Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Genetics. 2004 Dec; 168(4): 1899–1913.
PMCID: PMC1448725
PMID: 15371354

Role of the Unfolded Protein Response Pathway in Secretory Stress and Regulation of INO1 Expression in Saccharomyces cerevisiae

Abstract

The unfolded protein response pathway (UPR) enables the cell to cope with the buildup of unfolded proteins in the endoplasmic reticulum (ER). UPR loss-of-function mutants, hac1Δ and ire1Δ, are also inositol auxotrophs, a phenotype associated with defects in expression of INO1, the most highly regulated of a set of genes encoding enzymes of phospholipid metabolism. We now demonstrate that the UPR plays a functional role in membrane trafficking under conditions of secretory stress in yeast. Mutations conferring a wide range of membrane trafficking defects exhibited negative genetic interaction when combined with ire1Δ and hac1Δ. At semipermissive temperatures, carboxypeptidase Y transit time to the vacuole was slower in Sec cells containing an ire1Δ or hac1Δ mutation than in Sec cells with an intact UPR. The UPR was induced in Sec cells defective in subcellular membrane trafficking events ranging from ER vesicle trafficking to distal secretion and in erg6Δ cells challenged with brefeldin A. However, the high levels of UPR induction observed under these conditions were not correlated with elevated INO1 expression. Indeed, many of the Sec mutants that had elevated UPR expression at semipermissive growth temperatures failed to achieve wild-type levels of INO1 expression under these same conditions.

THE unfolded protein response pathway (UPR) is a stress response pathway that is activated when unfolded proteins accumulate in the endoplasmic reticulum (ER; Cox et al. 1993; Cox and Walter 1996; Mori et al. 1992, 1993). In yeast, the UPR consists of three components: Ire1p, Hac1p, and Rlg1p. Ire1p is a unique ER transmembrane spanning protein kinase/endoribonuclease. Hac1p is a transcription factor that is required for expression of UPR-responsive genes, including protein-folding chaperones, such as Kar2p (BiP; Kohno et al. 1993; Mori et al. 1992, 1993; Nikawa and Yamashita 1992). When improperly folded proteins accumulate in the ER, Ire1p autophosphorylates, thereby activating the Ire1p endoribonuclease activity, which catalyzes the splicing of the HAC1 mRNA (Cox and Walter 1996; Mori et al. 2000), followed by ligation by Rlg1p, a tRNA ligase (Sidrauski et al. 1996). Since only the spliced form of HAC1 mRNA is effectively translated (Chapman and Walter 1997; Kawahara et al. 1997), this regulated splicing leads to expression of Hac1p and subsequent activation of transcription of genes such as KAR2, containing the unfolded protein-responsive element (UPRE) in their promoters.

Cells carrying ire1Δ, hac1Δ, or rlg1-100 mutations are sensitive to drugs such as tunicamycin, which causes accumulation of misfolded proteins in the ER (Nikawa et al. 1996; Shamu and Walter 1996; Sidrauski et al. 1996; Cox et al. 1997). In addition, ire1Δ, hac1Δ, or rlg1-100 mutants are inositol auxotrophs (Nikawa and Yamashita 1992; Cox et al. 1993; Nikawa et al. 1996; Sidrauski et al. 1996), a phenotype associated with defects in expression of genes related to phospholipid metabolism, especially INO1, the structural gene encoding myo-inositol 3-phosphate synthase (for review see Henry and Patton-Vogt 1998). INO1 and coregulated genes of phospholipid metabolism contain the inositol-sensitive upstream activating sequence (UASINO) repeated element in their promoters and exhibit complex transcriptional regulation in response to a variety of environmental factors including the availability of soluble precursors of phospholipid metabolism such as inositol (Carman and Henry 1999). Wild-type yeast cells express INO1 and other UASINO-containing genes at a high level when inositol is limiting in the growth medium and repress these same genes when inositol is plentiful (Hirsch and Henry 1986; Greenberg and Lopes 1996; Carman and Henry 1999; Loewen et al. 2004). Cox et al. (1997) reported that the UPR is activated in the absence of inositol and suggested that the activation of the UPR might be directly involved in the mechanism by which INO1 transcription is activated when inositol is limiting.

In an earlier study, we explored the relationship between UPR induction and INO1 expression in response to signals generated due to altered phospholipid metabolism in mutants defective in the phosphatidylinositol (PI)/phosphatidylcholine (PC) transporter encoded by the SEC14 gene (Chang et al. 2002). While the SEC14 gene product is essential for viability, the growth and secretory defects of sec14 mutants can be suppressed by mutations in the cytidine 5′-diphosphate (CDP) choline pathway for PC biosynthesis, such as cki1Δ, pct1Δ, and cpt1Δ (Cleves et al. 1991). Double mutants, such as sec14ts cki1Δ, when shifted to a semipermissive or restrictive temperature for sec14ts, exhibit multiple abnormalities in phospholipid metabolism, including both elevated PC turnover via a phospholipase D catalyzed route and elevated expression of INO1 (Patton-Vogt et al. 1997; Sreenivas et al. 1998). Under these same conditions sec14ts cki1Δ cells also exhibit high levels of UPR activation (Chang et al. 2002). These observations are consistent with the hypothesis of Cox et al. (1997) that the UPR and inositol responses are linked and are branches of the same signaling pathway. However, when the UPR mutations ire1Δ and hac1Δ were crossed into the sec14ts cki1Δ background, elevated INO1 expression was still observed at the sec14ts semipermissive temperature, indicating that UPR activation is not required or responsible for the elevated INO1 expression observed in sec14ts cki1Δ cells (Chang et al. 2002).

The effects we observed in sec14ts cki1Δ cells led us to question whether defects in other steps in the secretory pathway might also activate the UPR and/or influence INO1 expression and, thus, provide further insights into the role of the UPR in the secretory pathway and in INO1 regulation. In this study, we have examined the role of the UPR in membrane trafficking and in INO1 expression under conditions of secretory stress induced by lesions in a number of transport steps of the secretory pathway. We report that cells defective in a wide range in membrane trafficking steps exhibit UPR activation and that a functional UPR plays a role in the survival of cells with inefficient membrane trafficking due to a wide range of secretory defects. Thus, the UPR appears to be essential to growth under conditions in which secretory capacity is diminished. However, Sec mutants grown under the conditions that result in UPR activation exhibit reduced, rather than elevated, INO1 expression. Thus, the relative levels of UPR activation and INO1 transcription are not correlated under conditions of secretory stress.

MATERIALS AND METHODS

Media and growth conditions:

Yeast extract peptone media with dextrose (YEPD) and synthetic medium, with (I+) or without (I) inositol, or containing tunicamycin (tm) with or without leucine or uracil were prepared as previously described (Chang et al. 2002). Media containing brefeldin A (BFA; Sigma, St. Louis) were prepared as described by Graham et al. (1993) and Vogel et al. (1993).

Yeast strains:

The genotypes and sources of strains are listed in Table 1. Yeast strains listed in Table 1 are of the S288C and W303 genetic backgrounds. Strains were constructed by standard tetrad analysis (Sherman et al. 1978; Rose et al. 1990). Briefly, HCY104, HCY105, HCY106, HCY107, HCY126, and HCY362 were obtained as meiotic spore colonies from diploid deletion strains obtained from the Research Genetics (Huntsville, AL) strain collection. To obtain an ire1Δ strain isogenic with the genetic background of Sec strains used in this study, the HCY405 strain was constructed using PCR-based gene disruption of IRE1 in the CKY8 genetic background. A 1.5-kb deletion cassette containing the bacterial KanMX gene was amplified by PCR using pFA6-KanMX4 as a template and IRE1-S1 and IRE1-S2 as the primers (Wach et al. 1994). The KanMX4-containing PCR fragment was included at each end, 45 bp (85–129) and 43 bp (3039–3081), of the IRE1 ORF, resulting in replacement of the IRE1 gene with KanMX. The 1.5-kb deletion cassette was transformed into CKY8 using the lithium acetate (LiAc) method as described previously (Hill et al. 1991). Transformants were selected on YEPD containing 200 mg/liter geneticin (YEPD + G418; Calbiochem, La Jolla, CA). To verify the correct ORF replacement, the size of the insert (1.5 kb) was checked by PCR using a set of primers flanking the disrupted gene (IRE1-S3 and IRE1-S4) and genomic DNA isolated from each of the putative deletion mutants as a template. HCY423 was obtained by mating HCY405 and CKY46. To obtain a hac1Δ mutant in the same genetic background, a hac1Δ meiotic spore generated from the heterozygous diploid strain collection (Research Genetics) was backcrossed to CKY49 to generate HCY126.

