Abstract
The endoplasmic reticulum (ER) is a cellular organelle that is physiologically responsible for protein folding, calcium homeostasis, and lipid biosynthesis. Pathological stimuli such as oxidative stress, ischemia, disruptions in calcium homeostasis, and increased production of normal and/or folding-defective proteins all contribute to the accumulation of misfolded proteins in the ER, causing ER stress. The adaptive response to ER stress is the activation of unfolded protein response (UPR), which affect a wide variety of cellular functions to maintain ER homeostasis or lead to apoptosis. Three different ER transmembrane sensors, including PKR-like ER kinase (PERK), activating transcription factor 6 (ATF6), and inositol-requiring enzyme-1 (IRE1), are responsible for initiating UPR. The UPR involves a variety of signal transduction pathways that reduce unfolded protein accumulation by boosting ER-resident chaperones, limiting protein translation, and accelerating unfolded protein degradation. ER is now acknowledged as a critical organelle in sensing dangers and determining cell life and death. On the other hand, UPR plays a critical role in the development and progression of several diseases such as cardiovascular diseases (CVD), metabolic disorders, chronic kidney diseases, neurological disorders, and cancer. Here, we critically analyze the most current knowledge of the master regulatory roles of ER stress particularly the PERK pathway as a conditional danger receptor, an organelle crosstalk regulator, and a regulator of protein translation. We highlighted that PERK is not only ER stress regulator by sensing UPR and ER stress but also a frontier sensor and direct senses for gut microbiota-generated metabolites. Our work also further highlighted the function of PERK as a central hub that leads to metabolic reprogramming and epigenetic modification which further enhanced inflammatory response and promoted trained immunity. Moreover, we highlighted the contribution of ER stress and PERK in the pathogenesis of several diseases such as cancer, CVD, kidney diseases, and neurodegenerative disorders. Finally, we discuss the therapeutic target of ER stress and PERK for cancer treatment and the potential novel therapeutic targets for CVD, metabolic disorders, and neurodegenerative disorders. Inhibition of ER stress, by the development of small molecules that target the PERK and UPR, represents a promising therapeutic strategy.
Keywords: unfolded protein response (UPR), ER stress, ER-mitochondrial contact sites, protein kinase R-like endoplasmic reticulum kinase (PERK), membraneless stress organelles
Introduction
The endoplasmic reticulum (ER) is a large vital intracellular organelle [1, 2] that contains the nuclear envelope, ER cisternae, and an interconnected tubular network [3]. ER has a folding enzymes, multifunctional integral membrane and luminal chaperones, as well as sensor molecules such as glucose-related protein 78 kDa (GRP78), also known as binding immunoglobulin protein (BiP), glucose-related protein 94 kDa (GRP94), inositol-1,4,5-trisphosphate (InsP3) receptor/calcium release channel, protein disulfide isomerases (PDI), lectins such as calreticulin and calnexin, sarco-ER calcium-ATPase (SERCA) pump, ryanodine receptors (RYRs), and ER-associated calcium sensors stromal interacting molecule (STIM) proteins [4]. ER provides an optimal environment for fundamental biological processes such as protein synthesis, protein folding, protein modification, ER retention, trafficking, and canonical secretome (secretory proteins with signal peptide) [5–7], as well as lipid metabolism and steroid synthesis [8], and calcium homeostasis and storage, making it an essential organelle for cell survival [9]. It has been reported that 11% of the 25,000 predicted human full-length open reading frames (ORFs) are secreted proteins, and 20% of single or multi-pass transmembrane proteins are trafficked into the ER lumen by their N-terminal signal sequences [10]. When a newly synthesized protein enters the ER, it undergoes a series of modifications and encounters a variety of molecular chaperones and folding enzymes, all of which aid in its proper folding and eventual release from the ER. ER dysfunction and the generation of large amounts of substances such as reactive oxygen species (ROS) [11] and calcium ions under stress conditions and disease risk factors/danger-associated molecular patterns (DAMPs) stimulation lead to reduced protein folding and the accumulation of unfolded or misfolded proteins in the ER lumen, which causes ER stress [12]. If the ER stress is mild and the inducers are temporary, the excess ER membranes and proteins are removed by autophagy in a process of ER remodeling and selective recycling called Reticulophagy to reduce ER stress [13]. However, the accumulation of unfolded proteins represents a key barrier in the secretory apparatus and is detrimental to cellular and organismal homeostasis. Thus, an evolutionarily conserved cellular checkpoint mechanism called the unfolding protein response serves to sense and promote adaptation or cellular execution in response to unfolding proteins [14].
