecular Microbiology, 47, 1149–61.CrossRefGoogle ScholarPubMed
Schröder, M., Chang, J. S. & Kaufman, R. J. (2000). The unfolded protein response represses nitrogen-starvation induced developmental differentiation in yeast. Genes and Development, 14, 2962–75.CrossRefGoogle Scholar
Schröder, M., Clark, R. & Kaufman, R. J. (2003). IRE1- and HAC1-independent transcriptional regulation in the unfolded protein response of yeast. Molecular Microbiology, 49, 591–606.CrossRefGoogle Scholar
Schröder, M., Clark, R., Liu, C. Y. & Kaufman, R. J. (2004). The unfolded protein response represses differentiation through the RPD3-SIN3 histone deacetylase. EMBO Journal, 23, 2281–92.CrossRefGoogle ScholarPubMed
Schröder, M. & Kaufman, R. J. (2005a). ER stress and the unfolded protein response. Mutation Research, 569, 29–63.CrossRefGoogle Scholar
Schröder, M. & Kaufman, R. J. (2005b). The mammalian unfolded protein response. Annual Review of Biochemistry, 74, 739–89.CrossRefGoogle Scholar
Schultz, L. D., Markus, H. Z., Hofmann, K. J., Montgomery, D. L., Dunwiddie, C. T., Kniskern, P. J., Freedman, R. B., Ellis, R. W. & Tuite, M. F. (1994). Using molecular genetics to improve the production of recombinant proteins by the yeast Saccharomyces cerevisiae. Annals of the New York Academy of Sciences, 721, 148–57.CrossRefGoogle ScholarPubMed
Shahinian, S., Dijkgraaf, G. J., Sdicu, A. M., Thomas, D. Y., Jakob, C. A., Aebi, M. & Bussey, H. (1998). Involvement of protein N-glycosyl chain glucosylation and processing in the biosynthesis of cell wall β-1,6-glucan of Saccharomyces cerevisiae. Genetics, 149, 843–56.Google ScholarPubMed
Shusta, E. V., Raines, R. T., Pluckthun, A. & Wittrup, K. D. (1998). Increasing the secretory capacity of Saccharomyces cerevisiae for production of single-chain antibody fragments. Nature Biotechnology, 16, 773–7.CrossRefGoogle ScholarPubMed
Sidrauski, C., Cox, J. S. & Walter, P. (1996). tRNA ligase is required for regulated mRNA splicing in the unfolded protein response. Cell, 87, 405–13.CrossRefGoogle ScholarPubMed
Spode, I., Maiwald, D., Hollenberg, C. P. & Suckow, M. (2002). ATF/CREB sites present in sub-telomeric regions of Saccharomyces cerevisiae chromosomes are part of promoters and act as UAS/URS of highly conserved COS genes. Journal of Molecular Biology, 319, 407–20.CrossRefGoogle ScholarPubMed
Steel, G. J., Fullerton, D. M., Tyson, J. R. & Stirling, C. J. (2004). Coordinated activation of Hsp70 chaperones. Science, 303, 98–101.CrossRefGoogle Scholar
Stevens, F. J. & Argon, Y. (1999). Protein folding in the ER. Seminars in Cell and Developmental Biology, 10, 443–54.CrossRefGoogle ScholarPubMed
Strich, R., Surosky, R. T., Steber, C., Dubois, E., Messenguy, F. & Esposito, R. E. (1994). UME6 is a key regulator of nitrogen repression and meiotic development. Genes and Development, 8, 796–810.CrossRefGoogle ScholarPubMed
Stroobants, A. K., Hettema, E. H., Berg, M. & Tabak, H. F. (1999). Enlargement of the endoplasmic reticulum membrane in Saccharomyces cerevisiae is not necessarily linked to the unfolded protein response via Ire1p. FEBS Letters, 453, 210–14.CrossRefGoogle Scholar
Tachibana, C. & Stevens, T. H. (1992). The yeast EUG1 gene encodes an endoplasmic reticulum protein that is functionally related to protein disulfide isomerase. Molecular and Cellular Biology, 12, 4601–11.CrossRefGoogle ScholarPubMed
Tirasophon, W., Lee, K., Callaghan, B., Welihinda, A. & Kaufman, R. J. (2000). The endoribonuclease activity of mammalian IRE1 autoregulates its mRNA and is required for the unfolded protein response. Genes and Development, 14, 2725–36.CrossRefGoogle ScholarPubMed