TABLE 1

Strains used in this study

StrainGenotypeSource/reference
BJ8928MATapep12::LEU2 his3 leu2 lys2 trp1 ura3Elizabeth Jones
RSY263MATα sec12ts-4 leu2-3,112 ura3-52Randy Schekman
RSY267MATα sec16ts-2 his4-619 ura3-52Randy Schekman
RSY269MATα sec17ts-1 his4-619 ura3-52Randy Schekman
RSY271MATα sec18ts-1 his4-619 ura3-52Randy Schekman
RSY279MATα sec22ts-3 his4-619 ura3-52Randy Schekman
RSY281MATα sec23ts-1 his4-619 ura3-52Randy Schekman
CKY480MATα sec13-Δ1 ade2-101 ade3-24 leu2-3, 112 ura3-52 pSEC13 (CEN4; URA3)Chris Kaiser
CKY46MATasec13ts-1 his4-619 ura3-52Chris Kaiser
CKY49MATasec13ts-4 leu2-3,112 ura3-52Chris Kaiser
CKY8MATα leu2-3,112 ura3-52Chris Kaiser
JBY318MATasec6ts-4 his4 leu2 ura3 GAL+Jeff Brodsky
RCY927MATα sec21ts-1 leu2-3,112 ura3-52Ruth Collins
RCY243MATα sec1ts-1 leu2-3,112 ura3-52Ruth Collins
RCY248MATα sec4ts-8 leu2-3,112 ura3-52Ruth Collins
RCY274MATα sec2ts-59 ura3-52Ruth Collins
RCY260MATasec15ts-1 leu2-3,112 ura3-52Ruth Collins
SHY625MATα sec14ts-3 his3 leu2 lys2 trp1 ura3Patton-Vogt et al. (1997)
HCY030MATα hac1::URA3 his3 lys2 leu2 trp1 ura3Chang et al. (2002)
HCY104MATahis3 leu2 ura3This study
HCY105MATα ire1::KanMX his3 leu2 ura3This study
HCY106MATα his3 leu2 lys2 ura3This study
HCY107MATaire1::KanMX his3 leu2 lys2 ura3This study
HCY126MATα hac1::KanMX his3 leu2 lys2 ura3This study
HCY178MATα sec13ts-4 hac1::KanMX leu2 ura3This study
HCY179MATahis3 leu2 ura3This study
HCY180MATα sec13ts-4 his3 leu2 lys2 ura3This study
HCY181MATahac1::KanMX leu2 lys2 ura3This study
HCY280MATα sec12ts-4 his4-619 ura3This study
HCY283MATα sec12ts-4 ire1::KanMX his4-619 leu2-3,112 lys2 ura3This study
HCY362MATaerg6::KanMX his3 leu2 met15 ura3This study
HCY401MATα ire1::TRP1 ade2 his3 leu2 trp1Chang et al. (2002)
HCY405MATα ire1::KanMX leu2-3,112 ura3-52This study
HCY422MATα sec13ts-1 ire1::KanMX his4-619 ura3This study
HCY423MATaire1::KanMX ura3-52This study
HCY424MATaleu2 ura3This study
HCY425MATα sec13ts-1 his4-619 leu2 ura3This study
HCY427MATα sec16ts-2 his4-619 ura3This study
HCY428MATasec16ts-2 ire1::KanMX his4-619 ura3This study
HCY435MATα sec17ts-1 ire1::KanMXhis4-619 ura3This study
HCY437MATα sec17ts-1 ura3This study
HCY442MATasec18ts-1 ire1::KanMX his4 ura3This study
HCY444MATα sec18ts-1 ura3This study
HCY446MATasec23ts-1 ura3This study
HCY448MATasec23ts-1 ire1::KanMXhis4 ura3This study
HCY455MATasec22ys-3 his4 ura3This study
HCY457MATα sec22ts-3 ire1::KanMXura3This study
HCY462MATahis3 leu2 met15 ura3This study
HCY463MATaerg6::KanMX ade2 his3 leu2This study
HCY464MATα erg6::KanMX ire1::TRP1 his3 leu2 met15 trp1This study
HCY465MATα ire1::TRP1 ade2 his3 leu2 trp1 ura3This study
HCY466MATaerg6::KanMX his3 leu2 met15 ura3This study
HCY467MATahac1::URA3 his3 leu2 lys2 met15 trp1 ura3This study
HCY468MATα erg6::KanMX hac1::URA3 his3 leu2 lys2 ura3This study
HCY469MATα his3 leu2 trp1 ura3This study
HCY470MATasec6ts-4 his4 leu2 ura3This study
HCY473MATα sec6ts-4 ire1::KanMXhis4 leu2 ura3This study
HCY495MATα sec1ts-1 his3 leu2 ura3This study
HCY496MATα sec1ts-1 ire1::KanMX leu2 lys2 ura3This study

Double mutants, Sec ire1Δ or Sec hac1Δ, were constructed by crossing the various ire1Δ or hac1Δ strains to the various Sec strains. HCY280 and HCY283 were obtained by mating HCY423 and RSY263. HCY427 and HCY428 were obtained by mating HCY423 and RSY267. HCY425 and HCY422 were obtained by mating HCY405 and CKY46. HCY437 and HCY435 were obtained by mating HCY423 and RSY269. HCY442 and HCY444 were obtained from the mating between HCY423 and RSY271. HCY455 and HCY457 were obtained from the cross of HCY423 and RCY279. HCY448 and HCY446 were obtained by mating HCY423 and RCY281. HCY470 and HCY473 were obtained by mating HCY405 and JBY318. HCY178, HCY179, HCY180, and HCY181 were obtained from the cross of HCY126 and CKY49. HCY462, HCY463, HCY465, and HCY464 were the products of a cross between HCY362 and HCY401. HCY495 and HCY496 were obtained from a mating between HCY107 and RCY243. Two sets of four spores containing the tetratype genotypes derived from a single tetrad (HCY462–HCY465 and HCY466–HCY469) were generated by crossing HCY362 (erg6Δ) to HCY401 (ire1Δ) and HCY030 (hac1Δ), respectively.

Transformation with pHAC1 and pHAC1i:

Various Sec strains were transformed with the HAC1- and HAC1i-containing plasmids (pHAC1 and pHAC1i, CEN/LEU2 marker; gifts from Dr. Kazutoshi Mori, Kyoto University, Kyoto, Japan). The sec13Δ strain, CKY480 (see Table 1 for the full genotype), carrying the wild-type SEC13 gene on a plasmid (pRS316; CEN; URA3 marker; provided by Dr. Chris Kaiser) was transformed with pHAC1i or an empty vector (YCp50; CEN; LEU2 marker). Since the SEC13 gene is essential for growth, the SEC13 null mutant is obligated to carry and express the wild-type SEC13 gene on a plasmid for its survival. The double transformants simultaneously carrying pSEC13 and an empty vector or pSEC13 and pHAC1i were grown for 1 day on plates containing synthetic medium lacking uracil and leucine and then replica plated onto plates lacking leucine and containing 5-fluoroorotic-acid (5-FOA) and incubated for an additional 3 days at room temperature. The expression of the URA3 gene is toxic to cells exposed to 5-FOA, forcing the transformants to lose the pSEC13 plasmid, thus obligating them to depend on expression of pHAC1i for survival.

β-Galactosidase assays:

To assay UPRE-CYC-lacZ expression, strains were transformed to uracil prototrophy with pJC104, containing UPRE-CYC-lacZ, provided by Dr. Peter Walter (Cox and Walter 1996). Transformants were precultured to midlogarithmic phase of growth at 25° in synthetic medium lacking uracil. Cells were collected by centrifugation, washed with sterile dH2O, and diluted to OD600 = 0.1 in medium lacking uracil and shifted to the designated temperature to grow for 3 additional hours. Samples (1 ml) were taken and analyzed for β-galactosidase activity using the Yeast β-galactosidase assay kit (Pierce, Rockford, IL). For UPRE-CYC-lacZ expression assay in BFA-treated erg6Δ cells, HCY104 and HCY362 were transformed with pJC104 (UPRE-CYC-lacZ). Cells were prepared as described above except that their initial OD600 was 0.5. Indicated amounts of BFA were added to medium lacking uracil and cells were grown at 30°, and β-galactosidase activity was measured as described above.

To assess INO1 expression, strains were transformed to uracil prototrophy with pJH359 (INO1-CYC-lacZ) and transformants were precultured at 25° to midlog phase of growth in synthetic medium with (I+) or without (I) inositol. Cells were then shifted to the indicated temperature for 3 hr and β-galactosidase was measured.

Northern blot analysis:

Cells were precultured overnight to midlogarithmic phase of growth at 25° in YEPD medium. Wild-type cells were shifted to YEPD medium at 30° in the presence or absence of 1 mm tunicamycin. Sec strains were shifted to YEPD medium at their semipermissive temperatures. Isolation of RNA was performed by hot phenol extraction. Northern analysis was performed by running 10-μg samples of RNA loaded onto 1% agarose, 6% formaldehyde, 1× MOPS gels and capillary transferred to Nylon Plus (QIAGEN, Valencia, CA) membrane. Prehybridization and hybridization conditions were as described in Hirsch and Henry (1986). Hybridization probes for ACT1 (Marykwas and Fox 1989) and HAC1 were prepared using a riboprobe in vitro transcription system. ACT1 was linearized with BamHI and riboprobe was synthesized with SP6 RNA polymerase according to manufacturer's instructions (Promega, Madison, WI). The 449-bp SalI fragment from YCp-HAC1 was subcloned into pGEM1 (Promega) through the SalI site (pGEM-HAC1). pGEM-HAC1 was linearized with HindIII. Riboprobe was synthesized with T7 RNA polymerase according to manufacturer's instructions.

Spheroplast pulse-chase labeling and immunoprecipitation:

Spheroplast labeling and immunoprecipitation were carried out essentially as described by Webb et al. (1997). All strains were grown at the temperatures indicated in Figure 5 and in the text to midlogarithmic phase (OD600 = 0.6–1.0) at 24° in either Wickerman's sulfate-free medium containing 200 μm MgSO4 and 0.2% yeast extract (wild type, sec13ts-1, and sec13ts-1 ire1Δ strains) or synthetic complete medium lacking leucine (sec13ts-4 hac1Δ pHAC1i transformants). Five OD600 units of cells per time point were collected and converted to spheroplasts in Wickerham's medium containing 1 m sorbitol and 1 mg/ml bovine serum albumin by digestion with 30 units/OD600 lyticase (Sigma) at room temperature for 45 min. Spheroplasts from wild-type, sec13ts-1, and sec13ts-1 ire1Δ strains were preincubated for 15 min at 30°; pulse labeled with 50 μCi Trans-35S labeling reagent (ICN Radiochemicals, Irvine, CA) in spheroplasting medium for 5 min at 30°; and chased by adding 5 mm methionine and 1 mm cysteine, 0.2% yeast extract, and 2% glucose and incubating for 0, 10, 20, or 30 min at 30°. Pulse-chase labeling of spheroplasts from sec13ts-4 hac1Δ pHAC1i transformants was carried out as described above except incubations were performed at 35°. Protein extracts were prepared by trichloracetic acid precipitation of cell pellets followed by resuspension and boiling with glass beads in suspension buffer (50 mm Tris-HCl pH 7.5, 1 mm EDTA, 1% SDS). After a fivefold dilution with IP dilution buffer (60 mm Tris-HCl pH 7.5, 6 mm EDTA, 190 mm NaCl, 1.25% Triton X-100), samples were precleared by centrifugation to remove insoluble material. Carboxypeptidase Y (CPY) was immunoprecipitated from each sample by incubating with rabbit anti-CPY antibodies (Rockland, Gilbertsville, PA) and protein A-sepharose CL-4B (Sigma) in IP buffer (50 mm Tris-HCl pH 7.5, 5 mm EDTA, 150 mm NaCl, 1% Triton X-100, 0.2% SDS). Washed immunoprecipitates were eluted with sample buffer, separated on 8% SDS polyacrylamide gels, and visualized by autoradiography. A digital image of the gel was acquired by scanning the autoradiograph. To obtain the signal intensity of each species of CPY, the digital image was analyzed by Kodak digital science 1D image analysis software (Eastman Kodak, Rochester, NY). The relative total labeling of CPY was estimated by summing up the signal intensity of p1, p2, and mature (mCPY) on each autoradiograph. The proportion of each species of was then calculated by dividing the separate signal intensity of p1, p2, or m by the sum of the intensity of the three bands.

