Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Cell Stress Chaperones. 2010 Sep; 15(5): 497–508.
Published online 2009 Dec 15. doi: 10.1007/s12192-009-0163-4
PMCID: PMC3006623
PMID: 20013084

IRE1α controls cyclin A1 expression and promotes cell proliferation through XBP-1

Associated Data

Supplementary Materials

Abstract

IRE1 is a conserved dual endoribonuclease/protein kinase that is indispensable for directing the endoplasmic reticulum (ER) stress response in yeast, flies, and worms. In mammalian systems, however, the precise biological activities carried out by IRE1α are unclear. Here, molecular and chemical genetic approaches were used to control IRE1 activity in a number of prostate cancer cell lines and the resulting impact on gene transcription, cell survival, and proliferation was examined. Modulating IRE1α activity had no transcriptional effect on the induction of genes classically associated with the ER stress response (Grp78 and CHOP) or cell survival when confronted with ER stress agents. Rather, IRE1α activity was positively correlated to proliferation. Since Xbp-1 mRNA is the sole known substrate for IRE1 endoribonuclease activity, the role of this transcription factor in mediating proliferation was examined. Repressing total Xbp-1 levels by siRNA techniques effectively slowed proliferation. In an effort to identify IRE1/XBP-1 targets responsible for the cell cycle response, genome-wide differential mRNA expression analysis was performed. Consistent with its ability to sense ER stress, IRE1α induction led to an enrichment of ER-Golgi, plasma membrane, and secretory gene products. An increase in cyclin A1 expression was the only differentially expressed cell cycle regulatory gene found. Greater cyclin A protein levels were consistently observed in cells with active IRE1α and were dependent on XBP-1. We conclude that IRE1α activity controls a subset of the ER stress response and mediates proliferation through tight control of Xbp-1 splicing.

Electronic supplementary material

The online version of this article (doi:10.1007/s12192-009-0163-4) contains supplementary material, which is available to authorized users.

Keywords: ER stress, Proliferation, Cyclin A, IRE1, XBP-1

Introduction

The endoplasmic reticulum (ER) possesses a unique capacity to sense and respond to various stress stimuli. The ER stress response (ESR) can be triggered by the accumulation of excessive or mis-folded protein, deprivation of glucose, calcium fluctuations, and numerous other cellular imbalances (Kaufman 1999; Mori 2000; Schroder and Kaufman 2005). The ESR attempts to restore homeostasis through a set of three primary mechanisms, together termed the unfolded protein response (UPR). First, the UPR can call for the production of additional chaperone proteins to accelerate protein folding (Rutkowski and Kaufman 2004). Second, activating proteasome-mediated degradation through a reverse-translocation process, referred to as ER-associated degradation, serves to reduce the protein load (Hampton 2002; Jarosch et al. 2003; Plemper and Wolf 1999). The third strategy is to reduce protein translation (Brostrom and Brostrom 1998; Harding et al. 2002). These UPR activities normally succeed in restoring ER homeostasis and ensure cellular survival. However, if the stress burden becomes great enough, then cell death can result (Breckenridge et al. 2003; Ferri and Kroemer 2001; Rao et al. 2004).

The genetic elements controlling the ESR have been best characterized in lower eukaryotes and, more recently, the ER-specific pathways responsible for metazoan stress responses have been investigated. In Saccharomyces cerevisiae, Drosophila melanogaster, and Caenorhabditis elegans, the ESR is mediated exclusively by IRE1 (Cox et al. 1993; Hollien and Weissman 2006; Mori et al. 1993; Shen et al. 2005), an ER-resident transmembrane dual endoribonuclease/kinase. The amino-terminal luminal region of IRE1 is responsible for sensing various ER stress stimuli. When properly induced, this signal causes dimerization, a subsequent conformational shift, autophosphorylation, and activation of the endoribonuclease activity (Shamu and Walter 1996; Welihinda and Kaufman 1996). The mRNA of the transcription factor Xbp-1 is the only known substrate for mammalian IRE1 activity (Niwa et al. 2005; Yoshida et al. 2001). Xbp-1 is processed through a unique unconventional endoribonuclease activity to remove 26 nucleotides within the coding region. This cleavage generates an alternative reading frame to be utilized. Translation of unspliced Xbp-1 results in a protein (XBP-1U) that contains a DNA-binding domain, while translation of spliced Xbp-1 (XBP-1S) generates a C-terminal transcriptional activation domain coded in the alternate reading frame adjacent to the DNA-binding region. XBP-1S has been shown to activate transcription (Yoshida et al. 2001). Although XBP-1U does not possess a functional transcriptional activation domain, this form has been proposed to regulate some biological activities (Tirosh et al. 2006; Yoshida et al. 2006).

An additional level of complexity arises in metazoan systems, as two other ER stress-sensing proteins are present that include the kinase PERK and transcription factor ATF6. Conflicting studies exist regarding how these players interact to regulate the mammalian ESR. A recent study found that IRE1 could modulate cell survival induced by the ER stress agents thapsigargin and tunicamycin (Lin et al. 2007). In contrast, using ATF6α null mouse embryonic fibroblasts (MEFs), ATF6α transcriptional induction was required for survival and for the expression of ER chaperones when confronted with ER stress (Wu et al. 2007; Yamamoto et al. 2007). Here, we engineered human epithelial cells to express human IRE1α constructs that either interfere with or induce the endoribonuclease activity in an effort to understand the role of IRE1 activity in mediating various facets of the ESR. We could find no role for this gene in mediating cell survival or initiating the ESR, but found that IRE1 and XBP-1 could regulate cell proliferation. Furthermore, by activating IRE1 it was possible to reproduce a physiologic change in Xbp-1 splicing and globally assess the resulting differentially expressed genes using human exon microarrays. These results identified cyclin A1 as an IRE1-dependent target. Together, these data suggest that IRE1 alone does not initiate a full ESR or confer survival in response to ER stress stimuli, but rather controls cyclin A levels and a subset of ER-resident gene products responsible for protein chaperone activity, detoxification, lipid synthesis, transcription, and anti-viral defenses.

Materials and methods

Cell culture

Human prostate cancer cell lines (LNCaP, DU145, and PC-3) and the Phoenix Amphotrophic retroviral packaging cell line were obtained from ATCC (Manassas, VA, USA). Prostate cancer cells were grown in low glucose (1.0 g/L) Dulbecco's modified Eagle's medium (DMEM) supplemented with 5% fetal bovine serum (FBS). Phoenix cells were grown in DMEM containing non-essential amino acids and 10% FBS. Velcade (Millenium Pharmaceuticals, Cambridge, MA, USA) was provided by the University of Kentucky Markey Cancer Center Pharmacy. Tunicamycin and thapsigargin were purchased from Sigma (St. Louis, MO, USA). MG132 and 1NM-PP1 were from EMD Biosciences (San Diego, CA, USA).