The unfolded protein response (UPR) is a cellular stress response triggered by the accumulation of unfolded proteins in the ER lumen in eukaryotic cells. The UPR is an adaptive signaling pathway which is highly controlled and integrates information about the type, intensity, and duration of the stress stimuli, therefore determines the fate of cells. Three transmembrane protein sensors regulate the UPR, including protein kinase R-like endoplasmic reticulum kinase (PERK) [15], inositol-requiring kinase 1 (IRE1) [16], and activating transcription factor 6 (ATF6) [17], which, in response to the accumulation of unfolded and misfolded proteins, determine cellular fate. In unstressed cells, these sensors are associated with GRP78/BiP, which consistently and dynamically binds and arrests those proteins in an inactive status in the ER membrane [18]. However, in response to ER stress, those ER resident sensors are dissociated from BiP, undergo conformational changes, and activate the UPR and downstream signaling (Figure 1).
Figure 1. ER stress regulators IRE1 and ATF6 signaling are mediated by mRNA processing and proteolytic processing, respectively, whereas PERK signaling is mediated by translational and posttranslational mechanisms.

Under normal conditions, the three proteins (sensors), including inositol-requiring enzyme 1 (IRE1), activating transcription factor 6 (ATF6), and protein kinase RNA (PKR) like ER kinase (PERK), bind to the molecular chaperone protein glucose-related protein 78 kDa (GRP78)/binding immunoglobulin protein (BiP) and inhibit their activation. Under stress conditions, GRP78/BiP dissociates from the three ER stress sensors and binds to unfolded proteins, leading to the activation of the three sensors. Each activation pathway has a different signal transduction mechanism. (A) IRE1 pathway: IRE1α splices X-box binding protein 1 (XBP1) mRNA to encode for the transcription factor XBP1s, which promotes the expression of genes involved in protein folding and induces chaperones, endoplasmic reticulum-associated protein degradation (ERAD), and lipid synthesis. (B) ATF6 pathway: under ER stress, ATF6 translocates from the ER to the Golgi apparatus. ATF6 is then cleaved to produce the amino terminus of ATF6 (ATF6-N), which migrates to the nucleus to increase the transcription activity of XBP1 and regulate ERAD and lipid metabolism. (C) PERK pathway: PERK (eukaryotic translation initiation factor 2 alpha kinase 3, EIF2AK3) undergoes oligomerization and auto-phosphorylation, which then promotes the phosphorylation of the eukaryotic initiation factor 2 alpha subunit (eIF2α), resulting in a general inhibition of protein translation. However, phosphorylated eIF2α selectively increases the translation of activating transcription factor 4 (ATF4), which upregulates CHOP and GADD34 mRNA. Created with BioRender.com
The main objective of the UPR is to maintain cellular homeostasis and restore ER function [19]. However, if ER stress is persistent or cellular homeostasis cannot be reestablished, the UPR promotes cell death [20]. The UPR can be induced by pathophysiological conditions such as viral infection, ischemia, hyperhomocysteinemia, hypoxia, hypoglycemia, proteotoxicity, excessive oxidative stress, and changes in membrane lipid composition [21–25] or experimentally by agents that decrease the folding capacity of the ER (e.g., tunicamycin, which inhibits glycosylation of asparagine residues, and thapsigargin, which inhibits SERCA (sarco/ER-Ca2+-ATPase), thereby reducing ER calcium)
(i). IRE1 pathway.
IRE1 is a type I transmembrane protein, identical to PERK. It has serine/threonine kinase and endoribonuclease (RNase) activities in the cytosol. Under non-stressful circumstances, the IRE1α cytosolic domain is bound to heat shock proteins 90 (HSP90) and 72 (HSP72). After ER stress, unfolded proteins bind to BiP and release IRE1α, which is autophosphorylated and rapidly activated. The activated IRE1α activates and removes the non-coding sequence of the X box-binding protein 1 (XBP1) gene [26], which regulates protein folding and trafficking, ER-associated degradation (ERAD), autophagy, and lipid biosynthesis. ERAD is a quality control system that regulates the degradation of unfolded and misfolded proteins and targets them for clearance [27]. It is also responsible for the regulation of the endogenous levels of physiologically important proteins such as 3-hydroxy-3-methylglutaryl coenzyme-A (HMG-CoA) reductase, a rate-limiting enzyme for cholesterol biosynthesis [28]. In addition, when the IRE1 signal is activated, the tumor necrosis factor receptor-associated factor 2 (TRAF2) and IRE1 combine to form a complex and activate the apoptosis signal-regulatory kinase 1 (ASK1), the c-Jun N-terminal kinase (JNK) [29], and the nuclear factor kappa B (NF-κB) [30, 31], which trigger cell death [9].