An external file that holds a picture, illustration, etc.
Object name is 2961f5.jpg

Expression of HAC1i partially suppresses the growth defect (ts) and Ino phenotypes of some Sec strains. (A) Growth at different temperatures of HCY180 (sec13ts-4) and HCY178 (sec13ts-4 hac1Δ) cells transformed with pYC50 (empty vector), pHAC1 (wild-type HAC1), and pHAC1i (intronless HAC1) is shown. (B) Growth at different temperatures of HCY496 (sec1ts-1 ire1Δ) cells transformed with pYC50, pHAC1, and pHAC1i. Transformants in both A and B were precultured to midlogarithmic phase of growth (OD600 = 0.5) at 20° in medium lacking leucine, spotted as a 10-fold dilution series on leucine-free plates, and incubated at the indicated temperatures for 4 days. (C) Effect of pHAC1 and pHAC1i on the Ino and ts phenotypes of sec13-1. sec13-1 cells were transformed with pYC50 (empty vector), pHAC1 (wild-type HAC1), and pHAC1i (intronless HAC1). Transformants were precultured to midlogarithmic phase of growth (OD600 = 0.5) at 20° in medium lacking leucine, spotted as a 10-fold dilution series on leucine-free plates containing (I+) or lacking (I) inositol, and incubated for 3 days at the indicated temperature.

RESULTS

Secretory stress activates the UPR:

Secretory mutants affecting various membrane transport steps (for review of secretory pathway mutant defects, see Kaiser et al. 1997) were chosen for analysis of UPR induction and INO1 expression under conditions of secretory stress. The strains analyzed include several carrying temperature-sensitive mutations conferring defects in formation of COPII vesicles affecting exit from the ER (sec12ts-4, sec13ts-1, sec13ts-4, sec16ts-2, and sec23ts-1), three with defects in vesicle docking and fusion processes affecting the early stages of the secretory pathway (sec17ts-1, sec18ts-1, and sec22ts-3), and one with a defect in COPI coatomer (sec21ts-1). Five mutations conferring temperature-sensitive defects in distal secretion to the plasma membrane (sec6ts-4, sec1ts-1, sec15ts-1, sec2ts-59, and sec4ts-8) and one deletion mutation conferring a nonlethal defect in vacuolar targeting, pep12Δ (Becherer et al. 1996), were also analyzed).

Activation of the UPR in wild-type and mutant strains (Figure 1A) was assayed using an UPRE-CYC-lacZ reporter gene, as described in materials and methods. Sec strains were assayed only under conditions at which they could grow continuously. Therefore, β-galactosidase activity is reported for only a subset of strains at 33° (Figure 1A). Wild-type cells exhibited the expected low levels of UPRE expression at 25° and 30°, and only a slight increase in UPRE expression was observed at 33° (Figure 1A). In contrast, virtually all of the temperature-sensitive Sec mutants showed elevated UPRE activation compared to wild type when grown at their semipermissive temperatures of 30° or 33° (Figure 1A).

An external file that holds a picture, illustration, etc.
Object name is 2961f1.jpg

Activation of the UPR in cells blocked in secretion and vacuolar targeting. (A) Transcription of UPRE is induced in mutants defective in membrane trafficking and transport. CKY8 (wt), RSY263 (sec12ts-4), CKY46 (sec13ts-1), CKY49 (sec13ts-4), RSY281 (sec23ts-1), RSY267 (sec16ts-2), RSY269 (sec17ts-1), RSY271 (sec18ts-1), RSY279 (sec22ts-3), RCY927 (sec21ts-1), RCY243 (sec1ts-1), RCY248 (sec4ts-8), JBY318 (sec6ts-4), RCY260 (sec15ts-1), and BJY8928 (pep12Δ) were transformed with pJC104 (UPRE-CYC-lacZ). Transformants were precultured at 20° to midlogarithmic phase of growth in medium lacking uracil and shifted to indicated temperatures. Samples were taken after 3 hr following the shift and analyzed for β-galactosidase activity as described in materials and methods. β-Galactosidase activity unit was defined as OD420/min/ml. (B) Transcription of UPRE is induced at 30° in BFA-treated erg6Δ cells. HCY 104 (wt) and HCY362 (erg6Δ) carrying pJC104 (UPRE-CYC-lacZ) were grown in uracil dropout medium containing 0, 10, 25, and 50 mm BFA. Samples were taken at 2 hr. β-Galactosidase activity unit was defined as OD420/min/ml. (C) Splicing of HAC1 transcripts is elevated in sec13ts-1 cells. Northern blot analysis was performed to monitor splicing of HAC1 mRNA in sec13ts-1 cells (CKY49) grown at 25° and 30°. As a control, splicing of HAC1 mRNA was also monitored in wild-type cells (CKY8) treated with 1 mm tunicamycin (Tm). Cells were precultured overnight in YEPD medium at 25° and shifted to indicated temperatures (for wild-type positive control, 1 mm Tm was added). Samples were taken after 3 hr following the temperature shift. ACT1 mRNA is shown as a loading control.

Several of the Sec mutants showed elevated UPRE expression even at 25°. For example, among the strains defective in COPII vesicle formation, UPRE-CYC-lacZ activity was elevated >11-fold in the sec16ts-2 strain and >3-fold in the sec12ts-4 and sec23ts-1 strains, in comparison to wild type, at the permissive temperature of 25° (Figure 1A). At the semipermissive temperature of 30°, UPRE expression 7- to 10-fold higher than that of wild type was observed in sec12ts-4, sec13ts-1, sec23ts-1, and sec16ts-2 mutants (Figure 1A). Strains carrying sec13ts-4 are capable of growing at temperatures up to 36°, and under these conditions, UPRE expression was elevated >7-fold compared to expression in the same strain at 25° (data not shown).

The sec17ts-1, sec18ts-1, and sec22ts-3 strains, which are defective in vesicle docking and fusion, also exhibited a five- to eightfold elevation in UPRE-driven β-galactosidase activity at 30° (Figure 1A). In the sec21ts-1 strain, which has a defect in COPI coatomer, UPRE expression was elevated more than fivefold at 30° and 33°, as compared to wild type (Figure 1A). UPRE expression three- to sevenfold higher than that of wild type was also observed at the semipermissive temperature of 30° or 33° in sec15ts-1, sec2ts-59, sec1ts-1, sec6ts-4, and sec4ts-8, which are defective in distal secretion. The pep12Δ strain, which is defective in vacuolar targeting and is not temperature sensitive, exhibited UPRE expression levels approximately fourfold higher than wild-type grown under identical conditions at all temperatures tested (Figure 1A). Unlike all of the other Sec strains examined here, the sec14ts-3 single-mutant strain exhibited levels of UPRE-CYC-lacZ expression indistinguishable from wild type when shifted to temperatures of 30° and 33° (data not shown), semipermissive temperatures at which we observed substantial UPRE induction in the sec14ts-3 cki1Δ strain in a previous study (Chang et al. 2002).

The mechanism of UPR activation involves splicing of the HAC1 transcript by Ire1p, following activation of Ire1p by autophosphorylation (Cox et al. 1993; Mori et al. 1993; Cox and Walter 1996). To determine whether HAC1 splicing was elevated under secretory stress, Northern blot analysis was performed on RNA extracted from sec13ts-1 cells. As a control, wild-type cells were treated with 1 mm tunicamycin, which has been shown to elicit the UPR (Cox and Walter 1996). Untreated wild-type and sec13ts-1 cells grown at 25° contained only the full-length (unspliced) HAC1 transcript (Figure 1C). As expected, tunicamycin-treated wild-type cells contained two distinctive sizes of HAC1 mRNA corresponding to unspliced (HAC1u) and spliced (HAC1i) mRNA (Figure 1C). sec13ts-1 cells grown at the semipermissive temperature of 30°, in the absence of tunicamycin, likewise contained both the spliced and unspliced forms of HAC1 transcript (Figure 1C), confirming that HAC1 mRNA is actively spliced in sec13ts-1 cells at 30°.

We observed a modest increase in the expression of β-galactosidase from the UPRE-CYC-lacZ construct after shifting wild-type cells to 36° (data not shown). Since changes in growth temperature are known to cause changes in metabolism, including alterations in the pattern of turnover of membrane lipids (Dowd et al. 2001), we examined activation of the UPR in strains treated with or without the fungal metabolite BFA at 30°. BFA blocks secretion, results in loss of COPI coats from the Golgi, and causes redistribution of Golgi enzymes to the ER (Lippincott-Schwartz et al. 1989; Klausner et al. 1992; Kreis et al. 1995; Sciaky et al. 1997; Seemann et al. 2000). Unlike mammalian cells, most Saccharomyces cerevisiae strains are insensitive to BFA. However, erg6 mutants, which are defective in biosynthesis of ergosterol, exhibit sensitivity to BFA presumably due to increased membrane permeability (Graham et al. 1993; Shah and Klausner 1993; Vogel et al. 1993). As shown in Figure 1B, wild-type cells that are impermeable to BFA exhibited no UPRE induction in the presence of BFA at 30°. erg6Δ cells, however, exhibited high levels of UPRE-driven β-galactosidase expression at 30° in medium containing 25 and 50 mm BFA (Figure 1B), concentrations that do not inhibit growth of erg6Δ. Expression driven by the UPRE construct in the erg6Δ strain was elevated ∼9-fold within 1 hr following exposure to 25 mm BFA (data not shown), compared to erg6Δ cells grown without BFA. After 2 hr of exposure to 25 mm BFA at 30°, UPRE expression was elevated ∼15-fold in the erg6Δ strain (Figure 1B). Similar elevated levels of UPRE-driven β-galactosidase expression were observed at equivalent time points in erg6Δ cells grown in medium containing 50 mm BFA (Figure 1B).

INO1 expression is reduced in Sec mutants experiencing secretory stress:

INO1 expression was assayed in a selection of the Sec mutants as described in materials and methods using an INO1-CYC-lacZ reporter gene. All of the mutants tested exhibited reduced INO1 expression compared to wild type when cultured at semipermissive growth temperatures in inositol-free (I) medium (Figure 2A). This effect was most pronounced in the sec13ts-1 and sec16ts-2 strains, where a three- to fivefold reduction in INO1-CYC-lacZ expression was observed in cells grown in I medium shifted to 30° compared to the same cells grown continuously at 25° (Figure 2A). The sec18ts-1, sec22ts-3, and sec21ts-1 mutants also exhibited levels of INO1 expression that were 50% of wild type or lower, even at the permissive temperature of 25° (Figure 2A). A further reduction in INO1-CYC-lacZ expression was observed in these strains when they were shifted to the semipermissive temperatures of 30° or 33°. Of the strains tested (Figure 2), the sec22ts-3 strain showed the lowest level of INO1 expression at 30° (∼10% of wild-type levels; Figure 2A). Somewhat reduced levels of INO1 expression were also observed at all temperatures in the sec1ts-1, sec6ts-4, and sec15ts-1 strains. In these cases, β-galactosidase activity from the INO1-CYC-lacZ reporter gene was reduced compared to wild type even at 25°, but was not significantly further reduced at 30° or 33°. In the sec14ts-3 strain, expression from the INO1 reporter gene was reduced compared to wild type at 25° and 30°. Expression of INO1-CYC-lacZ increased somewhat after a shift to 33°, but did not reach the level of expression observed in the wild-type strain grown under comparable circumstances (Figure 2A).