Cloning, mutagenesis, and mammalian protein expression

IRE1α (GenBank accession number NM_001433) was polymerase chain reaction (PCR) amplified from normal human prostate cDNA. Site-directed point mutations to IRE1α (K599A and I642G) were created by an overlapping PCR-based method and verified by sequence analysis. Stable cell lines were created by retroviral transduction as described (Schwarze et al. 2003).

Viability and proliferation assays

Cell viability was determined with cells (70,000/well) plated in a 24-well dish using a standard colorimetric MTT assay as described (Thorpe et al. 2008). Cell proliferation was determined by 5-Bromo-2-deoxy-uridine (BrdU) labeling. Cells were plated (500,000) in 6-cm dishes. The following day, the media were changed to contain 2.5% serum. BrdU labeling and flow cytometric analysis was carried out as previously described (Schwarze et al. 2003). The percentage of BrdU positive cells (10,000 gated events) and the cell fraction in G0/1 and G2/M determined using CellQuest software (B-D). All assays were carried out in replicates of three. All statistical assessments were made using a Student’s t test.

Xbp-1 splicing analysis

Total RNA was purified using the RNeasy purification kit (Qiagen). cDNA was generated with SuperScript II Reverse Transcriptase (Invitrogen). A PCR-based method was employed to quantify the status of Xbp-1 splicing (Shang 2005). The primers were designed to span the 26-nucleotide long excision characteristic of Xbp-1S. The PCR products were resolved by electrophoresis either through a 2% agarose gel in a Tris–acetate–EDTA (TAE) buffer, or through a 4% polyacrylamide gel in TAE. DNA was visualized by ethidium bromide staining and a digital imaged captured (Ultra-Violet Products, Upland, CA, USA). The percent of unspliced and spliced Xbp-1 transcript was quantified using ImageJ software.

siRNA-mediated Xbp-1 repression

An XBP-1 and non-targeting SMARTpool siRNA were purchased (Dharmacon, Lafayette, CO, USA). Cells were transfected with DharmaFECT (Dharmacon) in media containing 10% serum as described by the manufacturer. The next day, cells were fed with DMEM containing 10% serum or harvested to examine mRNA levels.

Real-time quantitative PCR

Total RNA was isolated using the RNeasy RNA isolation kit and cDNA was prepared. Quantitative RT-PCR was performed by monitoring the SYBR Green fluorescence using the 7300 Real-Time PCR System (Applied Biosystems, Foster City, CA, USA) as described (Morrison et al. 1998; Wittwer et al. 1997). Values were normalized to the relative amounts of 18S or cyclophilin B RNA. Primer sequences are available upon request.

Western blot analysis

Cell lysates were prepared and Western blot analysis carried out as described (Peterson et al. 2006). Antibodies were obtained from the following sources: IRE1a, 14C10, and cyclin A, BF683 (Cell Signaling Technology); GRP78, H-129 (Santa Cruz Biotechnology); GAPDH (Mab374), Millipore; and ß-actin, AC-74 (Sigma).

Microarray analysis

PC-3 cells (2.5 × 106) overexpressing wild-type IRE1 or I642G IRE1 were plated in quadruplicate and treated for 12 h with 5 µM 1NM-PP1. Total mRNA was isolated (Qiagen) and the mRNA labeled according to the manufacturer’s directions (GeneChip® Whole Transcript (WT) Sense Target Labeling Assay, Affymetrix, Santa Clara, CA, USA). Labeled mRNA was hybridized to GeneChip® Human Exon 1.0 ST Arrays (Affymetrix). The University of Kentucky Microarray Core performed the labeling and hybridization. Statistical analysis of variance (ANOVA) between groups was carried out with Partek Genomic Suites 6.4 Software (St. Louis, MO, USA).

Results

ESR induction in prostate epithelial cells

Our overriding goal was to examine the effect of IRE1 activity on the ER stress response using prostate epithelial cells as a model. First, the competency of several prostate cancer cell lines to engage the ER stress machinery was tested. LNCaP, DU145, and PC-3 cells were exposed to proteasome inhibitors (MG132 and Velcade), thapsigargin, and tunicamycin at concentrations known to induce the ESR. Cells were then examined for GRP78 induction, a hallmark of cells responding to ER stress, by Western blot analysis. All cell lines consistently increased steady-state GRP78 upon treatment with each agent (Fig. 1a). Cells were also examined for the induction of IRE1 activity by measuring the status of Xbp-1 splicing. All cell lines demonstrated a dose-dependent increase in Xbp-1S with Velcade and thapsigargin administration (Fig. 1b). These studies indicate that the LNCaP, DU145, and PC-3 cell lines possess the ability to mount an ESR.

An external file that holds a picture, illustration, etc.
Object name is 12192_2009_163_Fig1_HTML.jpg

Induction of the ER stress response in human prostate epithelial cells. a DU145, LNCaP, and PC-3 cells were exposed to MG132 (1 µM) for 24 h or tunicamycin (2 µg/ml) or thapsigargin (300 nM) for the indicated times and Western blot analysis was performed to detect GRP78 and ß-actin. b Cell lines were exposed to thapsigargin (300 nM) or the indicated concentrations of Velcade for 24 h. Total RNA was isolated and cDNA examined by RT-PCR and agarose gel electrophoresis to discern the Xbp-1 splicing status. u unspliced, s spliced

Blocking IRE1 activity impairs proliferation

The introduction of a mutated (K599A) kinase-defective IRE1α (IRE1) construct (Tirasophon et al. 2000) has been frequently used as a tool to diminish IRE1 activity and to block Xbp-1 splicing (Drogat et al. 2007; Hu et al. 2006; Kaneko et al. 2002). Here, K599A IRE1 was stably introduced into the PC-3 cell line by retroviral-mediated gene transduction. A polyclonal population was selected and found to readily overexpress the K599A IRE1 protein (Fig. 2a) and impair Xbp-1 splicing in response to tunicamycin and thapsigargin (Fig. 2b). First, the ability of the kinase-defective IRE1 to modulate the expression of two UPR-dependent hallmark genes upon ER stress was examined. Although Velcade and thapsigargin elevated CHOP and Grp78 expression, no significant change between the control PC-3 cells and the kinase-defective IRE1 was noted (Fig. 2c). The ability of IRE1-defective cells to modulate survival in response to ER stress agents was also studied. PC-3 cells were treated with Velcade or thapsigargin and cell viability was assayed. Both agents reduced viability, but no difference between PC-3 puro and the K599A IRE1 cells was noted (Fig. 2d). As a control, the effect of TRAIL, a cell death ligand that acts through a distinct receptor-mediated process, was also tested. Once again, no difference in cell viability between the two cell lines could be demonstrated (Fig. 2d).