Travers, K. J., Patil, C. K., Wodicka, L., Lockhart, D. J., Weissman, J. S. & Walter, P. (2000). Functional and genomic analyses reveal an essential coordination between the unfolded protein response and ER-associated degradation. Cell, 101, 249–58.CrossRefGoogle ScholarPubMed
Valkonen, M., Penttilä, M. & Saloheimo, M. (2003a). Effects of inactivation and constitutive expression of the unfolded-protein response pathway on protein production in the yeast Saccharomyces cerevisiae. Applied and Environmental Microbiology, 69, 2065–72.CrossRefGoogle Scholar
Valkonen, M., Penttilä, M. & Saloheimo, M. (2004). The ire1 and ptc2 genes involved in the unfolded protein response pathway in the filamentous fungus Trichoderma reesei. Molecular Genetics and Genomics, 272, 443–51.CrossRefGoogle ScholarPubMed
Valkonen, M., Ward, M., Wang, H., Penttilä, M. & Saloheimo, M. (2003b). Improvement of foreign-protein production in Aspergillus niger var. awamori by constitutive induction of the unfolded-protein response. Applied and Environmental Microbiology, 69, 6979–86.CrossRefGoogle Scholar
Gemeren, I. A., Beijersbergen, A., Hondel, C. A. M. J. J. & Verrips, C. T. (1998). Expression and secretion of defined cutinase variants by Aspergillus awamori. Applied and Environmental Microbiology, 64, 2794–9.Google ScholarPubMed
Vassileva, A., Chugh, D. A., Swaminathan, S. & Khanna, N. (2001). Effect of copy number on the expression levels of hepatitis B surface antigen in the methylotrophic yeast Pichia pastoris. Protein Expression and Purification, 21, 71–80.CrossRefGoogle ScholarPubMed
Verdoes, J. C., Punt, P. J., Schrickx, J. M., Verseveld, H. W., Stouthamer, A. H. & Hondel, C. A. M. J. J. (1993). Glucoamylase overexpression in Aspergillus niger: molecular genetic analysis of strains containing multiple copies of the glaA gene. Transgenic Research, 2, 84–92.CrossRefGoogle ScholarPubMed
Wang, H. & Ward, M. (2000). Molecular characterization of a PDI-related gene prpA in Aspergillus niger var. awamori. Current Genetics, 37, 57–64.CrossRefGoogle ScholarPubMed
Washburn, B. K. & Esposito, R. E. (2001). Identification of the Sin3-binding site in Ume6 defines a two-step process for conversion of Ume6 from a transcriptional repressor to an activator in yeast. Molecular and Cellular Biology, 21, 2057–69.CrossRefGoogle Scholar
Welihinda, A. A., Tirasophon, W., Green, S. R. & Kaufman, R. J. (1997). Gene induction in response to unfolded protein in the endoplasmic reticulum is mediated through Ire1p kinase interaction with a transcriptional coactivator complex containing Ada5p. Proceedings of the National Academy of Sciences of the United States of America, 94, 4289–94.CrossRefGoogle ScholarPubMed
Welihinda, A. A., Tirasophon, W., Green, S. R. & Kaufman, R. J. (1998). Protein serine/threonine phosphatase Ptc2p negatively regulates the unfolded-protein response by dephosphorylating Ire1p kinase. Molecular and Cellular Biology, 18, 1967–77.CrossRefGoogle ScholarPubMed
Welihinda, A. A., Tirasophon, W. & Kaufman, R. J. (2000). The transcriptional co-activator ADA5 is required for HAC1 mRNA processing in vivo. Journal of Biological Chemistry, 275, 3377–81.CrossRefGoogle ScholarPubMed
Zillmann, M., Gorovsky, M. A. & Phizicky, E. M. (1991). Conserved mechanism of tRNA splicing in eukaryotes. Molecular and Cellular Biology, 11, 5410–16.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×