An external file that holds a picture, illustration, etc.
Object name is 2961f2.jpg

INO1 expression is reduced in Sec cells under conditions of secretory stress. (A) Sec strains were transformed with pJH359 (INO1-CYC-lacZ) as described in materials and methods and transformants were precultured at 25° to midlogarithmic phase of growth in I medium lacking uracil and shifted to the indicated temperatures. Samples were taken after 3 hr following the temperature shift and analyzed for β-galactosidase activity. β-Galactosidase activity is expressed as OD420/min/ml. (B) sec13-1 cells carrying pJC104 (UPRE-CYC-lacZ) were precultured at 30° to midlogarithmic phase of growth in I+ medium lacking uracil. Cells were filtered and resuspended in I+ or I medium, which also lacked uracil, at an OD600nm = 0.5. The cultures were incubated for an additional 3 hr at 30° and then analyzed for β-galactosidase activity. (C) sec13ts-1 cells carrying pJH359 (INO1-CYC-lacZ) were precultured at 30° to midlogarithmic phase of growth in I+ medium lacking uracil. Cells were filtered and resuspended in I+ or I medium lacking uracil at an OD600nm = 0.5. Cultures were incubated at 30° for an additional 3 hr and analyzed for β-galactosidase activity.

Since the UPR is known to be activated under inositol-limiting conditions (Cox et al. 1997), we examined both UPRE-CYC-lacZ and INO1-CYC-lacZ expression in sec13ts-1 and wild-type cells shifted from inositol-containing (I+) to inositol-free (I) medium. Wild-type and sec13ts-1 cells, transformed with INO1-CYC-lacZ or UPRE-CYC-lacZ reporter constructs, were first grown to logarithmic phase at 30° in I+ medium and then shifted for 3 additional hours to I medium (Figure 2, B and C). Wild-type transformants treated in this fashion showed modest expression from the UPRE reporter gene following the shift to I medium at 30°, as well as derepression of the INO1 reporter construct (Figure 2, B and C), consistent with the report of Cox et al. (1997).

In contrast to wild type, at 30°, sec13ts-1 cells exhibited high levels of β-galactosidase expression from the UPRE reporter gene whether inositol was present or not (Figure 2B). In fact, the level of UPRE expression in sec13ts-1 cells grown at 30° in I+ medium, exceeded the level of UPRE expression in wild-type cells shifted to I medium by >8-fold. Following the shift to I medium, a further increase of ∼30% in expression from the UPRE-CYC-lacZ reporter gene was observed in sec13ts-1 cells (Figure 2B). Overall, the level of β-galactosidase activity driven by the UPRE construct in sec13ts-1 cells grown at 30° in I medium exceeded that observed under the same conditions in wild-type cells by ∼12-fold (Figure 2B). In contrast to the high level of UPRE expression observed in sec13ts-1 cells grown at 30° in I+ medium (Figure 2B), INO1-CYC-lacZ expression in the sec13ts-1 cells at 30° was fully repressed when inositol was present (Figure 2C). When sec13ts-1 cells were shifted to I medium at 30°, INO1-CYC-lacZ derepression occurred, but the level of β-galactosidase activity achieved was only about one-third of the level seen in the wild-type strain (Figure 2C).

The hac1Δ and ire1Δ mutations exhibit partial synthetic lethality with mutations conferring defects in membrane trafficking:

Sec ire1Δ and Sec hac1Δ strains were generated for each of the Sec mutants used in this study and, in most cases, double-mutant progeny were found to have growth defects that were more extreme than those of the corresponding parental Sec single mutant. Examples of Sec ire1Δ double mutants, from each of the categories of Sec mutants used for the temperature-shift experiments described above, are shown in Figure 3A. Sec hac1Δ double-mutant strains, which are not shown, had phenotypes very similar to the corresponding Sec ire1Δ strains. No change was detected in the growth pattern on YEPD plates in response to temperature of double mutants involving sec1ts-1 (Figure 3A) or pep12Δ with ire1Δ or hac1Δ, and slight, if any, difference was observed in the case of the double mutants involving sec4ts-8, sec14ts-3, and sec15ts-1, as compared to the corresponding single mutants (data not shown). Double mutants involving sec2ts-59 and sec21ts-1 were not generated. In all other cases, the double mutants, Sec ire1Δ or Sec hac1Δ, failed to grow at a temperature ∼2°–3° lower than that of the corresponding Sec parent (Figure 3A; Sec hac1Δ data not shown).

An external file that holds a picture, illustration, etc.
Object name is 2961f3a.jpg
An external file that holds a picture, illustration, etc.
Object name is 2961f3b.jpg

(A) Genetic interactions between ire1Δ and Sec mutations and Ino phenotypes of Sec Upr mutants. The ire1Δ (either HCY405 or HCY423) strain was crossed to Sec mutants (CKY46, RSY267, RSY269, and JBY318) to generate Sec ire1Δ double-mutant strains. From each cross, four spore colonies from a single tetratype ascus were used to assess the growth phenotype. For simplicity, only growth of wild-type and ire1Δ spores generated from a cross of CKY46 and HCY405 is shown here, since all other wild-type and ire1Δ spores generated from other crosses showed similar growth phenotypes. HCY424 (wt), HCY423 (ire1Δ), HCY425 (sec13ts-1), HCY422 (sec13ts-1 ire1Δ), HCY280 (sec12ts-4), HCY283 (sec12ts-4 ire1Δ), HCY427 (sec16ts-2), HCY428 (sec16ts-2 ire1Δ), HCY437 (sec17ts-1), HCY435 (sec17ts-1 ire1Δ), HCY470 (sec6ts-4), HCY473 (sec6ts-4 ire1Δ), HCY495 (sec1ts-1), and HCY496 (sec1ts-1 ire1Δ) were precultured to midlogarithmic phase of growth at 20° in YEPD medium. The concentration of the cells was adjusted to OD600 = 0.7 with sterile dH2O. Cells were initially diluted 1:100 using dH2O followed by 1:10 serial dilutions. Four microliters of cells from each dilution were spotted on YEPD plates and allowed to grow at the designated temperatures for 3 days. (B) Strains generated as described in A, above, were precultured to midlogarithmic phase of growth at 25° in I+ medium, washed twice, and spotted as a series of dilutions (1:10) on I+ or I medium and incubated at the temperatures shown. In the top, the phenotypes of four spore colonies from the cross of ire1Δ to sec17-1 are shown. Below, the wild-type and ire1Δ colonies are omitted.

Mutations affecting COPII vesicle formation exhibited some of the strongest negative interactions with ire1Δ and hac1Δ. For example, the sec13ts-1 ire1Δ and sec13ts-1 hac1Δ strains failed to grow at a temperature of 30° or higher, while the restrictive temperature for the sec13ts-1 strain is 33° (Figure 3A). The sec16ts-2 ire1Δ and sec16ts-2 hac1Δ strains grew poorly even at 25°, while the sec16ts-2 parental strain was able to grow, although poorly, at temperatures up to 30° (Figure 3A). Double mutants simultaneously defective in the UPR and sec12ts-4 (Figure 3A) or sec23ts-1 (data not shown) also failed to grow at temperatures ∼2° lower than the restrictive temperature for the corresponding parental Sec strain. The sec23ts-1 ire1Δ and sec23ts-1 hac1Δ double mutants exhibited particularly poor growth and could grow only at 20° or lower on YEPD plates. They could not be cultured in liquid YEPD medium even at 20° and they failed to grow entirely on YEPD plates at 25° or higher. The sec23ts-1 parental strain, by contrast, was capable of normal growth on YEPD plates at temperatures up to 28° (data not shown). Double mutants carrying ire1Δ or hac1Δ in combination with mutations affecting vesicle docking and fusion, sec17ts-1 (Figure 3A), sec18ts-1, or sec22ts-3 (data not shown), also failed to grow at temperatures 2°–3° lower than the minimum restrictive temperature of the corresponding parental Sec mutant. Among the mutations that affect distal secretion from the Golgi to the plasma membrane, only sec6ts-4 exhibited a clear negative interaction with ire1Δ or hac1Δ. The sec6ts-4 ire1Δ and sec6ts-4 hac1Δ strains failed to grow at 33°, whereas the sec6ts-4 parental strain has a restrictive temperature of 35° (Figure 3A).

The ire1Δ and hac1Δ mutations were also crossed into the erg6Δ genetic background to assess their effect on BFA sensitivity. While growth of the erg6Δ parental strain was completely inhibited by the presence of 100 mm BFA, as previously reported (Graham et al. 1993; Shah and Klausner 1993; Vogel et al. 1993), and was unaffected by 25 or 35 mm BFA in YEPD medium (Figure 4), the erg6Δ ire1Δ and erg6Δ hac1Δ double-mutant strains were inhibited by 35 mm BFA (Figure 4). Thus, the presence of a UPR mutation results in an increased sensitivity of erg6Δ to BFA.

An external file that holds a picture, illustration, etc.
Object name is 2961f4.jpg

Mutations in the UPR increase the sensitivity of erg6Δ cells to BFA. (A) As controls, HCY104 (wt), HCY105 (ire1Δ), HCY126 (hac1Δ), and HCY362 (erg6Δ) were tested for their ability to grow on YEPD medium containing 100 mm BFA. (B) Two sets of four spores containing the tetratype genotypes derived from a single tetrad (HCY462–HCY465 and HCY466–HCY469) were generated by crossing HCY362 (erg6Δ) to HCY401 (ire1Δ) or HCY030 (hac1Δ). Cells were precultured in YEPD medium at 30° to midlogarithmic phase (OD600 = 0.5) of growth, spotted as a series of dilutions (1:10) on YEPD plates containing 25 mm and 35 mm BFA, and allowed to grow at 30° for 4 days.

Growth of a subset of the Sec strains was also examined at the semipermissive temperature of 30° on media containing or lacking inositol. As reported by Nikawa and Yamashita (1992), ire1Δ (Figure 3B) and hac1Δ strains (data not shown) exhibit somewhat “leaky” inositol auxotrophy (Ino phenotype; Figure 3B). The sec13-1 strain also has a leaky Ino phenotype at its semipermissive temperature of 30° (Figure 3B) as reported by Gilstring et al. (1999). The sec14ts-3 strain also grows more poorly in I medium at the semipermissive temperature of 33° (Figure 3B), as reported by Kearns et al. (1997) and Chang et al. (2002). In a number of cases tested, the double mutants (i.e., Sec ire1Δ or Sec hac1Δ) exhibited somewhat tighter Ino phenotypes than either single-mutant parent did (Figure 3B). Growth of five such double mutants, sec14 ts-3, sec13ts-1, sec6ts-4, sec16ts-2, and sec17ts-1, in combination with ire1Δ is shown in Figure 3B.

Constitutive activation of the UPR by expression of the HAC1i gene rescues growth defects of some Sec mutants:

The sec1ts-1, sec4ts-8, sec13ts-1, sec13ts-4, sec14ts-3, sec15ts-1, and sec22ts-3 strains were transformed with the HAC1i construct, lacking the intron sequence. The lack of the intron in HAC1i results in constitutive activation of the UPR, even in ire1Δ strains (Cox and Walter 1996; Kawahara et al. 1997, 1998; Mori et al. 2000).