An external file that holds a picture, illustration, etc.
Object name is 12192_2009_163_Fig2_HTML.jpg

Reduction in IRE1 activity slows cells proliferation. a PC-3 cells stably expressing kinase-defective IRE1 (K599A) or the control pBabe puro (Puro) vector were examined by Western blot analysis for the indicated proteins. b PC-3 puro and K599A IRE1 expressing cells were exposed to tunicamycin (2 µg/ml) and thapsigargin (300 nM) for 24 h. Cells were analyzed for the abundance of unspliced (u) and spliced (s) Xbp-1. cCHOP and Grp78 mRNA expression was determined in the PC-3 cells assayed in b by qPCR following standardization to cyclophilin B or 18S rRNA. 18S standardization was used only for thapsigargin (Tg) as cyclophilin B was induced by this treatment. The average of three experiments ± the standard deviation is shown. d PC-3 cells (Puro and K599A IRE1) were treated for 24 h with the indicated concentrations of thapsigargin (left), Velcade (middle), or TRAIL (right) and viability was determined by a colorimetric MTT assay. e BrdU labeling was carried out on PC-3 (left) and LNCaP (right) cells expressing the K599A IRE1 construct or vector only (Puro). The average of three replicates ± the standard deviation is shown. *P = 0.002; **P = 0.01

Although no effect on the aforementioned ER stress-related parameters was noted, it was readily apparent that the K599A IRE1 overexpressing cells were dividing slower. Therefore, cell proliferation was assessed by quantifying BrdU incorporation into DNA. PC-3 cells expressing the kinase-defective IRE1 were found to proliferate 36% slower than the PC-3 puro control cells (Fig. 2e). LNCaP cells were also engineered to stably express K599A IRE1 (Supplementary material S1a) and also were observed to have a decrease in the rate of proliferation (Fig. 2e). We attempted to express K599A IRE1 in DU145 cells, but only obtained a twofold overexpression and noted no effect on Xbp-1 splicing or BrdU labeling (Supplementary material S1b). These data suggest IRE1 activity, and possibly basal XBP-1S expression, may contribute to normal cell growth.

IRE1 activity enhances Xbp-1 splicing and proliferation

IRE1 is a unique protein kinase as it changes conformation and becomes active upon nucleotide binding (Papa et al. 2003). By generating a point mutation (I642G) in the nucleotide binding cleft it is possible to induce the endoribonuclease activity with the addition of the bulky kinase inhibitor 1NM-PP1 (Lin et al. 2007). The ‘inducible’ I642G IRE1 construct was stably expressed in prostate cancer cells to examine the effect of IRE1 activation on cell survival, UPR gene induction, and proliferation. PC-3 cells readily overexpressed the I642G IRE1 construct and the addition of 1NM-PP1 conferred Xbp-1 splicing (Fig. 3a), indicating robust endoribonuclease activity. The I642G IRE1 overexpressing cells were tested for survival in response to ER stress stimuli. PC-3 I642G IRE1 and the paired control PC-3 puro cell populations were treated with Velcade and thapsigargin in the presence or absence of 1NM-PP1. Velcade and thapsiargin treatment led to a reduction in cell viability; however, no difference between the I642G IRE1 expressing and control PC-3 puro cells was observed (Fig. 3b). The ability of IRE1 to regulate the expression of ER stress hallmark genes was also studied. IRE1 activation was unable to induce CHOP and Grp78 mRNA levels (Fig. 3c), further suggesting that their expression is IRE1 independent. DU145 cells overexpressing the I642G IRE1 construct (Fig. 3d) also showed no difference in cell survival (Fig. 3e).

An external file that holds a picture, illustration, etc.
Object name is 12192_2009_163_Fig3_HTML.jpg

IRE1 I642G confers endoribonuclease activity and induces proliferation. a (top) Control (Puro) and PC-3 cells stably expressing the I642G IRE1 construct were subjected to Western blot analysis to detect IRE1 expression. (bottom) Control (Puro) and PC-3 cells stably expressing the I642G IRE1 construct were treated with DMSO, 1NM-PP1 (5 µM), or thapsigargin (Tg, 300 nM) for 24 h and Xbp-1 splicing examined. b PC-3 cells (Puro and IRE1-I642G) were treated with Velcade and thapsigargin in the presence or absence of 1NM-PP1 (5 µM) for 24 h and cell viability quantified with an MTT assay. Values depict the average ± the standard deviation. c Expression of CHOP and Grp78 mRNA in PC-3 cells (Puro and IRE1-I642G) treated with DMSO or 1NM-PP1 (5 µM) for 24 h as determined by quantitative real-time PCR. Values were standardized and compared to Puro cells treated with DMSO only. Error bars reflect the standard deviation. d (top) Control (Puro) and DU145 cells stably expressing the I642G IRE1 construct were subjected to Western blot analysis to detect IRE1 expression. (bottom) Puro and DU145 cells stably expressing I642G IRE1 were treated with DMSO, 1NM-PP1 (5 µM), or thapsigargin (Tg, 300 nM) for 24 h and Xbp-1 splicing examined. (e) DU145 cells (Puro and I642G IRE1) were treated and the results depicted as described in b. f BrdU labeling was carried out on PC-3 (left) and DU145 (right) cells expressing the I642G IRE1 construct or vector only (Puro). 1NM-PP1 (5 µM) was added to the PC-3 cells for 24 h. BrdU labeling in the DU145 cells as a function of serum concentration was conducted in the absence of 1NM-PP1. The average of three replicates ± the standard deviation is shown. gXbp-1 splicing status in PC-3 cells stably expressing the IRE1 K599A, or I642G, or empty pBabe Puro construct. PCR reactions were analyzed by polyacrylamide gel electrophoresis. Note, a reduction in sXbp-1 levels in the K599A cells and an induction of the spliced Xbp-1 message in the I642 cells compared to the Puro control

Since the kinase-defective IRE1 construct reduced cell proliferation, we hypothesized that the ‘inducible’ I642G IRE1 would have the opposite effect. The proliferation of cells expressing I642G IRE1 was tested in the presence and absence of 1NM-PP1. Surprisingly, the I642G IRE1 overexpressing PC-3 cells proliferated significantly faster even in the absence of the ATP analog (Fig. 3f). DU145 cells overexpressing I642G IRE1 in the absence of 1NM-PP1 also were proliferating faster regardless of the serum concentration (Fig. 3f). LNCaP cells stably overexpressing I642G IRE1 further confirmed these findings (Supplementary material S2).

Dominant negative and active IRE1 modulate basal endoribonuclease activity

The ability of the K599A and I642G IRE1 constructs to regulate proliferation suggested that IRE1 was active in the absence of exogenous ER stress stimuli. To detect endogenous IRE1 activity, Xbp-1 splicing PCR reactions were electrophoresed through 4% polyacrylamide gels instead of agarose gels. This method allowed the relatively low percentage of Xbp-1S (compared to Xbp-1U) to be resolved and revealed that the K599A IRE1 construct reduced basal Xbp-1 splicing, while the I642G mutation yielded a significantly higher basal Xbp-1S level compared to the control cell population (Fig. 3g). Therefore, IRE1 activity does exist in cell culture in the absence of ER stress agents, the I642G IRE1 construct is somewhat leaky, and the resulting Xbp-1S levels positively correlate with the proliferation rate.

