Different perturbations alter the function of the endoplasmic reticulum (ER), resulting in the accumulation of misfolded proteins in its lumen, a condition termed ER stress. To restore ER proteostasis, a highly conserved pathway is engaged, known as the unfolded protein response (UPR), triggering adaptive programs or apoptosis of terminally damaged cells. IRE1α (also known as ERN1), the most conserved UPR sensor, mediates the activation of responses to determine cell fate under ER stress. The complexity of IRE1α regulation and its signaling outputs is mediated in part by the assembly of a dynamic multi-protein complex, named the UPRosome, that regulates IRE1α activity and the crosstalk with other pathways. We discuss several studies identifying components of the UPRosome that have illuminated novel functions in cell death, autophagy, DNA damage, energy metabolism and cytoskeleton dynamics. Here, we provide a theoretical analysis to assess the biological significance of the UPRosome and present the results of a systematic bioinformatics analysis of the available IRE1α interactome data sets followed by functional enrichment clustering. This in silico approach decoded that IRE1α also interacts with proteins involved in the cell cycle, transport, differentiation, response to viral infection and immune response. Thus, defining the spectrum of IRE1α-binding partners will reveal novel signaling outputs and the relevance of the pathway to human diseases.

The endoplasmic reticulum (ER) is a highly dynamic and complex membranous network, responsible for a variety of crucial cellular functions, including protein synthesis and folding, and intracellular Ca2+ storage (Schwarz and Blower, 2016). A complex network of chaperones, foldases and cofactors, in addition to specific ionic and redox requirements, tightly control protein folding and quality within the ER lumen (Dubnikov et al., 2017). However, a significant amount of newly synthetized proteins do not reach their proper folding state and are delivered to the proteasome by the ER-associated degradation (ERAD) machinery (Hwang and Qi, 2018). Altered ER function can lead to the abnormal accumulation of unfolded or misfolded proteins, a condition known as ‘ER stress’ (Walter and Ron, 2011). ER stress triggers a series of adaptive mechanisms collectively known as the unfolded protein response (UPR) (Hetz, 2012; Walter and Ron, 2011). UPR signaling results in transcriptional and translational responses in order to increase the protein folding capacity of the cell and restore proteostasis (Oakes and Papa, 2015). If the UPR is unable to cope with protein misfolding stress, the pathway activates self-destruction programs to eliminate damaged cells by apoptosis (Urra et al., 2013).

Abnormal levels of ER stress are implicated in a variety of human diseases, including cancer, metabolic disorders, inflammation and neurodegenerative diseases (Wang and Kaufman, 2016). However, novel biological functions of the UPR beyond its traditional role in protein homeostasis are currently emerging. The UPR consists of three arms and the most conserved branch is initiated by the stress sensor inositol-requiring enzyme 1α (IRE1α; also known as ERN1) (Walter and Ron, 2011). The activation status of IRE1α signaling is regulated by the assembly of a multiprotein platform at the ER membrane, which we have previously termed the UPRosome (Hetz and Glimcher, 2009). The UPRosome also controls the crosstalk between the UPR and other stress pathways through the binding of adapter and signaling proteins and, in addition, might mediate non-canonical functions of the UPR (Hetz et al., 2020).

Here, we review all available protein–protein interaction studies to discuss emerging roles of the UPR in the control of cell function in addition to highlight novel regulatory aspects of IRE1α. We also present a new global analysis of available interactome data sets to speculate about possible novel functions of IRE1α in normal physiology and disease.

Under normal conditions, specialized secretory cells (i.e. pancreatic β-cells, B cell lymphocytes and salivary glands) require an active UPR to cope with the high demand for folded proteins, which generates abnormal levels of misfolded or unfolded intermediates. In addition, a number of conditions, such as hypoxia, nutrient deprivation, mutations in secretory cargoes and loss of Ca2+, redox or lipid homeostasis, can also result in altered ER protein homeostasis or ‘proteostasis’ (Walter and Ron, 2011). In the past 20 years, chronic ER stress and overactivation of the UPR have been proposed as a relevant contributor to the development of several diseases, including cancer, diabetes, neurodegeneration and inflammatory disorders, among others (Wang and Kaufman, 2016). Activation of the UPR reprograms the transcription of hundreds of genes involved in different aspects of the secretory pathway including the translocation of proteins into the ER, protein folding, glycosylation, redox metabolism, protein quality control, translation, ERAD and lipid biogenesis, among others (Hetz, 2012). If chronic ER stress results in irreversible cellular damage, UPR signaling switches from adaptive to pro-apoptotic programs through the engagement of several cell death mechanisms (Tabas and Ron, 2011; Urra et al., 2013).

The UPR is initiated by three types of ER transmembrane proteins that act as ER stress sensors and transducers, including IRE1α and IRE1β (ERN2), activating transcription factor 6 (ATF6; also known as ATF6α) and ATF6β (also known as ATF6B), and protein kinase RNA (PKR)-like ER kinase (PERK; also known as EIF2AK3) (Box 1). These sensors act in concert to regulate downstream transcription factors that engage adaptive or pro-apoptotic programs depending of the extent of cellular damage (Box 1).

Box 1. The unfolded protein response

The UPR is controlled by three ER transmembrane proteins named IRE1α, PERK and ATF6 (see figure). PERK is a type I protein kinase, which under ER stress phosphorylates the eukaryotic translation initiator factor-2 (eIF2α) at serine 51, resulting in the inhibition of general protein translation (Harding et al., 2000b). In addition, this event simultaneously triggers the selective translation of mRNAs that contain inhibitory upstream open reading frames (uORFs) within their 5′-untranslated region (UTR) such as the activating transcription factor 4 (ATF4) (Vattem and Wek, 2004). ATF4 controls the expression of genes involved in the antioxidant response, folding capacity, amino acid metabolism and autophagy (Harding et al., 2000a). Upon chronic ER stress, ATF4 upregulates the transcription factor CCAAT-enhancer-binding protein homologous protein (CHOP; also known as GADD153), resulting in the activation of pro-apoptotic programs (Tabas and Ron, 2011; Urra et al., 2013). For instance, CHOP induces the expression of the phosphatase subunit (GADD34 or PP1C) that results in eIF2α de-phosphorylation, thereby restoring protein synthesis and resulting in oxidative stress and proteotoxicity (Han et al., 2013; Marciniak et al., 2004). ATF6 is an ER transmembrane protein that contains a bZIP transcription factor in its cytosolic domain. Upon ER stress, ATF6 is cleaved in the Golgi apparatus by the site-1 and site-2 proteases (S1P and S2P), resulting in the release of its cytosolic region (ATF6f) (Haze et al., 1999). ATF6f then acts as a potent transcription factor, regulating many UPR target genes that are related to ERAD and protein folding (Adachi et al., 2008; Yamamoto et al., 2007).

The most conserved pathway of the UPR is initiated by the serine/threonine protein kinase IRE1α. Under ER stress, IRE1α dimerizes and auto-transphosphorylates, favoring a conformational change that activates its endoribonuclease domain (Karagöz et al., 2019). Active IRE1α catalyzes the excision of a 26-nucleotide intron in the mRNA encoding the transcriptional factor X-box binding protein-1 (XBP1) (Walter and Ron, 2011), which then is ligated by the tRNA ligase RTCB (Jurkin et al., 2014; Kosmaczewski et al., 2014; Lu et al., 2014). This processing event leads to a shift in the open reading frame of the XBP1 mRNA, generating a new C-terminal domain sequence (Calfon et al., 2002; Lee et al., 2002; Yoshida et al., 2001). The resulting spliced XBP1 protein (XBP1s) acts as a potent transcription factor that controls the upregulation of several UPR genes involved in protein folding, protein entry into the ER and ERAD (Hetz et al., 2011). In addition, the RNase domain of IRE1α can mediate the direct cleavage of multiple RNAs, in a process known as regulated IRE1α-dependent decay (RIDD) (Hollien et al., 2009; Hollien and Weissman, 2006). RIDD involves the cleavage of conserved RNA sequences that contain a specific secondary structure similar to the hairpins present in the mRNA of XBP1 (Maurel et al., 2014). The binding of IRE1α to the adaptor protein TNFR-associated factor 2 (TRAF2) can engage the activation of other signaling pathways, such as MAPK pathways, regulating cell death and autophagy (Nishitoh et al., 2002; Urano et al., 2000) (see figure in Box 1). Active IRE1α molecules can be found as dimers and oligomers, which might determine the regulation of XBP1 mRNA splicing and RIDD (Bouchecareilh et al., 2011; Ghosh et al., 2014; Wang et al., 2012). Nevertheless, the molecular mechanism underlying the selectivity for mRNA substrates is still controversial and poorly understood. In summary, the UPR represents a network of signaling pathways that orchestrate adaptive responses to ER stress on a dynamic and regulated manner.

For almost 20 years, it has been assumed that the chaperone BiP (also known as Grp78 and HSPA5) is the main mediator involved in the activation of all three UPR stress transducers (Bertolotti et al., 2000; Shen et al., 2002). However, accumulating evidence suggests that additional regulators and modes of activation exist. In yeast, Kar2 (a homolog of BiP) interacts with the luminal domain of IRE1p, keeping it in a monomeric and inactive form (Okamura et al., 2000). Interestingly, the direct binding of misfolded proteins to a MHC-like groove present in the luminal domain of IRE1p has been proposed to trigger its activation (Credle et al., 2005; Gardner and Walter, 2011).

For mammalian IRE1α, the mechanisms of activation are still under debate and three different models have been proposed (for extensive reviews see Adams et al., 2019; Karagöz et al., 2019; Preissler and Ron, 2019). Historically, it has been proposed that IRE1α monomers are prone to spontaneously forming dimers and become activated, a process inhibited by a physical interaction with BiP (Bertolotti et al., 2000). In this model, BiP binding destabilize IRE1α dimers, resulting in their inactivation, whereas, under ER stress, BiP preferentially binds unfolded proteins, releasing IRE1α, which then spontaneously activates (Amin-Wetzel et al., 2019, 2017). Alternatively, BiP has been suggested to act as a sensor where the binding of misfolded proteins to its substrate-binding domain (SBD) transduce signals to IRE1α through its ATPase domain (Carrara et al., 2015; Kopp et al., 2018). Once BiP is released from the IRE1α luminal domain, it interacts with substrates to assist the folding of non-native proteins at the ER (Adams et al., 2019). A recent report also reinforced the idea of a crucial role for BiP in the activation and clustering of IRE1α (Ricci et al., 2019) (Fig. 1). As for its yeast homolog, mammalian IRE1α also contains an MHC-like groove at the interface facing the ER lumen; however, the volume and positioning of this groove, as resolved in the crystal structure, was not compatible with peptide binding (Zhou et al., 2006). In contrast, in vitro studies have demonstrated that misfolded proteins can bind to the luminal domain of IRE1α, inducing a structural rearrangement that facilitates its dimerization (Karagöz et al., 2017) (Fig. 1). However, the experimental systems used in these studies are exclusively based on in vitro approaches and further work is needed to demonstrate the so-called ‘direct recognition model’. Besides, other studies have failed to detect the binding of misfolded proteins to the luminal domain of IRE1α (Oikawa et al., 2009). Interestingly, structural and functional analysis of the luminal domain of PERK suggest that it is also able to bind to misfolded proteins (Dalton et al., 2018 preprint; Wang et al., 2018b).

Fig. 1.

Sensing of ER stress by mammalian IRE1α. IRE1α is maintained in a repressed and monomeric state through an association with the ATPase domain of BiP (brighter green) through a recycling mechanism that involves ERdj4. ERdj4 binds to the IRE1α luminal domain, which leads to BiP recruitment. The interaction between ERdj4 and the substrate-binding domain (SBD) of BiP (in darker green) results in the activation of the ATPase domain of BiP and the dissociation or ERdj4, thus BiP–ERdj4-mediated repressing mechanism is crucial for maintaining IRE1α in a monomeric and inactive form. Then, BiP is released from the IRE1α luminal domain allowing the interaction with misfolded proteins. Upon ER stress, misfolded proteins dock to the SBD of BiP, triggering the dissociation from IRE1α through allosteric regulation (indirect model). In addition, misfolded proteins interact directly with the luminal domain of IRE1α, promoting IRE1α dimerization (direct model). These events induce the auto-transphosphorylation of IRE1α, which induces a conformational change in its RNase domain that allows its activation. Besides the BiP- and ERdj4-mediated repressing mechanism, IRE1α activation may be regulated by other ER chaperons, such as HSP47 and PDIA6, which can further boost IRE1α signaling, or by PDIA1 (also known as PDI) that can inhibit IRE1α activity. HSP47 can bind to and displace BiP from the luminal domain of IRE1α, thereby allowing the binding of misfolded proteins. During the attenuation phase, BiP and ERdj4 may operate as a buffer to sequester inactive IRE1α.

Fig. 1.

Sensing of ER stress by mammalian IRE1α. IRE1α is maintained in a repressed and monomeric state through an association with the ATPase domain of BiP (brighter green) through a recycling mechanism that involves ERdj4. ERdj4 binds to the IRE1α luminal domain, which leads to BiP recruitment. The interaction between ERdj4 and the substrate-binding domain (SBD) of BiP (in darker green) results in the activation of the ATPase domain of BiP and the dissociation or ERdj4, thus BiP–ERdj4-mediated repressing mechanism is crucial for maintaining IRE1α in a monomeric and inactive form. Then, BiP is released from the IRE1α luminal domain allowing the interaction with misfolded proteins. Upon ER stress, misfolded proteins dock to the SBD of BiP, triggering the dissociation from IRE1α through allosteric regulation (indirect model). In addition, misfolded proteins interact directly with the luminal domain of IRE1α, promoting IRE1α dimerization (direct model). These events induce the auto-transphosphorylation of IRE1α, which induces a conformational change in its RNase domain that allows its activation. Besides the BiP- and ERdj4-mediated repressing mechanism, IRE1α activation may be regulated by other ER chaperons, such as HSP47 and PDIA6, which can further boost IRE1α signaling, or by PDIA1 (also known as PDI) that can inhibit IRE1α activity. HSP47 can bind to and displace BiP from the luminal domain of IRE1α, thereby allowing the binding of misfolded proteins. During the attenuation phase, BiP and ERdj4 may operate as a buffer to sequester inactive IRE1α.