The sec15ts-1 and sec14ts-3 mutants transformed with pHAC1i appeared to grow more poorly than the same strains carrying vector alone (data not shown), a result consistent with the reports of Kawahara et al. (1997) and Mori et al. (2000), who observed that wild-type strains transformed with pHAC1i exhibited slower growth than controls. No change was observed in the growth of the sec1ts-1, sec22ts-3, or sec4ts-8 strains when transformed with pHAC1i (data not shown). While the growth of the sec1ts-1 strain was not affected by transformation with pHAC1i (data not shown), the sec1ts-1 ire1Δ strain transformed with pHAC1i was able to grow at 33° (Figure 5B), a temperature at which the sec1ts-1 single mutant is unable to grow (see Figure 3A). This result was not expected, since the sec1ts-1 ire1Δ and sec1ts-1 strains had similar phenotypes at 33° (Figure 3).

The most dramatic effects of expression of pHAC1i were obtained in the sec13ts-1 and sec13ts-4 strains. When transformed with pHAC1i, the sec13ts-4 and sec13ts-4 hac1Δ strains were able to grow at 37° (Figure 5A), a temperature that is restrictive for sec13ts-4. The sec13ts-1 strain was also able to grow at 32° when transformed with pHAC1i (data not shown), whereas sec13ts-1 will not grow above its semipermissive temperature of 30° when transformed with vector alone. HAC1i was also able to rescue growth of sec13ts-1 in the absence of inositol at 30° and 32° (Figure 5C).

However, HAC1i was not able to restore sec13ts-1 growth under any condition (with or without inositol) at temperatures of 33° or greater (Figure 5C), suggesting that HAC1i expression cannot suppress complete loss of Sec13p function. To test this idea, a sec13Δ strain carrying the wild-type SEC13 gene on plasmid pRS316 was transformed with pHAC1i or an empty vector (YCp50) and a plasmid “shuffling” experiment was performed, as described in materials and methods. When the transformants were forced to lose pSEC13, they failed to grow even when pHAC1i was present (data not shown), indicating that expression of the HAC1i gene cannot compensate for the total loss of Sec13p.

A functional UPR pathway improves the kinetics of transport to the vacuole in Sec mutants:

To assess the effect of the UPR on membrane trafficking, we monitored the kinetics of processing in several Sec mutants, with and without a deletion in either IRE1 or HAC1. These experiments were conducted at temperatures in which the UPR is activated in the corresponding Sec mutant, as determined in the experiments depicted in Figure 1A. In each case, the temperature selected was permissive or semipermissive for the corresponding Sec mutant, but restrictive in strains carrying the given Sec mutation in combination with ire1Δ or hac1Δ (i.e., the specific Secire1Δ or Sechac1Δ strains were incapable of sustained growth at the chosen temperature). The strains tested included sec13ts-1, sec13ts-4, sec22ts-3, and sec17ts-1 in combination with ire1Δ or hac1Δ. Data are shown in Figure 5 for strains carrying sec13ts-1 and sec13ts-4 mutations. Cells were briefly shifted to the elevated temperature, as described in materials and methods, and the kinetics of CPY maturation were monitored in strains using pulse-chase analysis followed by CPY immunoprecipitation and SDS-PAGE analysis followed by image analysis of the autoradiograms (Figure 6).

An external file that holds a picture, illustration, etc.
Object name is 2961f6.jpg

Comparison of processing in sec13 mutant strains when the UPR is absent or constitutively active. (A) The rate of conversion of to its mature form was determined by pulse labeling HCY424 (wt), HCY425 (sec13ts-1), and HCY422 (sec13ts-1 ire1Δ) cells for 5 min at 30° with [35S]methionine/cysteine, chasing with cold amino acids at 30° for the indicated times, and immunoprecipitating with anti-antibodies. ER-glycosylated (p1), Golgi-modified (p2) precursors, and vacuolar mature form (mCPY) are indicated. (B) The rate of conversion of in HCY178 (sec13ts-4 hac1Δ) transformants containing empty vector or expressing the HAC1i gene from a centromeric plasmid was measured by pulse labeling cells at 35° and chasing with cold amino acids for the indicated times. Processing of was analyzed following immunoprecipitation and SDS-PAGE. (C) Plotted percentages of mature at indicated times determined from autoradiograms shown in A and B, as described in materials and methods. Solid diamonds, wild type; solid squares, sec13ts-1; solid triangles, sec13ts-1 ire1Δ; open diamonds, sec13ts-4 hac1Δ + vector; open squares, sec13ts-4 hac1Δ + pHAC1i.

In the wild-type strain at 30°, the 67-kD ER form of CPY (Figure 6, p1) was rapidly converted to the 69-kD Golgi form (Figure 6, p2), which is proteolytically cleaved in the vacuole to the mature 61-kD form (Figure 6, mCPY). Thus, the appearance of mCPY signals the delivery of CPY to the vacuole (Stevens et al. 1982). The kinetics of CPY processing in the sec13ts-1 strain at 30°, which is a semipermissive temperature for this mutant allele of SEC13 (Figure 3A), were somewhat delayed compared to wild type (Figure 6A). The delay is most visible at the 10-min time point. At 10 min in sec13ts-1, about one-third of labeled CPY was recovered in mCPY, compared to about two-thirds in wild type, and the remaining label in CPY was distributed between the p1 and p2 forms (Figure 6, A and C). After 20 min in the sec13ts-1 strain, most of the labeled CPY was recovered in mCPY (Figure 6, A and C). By 30 min, there was very little difference between sec13ts-1 and wild type in the amount of m relative to the p2 Golgi form, but the p1 ER form was observed in the sec13ts-1 strain, even after 30 min. This observation is consistent with the reported ER to Golgi vesicular transport defect of sec13 mutants (Novick et al. 1980, 1981). In the sec13ts-1 ire1Δ strain, which is not capable of sustained growth at 30°, the kinetics of CPY processing at 30° were slowed still further as compared to both the wild-type and sec13ts-1 strains at 30°. In this strain, after 10 min, less than one-quarter of labeled CPY was present in mCPY (Figure 6, A and C). Comparable slowing in the appearance of mCPY was observed in the sec17ts-1 ire1Δ and sec22ts-3 ire1Δ strains relative to the sec17ts-1 and sec22ts-3 strains, respectively (data not shown), suggesting that slowing of CPY processing is not unique to the sec13ts-1 ire1Δ strain. Moreover, the kinetics of processing in an ire1Δ strain were identical to wild type (data not shown), indicating that the UPR pathway defect does not appreciably affect normal membrane trafficking in the absence of secretory stress.

We also measured the time course of CPY processing in the sec13ts-4 hac1Δ strain expressing the HAC1i gene from a centromeric plasmid and compared it to processing in the same strain transformed with an empty vector (Figure 6B). Transformants were briefly shifted to 35°, conditions in which the HAC1i gene product rescues the growth defect of sec13ts-4 hac1Δ strains (Figure 5A). Transformants were then subjected to pulse-chase analysis as described above. In the sec13ts-4 hac1Δ strain transformed with an empty vector, most of the labeled CPY remained in the ER (p1) form throughout the time course (Figure 6, B and C). However, in the sec13ts-4 hac1Δ strain transformed with the HAC1i gene, the kinetics of CPY processing were markedly faster than in the control transformed with vector alone (Figure 6, B and C). After 10 min, about one-third of the labeled CPY was present in mCPY compared to ∼ <20% in transformants containing vector alone (Figure 6, B and C). After 30 min, less than half of labeled CPY was processed to m in the transformants carrying vector alone, whereas in sec13ts-4 hac1Δ cells transformed with pHAC1i, ∼80% of the label was present in mCPY after 30 min.

DISCUSSION

In this study, we have documented a functional interaction between the UPR and the secretory pathway. The UPR is activated in Sec mutants defective in events extending from ER vesicle trafficking to distal secretion and in Pep mutants defective in vacuolar targeting (Figure 1). The UPR is also activated in erg6Δ cells challenged with BFA (Figure 1). Moreover, our results indicate that activation of the UPR plays a functional role during secretory stress, facilitating protein trafficking in cells experiencing partial impairment in secretory function. This conclusion is supported by the partial synthetic lethality of the ire1Δ and hac1Δ mutations in combination with a large majority of Sec mutants tested in this study (Figure 3), as well as the partial suppression of certain Sec phenotypes by HAC1i (Figure 5). The slowing of CPY processing in Sec strains carrying a UPR mutation and the rescue of these defects by HAC1i provide direct confirmation that UPR plays a role in facilitating protein processing under these conditions (Figure 6).

However, under conditions of secretory stress created by elevating Sec mutants to a semipermissive temperature, the observed activation of the UPR does not result in elevated INO1 expression compared to wild type, even when inositol is absent. Indeed, under conditions of secretory stress, in Sec cells exposed to semipermissive growth conditions, INO1 expression is actually reduced compared to wild type (Figure 2) and some strains, such as sec13ts-1 and sec14ts-3 grown at their semipermissive temperatures, actually require inositol for growth (Figure 3B). Thus, it appears that the activation of the UPR, which occurs under conditions of secretory stress, is not correlated with elevated INO1 expression.

The role of the UPR and the secretory pathway in expression and regulation of the INO1 gene:

INO1 is one of a large number of coregulated genes, which contain the promoter element UASINO and respond to the availability of inositol in the growth medium (Carman and Henry 1999). While the mechanism by which the cell regulates INO1 in response to inositol and other precursors of lipid metabolism has not been fully elucidated, it is known that the Ino2p and Ino4p transcription factors bind to UASINO to activate transcription (Carman and Henry 1989; Lopes et al. 1991; Ambroziak and Henry 1994; Schwank et al. 1995; Greenberg and Lopes 1996). Furthermore, the negative regulator, Opi1p, is required for repression of UASINO-containing genes in response to inositol (White et al. 1991). Recent evidence suggests that changes in the pattern of membrane phospholipid synthesis, produced in response to the incorporation of exogenous inositol, result in translocation of the negative regulator, Opi1p, to the nucleus from the endoplasmic reticulum and that Opi1p translocation coincides with repression of INO1 (Loewen et al. 2004).

Cox et al. (1997) suggested that activation of INO1 transcription in the absence of inositol might be linked to activation of the UPR signal transduction pathway. In support of this hypothesis, Cox et al. (1997) showed that the UPR is activated in wild-type cells growing in inositol-free medium and that deletion of OPI1 suppressed the Ino phenotype of ire1Δ and hac1Δ mutants. Previously, consistent with the hypothesis of Cox et al., we reported that expression of both UPR and INO1 is elevated in sec14-3ts strains carrying the cki1Δ suppressor (Chang et al. 2002). However, deletion of HAC1 or IRE1 did not eliminate overexpression of INO1 in sec14ts cki1Δ cells (Chang et al. 2002), suggesting that activation of the UPR, while correlated with INO1 expression, is not obligatory for INO1 activation.