Effect of wild-type IRE1 overexpression on proliferation

It remained possible that IRE1 overexpression contributed, at least in part, to the observed proliferative changes. Furthermore, Western blot examination of PC-3, DU145, and LNCaP cells revealed an increase in steady-state IRE1 protein levels with the addition of ER stress-inducing agents (Fig. 4a), suggesting a potential feedback mechanism may exist that controls biological activities mediated by IRE1. Therefore, cell populations overexpressing wild-type IRE1 were generated (Fig. 4b, e). The wild-type IRE1 overexpressing DU145 cells showed a minor impairment of basal Xbp-1 splicing (Fig. 4c) with a concomitant (15%) reduction in BrdU labeling (Fig. 4d). In contrast, IRE1 overexpression in PC-3 cells did not impair basal Xbp-1 splicing or proliferation (Fig. 4f, g). These data suggest that the ability of wild-type IRE1 overexpression to alter the endoribonuclease activity is minor and cell-type specific. Furthermore, the impairment of Xbp-1 splicing in the DU145 cells with a reduction in proliferation remains consistent with ability of IRE1 activity to regulate proliferation.

An external file that holds a picture, illustration, etc.
Object name is 12192_2009_163_Fig4_HTML.jpg

Effect of wild-type IRE1 overexpression on Xbp-1 splicing and proliferation. a Western blot analysis was carried out on the three cell lines stably expressing the indicated constructs to detect IRE1 and ß-actin expression. Cells were treated for 24 h with 1NM-PP1 (5 µM), Velcade (100 nM) or thapsigargin (Tg, 300 nM). b DU145 cells stably expressing wild-type IRE1 were examined by Western blot analysis to confirm protein overexpression. c DU145 cells overexpressing wild-type IRE1 were exposed to Velcade for 24 h and the Xbp-1 splicing status revealed by agarose gel electrophoresis (left) and quantified by NIH ImageJ software (right). d BrdU incorporation analysis was carried out on DU145 cells overexpressing wild-type IRE1 and the Puro control cells. Values reflect the average ± the standard deviation. e PC-3 cells stably overexpressing wild-type IRE1 were generated and protein overexpression analyzed by Western blot analysis. f PC-3 cells overexpressing the IRE1 constructs and control cells were exposed to Velcade for 4 h and the Xbp-1 splicing status revealed by agarose gel electrophoresis (left) and quantified by NIH ImageJ software (right). g BrdU incorporation analysis was carried out on control PC-3 cells (Puro) and those overexpressing wild-type IRE1. Values reflect the average ± the standard deviation

XBP-1 can regulate cell proliferation

Since Xbp-1 is the only known ribonuclease substrate for IRE1, this transcription factor was examined for its ability to regulate proliferation. First, XBP-1 expression was targeted by siRNA transfection. DU145 cells were chosen as a cell model to carry out these studies, as they were consistently the most sensitive to changes in IRE1 activity. DU145 cells were transfected with a siRNA pool to target total (spliced and unspliced) Xbp-1. This reagent reduced Xbp-1 expression by 77% (Fig. 5a). Cell proliferation examined 48-h post-transfection showed that targeting Xbp-1 reduced proliferation by 23% in the control Puro cells, but was less effective (14% decrease) with I642G IRE1 overexpression (Fig. 5b).

An external file that holds a picture, illustration, etc.
Object name is 12192_2009_163_Fig5_HTML.jpg

Regulation of cell proliferation by spliced XBP-1. a Percent Xbp-1 mRNA following siRNA-mediated knockdown in DU145 cells compared to a non-targeting control. b DU145 cells were transfected with siRNA to target Xbp-1 or a scrambled control. After 48 h, cells were analyzed by BrdU labeling. Values reflect the average ± the standard deviation. *P = 0.0002, **P = 0.0003

Genome-wide mRNA expression analysis upon IRE1 induction

To identify IRE1/XBP-1-dependent target genes associated with proliferation and other facets of the ESR, I642G IRE1 overexpressing PC-3 cells were compared to wild-type IRE1 overexpressing cells following 12 h of 1NM-PP1 treatment. The comparison between wild-type and the inducible I642G was utilized to obtain the greatest difference in Xbp-1 splicing (K599A IRE1 expressing cells spliced Xbp-1 following 1NM-PP1 addition; Supplementary material S3a). The 12-h time point was selected to provide an early assessment of XBP-1-dependent gene changes based on DNAJB9 levels (Supplementary material S3b). Hybridizations to Affymetrix Exon Arrays were carried out. After ANOVA analyses, 105 genes were found to be upregulated and 82 downregulated at P < 0.01 following IRE1 activation. The entire datasets are available online. Those genes with expression changes greater than 1.5-fold are shown (Tables 1 and and2).2). The gene ontology demonstrated a strong enrichment of ER/Golgi, plasma membrane, and secreted gene products (Fig. 6a). Of note, IRE1-dependent cyclin A1 induction was the only cell cycle regulatory gene significantly altered. Quantitative RT-PCR analysis confirmed this finding in separate samples (Fig. 6b). To further test if cyclin A1 expression could be induced by a second IRE1 stimulus, cells were treated with thapsigargin up to 24 h. XBP-1 splicing was observed upon thapsigargin treatment and preceded cyclin A1 induction (Fig. 6b). Cyclin A protein expression was subsequently probed in the wild-type and I642G IRE1 PC-3 cells in the presence and absence of 1NM-PP1. Consistent with the proliferative response, cyclin A protein expression was elevated in the I642G cell line, even in the absence of 1NM-PP1 (Fig. 6c). Cyclin A was also examined in cells expressing the K599A IRE1 construct; however, no difference in protein expression was observed in comparison to wild-type IRE1 (Fig. 6d). Finally, the dependence of XBP-1 expression on cyclin A levels was tested in the DU145 cell line. Consistent with the proliferative response (Fig. 5b), I642G expressing cells demonstrated higher cyclin A levels over control cells (Fig. 6e). Furthermore, Xbp-1 repression abrogated the cyclin A induction in both control and I642G IRE1 cell lines.