Recent studies have also suggested a new concept whereby IRE1α signaling is coupled to the function of a network of ER luminal chaperones. For example, the binding of BiP to IRE1α is mediated by the co-chaperone ERdj4 (also known as DNAJB9), functioning as a recycling loop to keep IRE1α in a monomeric state (Amin-Wetzel et al., 2017) (Fig. 1). In addition, using an interactome screen followed by functional validation we recently identified HSP47 (also known as SERPINH1), a collagen-specific chaperone, as a new IRE1α-binding partner that facilitates the dissociation of BiP from its luminal domain. HSP47 binding to IRE1α promotes its activation (Sepulveda et al., 2018) (Fig. 1). Since collagens are the most abundant cargo of the secretory pathway, these findings suggest that the demand for collagen production is tightly coupled to the activation of the UPR, possibly as a mechanism to improve the secretory capacity of the cell (Rojas-Rivera et al., 2018). Furthermore, depending on the experimental setting, the disulfide isomerase PDIA6 attenuates or enhances IRE1α signaling through an interaction with specific oxidized cysteine residues in the ER lumen (Eletto et al., 2016, 2014; Groenendyk et al., 2014). Recently, it has been demonstrated that phosphorylated PDI (also known as PDIA1) directly interacts with the IRE1α luminal domain, reducing its activity under ER stress (Yu et al., 2020) (Fig. 1). Thus, these few studies suggest that multiple ER chaperones are coupled to the IRE1α activation and signaling process to adjust the folding capacity of the cell.

Accumulating evidence indicates that IRE1α is also activated by lipid bilayer stress. Structural and functional studies have revealed that the IRE1α transmembrane domain is indispensable for its activation under lipid bilayer stress (Kono et al., 2017), while the luminal domain is not required (Volmer et al., 2013). A similar mechanism has also been proposed for ATF6 activation under these stress conditions (Tam et al., 2018). In addition to the transmembrane domain of IRE1α, an adjacent amphipathic helix is required to sense lipid stress (Halbleib et al., 2017). Thus, the UPR-sensing mechanisms might be coupled to other homeostatic fluctuations that are beyond its function in protein folding stress.

Although all three UPR stress sensors are simultaneously activated under ER stress, studying the composition of the UPRosome has revealed a new paradigm whereby the kinetics and amplitude of their individual responses can be regulated through the selective binding of specific cofactors. The concept of the UPRosome suggests the establishment of a dynamic multi-protein platform in which IRE1α associates with other components to regulate its signaling and crosstalk with other pathways (Hetz and Glimcher, 2009). Thus, UPR signaling intensity does not represent a direct measurement of stress levels since the amplitude and kinetics of IRE1α downstream responses can be tuned by several positive and negative regulatory steps.

Because most of the studies characterizing IRE1α-binding partners have only been performed once, or the characterizations use non-equivalent experimental conditions or were validated at different levels, here, we have compared the experimental conditions used in different studies to identify regulators of IRE1α signaling. Almost 95% of the interactors described in the literature were defined using overexpression systems; however, 48% of them were also validated at the level of endogenous proteins or in cell-free systems with purified components to detect direct interactions. In addition, in many studies, the possible effects of ER stress on the formation of protein complexes were tested.

As shown in Table 1, many of the IRE1α-binding partners have been described only once, with no follow-up studies confirming these interactions. Thus, readers must be aware of the differences between widely studied and confirmed interactors (i.e. BiP, PDIs and filamin A) and proteins that have been assessed only once. To this end we have summarized all the available data about IRE1α interactors (Table 1).

Table 1.

Possible components of the UPRosome

Possible components of the UPRosome
Possible components of the UPRosome

IRE1α has been described as a low-abundance protein in HeLa cells (Kulak et al., 2014), questioning the existence of multiple protein complexes or its function as a scaffold protein. The most probable scenario is that the interaction of IRE1α with UPRosome components is highly dynamic and transient (a ‘hit and run’ model), modulated by signaling events initiated by ER stress, or the interactions could depend on subpools of IRE1α segregated by their oligomerization and phosphorylation state (or other post-translational modifications). Supporting this notion, it was recently demonstrated that IRE1α cluster formation is highly dynamic, with changes within minutes upon ER stress (Belyy et al., 2020). An alternative hypothesis is that IRE1α interactions occurred in subdomains within the ER membrane where some UPRosome components are enriched. This is supported by recent findings indicating the enrichment of IRE1α and PERK in contacts between mitochondria and ER (Carreras-Sureda et al., 2019; Hayashi and Su, 2007; Mori et al., 2013) (see below). In addition, interactions of IRE1α with the translocon (Acosta-Alvear et al., 2018; Plumb et al., 2015; Sundaram et al., 2017) and the ERAD machinery (Yanagitani et al., 2009, 2011) have been reported, supporting the idea that there are distinct ER subdomains containing specific IRE1α-containing protein complexes. Moreover, IRE1α interacts with many proteins through adaptors that can physically bind to other signaling proteins. In the next three sections, we discuss the identification of different components of the UPRosome and their significance to the regulation and function of the UPR. We divide the discussion into three different categories: inhibitors, positive modulators and post-translational modifiers (see also Table 1 and Fig. 2).

Fig. 2.

Function of UPRosome components in IRE1α signaling. The activity of IRE1α is fine-tuned by several proteins that interact with it, forming a complex multiprotein platform referred to as the UPRosome. IRE1α-interacting proteins can be divided into several groups as shown in A–D. (A) Inhibitors of IRE1α signaling. Inhibitors such as BI-1 and fortilin, specifically decrease IRE1α signaling. In addition, PDIA6 and PDI can also inhibit IRE1α activity by interacting and regulating the oxidation status of IRE1α. (B) Positive regulators of IRE1α signaling. Several components of the BCL-2 family can control the balance between monomers and oligomers, such as BAX, BAK, BIM and PUMA. Sec61, ABL1 and NMIIB have also been described has specific inducers of IRE1α oligomers. In addition, components, such as HSP72 and PTP-1B, can specifically induce IRE1α activation upon ER stress. (C) Post-translational modifiers that affect the activation status of IRE1α. Besides IRE1α auto-transphosphorylation, PKA can also induce IRE1α phosphorylation at serine 752, inducing its activity. The balance between phosphorylated and total IRE1α is tightly regulated by several phosphatases, including RPAP2, PP2A and PP2Ce. In addition, IRE1α can also be ADP-ribosylated (PARP16) or ubiquitylated (CHIP), regulating the amplitude of activation. (D) Protein stability modifiers that alter IRE1α protein levels. Several proteins, such as HSP90 and UFM1, can regulate IRE1α protein stability. In addition, IRE1α can be targeted to autophagy through OPTN or degraded through ERAD by USP14, Sel1 L or SYVN1.

Fig. 2.

Function of UPRosome components in IRE1α signaling. The activity of IRE1α is fine-tuned by several proteins that interact with it, forming a complex multiprotein platform referred to as the UPRosome. IRE1α-interacting proteins can be divided into several groups as shown in A–D. (A) Inhibitors of IRE1α signaling. Inhibitors such as BI-1 and fortilin, specifically decrease IRE1α signaling. In addition, PDIA6 and PDI can also inhibit IRE1α activity by interacting and regulating the oxidation status of IRE1α. (B) Positive regulators of IRE1α signaling. Several components of the BCL-2 family can control the balance between monomers and oligomers, such as BAX, BAK, BIM and PUMA. Sec61, ABL1 and NMIIB have also been described has specific inducers of IRE1α oligomers. In addition, components, such as HSP72 and PTP-1B, can specifically induce IRE1α activation upon ER stress. (C) Post-translational modifiers that affect the activation status of IRE1α. Besides IRE1α auto-transphosphorylation, PKA can also induce IRE1α phosphorylation at serine 752, inducing its activity. The balance between phosphorylated and total IRE1α is tightly regulated by several phosphatases, including RPAP2, PP2A and PP2Ce. In addition, IRE1α can also be ADP-ribosylated (PARP16) or ubiquitylated (CHIP), regulating the amplitude of activation. (D) Protein stability modifiers that alter IRE1α protein levels. Several proteins, such as HSP90 and UFM1, can regulate IRE1α protein stability. In addition, IRE1α can be targeted to autophagy through OPTN or degraded through ERAD by USP14, Sel1 L or SYVN1.

We previously identified Bax inhibitor-1 (BI-1, also known as TMBIM6) as the first negative regulator of IRE1α signaling that associated with its cytosolic domain (Lisbona et al., 2009). This interaction was also shown to be important to regulate IRE1α function in models of ischemia, diabetes and liver steatosis (Bailly-Maitre et al., 2010, 2006; Lebeaupin et al., 2018). Another negative regulator of IRE1α signaling is fortilin (also known as TPT1), which can directly interact and inhibit both the kinase and RNase domains, protecting cells from ER stress-induced apoptosis (Pinkaew et al., 2017). Fortilin preferentially binds to phosphorylated and active IRE1α, modulating the attenuation phase under prolonged ER stress (Pinkaew et al., 2017). A recent report showed that the BH3-only protein BID also negatively regulates the RNase activity of IRE1α (Bashir and Majid Fazili, 2019 preprint). A common feature of the negative regulators of IRE1α is that they block its sustained activation, disabling the capacity to engage apoptotic programs under chronic ER stress (Fig. 2A).

The enhancement of IRE1α signaling by positive regulators is highly complex as these proteins can control several steps of IRE1α activation and signaling (Hetz et al., 2020) (Fig. 2B). The first identified enhancer of IRE1α signaling was the protein tyrosine phosphatase 1B (PTP-1B; also known as PTPN1) (Gu et al., 2004). PTP-1B promotes XBP1 mRNA splicing and JNK activation under ER stress, but not PERK signaling (Gu et al., 2004). However, a physical interaction between IRE1α and PTP-1B has not been reported.

We identified the first direct enhancers of IRE1α signaling as the pro-apoptotic BCL-2 family members BAX and BAK (also known as BAK1), two crucial components of the cell death machinery (Hetz et al., 2006a). BAX and BAK locate to the ER and fine-tune IRE1α activation and the amplitude of its signaling under ER stress. Similarly, two other BCL-2 family members, the BH3-only proteins BIM (BCL2L11) and PUMA (BBC3), have been described to repress the attenuation phase of IRE1α to sustain its signaling (Rodriguez et al., 2012). Thus, one possible scenario is that BAX and/or BAK might facilitate IRE1α dimer formation in the activation phase, whereas BH3-only proteins stabilize IRE1α dimers and/or oligomers during the attenuation phase after prolonged ER stress.

IRE1α oligomerization can be also regulated by specific factors. For instance, the non-muscle myosin heavy chain IIB (NMIIB) protein stabilizes the formation of IRE1α oligomers (He et al., 2012). In agreement with this finding, drugs that destabilize the actin cytoskeleton decrease IRE1α activation and cluster formation in mammalian and yeast cells (He et al., 2012; Ishiwata-Kimata et al., 2013). Remarkably, ER stress induction leads to the phosphorylation and activation of two important regulators of the actin cytoskeleton, the myosin regulatory light chain (RLC), a crucial activator of NMIIB (He et al., 2012), and filamin A (Urra et al., 2018), supporting the idea of an actin cytoskeleton crosstalk with the UPR. It has been also described that the tyrosine protein kinase ABL1 interacts with IRE1α and stabilizes the formation of oligomers, which may favor IRE1α RNase activation towards RIDD (Dufey et al., 2020; Morita et al., 2017). Finally, Yip1A (also known as YIPF5) regulates IRE1α auto-transphosphorylation and the formation of oligomers (Taguchi et al., 2015).

Sec61, a core component of the translocon, forms a complex with IRE1α at the ER membrane, recruiting the unspliced XBP1 (XBP1u) through a pseudo-transmembrane domain present in the nascent protein chain (Plumb et al., 2015). The close proximity between XBP1u and the IRE1α–Sec61 complex is crucial to optimize the XBP1 mRNA splicing process (Plumb et al., 2015). Other studies have suggested that the interaction between Sec61 and IRE1α maintains the balance between dimers and oligomers (Sundaram et al., 2017). In addition, this protein complex was also found in an unbiased interactome, showing that IRE1α binds to several proteins and ribosomal RNAs that are part of the signal recognition particle (SRP) and the ribosome (Acosta-Alvear et al., 2018), indicating that IRE1α is in close proximity with the ER protein translocation and targeting machineries. Recently, the protein kinase C substrate 80K-H (PRKCSH) was shown to physically interact with IRE1α, promoting its phosphorylation and oligomerization under ER stress (Shin et al., 2019). Taken together, this evidence indicates that different factors tune the balance between IRE1α dimers and oligomers to determine the amplitude and kinetics of UPR signaling.

Auto-transphosphorylation of IRE1α is a key event required for the activation of the RNase domain. Several phosphorylation sites have been identified in the cytosolic domain of human IRE1α, including serine residues 551 and 562 in the linker region, serine residues 724, 726 and 729 in the kinase activation loop, and threonine 973 in the C-terminal RNase domain (Prischi et al., 2014), whereas the phosphorylation in the serine residues located in the kinase domain are crucial for IRE1α RNase activation (Prischi et al., 2014). Homologous serine phosphorylation sites have also been described in yeast (Armstrong et al., 2017). Interestingly, the phosphorylation at serine 729 has been associated with the selective increase in RIDD activity in XBP1-deficient B cells (Tang et al., 2018).

Protein kinase A (PKA) has been suggested to directly phosphorylate serine 724 to enhance IRE1α signaling after fasting or glucagon stimulation (Mao et al., 2011). Similarly, in the brain, the activation of IRE1α by brain-derived growth factor (BDNF) (Martínez et al., 2016) is also mediated by PKA activation downstream of its receptor (Saito et al., 2018). Three phosphatases dephosphorylate IRE1α in different contexts. For instance, the association of the receptor for activated C kinase 1 (RACK1) to IRE1α recruits the phosphatase PP2A to the complex after glucose stimulation to sustain UPR signaling (Qiu et al., 2010). In addition, the ER-located phosphatase PP2Ce (also known as PPM1L) dephosphorylates IRE1α in the context of metabolic lipid control (Lu et al., 2013). Interestingly, in a siRNA screen targeting 273 human phosphatases, the RNA polymerase II-associated protein 2 (RPAP2) was found to de-phosphorylate IRE1α (Chang et al., 2018). Hence, the regulation of the phosphorylation status of IRE1α is another key step fine tuning its signaling.