It has been assumed that the Ino phenotype of UPR mutants, such as hac1Δ and ire1Δ, is due to the inability to maintain wild-type INO1 expression levels in the absence of UPR activation (Cox et al. 1997; Chang et al. 2002). Another possibility is that inositol deprivation results in a stress condition that elicits UPR activation and that UPR activation under these circumstances is essential in much the same way that it is in wild-type cells experiencing stress due to the buildup of unfolded proteins following exposure to tunicamycin. We propose that providing inositol to UPR mutants under such circumstances alleviates the underlying stress condition that necessitates UPR activation for survival, thus explaining the Ino phenotype. Since inositol limitation (Cox et al. 1997; Chang et al. 2002) and secretory stress (Figure 1) both result in UPR activation, we questioned whether these two stress conditions might have additive effects upon the level of UPRE and/or INO1 expression. Indeed, shifting the sec13ts-1 strain to inositol-free medium at the semipermissive temperature of 30° resulted in an increase of ∼30% in UPRE expression over the already elevated levels seen in this strain at 30° in the presence of inositol (Figure 2).

Inositol limitation and secretory stress, thus, affect UPR activation in an additive or synergistic fashion. In Sec cells growing near their restrictive temperatures, the additive stress caused by lack of inositol may, in some cases, exceed the stress tolerance of even those cells with an intact UPR. Provision of inositol under such circumstances would reduce the stress that results in UPR induction, potentially explaining the conditional Ino phenotype of sec13ts-1 cells near their restrictive temperature. Consistent with this idea, transformation of sec13ts-1 with HAC1i permits growth at temperatures up to 32° with or without inositol (Figure 5), whereas the parent sec13ts-1 strain transformed with the vector grows well only up to ∼30° and only if inositol is present (Figure 5).

While inositol limitation and secretory stress had an additive effect upon UPRE expression levels, no correlated additive effect on INO1-CYC-lacZ expression was observed in sec13ts-1 cells grown in the absence of inositol at semipermissive temperatures (Figure 2A). To the contrary, in sec13ts-1 cells shifted to 30° in the absence of inositol, INO1-CYC-lacZ expression was reduced to less than one-half of the level observed in the wild-type control grown under identical conditions (Figure 2C). Similar reductions in INO1-CYC-lacZ expression levels, relative to the wild-type control were observed in all of the Sec mutants assayed after growth in inositol-free media at semipermissive conditions (Figure 2A). The absence of a correlation between UPRE and INO1 expression was even more apparent in sec13ts-1 cells grown at 30° in the presence of inositol (compare Figure 2B and 2C). Under these growth conditions, UPRE expression was greatly elevated in sec13ts-1 cells (Figure 2B). Yet, INO1-lacZ expression in the sec13ts-1 strain at 30° was fully repressed when inositol was present (Figure 2C). Clearly, activation of the UPR during secretory stress does not result in activation of INO1 when inositol is present.

The sec14-3ts mutation, unlike the other Sec mutations analyzed here, does not affect an immediate component of the secretory apparatus itself. Rather, Sec14p is a lipid transfer protein that binds both phosphatidylinositol and phosphatidylcholine (Aitken et al. 1990). sec14 mutations result in a wide range of changes in both lipid metabolism and membrane trafficking (Novick et al. 1980, 1981; Bankaitis et al. 1989; Aitken et al. 1990; Cleves et al. 1991; Kearns et al. 1997; Patton-Vogt et al. 1997; Henry and Patton-Vogt 1998; Sreenivas et al. 1998; Xie et al. 1998). Surprisingly, despite the wide range of Sec mutants that exhibited UPRE activation at their semipermissive temperatures, we did not observe this effect in the sec14-3ts single mutant. The failure to observe UPRE activation in the sec14-3ts strain at the semipermissive temperatures of 30° and 33° was very surprising given that high levels of UPRE activation were previously observed at both 30° and 37° (Chang et al. 2002) in sec14-3ts strains carrying the bypass suppressor, cki1Δ. Inada and Guthrie (2004), however, recently reported active splicing of HAC1 mRNA, indicative of UPR activation, in the sec14-3ts mutant shifted to its restrictive temperature. Thus, it is possible that we failed to observe UPR activation in the sec14-3tsmutant because we used a less direct assay (i.e., a UPRE reporter construct vs. HAC1 mRNA splicing) and different growth conditions (i.e., semipermissive temperature vs. a transient shift to the restrictive condition) than did Inada and Guthrie (2004). At 30° and 37°, the sec14-3ts cki1Δ double mutant exhibits high levels of INO1 expression and inositol prototrophy and, in fact, overproduces and excretes inositol into the growth medium (Patton-Vogt et al. 1997; Chang et al. 2002). In contrast, the sec14-3ts single mutant exhibits inositol auxotrophy and lowered INO1 expression at the semipermissive temperature of 33°. The reduction in INO1 expression in sec14ts-3 under semipermissive growth conditions was similar to that in other Sec strains studied here, suggesting that lowered INO1 expression is a general response to impairment of secretory function. The elevated INO1 expression that we reported previously in the sec14ts-3 cki1Δ strain, and other sec14 strains carrying bypass suppressors affecting the CDP-choline pathway for phosphatidylcholine synthesis, is presumably a consequence of the specific mechanism of suppression (Patton-Vogt et al. 1997; Henry and Patton-Vogt 1998; Chang et al. 2002).

Regardless of the explanation for the general lack of correlation between UPR activation and INO1 expression in Sec mutants, it is clear that INO1 expression is reduced in Sec mutants experiencing secretory stress. The inhibitory effect of secretory stress on transcription of INO1 is similar to the effect of secretory stress on transcription of rRNA and ribosomal protein genes. Mizuta and Warner (1994) showed that the function of the entire secretory pathway is essential for ribosomal synthesis. Nierras and Warner (1999) subsequently demonstrated that rRNA transcription and ribosomal protein synthesis are slowed in cells undergoing secretory stress and that this response is not transduced by the UPR. Rather, they concluded that the effect on rRNA and ribosomal proteins genes in cells with secretory defects may be controlled by a mechanism involving the protein kinase C signal transduction pathway related to membrane stress. In similar fashion, under conditions of secretory stress, signals other than those generated by the UPR may take precedence in controlling expression of INO1 and other UASINO-containing genes. INO1 expression is known to be influenced by growth phase and nutrient availability Griac and Henry (1999) and the glucose response pathway has also been shown to influence the levels of INO1 expression (Ouyang et al. 1999; Shirra and Arndt 1999; Shirra et al. 2001). For these reasons, we are currently examining the relative roles of several other signal transduction pathways in transducing signals from the secretory pathway to the regulatory apparatus controlling INO1 transcription.

The UPR plays a functional role in cells experiencing secretory stress:

The idea that the UPR might play a functional role in exit from the ER and/or be induced by any slowing of ER-specific steps seems quite logical given that the UPR is known to regulate gene expression in response to stress in the ER. Several previous studies have suggested that the UPR might be activated in mutants having defects in specific early steps in membrane trafficking in yeast. For example, Rose et al. (1989) detected an increased level of KAR2 mRNA in the sec18ts-1 mutant, which is defective in vesicle fusion and exit from the ER and Semenza et al. (1990) reported secretion of Kar2p (BiP) from various Sec mutants, including sec22ts-3, sec17ts-1, sec20ts-1, and sec18ts-1, at their permissive temperatures. Belden and Barlowe (2001) demonstrated that the secretion of Kar2p in sec22ts-3 cells, reported by Semenza et al., is associated with strong UPR activation. Since Sec22p may function in retrograde transport of proteins from the Golgi to the ER, Belden and Barlowe (2001) speculated that accumulation of secretory proteins in the ER in sec22ts-3 cells at permissive temperatures could lead to proliferation of the ER, thereby activating the UPR.

The experiments we report here using BFA-treated erg6Δ cells and a wide range of secretory mutants suggest that the UPR can be triggered by secretory stress induced not only by mutations affecting exit from the ER, but also by mutations affecting a number of compartments of the secretory pathway. In the case of mutants such as sec12ts-4, sec13ts-1, sec13ts-4, sec16ts-2, and sec23ts-1, which have defects in COPII vesicle formation, and in erg6Δ cells treated with BFA, UPR activation might be triggered by disruption of the ER organization and accumulation of secretory cargo proteins as suggested by Belden and Barlowe. However, activation of the UPR in mutants defective in vacuolar targeting (pep12Δ) and the Sec6p complex (Figure 1A) indicates that secretory stress induced in compartments distal to the ER is also able to elicit the UPR through Ire1p, a kinase believed to be localized exclusively in the ER (Cox et al. 1993). In a related recent report, Leber et al. (2004) have shown that ER-distal stress boosts HAC1 mRNA abundance.

Travers et al. (2000) demonstrated that activation of the UPR affects expression of genes controlling a broad array of ER and secretory functions including ER-associated protein degradation (ERAD), ER-to-Golgi transport, Golgi-to-ER retrieval, vacuolar targeting, distal secretion, and cell wall biogenesis. Travers et al. also showed that the UPR was activated in ERAD mutants and that ire1Δ and hac1Δ mutants exhibit synthetic lethality with ERAD mutations. Moreover, Caldwell et al. (2001) and Vashist et al. (2001) demonstrated that degradation of soluble substrates by ERAD requires ER-Golgi transport, suggesting a functional relationship involving ERAD and membrane trafficking from the ER.

We have demonstrated similar synthetic lethality involving UPR and secretory mutations. In most cases, Sec Upr double mutants had more severe growth phenotypes than the corresponding Sec parent. These results are consistent with the hypothesis that UPR induction provides protection to cells experiencing stress resulting from partial impairment of membrane trafficking and/or inositol limitation. Also consistent with this hypothesis is the partial suppression (i.e., elevation of the restrictive temperature by several degrees and alleviation of the Ino phenotype) of sec13ts-1 by transformation with HAC1i (Figure 5). Consistent with the results reported here, Higashio and Kohno (2002) and Sato et al. (2002) recently reported partial suppression by transformation with pHAC1i of growth defects of the sec24ts-20 and sec12ts-4 mutants, respectively, which are defective in COPII vesicle formation. Furthermore, processing of was enhanced at the semipermissive temperature in Sec cells retaining an active UPR pathway or carrying HAC1i, as compared to Sec cells carrying a UPR mutation (Figure 6). Thus, the UPR not only is activated under secretory stress, but also plays a role in facilitating membrane trafficking and cell survival under these conditions.