Table 1

Categorical list of genes downregulated greater than 1.5-fold as determined by microarray analysis in I642G IRE1 PC-3 cells compared to wild-type IRE1 PC-3 cells in the presence of 1NM-PP1

SymbolGenbank accessionGene nameCellular componentFold changeP value
Enzymatic
DGAT2NM_032564Acyl CoA diacylglycerol acetyl transferaseER−1.9<0.001
SPTLC3NM_018327Serine palmitoyl transferaseGolgi mem−1.50.03
Signaling
TAS2R49NM_176889Taste receptor type 2, 49Plasma mem−2.40.002
LGR5NM_003667GPR49Plasma mem−1.80.001
OR51L1NM_001004755Olfactory receptor, family 51, subfamily L, 1Plasma mem−1.7<0.001
CDH6NM_004932K-cadherinPlasma mem−1.6<0.001
ODZ1NM_014253Odd Oz/ten-m homolog 1Plasma mem−1.6<0.001
SPARCNM_003118OsteonectinSecreted−1.5<0.001
Uncategorized
SCN3ANM_006922Voltage gated Na + channel, type IIIaPlasma mem−2.1<0.001
SERPINB2NM_002575Plasminogen activator inhibitor 2Secreted−1.70.002
SYT11NM_152280SynaptotagminPlasma mem−1.5<0.001
PROS1NM_000313Protein SSecreted−1.6<0.001

Mem membrane

Table 2

Categorical list of genes upregulated greater than 1.5-fold as determined by microarray analysis in I642G IRE1 PC-3 cells compared to wild-type IRE1 PC-3 cells in the presence of 1NM-PP1

SymbolGenbank accessionGene nameCellular componentFold changeP value
Adhesion
CLDN7NM_001307Claudin 7Plasma mem2.7<0.001
MPZL2NM_144765Myelin protein zero-like 2Plasma mem2.6<0.001
CDH1NM_004360E-cadherinPlasma mem2.2<0.001
ITGB6NM_000888Integrin ß6Plasma mem2.2<0.001
LAD1NM_005558Ladinin-1Secreted2.0<0.001
CNTN1NM_001843Contactin 1Plasma mem1.9<0.001
TNS4NM_032865Tensin 4Cytoplasm1.60.004
Cytoskeleton
MYO5BNM_001080467Myosin VßCytoskeleton1.9<0.001
ACTG2NM_001615Actin, gamma 2Cytoskeleton1.50.009
Signaling
TACSTD1NM_002354Tumor-associated calcium signal transducer 1Plasma mem1.8<0.001
TSPAN2NM_005725Tetraspanin 2Plasma mem1.50.008
Transcription factor
GRHL2BC069633Grainyhead-like 2Nucleus2.7<0.001
HOXB9NM_024017Homeobox B9Nucleus1.6<0.001
CREB3L2NM_194071cAMP responsive element binding protein 3-like 2ER/Nucleus1.50.004
Transport/carrier
AP1M2NM_005498Adaptor related protein, complex 1, mu 2Golgi2.5<0.001
CALB1NM_004929CalbindinCytosol1.9<0.001
TCN1NM_001062Transcobalamin ISecreted1.70.003
SLC9A2NM_003048Sodium/hydrogen exchanger 2Membrane1.50.002
Enzymatic
CYP1B1NM_000104Cytochrome P450 B1ER3.7<0.001
BCAT1NM_005504Branched chain aminotransferase 1Cytosol3.3<0.001
LPCAT2NM_017839Lysophosphatidylcholine acyltransferase 2ER/Golgi2.6<0.001
ST14NM_021978Membrane-type serine protease 1Plasma mem2.4<0.001
TMPRSS2NM_005656Transmembrane protease, serine 2Plasma mem1.50.003
TINAGL1BC064633Tubulointerstitial nephritis antigen-like 1Secreted1.5<0.001
Interferon inducible
IFITM1NM_003641Interferon-induced transmembrane protein 1Plasma mem2.00.006
IFI6NM_002038Interferon, α-inducible protein 6Membrane1.80.002
MX1NM_002462Myxovirus resistance 1Cytoplasm1.8<0.001
IRF6NM_006147Interferon regulatory factor 6Nucleus1.6<0.001
OAS2NM_0025352'-5'-oligoadenylate synthetase 2ER/mem1.50.008
Uncategorized
CCNA1NM_003914Cyclin A1Nucleus2.10.007
DNAJB9NM_012328Heat shock protein 40ER1.70.006

Mem membrane

An external file that holds a picture, illustration, etc.
Object name is 12192_2009_163_Fig6_HTML.jpg

Identification of cyclin A1 as an IRE1/XBP-1-dependent gene. a Gene ontology demonstrating the localization of genes found to be differentially expressed in IRE1 I642G PC-3 cells compared to IRE1 wild-type cells. b Quantitative PCR assessment of cyclin A1 gene expression in (left) I642G IRE1 PC-3 cells compared to wild-type IRE1 and (middle) in PC-3 cells following thapsigargin treatment compared to DMSO treated cells. Error bars reflect the fold change ± the standard deviation. (right) The Xbp-1 splicing status of the thapsigargin treated cells was assessed by PCR. c Cyclin A protein expression in wild-type and I642G IRE1 PC-3 cells as determined by Western blot analysis in the presence or absence of 1NM-PP1 as indicated. The numbers reflect the fold change in cyclin A expression compared to wild-type IRE1. d Cyclin A protein expression in wild-type, K599A, and I642G IRE1 PC-3 cells as determined by Western blot analysis in the absence of 1NM-PP1. The numbers reflect the fold change in cyclin A expression compared to wild-type IRE1. e DU145 cells expressing I642G IRE1 and Puro control cells were transfected for 48 h with a scrambled negative (SN) or XBP-1 siRNA and examined for cyclin A expression by Western blot analysis. The numbers reflect the fold change in cyclin A expression compared to the SN transfected PC-3 puro cells

Discussion

Two recent studies defining the sensory machinery responsible for mediating the ER stress response in mammalian cells have provided varying accounts on the role of IRE1 (Lin et al. 2007; Yamamoto et al. 2007). Here, evidence that IRE1 activity directs cell survival or ESR hallmark genes was not found. Instead, IRE1 activity was identified as a cell proliferation regulator. In every circumstance in which basal IRE1 activity (in the absence of ER stress agents) was impaired, cell proliferation was slowed; while elevating the basal IRE1 activity was able to stimulate cell proliferation in all three cell lines examined. Using a high-resolution polyacrylamide gel electrophoresis method, we discovered that the I642G IRE1 mutation possessed enhanced Xbp-1 splicing activity even in the absence of 1NM-PP1, while the K599A IRE1 mutation blocked the basal Xbp-1 splicing. Therefore, IRE1 endoribonuclease activity is present in the absence of ER stress stimuli and can promote cell proliferation. Furthermore, the use of siRNA that targeted total Xbp-1 provided a downstream mechanistic link to the basal IRE1 activity, as those cells with reduced Xbp-1 levels were impaired in their ability to proliferate. These data are consistent with a model in which IRE1 controls an arm of the ER stress response and highlight a biological function for IRE1 activity in cell proliferation control.