(ii). ATF6 pathway.
The leucine zipper protein transcription factor, ATF6, is a type-II transmembrane receptor with its C-terminal domain located in the ER lumen and its N-terminal DNA-binding domain located in the cytoplasm [32]. under non-stressful conditions, ATF6 is constitutively localized in the ER and binds to BiP. Under conditions of ER stress, ATF6 separates from BiP and moves into the Golgi apparatus. There, it’s cleaved by site-1 protease (S1P) and site-2 protease (S2P), releasing functional ATF6 fragments into the cytoplasm and transferring them into the nucleus to start transcription [33]. In the nucleus, the cleaved ATF6 induces the expression of genes involved in protein folding, components of ERAD, protein secretion, and lipid biogenesis. Furthermore, ATF6 induces the expression of BiP, XBP1, and CCAAT-enhancer binding protein (C/EBP) homologous protein (CHOP) [34].
(iii). PERK pathway.
PERK (1116 amino acids, https://www.ncbi.nlm.nih.gov/protein/NP_004827.4), also referred to as eukaryotic translation initiation factor-2α kinase 3 (eIF2AK3) [35], is a major sensor ER-resident type-I transmembrane protein that is a member of serine/threonine kinase family. Its N-terminal luminal domain (LD) is responsive to the upstream ER stress signal, whereas its C-terminal cytoplasmic domain directly phosphorylates eukaryotic translation initiation factor 2 alpha (eIF2α) and linked to cytosolic kinases (kinomes) [15]. Six PERK structural domains include the N-terminal signal peptide (amino acids (aa) 1–29), N-terminal ER luminal eIF2AK3 domain (aa 104–420), transmembrane domain (aa 515–535), cytosolic PKC-like domain (aa 587–663), cytosolic serine-threonine kinase domain (STKc-EIF2AK3-PERK, aa873–1075), and four cytosolic phosphorylation sites (aa 619, 715, 982, and 1094) (https://www.ncbi.nlm.nih.gov/protein/NP_004827.4). Under normal conditions, PERK binds to BiP to exist as an inactive monomer. PERK is activated as a homodimer under pathological conditions such as hypoxia and oxidative stress by separating from BiP, which enables PERK oligomerization and activation. Phosphorylated and activated PERK phosphorylates the downstream mediators (substrates), including eIF2α and nuclear factor erythroid-derived 2-like 2 (Nrf2), impairing global protein translation [35, 36]. Meanwhile, phosphorylated eIF2α induces the selective translation of activating transcription factor 4 (ATF4) and further activates the transcription of downstream UPR target genes [37, 38]. ATF4 has dual functions: under mild to moderate ER stress, it is responsible for the restoration of cellular homeostasis, but under severe and prolonged ER stress conditions, it induces the activation of CHOP-mediated apoptosis [39–41].
Recent studies have identified new PERK substrates, including lipids that signal the second messenger diacylglycerol (DAG) [42] and protein substrates such as the forkhead box O protein (FOXO) [43]. We recently reported that PERK activates 12 more kinases in the cytosol, including cAMP response element-binding protein kinase (CREB), p38 mitogen-activated protein kinase (MAPK) α (P38αMAPK), mammalian target of rapamycin (mTOR), extracellular signal-regulated kinase (ERK1/2MAPK), c-Jun N-terminal kinase (JNK1/2/3MAPK), protein kinase B (AKT1/2/3), protein kinase AMP-activated catalytic subunit alpha 1 (AMPKα1, PRKAA1), AMPKα2 (PRKAA2), glycogen synthase-3 (GSK-3α/β), mitogen- and stress-activated kinases (MSK1/2), platelet-derived growth factor receptor β (PDGFRβ), and epidermal growth factor receptor (EGFR) [15]. The control of the UPR and its activity under ER stress is greatly dependent on the PERK protein pathway, and loss of PERK activity severely affects the ability of cells to withstand ER stress. Because of its many branching pathways, PERK has been implicated in the development of several diseases [39]. PERK has 18 disease-related pathways, including many discussed in this review (https://www.ncbi.nlm.nih.gov/gene/9451). The significance of the PERK pathway has also been highlighted by findings from human genetic diseases. For example, Wolcott-Rallison syndrome has been identified as a PERK