By what mechanism might secretory stress signals generated in membrane compartments, other than the ER, result in UPR activation? One possibility is that disruption of post-Golgi trafficking perturbs the balance of anterograde and retrograde membrane transport pathways between the ER and Golgi by affecting the rate and/or efficiency of protein sorting at the Golgi. It has been suggested that for protein sorting in late-Golgi compartments, anterograde and retrograde transport pathways coexist and compete with each other (Cole et al. 1998). In mammalian cells, it has been shown that Golgi resident proteins continually cycle through the ER (Cole et al. 1998; Miles et al. 2001; Ward et al. 2001). It has been suggested that one function for this recycling is to allow Golgi residents to be periodically surveyed by the protein-folding machinery in the ER where they could be either refolded or degraded through the ERAD pathway (Cole et al. 1998; Ng et al. 2000). When anterograde pathways are blocked or slowed, retrograde pathways predominate, leading to a return of secretory and resident proteins to the ER through normal recycling pathways. The ER might sense the accumulation of these proteins, which, in turn, might account for activation of the UPR. Alternatively, there may be as yet unidentified signaling that occurs between compartments of the secretory pathway serving to coordinate the overall flow of materials between compartments. Clearly, whatever the mechanism, the UPR is activated under secretory stress generated in a number of membrane compartments and its activation provides functional protection, allowing Sec cells to grow at temperatures that are lethal in the absence of a functional UPR (Figure 3).

Acknowledgments

We are indebted to Elizabeth Jones, Randy Schekman, Chris Kaiser, Jeff Brodsky, and Ruth Collins for strains; Kazutoshi Mori for plasmids bearing HAC1 and HAC1i; and Chris Kaiser for the plasmid carrying SEC13. This report is taken in part from the Ph.D. Thesis of H.J.C. (Carnegie Mellon University, 2001). This work was supported by National Institutes of Health grant GM-19629 to S.A.H.