The independence of IRE1 activity on the expression of hallmark ESR genes in human cells was confirmed in this study. Recently, it has been demonstrated that the transcriptional induction of Grp78 and Grp94 was significantly impaired in ATF6α-deficient MEFs (Wu et al. 2007; Yamamoto et al. 2007). It appears, therefore, that ATF6 is necessary to drive the expression of these genes in the ESR. However, we cannot definitely exclude IRE1/XBP-1 involvement in their transcriptional control. Based on recent findings (Yamamoto et al. 2007), which showed dimerization between XBP-1 and ATF6, it is possible that if both IRE1 and ATF6 become activated simultaneously, then XBP-1 may participate in the transcription of additional genes unable to be modulated by IRE1 activation alone. As IRE1 and ATF6 are normally simultaneously induced upon the ESR, transcription co-regulation remains possible.

Gene expression array methodology was employed to provide insight into the mechanisms behind both cell cycle control and other IRE1-dependent functions. This screen was designed to identify genes regulated by both the basal induction of IRE1 activity and those with high IRE1 activity similar to that observed under an ESR. Cyclin A1 was the only gene differentially expressed with a direct role in cell division. Several pieces of additional data support a role for cyclin A in mediating proliferation by IRE1. First, elevated cyclin A protein expression was observed in both the PC-3 and DU145 I642G IRE1 cell lines. Second, Xbp-1 silencing repressed cyclin A levels and correlated with the decreased BrdU labeling index. Third, active IRE1 was associated with a decrease in the percent of cells in the G1 and G2 phase and Xbp-1 targeting increased G1 and G2 content. This cell cycle distribution is consistent with the characterized involvement of cyclin A in both G1/S and G2/M progression (Yang et al. 1999). Although it is impossible to know the function of several non-annotated transcripts and short-sighted to exclude the possibility of annotated genes in cell cycle control, cyclin A1 induction provides one mechanism for the observed proliferation by the IRE1/XBP-1 pathway. Importantly, we were able to use a second means of IRE1 activation, that of thapsigargin treatment, to induce cyclin A expression. Clearly, thapsigargin is a well-documented cytotoxic agent and has a negative effect on cell growth. However, if a situation would arise, in tumor cells for example, that could activate IRE1 exclusively, then it may be possible to activate cell growth through this pathway.

Consistent with IRE1 acting as an ER sensory molecule, examination of the gene products differentially expressed upon IRE1 activation revealed a strong enrichment for those that enter the ER and reside in this location, localize to the plasma membrane, or are destined for secretion. The specific gene identities revealed several biological functions that could be considered in agreement with mediating a stress response. The gene most significantly upregulated is the ER-resident detoxifying monooxygenase cytochrome P450 B1. CYP1B1 metabolizes procarcinogens, such as polycyclic aromatic hydrocarbons, and steroids (Pottenger and Jefcoate 1990). One of the highest activities is toward the substrate 17ß-estradiol, which is converted to the 4-hydroxyestradiol (Lee et al. 2003a, b). The induction of the chaperone DNAJB9 was observed in the microarray and confirmed the qPCR data. DNAJB9 has been reported to be driven by XBP-1S expression (Lee et al. 2003a, b), thus providing a link to the enhanced protein chaperone activity in the ESR. One entire class upregulated by IRE1 includes several genes classically induced by interferons. It is well established that numerous viruses including hepatitis, herpesvirus, and flavivirus induce an ESR (Bechill et al. 2006; Mulvey et al. 2007; Yu et al. 2006). Several of these studies have shown IRE1/XBP-1 is the sensor that recognizes the infection and participates in the biological response. Finally, a decrease in two rate-limiting enzymes involved in lipid synthesis, DGAT2 and SPTLC3, was observed. Pathological excesses of nutrients are known to stimulate an ESR (Gregor and Hotamisligil 2007) and mice haploinsufficient for Xbp-1 have an increase in adipose tissue compared to wild-type mice (Ozcan et al. 2004). Since DGAT2 is the primary mechanism by which triglycerides are formed and is essential for adipose formation (Bluher et al. 2002), it is possible that the IRE1/XBP-1 pathway prevents fat deposition though transcriptional repression of DGAT2. SPTLC3 is the rate-limiting step in de novo sphingolipid synthesis. A relationship between sphingolipid levels and the ESR has not been reported to date.

The contribution of ER stress pathways to tumorigenesis has become an intense area of interest. The extent to which the IRE1/XBP-1 pathway is modulated in human cancer is only beginning to emerge. Xbp-1 splicing has been observed both in hepatocellular carcinoma and in breast cancer (Lacroix and Leclercq 2004; Shuda et al. 2003). The mechanism behind the presumed induction of IRE1 activity is not clear, but a reduction in oxygen tension can increase Xbp-1 splicing (Koong et al. 2006). Although one recent report has linked IRE1 activity and pro-angiogenic processes (Drogat et al. 2007), few studies have explored how IRE1 activation may lead to tumor promotion. The microarray study carried out here revealed a number of changes that could be pro or anti-cancer. For example, the downregulation of proto-oncogenic K-cadherin and the induction of E-cadherin and several other genes with cell adhesion functions could point to a tumor or metastasis prevention function. However, the induction of a gene such as CYP1B1 could generate additional DNA damage. Functional studies will ultimately need to be performed to discern how each arm of the ESR alone and in combination modifies tumorigenesis.

Electronic supplementary materials

Below is the link to the electronic supplementary material.

Supplementary Fig. 1(60K, jpg)

Comparison of the…..(JPEG 60.2 kb)

Supplementary Fig. 2(35K, jpg)

Comparison of the…..(JPEG 34.6 kb)

Supplementary Fig. 3(41K, jpg)

Comparison of the…..(JPEG 40.7 kb)

Acknowledgements

This work was supported by the New York Academy of Medicine (SRS).