Additional post-translational modifications also modulate IRE1α signaling (Fig. 2C). IRE1α can be ADP-ribosylated under ER stress by the poly(ADP-ribose) polymerase 16 (PARP16), an event that is required for their full activation (Jwa and Chang, 2012). In addition, ubiquitylation of IRE1α at lysine residues 545 and 828 by the E3 ligase CHIP (also known as STUB1) regulates the phosphorylation status of IRE1α (Zhu et al., 2014). The E3 ligase MITOL (also known as MARCHF5) ubiquitylates IRE1α at lysine 481, inhibiting its signaling under prolonged ER stress (Takeda et al., 2019). S-nitrosylation of IRE1α at two conserved cysteine residues located in the RNase domain also attenuates its activity under metabolic stress and neurodegeneration (Nakato et al., 2015; Wang et al., 2018a; Yang et al., 2015).

Another important step modulating IRE1α signaling is the control of its stability (Fig. 2D). The E3 ubiquitin ligase synoviolin (SYVN1) ubiquitylates IRE1α to induce its degradation (Gao et al., 2008; Sun et al., 2015). Interestingly, ubiquitin D (UBD) can associate with IRE1α, impacting JNK signaling (Brozzi et al., 2016). Furthermore, DDRGK domain containing 1 (DDRGK1), a component of the ubiquitin-fold modifier 1 (Ufm1) system, interacts with and regulates IRE1α protein stability by inducing its ufmylation through a non-covalent modification (Liu et al., 2017). IRE1α can also be targeted to ERAD by USP14 (Nagai et al., 2009), or to the autophagy pathway by optineurin (Tschurtschenthaler et al., 2017). In addition, HSP90 can stabilize IRE1α by interacting with its cytosolic domain (Marcu et al., 2002; Ota and Wang, 2012). Interestingly, IRE1α can also be cleaved by caspases in some cancer models, resulting in the generation of a fragment that can inhibit cell death (Shemorry et al., 2019). Altogether, these reports suggest that many different factors tune IRE1α signaling at multiple levels.

Finally, although the UPRosome concept was originally conceived exclusively for IRE1α regulation (Hetz and Glimcher, 2009), recent studies have shown that ATF6 and PERK can be also modulated by specific interacting proteins (reviewed in Hetz et al., 2020). Thus, the concept of the UPRosome could be extended to PERK and ATF6.

Most of the studies in the field have addressed the biological function of IRE1α in the control of transcription and translation through XBP1s or RIDD, impacting multiple cellular processes (reviewed in Oakes and Papa, 2015; Wang and Kaufman, 2016). However, recent findings have pointed to novel functions executed through alternative signaling pathways (Fig. 3).

Fig. 3.

Possible novel outputs of the IRE1α UPRosome. (A) Venn diagram comparing the composition of all UPRosome proteins described in global references listed in Table 1 (56 proteins; blue circle), and two available screens using either yeast two-hybrid [32 proteins (index score >4); yellow circle; Urra et al., 2018] and immunoprecipitation followed by mass-spectrometry analysis [405 proteins (P>0.01); green circle; Acosta-Alvear et al., 2018]. Gene ontology (GO) analysis and functional clustering of all IRE1α-binding proteins was performed (see Table S2). (B) Cell death. IRE1α regulates the cell death process through multiple mechanisms that involve NFκB, JNK and RIDD. Our GO analysis also indicates that IRE1α interacts with clusters of proteins involved in cell death such as components of the BCL-2 family (BAX, BAK, BIM and PUMA), and BI-1 (TMBIM6) or α-catenein (CTNNA1), which can also be crucial components of the cell death machinery. (C) Protein degradation pathways. IRE1α associates with TRAF2 and BI-1/TMBIM6 to regulate JNK activation and autophagy. Our GO functional clustering indicates that IRE1α can also physically interact with components of the chaperone-mediated autophagy (HSPA8, HSP90AA1 and CHIP) or components of the ERAD pathway, among them SYVN1, USP14, the DERL1 family and SEL1L. Also, IRE1α interacts with OS9 and BiP, which also have been described to target proteins for degradation. (D) Cytoskeleton regulation. We have previously described that IRE1α regulates actin cytoskeleton through FLNA interaction. Here, we found that IRE1α also interacts with a cluster of proteins that are part of or regulate the cytoskeleton such as CORO1B and CORO1C, ARPC1A and ARPC1B, NMIIB, CDC42 and MYO1B. In addition, IRE1α binds to other proteins that can indirectly regulate the cytoskeleton, such as ABL1, BAIAP2, NCKAP1 and DPSYSL3. (E) Ca2+ homeostasis. IRE1α location at mitochondria-associated membranes (MAMs) regulates the docking of the IP3Rs (ITPR1, ITPR2 or ITPR3) at this subcompartment, promoting Ca2+ release from the ER. This mechanism enhances the transfer of Ca2+ to the mitochondrial matrix, promoting ATP production to tune bioenergetics. Our GO analysis reveals that IRE1α could affect this process by interacting with other proteins such as BAX, BAX and BI-1/TMBIM6 or through ABL1 and LYN. (F) Translation and ribosome function. IRE1α interacts with SEC61, keeping it in close proximity to all ribosome components. We postulate that IRE1α can interact with several ribosomal proteins (RPL proteins) and ribosome assembly proteins such as FTSJ1–FTSJ3 and URB1. In addition, IRE1α can bind to proteins involved in translation such as EIF4B–EIF4H. (G) DNA damage response. We recently described that upon DNA damage, IRE1α is specifically activated to induce the degradation of mRNAs (Ppp2r1a and Ruvbl1) through regulated IRE1α-dependent decay (RIDD), sustaining DNA repair and promoting survival. The molecular mechanism involves that downstream of ATM and ATR; ABL1 interacts with IRE1α promoting its oligomerization and RIDD. (H) Cell cycle control. It is known that IRE1α can regulate cyclin A expression through XBP1s promoting cell cycle progression. Our GO analysis reveals that IRE1α specifically binds to components that regulate cytokinesis such as CORO1B and CORO1C, PRC1, NUSAP1 and KPNB1. Also, IRE1α interacts with several proteins that can indirectly regulate cell cycle such as DCNT1, ACTR1A and PPP2R1A. (I) Vesicle transport and protein targeting. IRE1α can regulate the expression of secretory genes through XBP1s. We also indicate that IRE1α can affect vesicle transport through several other proteins such as the SEC24 family or components of vesicle formation and transport, such as SNAP23, SNAPIN and MYO1C. In addition, this process can be further modulated by other proteins including KIF5B, PKA, PSEN and DHX9. See Table S1 for the full names of proteins included here.

Fig. 3.

Possible novel outputs of the IRE1α UPRosome. (A) Venn diagram comparing the composition of all UPRosome proteins described in global references listed in Table 1 (56 proteins; blue circle), and two available screens using either yeast two-hybrid [32 proteins (index score >4); yellow circle; Urra et al., 2018] and immunoprecipitation followed by mass-spectrometry analysis [405 proteins (P>0.01); green circle; Acosta-Alvear et al., 2018]. Gene ontology (GO) analysis and functional clustering of all IRE1α-binding proteins was performed (see Table S2). (B) Cell death. IRE1α regulates the cell death process through multiple mechanisms that involve NFκB, JNK and RIDD. Our GO analysis also indicates that IRE1α interacts with clusters of proteins involved in cell death such as components of the BCL-2 family (BAX, BAK, BIM and PUMA), and BI-1 (TMBIM6) or α-catenein (CTNNA1), which can also be crucial components of the cell death machinery. (C) Protein degradation pathways. IRE1α associates with TRAF2 and BI-1/TMBIM6 to regulate JNK activation and autophagy. Our GO functional clustering indicates that IRE1α can also physically interact with components of the chaperone-mediated autophagy (HSPA8, HSP90AA1 and CHIP) or components of the ERAD pathway, among them SYVN1, USP14, the DERL1 family and SEL1L. Also, IRE1α interacts with OS9 and BiP, which also have been described to target proteins for degradation. (D) Cytoskeleton regulation. We have previously described that IRE1α regulates actin cytoskeleton through FLNA interaction. Here, we found that IRE1α also interacts with a cluster of proteins that are part of or regulate the cytoskeleton such as CORO1B and CORO1C, ARPC1A and ARPC1B, NMIIB, CDC42 and MYO1B. In addition, IRE1α binds to other proteins that can indirectly regulate the cytoskeleton, such as ABL1, BAIAP2, NCKAP1 and DPSYSL3. (E) Ca2+ homeostasis. IRE1α location at mitochondria-associated membranes (MAMs) regulates the docking of the IP3Rs (ITPR1, ITPR2 or ITPR3) at this subcompartment, promoting Ca2+ release from the ER. This mechanism enhances the transfer of Ca2+ to the mitochondrial matrix, promoting ATP production to tune bioenergetics. Our GO analysis reveals that IRE1α could affect this process by interacting with other proteins such as BAX, BAX and BI-1/TMBIM6 or through ABL1 and LYN. (F) Translation and ribosome function. IRE1α interacts with SEC61, keeping it in close proximity to all ribosome components. We postulate that IRE1α can interact with several ribosomal proteins (RPL proteins) and ribosome assembly proteins such as FTSJ1–FTSJ3 and URB1. In addition, IRE1α can bind to proteins involved in translation such as EIF4B–EIF4H. (G) DNA damage response. We recently described that upon DNA damage, IRE1α is specifically activated to induce the degradation of mRNAs (Ppp2r1a and Ruvbl1) through regulated IRE1α-dependent decay (RIDD), sustaining DNA repair and promoting survival. The molecular mechanism involves that downstream of ATM and ATR; ABL1 interacts with IRE1α promoting its oligomerization and RIDD. (H) Cell cycle control. It is known that IRE1α can regulate cyclin A expression through XBP1s promoting cell cycle progression. Our GO analysis reveals that IRE1α specifically binds to components that regulate cytokinesis such as CORO1B and CORO1C, PRC1, NUSAP1 and KPNB1. Also, IRE1α interacts with several proteins that can indirectly regulate cell cycle such as DCNT1, ACTR1A and PPP2R1A. (I) Vesicle transport and protein targeting. IRE1α can regulate the expression of secretory genes through XBP1s. We also indicate that IRE1α can affect vesicle transport through several other proteins such as the SEC24 family or components of vesicle formation and transport, such as SNAP23, SNAPIN and MYO1C. In addition, this process can be further modulated by other proteins including KIF5B, PKA, PSEN and DHX9. See Table S1 for the full names of proteins included here.

Two recent examples from our group have revealed that IRE1α influences cell physiology by acting as a scaffold protein, independently of its enzymatic activities (see below) (Carreras-Sureda et al., 2019; Urra et al., 2018). To further decode possible new IRE1α functions, here, we performed an extensive analysis of almost 40 IRE1α-binding partners reported in the literature (Table 1), followed by a gene ontology (GO) analysis and functional clustering. In addition, we also included other IRE1α-binding proteins presented in data sets available from different public interactome screenings (Acosta-Alvear et al., 2018; Urra et al., 2018) (Table S1). This comparison revealed eight common UPRosome components in at least two data sets (Fig. 3A). Furthermore, our GO analyses of all the 495 proteins identified showed that almost all significant associations fall into 20 different biological processes (Table S2). This functional analysis can be further minimized to five major clusters of proteins involved in cell cycle, cytoskeleton regulation, protein targeting, ribosome biogenesis and protein translation (Table S2). In the next sections, we discuss available evidence that suggests the involvement of IRE1a in different biological processes by IRE1α through the interaction with specific factors.

A number of studies support a direct molecular connection between IRE1α and several signaling pathways regulating cell death. For instance, binding of IRE1α to SH2/SH3-containing adaptor protein Nck can modulate p38 MAPK family and ERK1 (MAPK3) signaling, impacting cell survival (Nguyên et al., 2004). In addition, IRE1α can enhance the activation of NF-κB by interacting with TRAF2 and the inhibitor of nuclear factor κβ kinase (IKK), promoting cell death (Hu et al., 2006). Remarkably, the IRE1α–TRAF2 interaction also triggers the activation of the apoptosis signal-regulating kinase 1 (ASK1) and Jun-N terminal kinase (JNK) pathway (Nishitoh et al., 2002; Urano et al., 2000). The IRE1α–MAPK signaling branch is also modulated by other components of the UPRosome, such as AIP1 (Luo et al., 2008), RNF13 (Arshad et al., 2013), UBD (Brozzi et al., 2016), BI-1 (Castillo et al., 2011) and N-Myc interactor (Brozzi et al., 2014). However, the mechanisms of JNK activation downstream of IRE1α and its consequences for cell fate under ER stress are not completely understood.

Our GO enrichment analyses and functional clustering revealed a group of IRE1α-binding partners that are involved in apoptosis, including pro-apoptotic proteins of the BCL-2 family (BAX, BAK, BIM and PUMA), α-catenin and BI-1 (Fig. 3B). We speculate that this association may represent a molecular interface to monitor alterations in proteostasis and engage cell death programs when cell damage is deemed irreversible.

One of the outputs of UPR activation under ER stress is the engagement of the autophagy and ERAD pathways, contributing to stress mitigation (Rashid et al., 2015). The interplay between autophagy and the UPR affects the progression of several human diseases, such as neurodegenerative diseases (Hetz et al., 2009; Vidal et al., 2012), cancer (Hart et al., 2012) and auditory loss (Kishino et al., 2017). Interestingly, the formation of the IRE1α–TRAF2 complex can also induce autophagy in a JNK-dependent manner (Castillo et al., 2011). Indeed, JNK activation mediates the phosphorylation of BCL-2 (Pattingre et al., 2009); this disrupts the Beclin-1–BCL-2 complex, facilitating phagosome initiation and autophagy (Pattingre et al., 2005). The IRE1α–TRAF2 interaction can be further modulated by BI-1, displacing TRAF2 from the complex, blocking autophagy flux (Castillo et al., 2011).

Our GO analysis indicates that IRE1α interacts with three proteins involved in targeting substrates to chaperone-mediated autophagy (CMA) – HSPA8, HSP90AA1 and STUB1 (Kaushik et al., 2011). IRE1α also associates with factors involved in ERAD, including SYVN1, a crucial E3 ubiquitin-protein ligase of the ERAD pathway, DERL1 and DERL3, proteins regulating the retrotranslocation of misfolded proteins to the cytosol, and the proteasome-associated deubiquitylase USP14, which controls the ubiquitylation status of proteasome-targeted proteins (Fig. 3C). Hence, IRE1α interacts with components of autophagy and ERAD machineries, suggesting that IRE1α could be involved in the regulation of these pathways.