mutation (https://databases.lovd.nl/shared/genes/EIF2AK3), which is an uncommon autosomal recessive disease that causes permanent neonatal or early infancy insulin-dependent diabetes (https://omim.org/entry/226980). Additional human phenotype ontology includes abnormalities of body height, kidney, urinary system, and renal insufficiency (https://www.genecards.org/cgi-bin/carddisp.pl?gene=EIF2AK3&keywords=EIF2AK3). Moreover, the results from gene-deficient mice have also significantly enhanced our understanding. PERK knockout cells are hypersensitive to the lethal effects of ER stress-inducing toxins such as thapsigargin (a sarcoplasmic and ER Ca2+-ATPase (SERCA) inhibitor, leading to ER calcium depletion and increased cytosolic calcium concentrations) and tunicamycin (a protein glycosylation inhibitor), which cause ER stress by reducing the protein folding capacity of the ER [44]. In addition to serving as a secondary (indirect) DAMP sensor for sensing ER stress induced by endogenous DAMPs such as ROS [11] and calcium ions under stress conditions, PERK can also become the primary (or frontier) DAMP sensor to directly sense exogenous pathogen-associated molecular patterns (PAMPs). Recently, Dr. Biddinger’s team, our team, and others reported that the gut microbiota-generated uremic toxin trimethylamine-N-oxide (TMAO) binds to the N-terminal luminal domain (LD) of PERK and activates PERK [15, 45–48].
There are several studies that indicate that the ER stress-related signaling pathways could serve as a promising target for developing innovative therapeutic approaches [49]. These approaches may help to decrease the adaptation of cancer cells to hypoxia, inflammation, and angiogenesis, thereby overcoming adaptation-related drug resistance [50]. The PERK-mediated UPR signaling pathways have a crucial role in promoting adaptation and survival during ER stress conditions resulting from hypoxia. Therefore, it should be the future target of anticancer therapeutic intervention [51]. In summary, the role of ER stress, particularly the PERK pathway, in several diseases, including CVDs, neurodegenerative diseases, and cancers, has been reviewed previously; however, the role of a protein-rich diet, a seafood diet, and gut microbiota-generated uremic toxins TMAO in PERK activation was not highlighted in detail in the previous reviews. In addition, how PERK activation affects the innate immune response and further exacerbates the inflammatory response was not discussed in previous publications. Therefore, in this review, we summarize the functions of the ER stress sensors, particularly PERK [52] as a conditional DAMP [53, 54] receptor for gut microbiota-generated uremic toxin TMAO [47], and focus on its role in the regulation of protein translation and organelle crosstalk regulation [1]. We highlight that PERK is not only an ER stress regulator by sensing UPR and ER stress but also a frontier sensor that directly senses gut microbiota-generated metabolites. We further highlight the function of PERK as a central hub that leads to metabolic reprogramming and epigenetic modification [55], which further enhance the inflammatory response [56, 57] and promote trained immunity (innate immune memory) [58–63], innate immunity [64, 65], cell death [66–68], reticulophagy (ER autophagy) [69, 70], and fibrogenesis [71, 72]. Moreover, we highlight the contribution of ER stress and PERK in the pathogenesis of several diseases, including cardiovascular diseases [72–74], inflammations [52], chronic kidney disease [75], Alzheimer’s disease and neurodegenerative diseases [76–78], type II diabetes [79], and cancers [8, 80–85]. Finally, we discuss the therapeutic targets of ER stress and PERK for cancer treatment and potential novel therapeutic targets for cardiovascular disease, metabolic disorders, chronic kidney diseases, and neurodegenerative disorders. Inhibition of ER stress/UPR activation and the PERK pathway by the development of small molecules that target the PERK and UPR represents a promising therapeutic strategy.
ER stress sensor PERK is a conditional danger receptor for the microbiota-generated uremic toxin TMAO.