References

  • Bechill J, Chen Z, Brewer JW, Baker SC. Mouse hepatitis virus infection activates the Ire1/XBP1 pathway of the unfolded protein response. Adv Exp Med Biol. 2006;581:139–144. doi: 10.1007/978-0-387-33012-9_24. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Bluher M, Michael MD, Peroni OD, Ueki K, Carter N, Kahn BB, Kahn CR. Adipose tissue selective insulin receptor knockout protects against obesity and obesity-related glucose intolerance. Dev Cell. 2002;3:25–38. doi: 10.1016/S1534-5807(02)00199-5. [PubMed] [CrossRef] [Google Scholar]
  • Breckenridge DG, Germain M, Mathai JP, Nguyen M, Shore GC. Regulation of apoptosis by endoplasmic reticulum pathways. Oncogene. 2003;22:8608–8618. doi: 10.1038/sj.onc.1207108. [PubMed] [CrossRef] [Google Scholar]
  • Brostrom CO, Brostrom MA. Regulation of translational initiation during cellular responses to stress. Prog Nucleic Acid Res Mol Biol. 1998;58:79–125. doi: 10.1016/S0079-6603(08)60034-3. [PubMed] [CrossRef] [Google Scholar]
  • Cox JS, Shamu CE, Walter P. Transcriptional induction of genes encoding endoplasmic reticulum resident proteins requires a transmembrane protein kinase. Cell. 1993;73:1197–1206. doi: 10.1016/0092-8674(93)90648-A. [PubMed] [CrossRef] [Google Scholar]
  • Drogat B, Auguste P, Nguyen DT, Bouchecareilh M, Pineau R, Nalbantoglu J, Kaufman RJ, Chevet E, Bikfalvi A, Moenner M. IRE1 signaling is essential for ischemia-induced vascular endothelial growth factor-A expression and contributes to angiogenesis and tumor growth in vivo. Cancer Res. 2007;67:6700–6707. doi: 10.1158/0008-5472.CAN-06-3235. [PubMed] [CrossRef] [Google Scholar]
  • Ferri KF, Kroemer G. Organelle-specific initiation of cell death pathways. Nat Cell Biol. 2001;3:E255–E263. doi: 10.1038/ncb1101-e255. [PubMed] [CrossRef] [Google Scholar]
  • Gregor MG, Hotamisligil GS. Adipocyte stress: the endoplasmic reticulum and metabolic disease. J Lipid Res. 2007;48:1905–1914. doi: 10.1194/jlr.R700007-JLR200. [PubMed] [CrossRef] [Google Scholar]
  • Hampton RY. ER-associated degradation in protein quality control and cellular regulation. Curr Opin Cell Biol. 2002;14:476–482. doi: 10.1016/S0955-0674(02)00358-7. [PubMed] [CrossRef] [Google Scholar]
  • Harding HP, Calfon M, Urano F, Novoa I, Ron D. Transcriptional and translational control in the mammalian unfolded protein response. Annu Rev Cell Dev Biol. 2002;18:575–599. doi: 10.1146/annurev.cellbio.18.011402.160624. [PubMed] [CrossRef] [Google Scholar]
  • Hollien J, Weissman JS. Decay of endoplasmic reticulum-localized mRNAs during the unfolded protein response. Science. 2006;313:104–107. doi: 10.1126/science.1129631. [PubMed] [CrossRef] [Google Scholar]
  • Hu P, Han Z, Couvillon AD, Kaufman RJ, Exton JH. Autocrine tumor necrosis factor alpha links endoplasmic reticulum stress to the membrane death receptor pathway through IRE1alpha-mediated NF-kappaB activation and down-regulation of TRAF2 expression. Mol Cell Biol. 2006;26:3071–3084. doi: 10.1128/MCB.26.8.3071-3084.2006. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Jarosch E, Lenk U, Sommer T. Endoplasmic reticulum-associated protein degradation. Int Rev Cytol. 2003;223:39–81. doi: 10.1016/S0074-7696(05)23002-4. [PubMed] [CrossRef] [Google Scholar]
  • Kaneko M, Ishiguro M, Niinuma Y, Uesugi M, Nomura Y. Human HRD1 protects against ER stress-induced apoptosis through ER-associated degradation. FEBS Lett. 2002;532:147–152. doi: 10.1016/S0014-5793(02)03660-8. [PubMed] [CrossRef] [Google Scholar]
  • Kaufman RJ. Stress signaling from the lumen of the endoplasmic reticulum: coordination of gene transcriptional and translational controls. Genes Dev. 1999;13:1211–1233. doi: 10.1101/gad.13.10.1211. [PubMed] [CrossRef] [Google Scholar]
  • Koong AC, Chauhan V, Romero-Ramirez L. Targeting XBP-1 as a novel anti-cancer strategy. Cancer Biol Ther. 2006;5:756–759. [PubMed] [Google Scholar]
  • Lacroix M, Leclercq G. About GATA3, HNF3A, and XBP1, three genes co-expressed with the oestrogen receptor-alpha gene (ESR1) in breast cancer. Mol Cell Endocrinol. 2004;219:1–7. doi: 10.1016/j.mce.2004.02.021. [PubMed] [CrossRef] [Google Scholar]
  • Lee AH, Iwakoshi NN, Glimcher LH. XBP-1 regulates a subset of endoplasmic reticulum resident chaperone genes in the unfolded protein response. Mol Cell Biol. 2003;23:7448–7459. doi: 10.1128/MCB.23.21.7448-7459.2003. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Lee AJ, Cai MX, Thomas PE, Conney AH, Zhu BT. Characterization of the oxidative metabolites of 17beta-estradiol and estrone formed by 15 selectively expressed human cytochrome p450 isoforms. Endocrinology. 2003;144:3382–3398. doi: 10.1210/en.2003-0192. [PubMed] [CrossRef] [Google Scholar]
  • Lin JH, Li H, Yasumura D, Cohen HR, Zhang C, Panning B, Shokat KM, Lavail MM, Walter P. IRE1 signaling affects cell fate during the unfolded protein response. Science. 2007;318:944–949. doi: 10.1126/science.1146361. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Mori K. Tripartite management of unfolded proteins in the endoplasmic reticulum. Cell. 2000;101:451–454. doi: 10.1016/S0092-8674(00)80855-7. [PubMed] [CrossRef] [Google Scholar]
  • Mori K, Ma W, Gething MJ, Sambrook J. A transmembrane protein with a cdc2+/CDC28-related kinase activity is required for signaling from the ER to the nucleus. Cell. 1993;74:743–756. doi: 10.1016/0092-8674(93)90521-Q. [PubMed] [CrossRef] [Google Scholar]
  • Morrison TB, Weis JJ, Wittwer CT. Quantification of low-copy transcripts by continuous SYBR Green I monitoring during amplification. Biotechniques. 1998;24:954–958. [PubMed] [Google Scholar]
  • Mulvey M, Arias C, Mohr I. Maintenance of endoplasmic reticulum (ER) homeostasis in herpes simplex virus type 1-infected cells through the association of a viral glycoprotein with PERK, a cellular ER stress sensor. J Virol. 2007;81:3377–3390. doi: 10.1128/JVI.02191-06. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Niwa M, Patil CK, DeRisi J, Walter P. Genome-scale approaches for discovering novel nonconventional splicing substrates of the Ire1 nuclease. Genome Biol. 2005;6:R3. doi: 10.1186/gb-2004-6-1-r3. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Ozcan U, Cao Q, Yilmaz E, Lee AH, Iwakoshi NN, Ozdelen E, Tuncman G, Gorgun C, Glimcher LH, Hotamisligil GS. Endoplasmic reticulum stress links obesity, insulin action, and type 2 diabetes. Science. 2004;306:457–461. doi: 10.1126/science.1103160. [PubMed] [CrossRef] [Google Scholar]