Using a yeast two-hybrid screen, we recently identified almost 70 candidate proteins that might directly bind to the cytosolic region of IRE1α, where filamin A was the strongest hit (Urra et al., 2018). Filamin A was also validated as a strong IRE1α interactor in a second unbiased screen (Tavernier et al., 2018) and, in addition, in a recent study (Augusto et al., 2020). Filamin A is an actin-filament-crosslinking protein that regulates cytoskeleton dynamics, impacting different processes, such as cell migration, adhesion and mechanotransduction (Kim and McCulloch, 2011; Lynch and Sheetz, 2011). We showed that IRE1α acts as a scaffold to recruit filamin A to the ER membrane and facilitate its phosphorylation to modulate cytoskeleton dynamics (Urra et al., 2018; Fig. 3D). In this model, PKCα mediates the phosphorylation of many filamin A proteins at serine 2152 in an IRE1α-dependent manner (Urra et al., 2018).

The interaction between IRE1α and filamin A requires a proline-rich domain located at the distal C-terminal region of IRE1α (Urra et al., 2018). Interestingly, disruption of this domain does not affect XBP1 mRNA splicing, whereas it selectively inhibits the regulation of the actin cytoskeleton, fully separating both cellular functions (Urra et al., 2018). In vivo studies demonstrated that IRE1α ablation decreases cortical neuronal migration in the mouse brain (Urra et al., 2018). Since filamin A mutations in humans are the main cause of periventricular nodular heterotopia (PVNH), a genetic disorder characterized by abnormal neuronal migration (Fox et al., 1998; Sarkisian et al., 2008), we speculate that IRE1α loss of function might contribute to PVNH as reported for other filamin A regulators (Sarkisian et al., 2008; Zhou et al., 2010). Recently, it has been reported that the IRE1α and FLNA interaction is also implicated in the migration of cells infected with intracellular parasites (Augusto et al., 2020).

According to our GO analysis, IRE1α interacts with other actin-binding proteins such as CORO1B and CORO1C, which are involved in actin remodeling and lamellipodia formation in the migrating leading edge (Cai et al., 2007) and in ER-associated endosome fission (Hoyer et al., 2018) (Fig. 3D). In addition, we found an important cluster of regulators of actin and microtubule dynamics, suggesting that IRE1α may affect cytoskeleton remodeling through several mechanisms.

IRE1α and PERK have been reported to be enriched at the ER–mitochondrial contact sites (termed mitochondria-associated membranes; MAMs) (Carreras-Sureda et al., 2019; Mori et al., 2013; van Vliet et al., 2017). It was suggested that the specific localization of UPR stress sensors at this ER membrane subdomain set the threshold for UPR activation (Mori et al., 2013). However, the actual role of the IRE1α pool located at MAMs in mitochondrial physiology was not defined until very recently. Under ER stress, the inositol-1,4,5-trisphosphate receptors (IP3Rs) trigger the exit of Ca2+ from the ER to the cytosol, regulating autophagy and apoptosis (Kiviluoto et al., 2013). ER Ca2+ release is pivotal to tune mitochondrial function to produce ATP, because different enzymes of the tricarboxylic acid cycle (TCA) are Ca2+ dependent (Giorgi et al., 2018). The ER-to-mitochondria Ca2+ transfer process requires local microdomains of high Ca2+ concentrations to be efficient (Csordás et al., 2010; Giacomello et al., 2010). We recently reported a novel function of IRE1α in sustaining ER–mitochondrial communication and bioenergetics by demonstrating that IRE1α regulates the distribution of IP3Rs at MAMs through a direct interaction (Carreras-Sureda et al., 2019). IRE1α localization at MAMs recruits and docks IP3Rs, improving mitochondria calcium transfer to burst energy metabolism. In addition, besides IP3R localization, IRE1α expression alters the protein composition of MAMs (Fig. 3E). The GO analysis also suggested that IRE1α interacts with other proteins that can regulate calcium homeostasis, such as BAX, BAK and BI-1 (Fig. 3E).

As discussed above, the XBP1u mRNA is recruited to the ER by the formation of a complex between IRE1α and Sec61 (Plumb et al., 2015). However, the mechanisms explaining how RIDD substrates are made available to the ER membrane are still unknown. Using photoactivatable ribonucleoside crosslinking and immunoprecipitation (PAR-CLIP), it was shown that IRE1α is part of a large complex containing almost 400 different RNAs, including mRNAs, ncRNA and rRNA (Acosta-Alvear et al., 2018). Furthermore, IRE1α interacts with the exposed surface of translating ribosomes, and with components of the SRP and translocon (Acosta-Alvear et al., 2018). Therefore, IRE1α could function as a scaffold to recruit and maintain the protein targeting and translation machineries in close proximity under ER stress conditions (Acosta-Alvear et al., 2018) (Fig. 3F). Our analysis suggests that IRE1α could not only regulate mRNA translation by interacting with several ribosomal proteins, but also through the regulation of ribosome biogenesis or their assembly since IRE1α interacts with FTSJ3 and URB1 (Acosta-Alvear et al., 2018), two important regulators of ribosomal biogenesis (Morello et al., 2011). Other IRE1α-binding proteins are involved in the translation machinery for rRNA, ncRNA and mRNAs, suggesting that IRE1α could also regulate translation through the EIF4 complex. Hence, IRE1α associates with ribosome and translocon components; however, the functional consequences of these interactions need to be defined.

Although it was not found in our dataset analysis, we recently reported that ABL1, a previously characterized IRE1α-interacting protein (Morita et al., 2017), induces the selective activation of IRE1α under genotoxic stress to promote DNA repair programs. Upon DNA damage, ABL1 interacts with IRE1α, selectively activating RIDD in the absence of XBP1 mRNA splicing (Fig. 3G). This activity results in the degradation of Ppp2r1a and Ruvbl1 mRNAs, two factors that are involved in the DNA damage response (DDR) by regulating the de-phosphorylation of checkpoint kinase 1 and 2 (CHK1/2; also known as CHEK1/2) and histone 2AX (H2AX) (Chowdhury et al., 2005; Jha et al., 2008). The specific activation of IRE1α by genotoxic stress promotes DNA repair and cell cycle progression, resulting in cell survival. This function is conserved in evolution as demonstrated in fly and mouse models of DNA damage (Dufey et al., 2020). Other studies have also linked the DDR with the UPR, because XBP1u controls the levels of p53, a central regulator of the DNA damage response, through its ubiquitylation and degradation (Huang et al., 2017). In addition, ER stress in tumor cells sensitizes cells to death induced by DNA-damaging agents (e.g. chemotherapy and radiotherapy) (Bobrovnikova-Marjon et al., 2010; Weatherbee et al., 2016; Yamamori et al., 2013). These studies suggest that the UPR functionally interacts with pathways that maintain genome stability.

In addition to the above examples, our GO enrichment analyses identified some interesting clusters of proteins involved in additional functions, including cell cycle control and protein transport (see Box 2 and Fig. 3H,I). Taken together, the analysis of the available interactome data sets reinforce the notion that the UPRosome is implicated in several biological processes that need to be further studied.

Box 2. Deciphering novel physiological functions of IRE1α

Our GO enrichment analyses and functional clustering of IRE1α-binding partners identified some interesting clusters of proteins involved in different functions including cell cycle control and cytokinesis (Fig. 3H). IRE1α interacts with CORO1B and CORO1C, two coronins that together with the Arp2/3 complex control actin rearrangement and cell cycle progression (Molinie et al., 2019). PRC1 and NUSAP1, two important players involved in cytokinesis regulation and cancer progression, were also identified (Li et al., 2018). In addition, our analysis also revealed a cluster of proteins regulating vesicle transport and protein targeting (Fig. 3I). For example, distinct isoforms of Sec24 are observed that are part of the coat protein complex II (COPII), which is involved in the selection of cargo proteins and the formation of transport vesicles at the ER (Mancias and Goldberg, 2007). In addition, proteins involved in vesicle transport and fusion, like SNAP23 and SNAPIN, are also detected in IRE1α-containing protein complexes (van Niel et al., 2018). Several molecular motors involved in mRNA trafficking, such as KIF5B and FXR2 (Hirokawa, 2006), are detected in IRE1α-containing protein complexes that are crucial for axonal transport, promoting axonal growth cone and dendrite morphogenesis (Fig. 3I). IRE1α can also interact with proteins involved in cell differentiation, including the myeloid, hematopoietic and osteoblastic lineage (Table S2). In addition, IRE1α associates with crucial components involved in viral transcription, replication and defense (Table S2). Viral infection induces ER stress (Verchot, 2016); however, the actual impact of IRE1α on viral infection and replication is poorly defined. Our analysis reinforces the idea that IRE1α interacts with partners involved in immune responses including antigen processing and presentation via MHC class I, cytokine production and secretion, and phagocytosis. This is in agreement with functional studies indicating that IRE1α can regulate some of these processes by mechanism involving RIDD or XBP1s (for an extensive review, see Osorio et al., 2018).

The UPR is a central signal transduction pathway that acts to restore ER function and proteostasis, and engages a series of adaptive responses or triggers apoptosis under acute or sustained ER stress. The mechanisms that explain this transition from adaptive to pro-death are not completely defined, but are essential to understanding how cell fate is determined under ER stress. One possible explanation may involve the dynamic nature of the UPRosome as a mechanism to regulate IRE1α signaling outputs. BiP is the most validated IRE1α-binding partner and has a fundamental role in adjusting the capacity of a cell to trigger the UPR. Importantly, several single studies suggest that UPR activation, and the threshold of ER stress to induce IRE1α activation and attenuation, is regulated by other ER chaperones and foldases including HSP47, PDIA6 and PDIA1. We speculate that these observations represent the tip of the iceberg, suggesting that the ER stress surveillance process is broadly coupled to the folding machinery involving a complex network of chaperones.

After the proposition of the UPRosome concept more than 10 years ago (Hetz and Glimcher, 2009), many new IRE1α-binding partners have been identified that modulate its activity, conformation and stability. As stated in Table 1, many of these interactions were determined using artificial overexpression systems and validated only once. Although the number of IRE1α-binding partners is quite large (∼47 proteins), with dozens of laboratories confirming the concept of the UPRosome, readers must be cautious when interpreting these data sets since further independent validation is needed. Despite the fact some of these interactors were defined in single studies, several binders were properly validated using different complementary approaches. Here, we provide a new layer of analysis by performing a broad bioinformatics assessment using most data sets available, highlighting clusters of proteins sharing similar functions, suggesting novel functions of the UPR through protein–protein interactions and eliminating the bias of one-time interactors or reports.

Overall, available studies assessing the nature of the IRE1α interactome have revealed unexpected biological functions of the pathway in diverse processes such as autophagy, cell migration and cytoskeleton dynamics, protein translation and energy metabolism. Some of these emerging roles are mediated by non-canonical signaling modes independent of the canonical IRE1α RNase domain, where IRE1α operates as a scaffold that recruits proteins involved in multiple cellular processes. Thus, we propose that a comprehensive analysis of the IRE1α interactome is required to understand the global impact of the pathway to cell physiology and disease. We expect that, depending on the cell type and stimuli analyzed, distinct UPRosomes are assembled in a dynamic manner. There might be even different IRE1α-containing complexes in different ER compartments, as we proposed for the MAM subdomains. Since IRE1α function and dysfunction has been extensively linked to different human diseases, strategies to fine tune the composition and activity of the UPRosome may have important therapeutic applications.

We thank R. Aravena for table compilation.

Funding

This work was funded by Fondo Nacional de Desarrollo Científico y Tecnológico (FONDECYT) 11180825 (H.U.), FONDECYT 1180186 (C.H.), FONDECYT 3200716 (P.P.) ANID/FONDAP/15150012 (C.H.), ECOS Comisión Nacional de Investigación Científica y Tecnológica (CONICYT) Cooperation grant Chile-France ECOS170032 (H.U., C.H.), Millennium Institute P09-015-F (C.H.) and European Commission R&D MSCA-RISE 734749 (C.H.). We also thank the support from Michael J Fox Foundation for Parkinson's Research – Target Validation grant 9277, FONDEF ID16I10223, FONDEF D11E1007 (C.H.).