References

  • Aitken, J. F., G. P. vanHeusden, M. Temkin and W. Dowhan, 1990. The gene encoding the phosphatidylinositol transfer protein is essential for cell growth. J. Biol. Chem. 265: 4711–4717. [PubMed] [Google Scholar]
  • Ambroziak, J., and S. A. Henry, 1994. INO2 and INO4 gene products, positive regulators of phospholipid biosynthesis in Saccharomyces cerevisiae, form a complex that binds to the INO1 promoter. J. Biol. Chem. 269: 15344–15349. [PubMed] [Google Scholar]
  • Bankaitis, V. A., D. E. Malehorn, S. D. Emr and R. Greene, 1989. The Saccharomyces cerevisiae SEC14 gene encodes a cytosolic factor that is required for transport of secretory proteins from the yeast Golgi complex. J. Cell Biol. 108: 1271–1281. [PMC free article] [PubMed] [Google Scholar]
  • Becherer, K. A., S. E. Rieder, S. D. Emr and E. W. Jones, 1996. Novel syntaxin homologue, Pep12p, required for the sorting of lumenal hydrolases to the lysosome-like vacuole in yeast. Mol. Cell. Biol. 7: 579–594. [PMC free article] [PubMed] [Google Scholar]
  • Belden, W. J., and C. Barlowe, 2001. Deletion of yeast p24 genes activates the unfolded protein response. Mol. Biol. Cell 12: 957–969. [PMC free article] [PubMed] [Google Scholar]
  • Caldwell, S. R., K. J. Hill and A. A. Cooper, 2001. Degradation of endoplasmic reticulum (ER) quality control substrates requires transport between the ER and Golgi. J. Biol. Chem. 276: 23296–23303. [PubMed] [Google Scholar]
  • Carman, G. M., and S. A. Henry, 1989. Phospholipid biosynthesis in yeast. Annu. Rev. Biochem. 58: 635–669. [PubMed] [Google Scholar]
  • Carman, G. M., and S. A. Henry, 1999. Phospholipid biosynthesis in the yeast Saccharomyces cerevisiae and interrelationship with other metabolic processes. Prog. Lipid Res. 38: 361–399. [PubMed] [Google Scholar]
  • Chang, H. J., E. W. Jones and S. A. Henry, 2002. Role of the unfolded protein response pathway in regulation of INO1 and in the sec14 bypass mechanism in Saccharomyces cerevisiae. Genetics 162: 27–43. [PMC free article] [PubMed] [Google Scholar]
  • Chapman, R. E., and P. Walter, 1997. Translational attenuation mediated by an mRNA intron. Curr. Biol. 7: 850–859. [PubMed] [Google Scholar]
  • Cleves, A. E., T. P. McGee, E. A. Whitters, K. M. Champion, J. R. Aitken et al., 1991. Mutations in the CDP-choline pathway for phospholipid biosynthesis bypass the requirement for an essential phospholipid transfer protein. Cell 64: 789–800. [PMC free article] [PubMed] [Google Scholar]
  • Cole, N. B., J. Ellenberg, J. Song, D. DiEuliis and J. Lippincott-Schwartz, 1998. Retrograde transport of Golgi-localized proteins to the ER. J. Cell Biol. 140: 1–15. [PMC free article] [PubMed] [Google Scholar]
  • Cox, J. S., and P. Walter, 1996. A novel mechanism for regulating activity of a transcription factor that controls the unfolded protein response. Cell 87: 391–404. [PubMed] [Google Scholar]
  • Cox, J. S., C. E. Shamu and P. Walter, 1993. Transcriptional induction of genes encoding endoplasmic reticulum resident proteins requires a transmembrane protein kinase. Cell 73: 1197–1206. [PubMed] [Google Scholar]
  • Cox, J. S., R. E. Chapman and P. Walter, 1997. The unfolded protein response coordinates the production of endoplasmic reticulum protein and endoplasmic reticulum membrane. Mol. Biol. Cell 8: 1805–1814. [PMC free article] [PubMed] [Google Scholar]
  • Dowd, S. R., M. E. Bier and J. L. Patton-Vogt, 2001. Turnover of phosphatidylcholine in Saccharomyces cerevisiae. J. Biol. Chem. 276: 3756–3763. [PubMed] [Google Scholar]
  • Gilstring, C. F., M. Melin-Larsson and P. O. Ljungdahl, 1999. Shr3p mediates specific COPII coatomer-cargo interactions required for the packaging of amino acid permeases into ER-derived transport vesicles. Mol. Biol. Cell 10: 3549–3565. [PMC free article] [PubMed] [Google Scholar]
  • Graham, T. R., P. A. Scott and S. D. Emr, 1993. Brefeldin A reversibly blocks early but not late protein transport steps in the yeast secretory pathway. EMBO J. 12: 869–877. [PMC free article] [PubMed] [Google Scholar]
  • Greenberg, M. L., and J. M. Lopes, 1996. Genetic regulation of phospholipid biosynthesis in Saccharomyces cerevisiae. Microbiol. Rev. 60: 1–20. [PMC free article] [PubMed] [Google Scholar]
  • Griac, P., and S. A. Henry, 1999. The yeast inositol sensitive upstream activating sequence, UASINO, responds to nitrogen availability. Nucleic Acids Res. 27: 2043–2050. [PMC free article] [PubMed] [Google Scholar]
  • Henry, S. A., and J. L. Patton-Vogt, 1998 Genetic regulation of phospholipid metabolism: yeast as a model eukaryote, pp. 133–179 in Progress in Nucleic Acid Research and Molecular Biology, edited by W. E. Cohn and K. Moldave. Academic Press, San Diego. [PubMed]
  • Higashio, H., and K. Kohno, 2002. A genetic link between the unfolded protein response and vesicle formation from the endoplasmic reticulum. Biochem. Biophys. Res. Commun. 296: 568–574. [PubMed] [Google Scholar]
  • Hill, J., K. A. Donal, D. E. Griffiths and G. Donald, 1991. DMSO-enhanced whole cell yeast transformation. Nucleic Acids Res. 19: 5791. [PMC free article] [PubMed] [Google Scholar]
  • Hirsch, J. P., and S. A. Henry, 1986. Expression of the Saccharomyces cerevisiae inositol-1-phosphate synthase (INO1) gene is regulated by factors that affect phospholipid synthesis. Mol. Cell. Biol. 6: 3320–3328. [PMC free article] [PubMed] [Google Scholar]
  • Inada, M., and C. Guthrie, 2004. Identification of Lhp1p-associated RNAs by microarray analysis in Saccharomyces cerevisiae reveals association with coding and noncoding RNAs. Proc. Natl. Acad. Sci. USA 101: 434–439. [PMC free article] [PubMed] [Google Scholar]
  • Kaiser, C. A., R. E. Gimeno and D. A. Shaywitz, 1997 Protein secretion, membrane biogenesis, and endocytosis, pp. 91–227 in Yeast III, edited by J. R. Pringle, J. R. Broach and E. W. Jones. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.
  • Kawahara, T., H. Yanagi, T. Yara and K. Mori, 1997. Endoplasmic reticulum stress-induced mRNA splicing permits synthesis of transcription factor Hac1p/Ern4p that activates the unfolded protein response. Mol. Biol. Cell 8: 1845–1862. [PMC free article] [PubMed] [Google Scholar]
  • Kawahara, T., H. Yanagi, T. Yura and K. Mori, 1998. Unconventional splicing of HAC1/ERN4 mRNA required for the unfolded protein response. Sequence-specific and non-sequential cleavage of the splice sites. J. Biol. Chem. 273: 1802–1807. [PubMed] [Google Scholar]
  • Kearns, B. G., T. P. McGee, P. Mayinger, A. Gedvilaite, S. E. Phillips et al., 1997. Essential role for diacylglycerol in protein transport from the yeast Golgi complex. Nature 387: 101–105. [PMC free article] [PubMed] [Google Scholar]
  • Klausner, R. D., J. G. Donaldson and J. Lippincott-Schwartz, 1992. Brefeldin A: insights into the control of membrane traffic and organelle structure. J. Cell Biol. 116: 1071–1080. [PMC free article] [PubMed] [Google Scholar]
  • Kohno, K., K. Normington, J. Sambrook, M. J. Gething and K. Mori, 1993. The promoter region of the yeast KAR2 (BiP) gene contains a regulatory domain that responds to the presence of unfolded proteins in the endoplasmic reticulum. Mol. Cell. Biol. 13: 877–890. [PMC free article] [PubMed] [Google Scholar]
  • Kreis, T. E., M. Lowe and R. Pepperkok, 1995. COPs regulating membrane traffic. Annu. Rev. Cell Dev. Biol. 11: 677–706. [PubMed] [Google Scholar]
  • Leber, J. H., S. Bernales and P. Walter, 2004. IRE1-independent gain control of the unfolded protein response. PLoS Biol. 2: 1197–1207. [PMC free article] [PubMed] [Google Scholar]
  • Lippincott-Schwartz, J., L. C. Yuan, J. S. Bonifacino and R. D. Klausner, 1989. Rapid redistribution of Golgi proteins into the ER in cells treated with brefeldin A: evidence for membrane cycling from Golgi to ER. Cell 56: 801–813. [PMC free article] [PubMed] [Google Scholar]
  • Loewen, C. J. R., M. L. Gaspar, S. A. Jesch, C. Delon, N. T. Ktistakis et al., 2004. Phospholipid metabolism regulated by a transcription factor sensing phosphatidic acid. Science 304: 1644–1647. [PubMed] [Google Scholar]
  • Lopes, J. M., J. P. Hirsch, P. A. Chorgo, K. L. Schulze and S. A. Henry, 1991. Analysis of sequences in the INO1 promoter that are involved in its regulation by phospholipid precursors. Nucleic Acids Res. 19: 1687–1693. [PMC free article] [PubMed] [Google Scholar]
  • Marykwas, D. L., and T. D. Fox, 1989. Control of the Saccharomyces cerevisiae regulatory gene PET494: transcriptional repression by glucose and translational induction by oxygen. Mol. Cell. Biol. 9: 484–491. [PMC free article] [PubMed] [Google Scholar]
  • Miles, S., H. McManus, K. E. Forsten and B. Storrie, 2001. Evidence that the entire Golgi apparatus cycles in interphase HeLa cells: sensitivity of Golgi matrix proteins to an ER exit block. J. Cell Biol. 155: 543–555. [PMC free article] [PubMed] [Google Scholar]
  • Mizuta, K., and J. R. Warner, 1994. Continued functioning of the secretory pathway is essential for ribosome synthesis. Mol. Cell. Biol. 14: 2493–2502. [PMC free article] [PubMed] [Google Scholar]
  • Mori, K., A. Sant, K. Kohno, K. Normington, M. J. Gething et al., 1992. A 22 bp cis-acting element is necessary and sufficient for the induction of the yeast KAR2 (BiP) gene by unfolded proteins. EMBO J. 11: 2583–2593. [PMC free article] [PubMed] [Google Scholar]
  • Mori, K., W. Ma, M. J. Gething and J. F. Sambrook, 1993. A transmembrane protein with a cdc2+/CDC28-related kinase activity is required for signaling from the ER to the nucleus. Cell 74: 743–756. [PubMed] [Google Scholar]
  • Mori, K., N. Ogawa, T. Kawahara, H. Yanagi and T. Yura, 2000. mRNA splicing-mediated C-terminal replacement of transcription factor Hac1p is required for efficient activation of the unfolded protein response. Proc. Natl. Acad. Sci. USA 97: 4660–4665. [PMC free article] [PubMed] [Google Scholar]
  • Ng, D. T., E. D. Spear and P. Walter, 2000. The unfolded protein response regulates multiple aspects of secretory and membrane protein biogenesis and endoplasmic reticulum quality control. J. Cell Biol. 150: 77–88. [PMC free article] [PubMed] [Google Scholar]
  • Nierras, C. R., and J. R. Warner, 1999. Protein kinase C enables the regulatory circuit that connects membrane synthesis to ribosome synthesis in Saccharomyces cerevisiae. J. Biol. Chem. 274: 13235–13241. [PubMed] [Google Scholar]
  • Nikawa, J. I., and S. Yamashita, 1992. IRE1 encodes a putative protein kinase containing a membrane-spanning domain and is required for inositol phototrophy in Saccharomyces cerevisiae. Mol. Microbiol. 6: 1441–1446. [PubMed] [Google Scholar]
  • Nikawa, J.-I., M. Akiyoshi, S. Hirata and T. Fukuda, 1996. Saccharomyces cerevisiae IRE2/HAC1 is involved in IRE1-mediated KAR2 expression. Nucleic Acids Res. 24: 4222–4226. [PMC free article] [PubMed] [Google Scholar]
  • Novick, P., C. Field and R. Schekman, 1980. Identification of 23 complementation groups required for post-translational events in the yeast secretory pathway. Cell 21: 205–215. [PubMed] [Google Scholar]
  • Novick, P., S. Ferro and R. Schekman, 1981. Order of events in the yeast secretory pathway. Cell 25: 461–469. [PubMed] [Google Scholar]
  • Ouyang, Q., M. Ruiz-Noriega and S. A. Henry, 1999. The REG1 gene product is required for repression of INO1 and other inositol-sensitive upstream activating sequence-containing genes of yeast. Genetics 152: 89–100. [PMC free article] [PubMed] [Google Scholar]
  • Patton-Vogt, J. L., P. Griac, A. Sreenivas, V. Bruno, S. Dowd et al., 1997. Role of the yeast phosphatidylinositol/phosphatidylcholine transfer protein (Sec14p) in phosphatidylcholine turnover and INO1 regulation. J. Biol. Chem. 272: 20873–20883. [PubMed] [Google Scholar]
  • Rose, M. D., L. M. Misra and J. P. Vogel, 1989. KAR2, a karyogamy gene, is the yeast homolog of the mammalian BiP/GRP78 gene. Cell 57: 1211–1221. [PubMed] [Google Scholar]
  • Rose, M. D., F. Winston and P. Hieter, 1990 Methods in Yeast Genetics. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.
  • Sato, M., K. Sato and A. Nakano, 2002. Evidence for the intimate relationship between vesicle budding from the ER and the unfolded protein response. Biochem. Biophys. Res. Commun. 296: 560–567. [PubMed] [Google Scholar]
  • Schwank, S., R. Ebbert, K. Rautenstrauss, E. Schweizer and H. Schuller, 1995. Yeast transcriptional activator INO2 interacts as an Ino2p/Ino4p basic helix-loop-helix heteromeric complex with the inositol/choline-responsive element necessary for expression of phospholipid biosynthetic genes in Saccharomyces cerevisiae. Nucleic Acids Res. 23: 230–237. [PMC free article] [PubMed] [Google Scholar]
  • Sciaky, N., J. Presley, C. Smith, K. J. Zaal, N. Cole et al., 1997. Golgi tubule traffic and the effects of brefeldin A visualized in living cells. J. Cell Biol. 139: 1137–1155. [PMC free article] [PubMed] [Google Scholar]
  • Seemann, J., E. Jokitalo, M. Pypaert and G. Warren, 2000. Matrix proteins can generate the higher order architecture of the Golgi apparatus. Nature 407: 1022–1026. [PubMed] [Google Scholar]
  • Semenza, J. C., K. G. Hardwick, N. Dean and H. R. Pelham, 1990. ERD2, a yeast gene required for the receptor-mediated retrieval of luminal ER proteins from the secretory pathway. Cell 61: 1349–1357. [PubMed] [Google Scholar]
  • Shah, N., and R. D. Klausner, 1993. Brefeldin A reversibly inhibits secretion in Saccharomyces cerevisiae. J. Biol. Chem. 268: 5345–5348. [PubMed] [Google Scholar]
  • Shamu, C. E., and P. Walter, 1996. Oligomerization and phosphorylation of the Ire1p kinase during intracellular signaling from the endoplasmic reticulum to the nucleus. EMBO J. 15: 3028–3039. [PMC free article] [PubMed] [Google Scholar]
  • Sherman, F., G. R. Fink and C. W. Lawrence, 1978 Methods in Yeast Genetics. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.
  • Shirra, M. K., and K. M. Arndt, 1999. Evidence for the involvement of the Glc7-Reg1 phosphatase and the Snf1-Snf4 kinase in the regulation of INO1 transcription in Saccharomyces cerevisiae. Genetics 152: 73–87. [PMC free article] [PubMed] [Google Scholar]
  • Shirra, M. K., J. Patton-Vogt, A. Ulrich, O. Liuta-Tehlivets, S. D. Kohlwein et al., 2001. Inhibition of acetyl coenzyme A carboxylase activity restores expression of the INO1 gene in a snf1 mutant strain of Saccharomyces cerevisiae. Mol. Cell. Biol. 17: 5710–5722. [PMC free article] [PubMed] [Google Scholar]
  • Sidrauski, C., J. S. Cox and P. Walter, 1996. tRNA ligase is required for regulated mRNA splicing in the unfolded protein response. Cell 87: 405–413. [PubMed] [Google Scholar]
  • Sreenivas, A., J. L. Patton-Vogt, V. Bruno, P. Griac and S. A. Henry, 1998. A role for phospholipase D (Pld1p) in growth, secretion, and regulation of membrane lipid synthesis in yeast. J. Biol. Chem. 273: 16635–16638. [PubMed] [Google Scholar]
  • Stevens, T., B. Esmon and R. Schekman, 1982. Early stages in the yeast secretory pathway are required for transport of carboxypeptidase Y to the vacuole. Cell 30: 439–448. [PubMed] [Google Scholar]
  • Travers, K. J., C. K. Patil, L. Wodicka, D. J. Lockhart, J. S. Weissman et al., 2000. Functional and genomic analyses reveal an essential coordination between the unfolded protein response and ER-associated degradation. Cell 101: 249–258. [PubMed] [Google Scholar]
  • Vashist, S., W. Kim, W. J. Belden, E. D. Spear, C. Barlowe et al., 2001. Distinct retrieval and retention mechanisms are required for the quality control of endophasmic reticulum protein folding. J. Cell Biol. 155: 355–367. [PMC free article] [PubMed] [Google Scholar]
  • Vogel, J. P., J. N. Lee, D. R. Kirsch, M. D. Rose and E. S. Sztul, 1993. Brefeldin A causes a defect in secretion in Saccharomyces cerevisiae. J. Biol. Chem. 268: 3040–3043. [PubMed] [Google Scholar]
  • Wach, A., A. Brachat, R. Pohlmann and P. Philippsen, 1994. New heterologous modules for classical or PCR-based gene disruptions in Saccharomyces cerevisiae. Yeast 10: 1793–1808. [PubMed] [Google Scholar]
  • Ward, T. H., R. S. Polishchuk, S. Caplan, K. Hirschberg and J. Lippincott-Schwartz, 2001. Maintenance of Golgi structure and function depends on the integrity of ER export. J. Cell Biol. 155: 557–570. [PMC free article] [PubMed] [Google Scholar]
  • Webb, G. C., J. Zhang, S. J. Garlow, A. Wesp, H. Riezman et al., 1997. Pep7p provides a novel protein that functions in vesicle-mediated transport between the yeast Golgi and endosome. Mol. Biol. Cell 8: 871–895. [PMC free article] [PubMed] [Google Scholar]
  • White, M. J., J. P. Hirsch and S. A. Henry, 1991. The OPI1 gene of Saccharomyces cerevisiae, a negative regulator of phospholipid biosynthesis, encodes a protein containing polyglutamine tracts and a leucine zipper. J. Biol. Chem. 266: 863–872. [PubMed] [Google Scholar]
  • Xie, Z., M. Fang, M. Rivas, A. J. Faulkner, P. Sterweis et al., 1998. Phospholipase D activity is required for suppression of yeast phosphatidylinositol transfer protein defects. Proc. Natl. Acad. Sci. USA 95: 12346–12351. [PMC free article] [PubMed] [Google Scholar]

Articles from Genetics are provided here courtesy of Oxford University Press

-