  • Papa FR, Zhang C, Shokat K, Walter P. Bypassing a kinase activity with an ATP-competitive drug. Science. 2003;302:1533–1537. doi: 10.1126/science.1090031. [PubMed] [CrossRef] [Google Scholar]
  • Peterson FC, Thorpe JA, Harder AG, Volkman BF, Schwarze SR. Structural determinants involved in the regulation of CXCL14/BRAK expression by the 26 S proteasome. J Mol Biol. 2006;363:813–822. doi: 10.1016/j.jmb.2006.08.057. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Plemper RK, Wolf DH. Retrograde protein translocation: ERADication of secretory proteins in health and disease. Trends Biochem Sci. 1999;24:266–270. doi: 10.1016/S0968-0004(99)01420-6. [PubMed] [CrossRef] [Google Scholar]
  • Pottenger LH, Jefcoate CR. Characterization of a novel cytochrome P450 from the transformable cell line, C3H/10T1/2. Carcinogenesis. 1990;11:321–327. doi: 10.1093/carcin/11.2.321. [PubMed] [CrossRef] [Google Scholar]
  • Rao RV, Ellerby HM, Bredesen DE. Coupling endoplasmic reticulum stress to the cell death program. Cell Death Differ. 2004;11:372–380. doi: 10.1038/sj.cdd.4401378. [PubMed] [CrossRef] [Google Scholar]
  • Rutkowski DT, Kaufman RJ. A trip to the ER: coping with stress. Trends Cell Biol. 2004;14:20–28. doi: 10.1016/j.tcb.2003.11.001. [PubMed] [CrossRef] [Google Scholar]
  • Schroder M, Kaufman RJ. The mammalian unfolded protein response. Annu Rev Biochem. 2005;74:739–789. doi: 10.1146/annurev.biochem.73.011303.074134. [PubMed] [CrossRef] [Google Scholar]
  • Schwarze SR, Fu VX, Jarrard DF. Cdc37 enhances proliferation and is necessary for normal human prostate epithelial cell survival. Cancer Res. 2003;63:4614–4619. [PubMed] [Google Scholar]
  • Shamu CE, Walter P. Oligomerization and phosphorylation of the Ire1p kinase during intracellular signaling from the endoplasmic reticulum to the nucleus. Embo J. 1996;15:3028–3039. [PMC free article] [PubMed] [Google Scholar]
  • Shang J. Quantitative measurement of events in the mammalian unfolded protein response. Methods. 2005;35:390–394. doi: 10.1016/j.ymeth.2004.10.012. [PubMed] [CrossRef] [Google Scholar]
  • Shen X, Ellis RE, Sakaki K, Kaufman RJ. Genetic interactions due to constitutive and inducible gene regulation mediated by the unfolded protein response in C. elegans. PLoS Genet. 2005;1:e37. doi: 10.1371/journal.pgen.0010037. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Shuda M, Kondoh N, Imazeki N, Tanaka K, Okada T, Mori K, Hada A, Arai M, Wakatsuki T, Matsubara O, Yamamoto N, Yamamoto M. Activation of the ATF6, XBP1 and grp78 genes in human hepatocellular carcinoma: a possible involvement of the ER stress pathway in hepatocarcinogenesis. J Hepatol. 2003;38:605–614. doi: 10.1016/S0168-8278(03)00029-1. [PubMed] [CrossRef] [Google Scholar]
  • Thorpe JA, Christian PA, Schwarze SR. Proteasome inhibition blocks caspase-8 degradation and sensitizes prostate cancer cells to death receptor-mediated apoptosis. Prostate. 2008;68:200–209. doi: 10.1002/pros.20706. [PubMed] [CrossRef] [Google Scholar]
  • Tirasophon W, Lee K, Callaghan B, Welihinda A, Kaufman RJ. The endoribonuclease activity of mammalian IRE1 autoregulates its mRNA and is required for the unfolded protein response. Genes Dev. 2000;14:2725–2736. doi: 10.1101/gad.839400. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Tirosh B, Iwakoshi NN, Glimcher LH, Ploegh HL. Rapid turnover of unspliced Xbp-1 as a factor that modulates the unfolded protein response. J Biol Chem. 2006;281:5852–5860. doi: 10.1074/jbc.M509061200. [PubMed] [CrossRef] [Google Scholar]
  • Welihinda AA, Kaufman RJ. The unfolded protein response pathway in Saccharomyces cerevisiae. Oligomerization and trans-phosphorylation of Ire1p (Ern1p) are required for kinase activation. J Biol Chem. 1996;271:18181–18187. doi: 10.1074/jbc.271.30.18181. [PubMed] [CrossRef] [Google Scholar]
  • Wittwer CT, Herrmann MG, Moss AA, Rasmussen RP. Continuous fluorescence monitoring of rapid cycle DNA amplification. Biotechniques. 1997;22(130–131):134–138. [PubMed] [Google Scholar]
  • Wu J, Rutkowski DT, Dubois M, Swathirajan J, Saunders T, Wang J, Song B, Yau GD, Kaufman RJ. ATF6alpha optimizes long-term endoplasmic reticulum function to protect cells from chronic stress. Dev Cell. 2007;13:351–364. doi: 10.1016/j.devcel.2007.07.005. [PubMed] [CrossRef] [Google Scholar]
  • Yamamoto K, Sato T, Matsui T, Sato M, Okada T, Yoshida H, Harada A, Mori K. Transcriptional induction of mammalian ER quality control proteins is mediated by single or combined action of ATF6alpha and XBP1. Dev Cell. 2007;13:365–376. doi: 10.1016/j.devcel.2007.07.018. [PubMed] [CrossRef] [Google Scholar]
  • Yang R, Muller C, Huynh V, Fung YK, Yee AS, Koeffler HP. Functions of cyclin A1 in the cell cycle and its interactions with transcription factor E2F–1 and the Rb family of proteins. Mol Cell Biol. 1999;19:2400–2407. [PMC free article] [PubMed] [Google Scholar]
  • Yoshida H, Matsui T, Yamamoto A, Okada T, Mori K. XBP1 mRNA is induced by ATF6 and spliced by IRE1 in response to ER stress to produce a highly active transcription factor. Cell. 2001;107:881–891. doi: 10.1016/S0092-8674(01)00611-0. [PubMed] [CrossRef] [Google Scholar]
  • Yoshida H, Oku M, Suzuki M, Mori K. pXBP1(U) encoded in XBP1 pre-mRNA negatively regulates unfolded protein response activator pXBP1(S) in mammalian ER stress response. J Cell Biol. 2006;172:565–575. doi: 10.1083/jcb.200508145. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Yu CY, Hsu YW, Liao CL, Lin YL. Flavivirus infection activates the XBP1 pathway of the unfolded protein response to cope with endoplasmic reticulum stress. J Virol. 2006;80:11868–11880. doi: 10.1128/JVI.00879-06. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Articles from Cell Stress & Chaperones are provided here courtesy of Elsevier

-