Acosta-Alvear
,
D.
,
Karagöz
,
G. E.
,
Fröhlich
,
F.
,
Li
,
H.
,
Walther
,
T. C.
and
Walter
,
P.
(
2018
).
The unfolded protein response and endoplasmic reticulum protein targeting machineries converge on the stress sensor IRE1
.
eLife
7
,
e43036
.
Adachi
,
Y.
,
Yamamoto
,
K.
,
Okada
,
T.
,
Yoshida
,
H.
,
Harada
,
A.
and
Mori
,
K.
(
2008
).
ATF6 is a transcription factor specializing in the regulation of quality control proteins in the endoplasmic reticulum
.
Cell Struct. Funct.
33
,
75
-
89
.
Adams
,
C. J.
,
Kopp
,
M. C.
,
Larburu
,
N.
,
Nowak
,
P. R.
and
Ali
,
M. M. U.
(
2019
).
Structure and molecular mechanism of ER stress signaling by the unfolded protein response signal activator IRE1
.
Front. Mol. Biosci.
6
,
11
.
Amin-Wetzel
,
N.
,
Saunders
,
R. A.
,
Kamphuis
,
M. J.
,
Rato
,
C.
,
Preissler
,
S.
,
Harding
,
H. P.
and
Ron
,
D.
(
2017
).
A J-protein co-chaperone recruits BiP to monomerize IRE1 and repress the unfolded protein response
.
Cell
171
,
1625
-
1637.e13
.
Amin-Wetzel
,
N.
,
Neidhardt
,
L.
,
Yan
,
Y.
,
Mayer
,
M. P.
and
Ron
,
D.
(
2019
).
Unstructured regions in IRE1alpha specify BiP-mediated destabilisation of the luminal domain dimer and repression of the UPR
.
eLife
8
,
e50793
.
Armstrong
,
M. C.
,
Šestak
,
S.
,
Ali
,
A. A.
,
Sagini
,
H. A. M.
,
Brown
,
M.
,
Baty
,
K.
,
Treumann
,
A.
and
Schröder
,
M.
(
2017
).
Bypass of activation loop phosphorylation by aspartate 836 in activation of the endoribonuclease activity of Ire1
.
Mol. Cell. Biol.
37
,
e00655-16
.
Arshad
,
M.
,
Ye
,
Z.
,
Gu
,
X.
,
Wong
,
C. K.
,
Liu
,
Y.
,
Li
,
D.
,
Zhou
,
L.
,
Zhang
,
Y.
,
Bay
,
W. P.
,
Yu
,
V. C.
, et al. 
(
2013
).
RNF13, a RING finger protein, mediates endoplasmic reticulum stress-induced apoptosis through the inositol-requiring enzyme (IRE1alpha)/c-Jun NH2-terminal kinase pathway
.
J. Biol. Chem.
288
,
8726
-
8736
.
Augusto
,
L.
,
Martynowicz
,
J.
,
Amin
,
P. H.
,
Alakhras
,
N. S.
,
Kaplan
,
M. H.
,
Wek
,
R. C.
and
Sullivan
,
W. J.
, Jr.
(
2020
).
Toxoplasma gondii co-opts the unfolded protein response to enhance migration and dissemination of infected host cells
.
mBio
11
,
e00915
-
20
.
Bailly-Maitre
,
B.
,
Fondevila
,
C.
,
Kaldas
,
F.
,
Droin
,
N.
,
Luciano
,
F.
,
Ricci
,
J. E.
,
Croxton
,
R.
,
Krajewska
,
M.
,
Zapata
,
J. M.
,
Kupiec-Weglinski
,
J. W.
, et al. 
(
2006
).
Cytoprotective gene bi-1 is required for intrinsic protection from endoplasmic reticulum stress and ischemia-reperfusion injury
.
Proc. Natl. Acad. Sci. USA
103
,
2809
-
2814
.
Bailly-Maitre
,
B.
,
Belgardt
,
B. F.
,
Jordan
,
S. D.
,
Coornaert
,
B.
,
von Freyend
,
M.
,
Kleinridders
,
A.
,
Mauer
,
J.
,
Cuddy
,
M.
,
Kress
,
C. L.
,
Willmes
,
D.
, et al. 
(
2010
).
Hepatic Bax inhibitor-1 inhibits IRE1alpha and protects from obesity-associated insulin resistance and glucose intolerance
.
J. Biol. Chem.
285
,
6198
-
6207
.
Bashir
,
S.
and
Majid Fazili
,
K.
(
2019
).
Bid as a novel interacting partner of IRE1 differentially regulating its RNAse activity
.
bioRxiv
.
Belyy
,
V.
,
Tran
,
N. H.
and
Walter
,
P.
(
2020
).
Quantitative microscopy reveals dynamics and fate of clustered IRE1alpha
.
Proc. Natl. Acad. Sci. USA
117
,
1533
-
1542
.
Bertolotti
,
A.
,
Zhang
,
Y.
,
Hendershot
,
L. M.
,
Harding
,
H. P.
and
Ron
,
D.
(
2000
).
Dynamic interaction of BiP and ER stress transducers in the unfolded-protein response
.
Nat. Cell Biol.
2
,
326
-
332
.
Bobrovnikova-Marjon
,
E.
,
Grigoriadou
,
C.
,
Pytel
,
D.
,
Zhang
,
F.
,
Ye
,
J.
,
Koumenis
,
C.
,
Cavener
,
D.
and
Diehl
,
J. A.
(
2010
).
PERK promotes cancer cell proliferation and tumor growth by limiting oxidative DNA damage
.
Oncogene
29
,
3881
-
3895
.
Bouchecareilh
,
M.
,
Higa
,
A.
,
Fribourg
,
S.
,
Moenner
,
M.
and
Chevet
,
E.
(
2011
).
Peptides derived from the bifunctional kinase/RNase enzyme IRE1alpha modulate IRE1alpha activity and protect cells from endoplasmic reticulum stress
.
FASEB J.
25
,
3115
-
3129
.
Brozzi
,
F.
,
Gerlo
,
S.
,
Grieco
,
F. A.
,
Nardelli
,
T. R.
,
Lievens
,
S.
,
Gysemans
,
C.
,
Marselli
,
L.
,
Marchetti
,
P.
,
Mathieu
,
C.
,
Tavernier
,
J.
, et al. 
(
2014
).
A combined “omics” approach identifies N-Myc interactor as a novel cytokine-induced regulator of IRE1alpha protein and c-Jun N-terminal kinase in pancreatic beta cells
.
J. Biol. Chem.
289
,
20677
-
20693
.
Brozzi
,
F.
,
Gerlo
,
S.
,
Grieco
,
F. A.
,
Juusola
,
M.
,
Balhuizen
,
A.
,
Lievens
,
S.
,
Gysemans
,
C.
,
Bugliani
,
M.
,
Mathieu
,
C.
,
Marchetti
,
P.
, et al. 
(
2016
).
Ubiquitin D regulates IRE1alpha/c-Jun N-terminal kinase (JNK) protein-dependent apoptosis in pancreatic beta cells
.
J. Biol. Chem.
291
,
12040
-
12056
.
Cai
,
L.
,
Marshall
,
T. W.
,
Uetrecht
,
A. C.
,
Schafer
,
D. A.
and
Bear
,
J. E.
(
2007
).
Coronin 1B coordinates Arp2/3 complex and cofilin activities at the leading edge
.
Cell
128
,
915
-
929
.
Calfon
,
M.
,
Zeng
,
H.
,
Urano
,
F.
,
Till
,
J. H.
,
Hubbard
,
S. R.
,
Harding
,
H. P.
,
Clark
,
S. G.
and
Ron
,
D.
(
2002
).
IRE1 couples endoplasmic reticulum load to secretory capacity by processing the XBP-1 mRNA
.
Nature
415
,
92
-
96
.
Carrara
,
M.
,
Prischi
,
F.
,
Nowak
,
P. R.
,
Kopp
,
M. C.
and
Ali
,
M. M.
(
2015
).
Noncanonical binding of BiP ATPase domain to Ire1 and Perk is dissociated by unfolded protein CH1 to initiate ER stress signaling
.
eLife
4
,
e03522
.
Carreras-Sureda
,
A.
,
Jana
,
F.
,
Urra
,
H.
,
Durand
,
S.
,
Mortenson
,
D. E.
,
Sagredo
,
A.
,
Bustos
,
G.
,
Hazari
,
Y.
,
Ramos-Fernandez
,
E.
,
Sassano
,
M. L.
, et al. 
(
2019
).
Non-canonical function of IRE1alpha determines mitochondria-associated endoplasmic reticulum composition to control calcium transfer and bioenergetics
.
Nat. Cell Biol.
21
,
755
-
767
.
Castillo
,
K.
,
Rojas-Rivera
,
D.
,
Lisbona
,
F.
,
Caballero
,
B.
,
Nassif
,
M.
,
Court
,
F. A.
,
Schuck
,
S.
,
Ibar
,
C.
,
Walter
,
P.
,
Sierralta
,
J.
, et al. 
(
2011
).
BAX inhibitor-1 regulates autophagy by controlling the IRE1alpha branch of the unfolded protein response
.
EMBO J.
30
,
4465
-
4478
.
Chang
,
T. K.
,
Lawrence
,
D. A.
,
Lu
,
M.
,
Tan
,
J.
,
Harnoss
,
J. M.
,
Marsters
,
S. A.
,
Liu
,
P.
,
Sandoval
,
W.
,
Martin
,
S. E.
and
Ashkenazi
,
A.
(
2018
).
Coordination between two branches of the unfolded protein response determines apoptotic cell fate
.
Mol. Cell
71
,
629
-
636.e5
.
Chowdhury
,
D.
,
Keogh
,
M. C.
,
Ishii
,
H.
,
Peterson
,
C. L.
,
Buratowski
,
S.
and
Lieberman
,
J.
(
2005
).
gamma-H2AX dephosphorylation by protein phosphatase 2A facilitates DNA double-strand break repair
.
Mol. Cell
20
,
801
-
809
.
Credle
,
J. J.
,
Finer-Moore
,
J. S.
,
Papa
,
F. R.
,
Stroud
,
R. M.
and
Walter
,
P.
(
2005
).
On the mechanism of sensing unfolded protein in the endoplasmic reticulum
.
Proc. Natl. Acad. Sci. USA
102
,
18773
-
18784
.
Csordás
,
G.
,
Várnai
,
P.
,
Golenár
,
T.
,
Roy
,
S.
,
Purkins
,
G.
,
Schneider
,
T. G.
,
Balla
,
T.
and
Hajnóczky
,
G.
(
2010
).
Imaging interorganelle contacts and local calcium dynamics at the ER-mitochondrial interface
.
Mol. Cell
39
,
121
-
132
.
Dalton
,
R. P.
,
Karagöz
,
G. E.
,
Kahiapo
,
J.
,
Sharma
,
R.
,
Bashkirova
,
L. E.
,
Lyons
,
D. B.
,
Matsunami
,
H.
and
Walter
,
P.
(
2018
).
Olfactory and vomeronasal receptor feedback employ divergent mechanisms of PERK Activation
.
bioRxiv
,
239830
.
Dubnikov
,
T.
,
Ben-Gedalya
,
T.
and
Cohen
,
E.
(
2017
).
Protein quality control in health and disease
.
Cold Spring Harb. Perspect. Biol.
9
,
a023523
.
Dufey
,
E.
,
Bravo-San Pedro
,
J. M.
,
Eggers
,
C.
,
González-Quiroz
,
M.
,
Urra
,
H.
,
Sagredo
,
A. I.
,
Sepulveda
,
D.
,
Pihán
,
P.
,
Carreras-Sureda
,
A.
,
Hazari
,
Y.
, et al. 
(
2020
).
Genotoxic stress triggers the activation of IRE1alpha-dependent RNA decay to modulate the DNA damage response
.
Nat. Commun.
11
,
2401
.
Eletto
,
D.
,
Dersh
,
D.
,
Gidalevitz
,
T.
and
Argon
,
Y.
(
2014
).
Protein disulfide isomerase A6 controls the decay of IRE1alpha signaling via disulfide-dependent association
.
Mol. Cell
53
,
562
-
576
.
Eletto
,
D.
,
Boyle
,
S.
and
Argon
,
Y.
(
2016
).
PDIA6 regulates insulin secretion by selectively inhibiting the RIDD activity of IRE1
.
FASEB J.
30
,
653
-
665
.
Estornes
,
Y.
,
Aguileta
,
M. A.
,
Dubuisson
,
C.
,
De Keyser
,
J.
,
Goossens
,
V.
,
Kersse
,
K.
,
Samali
,
A.
,
Vandenabeele
,
P.
and
Bertrand
,
M. J.
(
2014
).
RIPK1 promotes death receptor-independent caspase-8-mediated apoptosis under unresolved ER stress conditions
.
Cell Death Dis.
5
,
e1555
.
Fox
,
J. W.
,
Lamperti
,
E. D.
,
Eksioğlu
,
Y. Z.
,
Hong
,
S. E.
,
Feng
,
Y.
,
Graham
,
D. A.
,
Scheffer
,
I. E.
,
Dobyns
,
W. B.
,
Hirsch
,
B. A.
,
Radtke
,
R. A.
, et al. 
(
1998
).
Mutations in filamin 1 prevent migration of cerebral cortical neurons in human periventricular heterotopia
.
Neuron
21
,
1315
-
1325
.
Gao
,
B.
,
Lee
,
S. M.
,
Chen
,
A.
,
Zhang
,
J.
,
Zhang
,
D. D.
,
Kannan
,
K.
,
Ortmann
,
R. A.
and
Fang
,
D.
(
2008
).
Synoviolin promotes IRE1 ubiquitination and degradation in synovial fibroblasts from mice with collagen-induced arthritis
.
EMBO Rep.
9
,
480
-
485
.
Gardner
,
B. M.
and
Walter
,
P.
(
2011
).
Unfolded proteins are Ire1-activating ligands that directly induce the unfolded protein response
.
Science
333
,
1891
-
1894
.
Ghosh
,
R.
,
Wang
,
L.
,
Wang
,
E. S.
,
Perera
,
B. G.
,
Igbaria
,
A.
,
Morita
,
S.
,
Prado
,
K.
,
Thamsen
,
M.
,
Caswell
,
D.
,
Macias
,
H.
, et al. 
(
2014
).
Allosteric inhibition of the IRE1alpha RNase preserves cell viability and function during endoplasmic reticulum stress
.
Cell
158
,
534
-
548
.
Giacomello
,
M.
,
Drago
,
I.
,
Bortolozzi
,
M.
,
Scorzeto
,
M.
,
Gianelle
,
A.
,
Pizzo
,
P.
and
Pozzan
,
T.
(
2010
).
Ca2+ hot spots on the mitochondrial surface are generated by Ca2+ mobilization from stores, but not by activation of store-operated Ca2+ channels
.
Mol. Cell
38
,
280
-
290
.
Giorgi
,
C.
,
Marchi
,
S.
and
Pinton
,
P.
(
2018
).
The machineries, regulation and cellular functions of mitochondrial calcium
.
Nat. Rev. Mol. Cell Biol.
19
,
713
-
730
.
Groenendyk
,
J.
,
Peng
,
Z.
,
Dudek
,
E.
,
Fan
,
X.
,
Mizianty
,
M. J.
,
Dufey
,
E.
,
Urra
,
H.
,
Sepulveda
,
D.
,
Rojas-Rivera
,
D.
,
Lim
,
Y.
, et al. 
(
2014
).
Interplay between the oxidoreductase PDIA6 and microRNA-322 controls the response to disrupted endoplasmic reticulum calcium homeostasis
.
Sci. Signal.
7
,
ra54
.
Gu
,
F.
,
Nguyên
,
D. T.
,
Stuible
,
M.
,
Dubé
,
N.
,
Tremblay
,
M. L.
and
Chevet
,
E.
(
2004
).
Protein-tyrosine phosphatase 1B potentiates IRE1 signaling during endoplasmic reticulum stress
.
J. Biol. Chem.
279
,
49689
-
49693
.
Gupta
,
S.
,
Deepti
,
A.
,
Deegan
,
S.
,
Lisbona
,
F.
,
Hetz
,
C.
and
Samali
,
A.
(
2010
).
HSP72 protects cells from ER stress-induced apoptosis via enhancement of IRE1alpha-XBP1 signaling through a physical interaction
.
PLoS Biol.
8
,
e1000410
.
Halbleib
,
K.
,
Pesek
,
K.
,
Covino
,
R.
,
Hofbauer
,
H. F.
,
Wunnicke
,
D.
,
Hänelt
,
I.
,
Hummer
,
G.
and
Ernst
,
R.
(
2017
).
Activation of the unfolded protein response by lipid bilayer stress
.
Mol. Cell
67
,
673
-
684.e8
.
Han
,
J.
,
Back
,
S. H.
,
Hur
,
J.
,
Lin
,
Y.-H.
,
Gildersleeve
,
R.
,
Shan
,
J.
,
Yuan
,
C. L.
,
Krokowski
,
D.
,
Wang
,
S.
,
Hatzoglou
,
M.
, et al. 
(
2013
).
ER-stress-induced transcriptional regulation increases protein synthesis leading to cell death
.
Nat. Cell Biol.
15
,
481
-
490
.
Harding
,
H. P.
,
Novoa
,
I.
,
Zhang
,
Y.
,
Zeng
,
H.
,
Wek
,
R.
,
Schapira
,
M.
and
Ron
,
D.
(
2000a
).
Regulated translation initiation controls stress-induced gene expression in mammalian cells
.
Mol. Cell
6
,
1099
-
1108
.
Harding
,
H. P.
,
Zhang
,
Y.
,
Bertolotti
,
A.
,
Zeng
,
H.
and
Ron
,
D.
(
2000b
).
Perk is essential for translational regulation and cell survival during the unfolded protein response
.
Mol. Cell
5
,
897
-
904
.
Hart
,
L. S.
,
Cunningham
,
J. T.
,
Datta
,
T.
,
Dey
,
S.
,
Tameire
,
F.
,
Lehman
,
S. L.
,
Qiu
,
B.
,
Zhang
,
H.
,
Cerniglia
,
G.
,
Bi
,
M.
, et al. 
(
2012
).
ER stress-mediated autophagy promotes Myc-dependent transformation and tumor growth
.
J. Clin. Invest.
122
,
4621
-
4634
.
Hayashi
,
T.
and
Su
,
T.-P.
(
2007
).
Sigma-1 receptor chaperones at the ER-mitochondrion interface regulate Ca2+ signaling and cell survival
.
Cell
131
,
596
-
610
.
Haze
,
K.
,
Yoshida
,
H.
,
Yanagi
,
H.
,
Yura
,
T.
and
Mori
,
K.
(
1999
).
Mammalian transcription factor ATF6 is synthesized as a transmembrane protein and activated by proteolysis in response to endoplasmic reticulum stress
.
Mol. Biol. Cell
10
,
3787
-
3799
.
He
,
Y.
,
Beatty
,
A.
,
Han
,
X.
,
Ji
,
Y.
,
Ma
,
X.
,
Adelstein
,
R. S.
,
Yates
,
J. R.
, III
,
Kemphues
,
K.
and
Qi
,
L.
(
2012
).
Nonmuscle myosin IIB links cytoskeleton to IRE1alpha signaling during ER stress
.
Dev. Cell
23
,
1141
-
1152
.
Hetz
,
C.
(
2012
).
The unfolded protein response: controlling cell fate decisions under ER stress and beyond
.
Nat. Rev. Mol. Cell Biol.
13
,
89
-
102
.
Hetz
,
C.
and
Glimcher
,
L. H.
(
2009
).
Fine-tuning of the unfolded protein response: assembling the IRE1alpha interactome
.
Mol. Cell
35
,
551
-
561
.
Hetz
,
C.
and
Papa
,
F. R.
(
2018
).
The unfolded protein response and cell fate control
.
Mol. Cell
69
,
169
-
181
.
Hetz
,
C.
,
Bernasconi
,
P.
,
Fisher
,
J.
,
Lee
,
A. H.
,
Bassik
,
M. C.
,
Antonsson
,
B.
,
Brandt
,
G. S.
,
Iwakoshi
,
N. N.
,
Schinzel
,
A.
,
Glimcher
,
L. H.
, et al. 
(
2006
).
Proapoptotic BAX and BAK modulate the unfolded protein response by a direct interaction with IRE1alpha
.
Science
312
,
572
-
576
.
Hetz
,
C.
,
Thielen
,
P.
,
Matus
,
S.
,
Nassif
,
M.
,
Court
,
F.
,
Kiffin
,
R.
,
Martinez
,
G.
,
Cuervo
,
A. M.
,
Brown
,
R. H.
and
Glimcher
,
L. H.
(
2009
).
XBP-1 deficiency in the nervous system protects against amyotrophic lateral sclerosis by increasing autophagy
.
Genes Dev.
23
,
2294
-
2306
.
Hetz
,
C.
,
Martinon
,
F.
,
Rodriguez
,
D.
and
Glimcher
,
L. H.
(
2011
).
The unfolded protein response: integrating stress signals through the stress sensor IRE1α
.
Physiol. Rev.
91
,
1219
-
1243
.
Hetz
,
C.
,
Zhang
,
K.
and
Kaufman
,
R. J.
(
2020
).
Mechanisms, regulation and functions of the unfolded protein response
.
Nat. Rev. Mol. Cell Biol.
21
,
421
-
438
.
Hirokawa
,
N.
(
2006
).
mRNA transport in dendrites: RNA granules, motors, and tracks
.
J. Neurosci.
26
,
7139
-
7142
.
Hollien
,
J.
and
Weissman
,
J. S.
(
2006
).
Decay of endoplasmic reticulum-localized mRNAs during the unfolded protein response
.
Science
313
,
104
-
107
.
Hollien
,
J.
,
Lin
,
J. H.
,
Li
,
H.
,
Stevens
,
N.
,
Walter
,
P.
and
Weissman
,
J. S.
(
2009
).
Regulated Ire1-dependent decay of messenger RNAs in mammalian cells
.
J. Cell Biol.
186
,
323
-
331
.
Hoyer
,
M. J.
,
Chitwood
,
P. J.
,
Ebmeier
,
C. C.
,
Striepen
,
J. F.
,
Qi
,
R. Z.
,
Old
,
W. M.
and
Voeltz
,
G. K.
(
2018
).
A novel class of ER membrane proteins regulates ER-associated endosome fission
.
Cell
175
,
254
-
265.e14
.
Hu
,
P.
,
Han
,
Z.
,
Couvillon
,
A. D.
,
Kaufman
,
R. J.
and
Exton
,
J. H.
(
2006
).
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.
26
,
3071
-
3084
.
Huang
,
C.
,
Wu
,
S.
,
Ji
,
H.
,
Yan
,
X.
,
Xie
,
Y.
,
Murai
,
S.
,
Zhao
,
H.
,
Miyagishi
,
M.
and
Kasim
,
V.
(
2017
).
Identification of XBP1-u as a novel regulator of the MDM2/p53 axis using an shRNA library
.
Sci. Adv.
3
,
e1701383
.
Hwang
,
J.
and
Qi
,
L.
(
2018
).
Quality control in the endoplasmic reticulum: crosstalk between ERAD and UPR pathways
.
Trends Biochem. Sci.
43
,
593
-
605
.
Ishiwata-Kimata
,
Y.
,
Yamamoto
,
Y. H.
,
Takizawa
,
K.
,
Kohno
,
K.
and
Kimata
,
Y.
(
2013
).
F-actin and a type-II myosin are required for efficient clustering of the ER stress sensor Ire1
.
Cell Struct. Funct.
38
,
135
-
143
.
Jha
,
S.
,
Shibata
,
E.
and
Dutta
,
A.
(
2008
).
Human Rvb1/Tip49 is required for the histone acetyltransferase activity of Tip60/NuA4 and for the downregulation of phosphorylation on H2AX after DNA damage
.
Mol. Cell. Biol.
28
,
2690
-
2700
.
Jurkin
,
J.
,
Henkel
,
T.
,
Nielsen
,
A. F.
,
Minnich
,
M.
,
Popow
,
J.
,
Kaufmann
,
T.
,
Heindl
,
K.
,
Hoffmann
,
T.
,
Busslinger
,
M.
and
Martinez
,
J.
(
2014
).
The mammalian tRNA ligase complex mediates splicing of XBP1 mRNA and controls antibody secretion in plasma cells
.
EMBO J.
33
,
2922
-
2936
.
Jwa
,
M.
and
Chang
,
P.
(
2012
).
PARP16 is a tail-anchored endoplasmic reticulum protein required for the PERK- and IRE1alpha-mediated unfolded protein response
.
Nat. Cell Biol.
14
,
1223
-
1230
.
Karagöz
,
G. E.
,
Acosta-Alvear
,
D.
,
Nguyen
,
H. T.
,
Lee
,
C. P.
,
Chu
,
F.
and
Walter
,
P.
(
2017
).
An unfolded protein-induced conformational switch activates mammalian IRE1
.
eLife
6
,
e30700
.
Karagöz
,
G. E.
,
Acosta-Alvear
,
D.
and
Walter
,
P.
(
2019
).
The unfolded protein response: detecting and responding to fluctuations in the protein-folding capacity of the endoplasmic reticulum
.
Cold Spring Harb. Perspect. Biol.
11
,
a033886
.
Katayama
,
T.
,
Imaizumi
,
K.
,
Sato
,
N.
,
Miyoshi
,
K.
,
Kudo
,
T.
,
Hitomi
,
J.
,
Morihara
,
T.
,
Yoneda
,
T.
,
Gomi
,
F.
,
Mori
,
Y.
, et al. 
(
1999
).
Presenilin-1 mutations downregulate the signalling pathway of the unfolded-protein response
.
Nat. Cell Biol.
1
,
479
-
485
.
Kaushik
,
S.
,
Bandyopadhyay
,
U.
,
Sridhar
,
S.
,
Kiffin
,
R.
,
Martinez-Vicente
,
M.
,
Kon
,
M.
,
Orenstein
,
S. J.
,
Wong
,
E.
and
Cuervo
,
A. M.
(
2011
).
Chaperone-mediated autophagy at a glance
.
J. Cell Sci.
124
,
495
-
499
.
Kim
,
H.
and
McCulloch
,
C. A.
(
2011
).
Filamin A mediates interactions between cytoskeletal proteins that control cell adhesion
.
FEBS Lett.
585
,
18
-
22
.
Kishino
,
A.
,
Hayashi
,
K.
,
Hidai
,
C.
,
Masuda
,
T.
,
Nomura
,
Y.
and
Oshima
,
T.
(
2017
).
XBP1-FoxO1 interaction regulates ER stress-induced autophagy in auditory cells
.
Sci. Rep.
7
,
4442
.
Kiviluoto
,
S.
,
Vervliet
,
T.
,
Ivanova
,
H.
,
Decuypere
,
J. P.
,
De Smedt
,
H.
,
Missiaen
,
L.
,
Bultynck
,
G.
and
Parys
,
J. B.
(
2013
).
Regulation of inositol 1,4,5-trisphosphate receptors during endoplasmic reticulum stress
.
Biochim. Biophys. Acta
1833
,
1612
-
1624
.
Kono
,
N.
,
Amin-Wetzel
,
N.
and
Ron
,
D.
(
2017
).
Generic membrane-spanning features endow IRE1alpha with responsiveness to membrane aberrancy
.
Mol. Biol. Cell
28
,
2318
-
2332
.
Kopp
,
M. C.
,
Nowak
,
P. R.
,
Larburu
,
N.
,
Adams
,
C. J.
and
Ali
,
M. M.
(
2018
).
In vitro FRET analysis of IRE1 and BiP association and dissociation upon endoplasmic reticulum stress
.
eLife
7
,
e30257
.
Kosmaczewski
,
S. G.
,
Edwards
,
T. J.
,
Han
,
S. M.
,
Eckwahl
,
M. J.
,
Meyer
,
B. I.
,
Peach
,
S.
,
Hesselberth
,
J. R.
,
Wolin
,
S. L.
and
Hammarlund
,
M.
(
2014
).
The RtcB RNA ligase is an essential component of the metazoan unfolded protein response
.
EMBO Rep.
15
,
1278
-
1285
.
Kulak
,
N. A.
,
Pichler
,
G.
,
Paron
,
I.
,
Nagaraj
,
N.
and
Mann
,
M.
(
2014
).
Minimal, encapsulated proteomic-sample processing applied to copy-number estimation in eukaryotic cells
.
Nat. Methods
11
,
319
-
324
.
Lebeaupin
,
C.
,
Vallée
,
D.
,
Rousseau
,
D.
,
Patouraux
,
S.
,
Bonnafous
,
S.
,
Adam
,
G.
,
Luciano
,
F.
,
Luci
,
C.
,
Anty
,
R.
,
Iannelli
,
A.
, et al. 
(
2018
).
Bax inhibitor-1 protects from nonalcoholic steatohepatitis by limiting inositol-requiring enzyme 1 alpha signaling in mice
.
Hepatology
68
,
515
-
532
.
Lee
,
K.
,
Tirasophon
,
W.
,
Shen
,
X.
,
Michalak
,
M.
,
Prywes
,
R.
,
Okada
,
T.
,
Yoshida
,
H.
,
Mori
,
K.
and
Kaufman
,
R. J.
(
2002
).
IRE1-mediated unconventional mRNA splicing and S2P-mediated ATF6 cleavage merge to regulate XBP1 in signaling the unfolded protein response
.
Genes Dev.
16
,
452
-
466
.
Li
,
J.
,
Dallmayer
,
M.
,
Kirchner
,
T.
,
Musa
,
J.
and
Grünewald
,
T. G. P.
(
2018
).
PRC1: linking cytokinesis, chromosomal instability, and cancer evolution
.
Trends Cancer
4
,
59
-
73
.
Lisbona
,
F.
,
Rojas-Rivera
,
D.
,
Thielen
,
P.
,
Zamorano
,
S.
,
Todd
,
D.
,
Martinon
,
F.
,
Glavic
,
A.
,
Kress
,
C.
,
Lin
,
J. H.
,
Walter
,
P.
, et al. 
(
2009
).
BAX inhibitor-1 is a negative regulator of the ER stress sensor IRE1alpha
.
Mol. Cell
33
,
679
-
691
.
Liu
,
J.
,
Wang
,
Y.
,
Song
,
L.
,
Zeng
,
L.
,
Yi
,
W.
,
Liu
,
T.
,
Chen
,
H.
,
Wang
,
M.
,
Ju
,
Z.
and
Cong
,
Y. S.
(
2017
).
A critical role of DDRGK1 in endoplasmic reticulum homoeostasis via regulation of IRE1alpha stability
.
Nat. Commun.
8
,
14186
.
Lu
,
G.
,
Ota
,
A.
,
Ren
,
S.
,
Franklin
,
S.
,
Rau
,
C. D.
,
Ping
,
P.
,
Lane
,
T. F.
,
Zhou
,
Z. H.
,
Reue
,
K.
,
Lusis
,
A. J.
, et al. 
(
2013
).
PPM1l encodes an inositol requiring-protein 1 (IRE1) specific phosphatase that regulates the functional outcome of the ER stress response
.
Mol. Metab.
2
,
405
-
416
.
Lu
,
Y.
,
Liang
,
F. X.
and
Wang
,
X.
(
2014
).
A synthetic biology approach identifies the mammalian UPR RNA ligase RtcB
.
Mol. Cell
55
,
758
-
770
.
Luo
,
D.
,
He
,
Y.
,
Zhang
,
H.
,
Yu
,
L.
,
Chen
,
H.
,
Xu
,
Z.
,
Tang
,
S.
,
Urano
,
F.
and
Min
,
W.
(
2008
).
AIP1 is critical in transducing IRE1-mediated endoplasmic reticulum stress response
.
J. Biol. Chem.
283
,
11905
-
11912
.
Lynch
,
C. D.
and
Sheetz
,
M. P.
(
2011
).
Cellular mechanotransduction: filamin a strains to regulate motility
.
Curr. Biol.
21
,
R916
-
R918
.
Mancias
,
J. D.
and
Goldberg
,
J.
(
2007
).
The transport signal on Sec22 for packaging into COPII-coated vesicles is a conformational epitope
.
Mol. Cell
26
,
403
-
414
.
Mao
,
T.
,
Shao
,
M.
,
Qiu
,
Y.
,
Huang
,
J.
,
Zhang
,
Y.
,
Song
,
B.
,
Wang
,
Q.
,
Jiang
,
L.
,
Liu
,
Y.
,
Han
,
J.-D. J.
, et al. 
(
2011
).
PKA phosphorylation couples hepatic inositol-requiring enzyme 1alpha to glucagon signaling in glucose metabolism
.
Proc. Natl. Acad. Sci. USA
108
,
15852
-
15857
.
Marciniak
,
S. J.
,
Yun
,
C. Y.
,
Oyadomari
,
S.
,
Novoa
,
I.
,
Zhang
,
Y.
,
Jungreis
,
R.
,
Nagata
,
K.
,
Harding
,
H. P.
and
Ron
,
D.
(
2004
).
CHOP induces death by promoting protein synthesis and oxidation in the stressed endoplasmic reticulum
.
Genes Dev.
18
,
3066
-
3077
.
Marcu
,
M. G.
,
Doyle
,
M.
,
Bertolotti
,
A.
,
Ron
,
D.
,
Hendershot
,
L.
and
Neckers
,
L.
(
2002
).
Heat shock protein 90 modulates the unfolded protein response by stabilizing IRE1alpha
.
Mol. Cell. Biol.
22
,
8506
-
8513
.
Martínez
,
G.
,
Vidal
,
R. L.
,
Mardones
,
P.
,
Serrano
,
F. G.
,
Ardiles
,
A. O.
,
Wirth
,
C.
,
Valdés
,
P.
,
Thielen
,
P.
,
Schneider
,
B. L.
,
Kerr
,
B.
, et al. 
(
2016
).
Regulation of memory formation by the transcription factor XBP1
.
Cell Rep.
14
,
1382
-
1394
.
Martínez
,
G.
,
Khatiwada
,
S.
,
Costa-Mattioli
,
M.
and
Hetz
,
C.
(
2018
).
ER proteostasis control of neuronal physiology and synaptic function
.
Trends Neurosci.
41
,
610
-
624
.
Maurel
,
M.
,
Chevet
,
E.
,
Tavernier
,
J.
and
Gerlo
,
S.
(
2014
).
Getting RIDD of RNA: IRE1 in cell fate regulation
.
Trends Biochem. Sci.
39
,
245
-
254
.
Molinie
,
N.
,
Rubtsova
,
S. N.
,
Fokin
,
A.
,
Visweshwaran
,
S. P.
,
Rocques
,
N.
,
Polesskaya
,
A.
,
Schnitzler
,
A.
,
Vacher
,
S.
,
Denisov
,
E. V.
,
Tashireva
,
L. A.
, et al. 
(
2019
).
Cortical branched actin determines cell cycle progression
.
Cell Res.
29
,
432
-
445
.
Morello
,
L. G.
,
Coltri
,
P. P.
,
Quaresma
,
A. J.
,
Simabuco
,
F. M.
,
Silva
,
T. C.
,
Singh
,
G.
,
Nickerson
,
J. A.
,
Oliveira
,
C. C.
,
Moore
,
M. J.
and
Zanchin
,
N. I.
(
2011
).
The human nucleolar protein FTSJ3 associates with NIP7 and functions in pre-rRNA processing
.
PLoS ONE
6
,
e29174
.
Mori
,
T.
,
Hayashi
,
T.
,
Hayashi
,
E.
and
Su
,
T. P.
(
2013
).
Sigma-1 receptor chaperone at the ER-mitochondrion interface mediates the mitochondrion-ER-nucleus signaling for cellular survival
.
PLoS ONE
8
,
e76941
.
Morita
,
S.
,
Villalta
,
S. A.
,
Feldman
,
H. C.
,
Register
,
A. C.
,
Rosenthal
,
W.
,
Hoffmann-Petersen
,
I. T.
,
Mehdizadeh
,
M.
,
Ghosh
,
R.
,
Wang
,
L.
,
Colon-Negron
,
K.
, et al. 
(
2017
).
Targeting ABL-IRE1alpha signaling spares ER-stressed pancreatic beta cells to reverse autoimmune diabetes
.
Cell Metab.
25
,
883
-
897.e8
.
Nagai
,
A.
,
Kadowaki
,
H.
,
Maruyama
,
T.
,
Takeda
,
K.
,
Nishitoh
,
H.
and
Ichijo
,
H.
(
2009
).
USP14 inhibits ER-associated degradation via interaction with IRE1alpha
.
Biochem. Biophys. Res. Commun.
379
,
995
-
1000
.
Nakato
,
R.
,
Ohkubo
,
Y.
,
Konishi
,
A.
,
Shibata
,
M.
,
Kaneko
,
Y.
,
Iwawaki
,
T.
,
Nakamura
,
T.
,
Lipton
,
S. A.
and
Uehara
,
T.
(
2015
).
Regulation of the unfolded protein response via S-nitrosylation of sensors of endoplasmic reticulum stress
.
Sci. Rep.
5
,
14812
.
Nguyên
,
D. T.
,
Kebache
,
S.
,
Fazel
,
A.
,
Wong
,
H. N.
,
Jenna
,
S.
,
Emadali
,
A.
,
Lee
,
E. H.
,
Bergeron
,
J. J.
,
Kaufman
,
R. J.
,
Larose
,
L.
, et al. 
(
2004
).
Nck-dependent activation of extracellular signal-regulated kinase-1 and regulation of cell survival during endoplasmic reticulum stress
.
Mol. Biol. Cell
15
,
4248
-
4260
.
Nishitoh
,
H.
,
Matsuzawa
,
A.
,
Tobiume
,
K.
,
Saegusa
,
K.
,
Takeda
,
K.
,
Inoue
,
K.
,
Hori
,
S.
,
Kakizuka
,
A.
and
Ichijo
,
H.
(
2002
).
ASK1 is essential for endoplasmic reticulum stress-induced neuronal cell death triggered by expanded polyglutamine repeats
.
Genes Dev.
16
,
1345
-
1355
.
Oakes
,
S. A.
and
Papa
,
F. R.
(
2015
).
The role of endoplasmic reticulum stress in human pathology
.
Annu. Rev. Pathol.
10
,
173
-
194
.
Oikawa
,
D.
,
Kimata
,
Y.
,
Kohno
,
K.
and
Iwawaki
,
T.
(
2009
).
Activation of mammalian IRE1alpha upon ER stress depends on dissociation of BiP rather than on direct interaction with unfolded proteins
.
Exp. Cell Res.
315
,
2496
-
2504
.
Okamura
,
K.
,
Kimata
,
Y.
,
Higashio
,
H.
,
Tsuru
,
A.
and
Kohno
,
K.
(
2000
).
Dissociation of Kar2p/BiP from an ER sensory molecule, Ire1p, triggers the unfolded protein response in yeast
.
Biochem. Biophys. Res. Commun.
279
,
445
-
450
.
Oono
,
K.
,
Yoneda
,
T.
,
Manabe
,
T.
,
Yamagishi
,
S.
,
Matsuda
,
S.
,
Hitomi
,
J.
,
Miyata
,
S.
,
Mizuno
,
T.
,
Imaizumi
,
K.
,
Katayama
,
T.
, et al. 
(
2004
).
JAB1 participates in unfolded protein responses by association and dissociation with IRE1
.
Neurochem. Int.
45
,
765
-
772
.
Osorio
,
F.
,
Lambrecht
,
B. N.
and
Janssens
,
S.
(
2018
).
Antigen presentation unfolded: identifying convergence points between the UPR and antigen presentation pathways
.
Curr. Opin. Immunol.
52
,
100
-
107
.
Ota
,
A.
and
Wang
,
Y.
(
2012
).
Cdc37/Hsp90 protein-mediated regulation of IRE1alpha protein activity in endoplasmic reticulum stress response and insulin synthesis in INS-1 cells
.
J. Biol. Chem.
287
,
6266
-
6274
.
Pattingre
,
S.
,
Tassa
,
A.
,
Qu
,
X.
,
Garuti
,
R.
,
Liang
,
X. H.
,
Mizushima
,
N.
,
Packer
,
M.
,
Schneider
,
M. D.
and
Levine
,
B.
(
2005
).
Bcl-2 antiapoptotic proteins inhibit Beclin 1-dependent autophagy
.
Cell
122
,
927
-
939
.
Pattingre
,
S.
,
Bauvy
,
C.
,
Carpentier
,
S.
,
Levade
,
T.
,
Levine
,
B.
and
Codogno
,
P.
(
2009
).
Role of JNK1-dependent Bcl-2 phosphorylation in ceramide-induced macroautophagy
.
J. Biol. Chem.
284
,
2719
-
2728
.
Pinkaew
,
D.
,
Chattopadhyay
,
A.
,
King
,
M. D.
,
Chunhacha
,
P.
,
Liu
,
Z.
,
Stevenson
,
H. L.
,
Chen
,
Y.
,
Sinthujaroen
,
P.
,
McDougal
,
O. M.
and
Fujise
,
K.
(
2017
).
Fortilin binds IRE1alpha and prevents ER stress from signaling apoptotic cell death
.
Nat. Commun.
8
,
18
.
Plumb
,
R.
,
Zhang
,
Z. R.
,
Appathurai
,
S.
and
Mariappan
,
M.
(
2015
).
A functional link between the co-translational protein translocation pathway and the UPR
.
eLife
4
,
e07426
.
Preissler
,
S.
and
Ron
,
D.
(
2019
).
Early events in the endoplasmic reticulum unfolded protein response
.
Cold Spring Harb. Perspect. Biol.
11
,
a033894
.
Prischi
,
F.
,
Nowak
,
P. R.
,
Carrara
,
M.
and
Ali
,
M. M.
(
2014
).
Phosphoregulation of Ire1 RNase splicing activity
.
Nat. Commun.
5
,
3554
.
Qiu
,
Y.
,
Mao
,
T.
,
Zhang
,
Y.
,
Shao
,
M.
,
You
,
J.
,
Ding
,
Q.
,
Chen
,
Y.
,
Wu
,
D.
,
Xie
,
D.
,
Lin
,
X.
, et al. 
(
2010
).
A crucial role for RACK1 in the regulation of glucose-stimulated IRE1alpha activation in pancreatic beta cells
.
Sci. Signal.
3
,
ra7
.
Rashid
,
H. O.
,
Yadav
,
R. K.
,
Kim
,
H. R.
and
Chae
,
H. J.
(
2015
).
ER stress: autophagy induction, inhibition and selection
.
Autophagy
11
,
1956
-
1977
.
Ricci
,
D.
,
Marrocco
,
I.
,
Blumenthal
,
D.
,
Dibos
,
M.
,
Eletto
,
D.
,
Vargas
,
J.
,
Boyle
,
S.
,
Iwamoto
,
Y.
,
Chomistek
,
S.
,
Paton
,
J. C.
, et al. 
(
2019
).
Clustering of IRE1alpha depends on sensing ER stress but not on its RNase activity
.
FASEB J.
33
,
9811
-
9827
.
Rodriguez
,
D. A.
,
Zamorano
,
S.
,
Lisbona
,
F.
,
Rojas-Rivera
,
D.
,
Urra
,
H.
,
Cubillos-Ruiz
,
J. R.
,
Armisen
,
R.
,
Henriquez
,
D. R.
,
Cheng
,
E. H.
,
Letek
,
M.
, et al. 
(
2012
).
BH3-only proteins are part of a regulatory network that control the sustained signalling of the unfolded protein response sensor IRE1alpha
.
EMBO J.
31
,
2322
-
2335
.
Rojas-Rivera
,
D.
,
Rodriguez
,
D. A.
,
Sepulveda
,
D.
and
Hetz
,
C.
(
2018
).
ER stress sensing mechanism: putting off the brake on UPR transducers
.
Oncotarget
9
,
19461
-
19462
.
Saito
,
A.
,
Cai
,
L.
,
Matsuhisa
,
K.
,
Ohtake
,
Y.
,
Kaneko
,
M.
,
Kanemoto
,
S.
,
Asada
,
R.
and
Imaizumi
,
K.
(
2018
).
Neuronal activity-dependent local activation of dendritic unfolded protein response promotes expression of brain-derived neurotrophic factor in cell soma
.
J. Neurochem.
144
,
35
-
49
.
Sarkisian
,
M. R.
,
Bartley
,
C. M.
and
Rakic
,
P.
(
2008
).
Trouble making the first move: interpreting arrested neuronal migration in the cerebral cortex
.
Trends Neurosci.
31
,
54
-
61
.
Schwarz
,
D. S.
and
Blower
,
M. D.
(
2016
).
The endoplasmic reticulum: structure, function and response to cellular signaling
.
Cell. Mol. Life Sci.
73
,
79
-
94
.
Sepulveda
,
D.
,
Rojas-Rivera
,
D.
,
Rodríguez
,
D. A.
,
Groenendyk
,
J.
,
Köhler
,
A.
,
Lebeaupin
,
C.
,
Ito
,
S.
,
Urra
,
H.
,
Carreras-Sureda
,
A.
,
Hazari
,
Y.
, et al. 
(
2018
).
Interactome screening identifies the ER luminal chaperone Hsp47 as a regulator of the unfolded protein response transducer IRE1alpha
.
Mol. Cell
69
,
238
-
252.e7
.
Shemorry
,
A.
,
Harnoss
,
J. M.
,
Guttman
,
O.
,
Marsters
,
S. A.
,
Kőműves
,
L. G.
,
Lawrence
,
D. A.
and
Ashkenazi
,
A.
(
2019
).
Caspase-mediated cleavage of IRE1 controls apoptotic cell commitment during endoplasmic reticulum stress
.
eLife
8
,
e47084
.
Shen
,
J.
,
Chen
,
X.
,
Hendershot
,
L.
and
Prywes
,
R.
(
2002
).
ER stress regulation of ATF6 localization by dissociation of BiP/GRP78 binding and unmasking of Golgi localization signals
.
Dev. Cell
3
,
99
-
111
.
Shin
,
G. C.
,
Moon
,
S. U.
,
Kang
,
H. S.
,
Choi
,
H. S.
,
Han
,
H. D.
and
Kim
,
K. H.
(
2019
).
PRKCSH contributes to tumorigenesis by selective boosting of IRE1 signaling pathway
.
Nat. Commun.
10
,
3185
.
Sun
,
S.
,
Shi
,
G.
,
Sha
,
H.
,
Ji
,
Y.
,
Han
,
X.
,
Shu
,
X.
,
Ma
,
H.
,
Inoue
,
T.
,
Gao
,
B.
,
Kim
,
H.
, et al. 
(
2015
).
IRE1alpha is an endogenous substrate of endoplasmic-reticulum-associated degradation
.
Nat. Cell Biol.
17
,
1546
-
1555
.
Sundaram
,
A.
,
Plumb
,
R.
,
Appathurai
,
S.
and
Mariappan
,
M.
(
2017
).
The Sec61 translocon limits IRE1alpha signaling during the unfolded protein response
.
eLife
6
,
e27187
.
Tabas
,
I.
and
Ron
,
D.
(
2011
).
Integrating the mechanisms of apoptosis induced by endoplasmic reticulum stress
.
Nat. Cell Biol.
13
,
184
-
190
.
Taguchi
,
Y.
,
Imaoka
,
K.
,
Kataoka
,
M.
,
Uda
,
A.
,
Nakatsu
,
D.
,
Horii-Okazaki
,
S.
,
Kunishige
,
R.
,
Kano
,
F.
and
Murata
,
M.
(
2015
).
Yip1A, a novel host factor for the activation of the IRE1 pathway of the unfolded protein response during Brucella infection
.
PLoS Pathog.
11
,
e1004747
.
Takeda
,
K.
,
Nagashima
,
S.
,
Shiiba
,
I.
,
Uda
,
A.
,
Tokuyama
,
T.
,
Ito
,
N.
,
Fukuda
,
T.
,
Matsushita
,
N.
,
Ishido
,
S.
,
Iwawaki
,
T.
, et al. 
(
2019
).
MITOL prevents ER stress-induced apoptosis by IRE1alpha ubiquitylation at ER-mitochondria contact sites
.
EMBO J.
38
,
e100999
.
Tam
,
A. B.
,
Roberts
,
L. S.
,
Chandra
,
V.
,
Rivera
,
I. G.
,
Nomura
,
D. K.
,
Forbes
,
D. J.
and
Niwa
,
M.
(
2018
).
The UPR activator ATF6 responds to proteotoxic and lipotoxic stress by distinct mechanisms
.
Dev. Cell
46
,
327
-
343.e7
.
Tang
,
C. H.
,
Chang
,
S.
,
Paton
,
A. W.
,
Paton
,
J. C.
,
Gabrilovich
,
D. I.
,
Ploegh
,
H. L.
,
Del Valle
,
J. R.
and
Hu
,
C. C.
(
2018
).
Phosphorylation of IRE1 at S729 regulates RIDD in B cells and antibody production after immunization
.
J. Cell Biol.
217
,
1739
-
1755
.
Tavernier
,
Q.
,
Bennana
,
E.
,
Poindessous
,
V.
,
Schaeffer
,
C.
,
Rampoldi
,
L.
,
Pietrancosta
,
N.
and
Pallet
,
N.
(
2018
).
Regulation of IRE1 RNase activity by the ribonuclease inhibitor 1 (RNH1)
.
Cell Cycle
17
,
1901
-
1916
.
Tschurtschenthaler
,
M.
,
Adolph
,
T. E.
,
Ashcroft
,
J. W.
,
Niederreiter
,
L.
,
Bharti
,
R.
,
Saveljeva
,
S.
,
Bhattacharyya
,
J.
,
Flak
,
M. B.
,
Shih
,
D. Q.
,
Fuhler
,
G. M.
, et al. 
(
2017
).
Defective ATG16L1-mediated removal of IRE1alpha drives Crohn's disease-like ileitis
.
J. Exp. Med.
214
,
401
-
422
.
Urano
,
F.
,
Wang
,
X.
,
Bertolotti
,
A.
,
Zhang
,
Y.
,
Chung
,
P.
,
Harding
,
H. P.
and
Ron
,
D.
(
2000
).
Coupling of stress in the ER to activation of JNK protein kinases by transmembrane protein kinase IRE1
.
Science
287
,
664
-
666
.
Urra
,
H.
,
Dufey
,
E.
,
Lisbona
,
F.
,
Rojas-Rivera
,
D.
and
Hetz
,
C.
(
2013
).
When ER stress reaches a dead end
.
Biochim. Biophys. Acta
1833
,
3507
-
3517
.
Urra
,
H.
,
Henriquez
,
D. R.
,
Cánovas
,
J.
,
Villarroel-Campos
,
D.
,
Carreras-Sureda
,
A.
,
Pulgar
,
E.
,
Molina
,
E.
,
Hazari
,
Y. M.
,
Limia
,
C. M.
,
Alvarez-Rojas
,
S.
, et al. 
(
2018
).
IRE1alpha governs cytoskeleton remodelling and cell migration through a direct interaction with filamin A
.
Nat. Cell Biol.
20
,
942
-
953
.
van Niel
,
G.
,
D'Angelo
,
G.
and
Raposo
,
G.
(
2018
).
Shedding light on the cell biology of extracellular vesicles
.
Nat. Rev. Mol. Cell Biol.
19
,
213
-
228
.
van Vliet
,
A. R.
,
Giordano
,
F.
,
Gerlo
,
S.
,
Segura
,
I.
,
Van Eygen
,
S.
,
Molenberghs
,
G.
,
Rocha
,
S.
,
Houcine
,
A.
,
Derua
,
R.
,
Verfaillie
,
T.
, et al. 
(
2017
).
The ER stress sensor PERK coordinates ER-plasma membrane contact site formation through interaction with filamin-A and F-actin remodeling
.
Mol. Cell
65
,
885
-
899.e6
.
Vattem
,
K. M.
and
Wek
,
R. C.
(
2004
).
Reinitiation involving upstream ORFs regulates ATF4 mRNA translation in mammalian cells
.
Proc. Natl. Acad. Sci. USA
101
,
11269
-
11274
.
Verchot
,
J.
(
2016
).
How does the stressed out ER find relief during virus infection?
Curr. Opin. Virol.
17
,
74
-
79
.
Vidal
,
R. L.
,
Figueroa
,
A.
,
Court
,
F. A.
,
Thielen
,
P.
,
Molina
,
C.
,
Wirth
,
C.
,
Caballero
,
B.
,
Kiffin
,
R.
,
Segura-Aguilar
,
J.
,
Cuervo
,
A. M.
, et al. 
(
2012
).
Targeting the UPR transcription factor XBP1 protects against Huntington's disease through the regulation of FoxO1 and autophagy
.
Hum. Mol. Genet.
21
,
2245
-
2262
.
Volmer
,
R.
,
van der Ploeg
,
K.
and
Ron
,
D.
(
2013
).
Membrane lipid saturation activates endoplasmic reticulum unfolded protein response transducers through their transmembrane domains
.
Proc. Natl. Acad. Sci. USA
110
,
4628
-
4633
.
Walter
,
P.
and
Ron
,
D.
(
2011
).
The unfolded protein response: from stress pathway to homeostatic regulation
.
Science
334
,
1081
-
1086
.
Wang
,
M.
and
Kaufman
,
R. J.
(
2016
).
Protein misfolding in the endoplasmic reticulum as a conduit to human disease
.
Nature
529
,
326
-
335
.
Wang
,
L.
,
Perera
,
B. G. K.
,
Hari
,
S. B.
,
Bhhatarai
,
B.
,
Backes
,
B. J.
,
Seeliger
,
M. A.
,
Schürer
,
S. C.
,
Oakes
,
S. A.
,
Papa
,
F. R.
and
Maly
,
D. J.
(
2012
).
Divergent allosteric control of the IRE1alpha endoribonuclease using kinase inhibitors
.
Nat. Chem. Biol.
8
,
982
-
989
.
Wang
,
J.-M.
,
Qiu
,
Y.
,
Yang
,
Z.
,
Kim
,
H.
,
Qian
,
Q.
,
Sun
,
Q.
,
Zhang
,
C.
,
Yin
,
L.
,
Fang
,
D.
,
Back
,
S. H.
, et al. 
(
2018a
).
IRE1alpha prevents hepatic steatosis by processing and promoting the degradation of select microRNAs
.
Sci. Signal.
11
,
eaao4617
.
Wang
,
P.
,
Li
,
J.
,
Tao
,
J.
and
Sha
,
B.
(
2018b
).
The luminal domain of the ER stress sensor protein PERK binds misfolded proteins and thereby triggers PERK oligomerization
.
J. Biol. Chem.
293
,
4110
-
4121
.
Weatherbee
,
J. L.
,
Kraus
,
J. L.
and
Ross
,
A. H.
(
2016
).
ER stress in temozolomide-treated glioblastomas interferes with DNA repair and induces apoptosis
.
Oncotarget
7
,
43820
-
43834
.
Yamamori
,
T.
,
Meike
,
S.
,
Nagane
,
M.
,
Yasui
,
H.
and
Inanami
,
O.
(
2013
).
ER stress suppresses DNA double-strand break repair and sensitizes tumor cells to ionizing radiation by stimulating proteasomal degradation of Rad51
.
FEBS Lett.
587
,
3348
-
3353
.
Yamamoto
,
K.
,
Sato
,
T.
,
Matsui
,
T.
,
Sato
,
M.
,
Okada
,
T.
,
Yoshida
,
H.
,
Harada
,
A.
and
Mori
,
K.
(
2007
).
Transcriptional induction of mammalian ER quality control proteins is mediated by single or combined action of ATF6alpha and XBP1
.
Dev. Cell
13
,
365
-
376
.
Yanagitani
,
K.
,
Imagawa
,
Y.
,
Iwawaki
,
T.
,
Hosoda
,
A.
,
Saito
,
M.
,
Kimata
,
Y.
and
Kohno
,
K.
(
2009
).
Cotranslational targeting of XBP1 protein to the membrane promotes cytoplasmic splicing of its own mRNA
.
Mol. Cell
34
,
191
-
200
.
Yanagitani
,
K.
,
Kimata
,
Y.
,
Kadokura
,
H.
and
Kohno
,
K.
(
2011
).
Translational pausing ensures membrane targeting and cytoplasmic splicing of XBP1u mRNA
.
Science
331
,
586
-
589
.
Yang
,
L.
,
Calay
,
E. S.
,
Fan
,
J.
,
Arduini
,
A.
,
Kunz
,
R. C.
,
Gygi
,
S. P.
,
Yalcin
,
A.
,
Fu
,
S.
and
Hotamisligil
,
G. S.
(
2015
).
METABOLISM. S-Nitrosylation links obesity-associated inflammation to endoplasmic reticulum dysfunction
.
Science
349
,
500
-
506
.
Yoneda
,
T.
,
Imaizumi
,
K.
,
Oono
,
K.
,
Yui
,
D.
,
Gomi
,
F.
,
Katayama
,
T.
and
Tohyama
,
M.
(
2001
).
Activation of caspase-12, an endoplastic reticulum (ER) resident caspase, through tumor necrosis factor receptor-associated factor 2-dependent mechanism in response to the ER stress
.
J. Biol. Chem.
276
,
13935
-
13940
.
Yoshida
,
H.
,
Matsui
,
T.
,
Yamamoto
,
A.
,
Okada
,
T.
and
Mori
,
K.
(
2001
).
XBP1 mRNA is induced by ATF6 and spliced by IRE1 in response to ER stress to produce a highly active transcription factor
.
Cell
107
,
881
-
891
.
Yu
,
J.
,
Li
,
T.
,
Liu
,
Y.
,
Wang
,
X.
,
Zhang
,
J.
,
Wang
,
X.
,
Shi
,
G.
,
Lou
,
J.
,
Wang
,
L.
,
Wang
,
C.-C.
, et al. 
(
2020
).
Phosphorylation switches protein disulfide isomerase activity to maintain proteostasis and attenuate ER stress
.
EMBO J.
39
,
e103841
.
Zhou
,
J.
,
Liu
,
C. Y.
,
Back
,
S. H.
,
Clark
,
R. L.
,
Peisach
,
D.
,
Xu
,
Z.
and
Kaufman
,
R. J.
(
2006
).
The crystal structure of human IRE1 luminal domain reveals a conserved dimerization interface required for activation of the unfolded protein response
.
Proc. Natl. Acad. Sci. USA
103
,
14343
-
14348
.
Zhou
,
A.-X.
,
Hartwig
,
J. H.
and
Akyürek
,
L. M.
(
2010
).
Filamins in cell signaling, transcription and organ development
.
Trends Cell Biol.
20
,
113
-
123
.
Zhu
,
X.
,
Zhang
,
J.
,
Sun
,
H.
,
Jiang
,
C.
,
Dong
,
Y.
,
Shan
,
Q.
,
Su
,
S.
,
Xie
,
Y.
,
Xu
,
N.
,
Lou
,
X.
, et al. 
(
2014
).
Ubiquitination of inositol-requiring enzyme 1 (IRE1) by the E3 ligase CHIP mediates the IRE1/TRAF2/JNK pathway
.
J. Biol. Chem.
289
,
30567
-
30577
.

Competing interests

The authors declare no competing or financial interests.

Supplementary information