Pathogen-associated molecular patterns (PAMPs) and DAMPs are molecules associated with groups of pathogens [86] and elevated endogenous metabolites [48, 53] and have an exceptionally high potential for triggering inflammatory responses [86–88]. Several types of molecules can serve as PAMPs, including lipopolysaccharides (LPS), endotoxins found on the cell membranes of gram-negative bacteria [89], which are considered to be the prototypical class of PAMPs and specifically recognized by toll-like receptor-4 (TLR4), a receptor of the innate immune system [60, 61, 86, 90]. DAMPs, also known as alarmins, are endogenous danger signals released by stressed cells that induce and amplify the immunological and inflammatory response [91]. These DAMPs include heat-shock proteins, high-mobility group box 1 (HMGB1) [5], and reactive oxygen intermediates [92]. DAMPs can be recognized by classical DAMP receptors, including Toll-like receptors (TLRs) and other pattern-recognition receptors. Nucleotide-binding oligomerization domain (NOD)-like receptors (NLRs) play essential roles in innate immunity by detecting intracellular PAMPs/DAMPs [87, 93–95]. Other DAMP receptors have been identified, including the receptor for advanced glycation end products (RAGE), a receptor for HMGB1, a retinoid acid-inducible gene I (RIG-I), transmembrane C-type lectin receptors, and the absent in melanoma 2 (AIM2) [96, 97]. Our team previously reported that the uremic toxins serve as conditional DAMPs and homeostasis-associated molecular patterns (HAMPs) that modulate inflammation [48, 54, 98]. In addition, we identified lysophospholipid receptors, G-protein-coupled receptors (GPCRs), as novel conditional DAMP receptors for the endogenous metabolite lysophospholipid [53, 99]. We also proposed that metabolism-associated danger signal (MADS) recognition is a novel mechanism for metabolic risk factor-induced inflammatory responses [100] and that the homocysteine-methionine (HM) cycle is a key metabolic sensor system controlling methylation-regulated pathological signaling [101]. Furthermore, we proposed the new concept that ROS systems create an integrated metabolic sensor network for sensing metabolic stresses and metabolic homeostasis [11, 102].
ER stress responses result in upregulation and ligand-independent activation of tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) receptors 1 and 2 (also known as DR4 and DR5), leading to a FAS-associated death domain protein (FADD)/caspase-8/receptor-interacting serine/threonine protein kinase 1 (RIPK1) (FADDosome)-dependent pathway to NF-κB activation and inflammatory cytokine production, implying that TRAIL receptors behave as “stress-associated molecular patterns (SAMPs)” that link ER stress to NF-κB-dependent inflammation [103]. Interestingly, we and others recently reported that gut microbiota-generated uremic toxin TMAO is a conditional DAMP that binds to and activates its intrinsic receptor PERK rather than commonly shared DAMP receptors to initiate pathological signaling [15, 47], which is the first DAMP receptor localized in the ER [97]. TMAO is significantly increased in pathological conditions such as end-stage renal disease (ESRD), and induced endothelial cell dysfunction, and inflammation [104, 105]. Protein-rich, sea food, and gut microbiota-generated metabolites [46] has been linked to ER stress response [106]. In conclusion, conditional DAMPs, including lysophospholipids and uremic toxins, have several important features. They are (a) endogenous metabolites; (b) pathologically elevated; (c) have a physiological signaling function; and (d) bind to a unique receptor and amplify the signals rather than shared DAMP receptors such as TLRs and NLRs. As far as we are aware, our study was the first to demonstrate that PERK, as the sensor for UPR/ER response and mitochondrial UPR [107] and also as a sensor/receptor for gut microbiota-generated uremic toxin TMAO, serves as a significant driver for metabolic reprogramming and trained immunity (also termed innate immune memory) [15, 47, 60, 108].
ER stress sensors are organelle crosstalk regulators, enriched in MAMs, and interact with MAM-tethering proteins to regulate metabolite transfer and affect mitochondrial unfolded protein responses through inter-organelle signaling.
The ER has multiple contact sites with every membrane-bound organelle, including mitochondria, plasma membrane, Golgi, endosomes, and peroxisomes [1, 3]. Connections between the ER and other organelles are required for the integration of cellular signals originating from other cellular compartments. The most extensively studied structural connectivity is the junction between ER-mitochondria and the ER-plasma membrane. Mitochondria are highly sensitive to a variety of cellular insults, including unfolded proteins in the ER, calcium, lipids, and ROS [109–115], and the exchange of these metabolites from the ER to the mitochondria causes mitochondrial dysfunction, accumulation of mitochondrial unfolded protein response (UPRmt) [116] and determines ultimate cell fate [107]. The sites of ER that are in contact with mitochondria are known as the mitochondria-associated ER membranes (MAM). Recently, MAMs have received much attention due to their functional significance in organelle communication at membrane contact sites. It is tightly tethered by different multifunctional proteins that integrally regulate lipid metabolism, calcium signaling, and ROS homeostasis [117, 118].
Several studies have shown that different mitochondrial proteins involved in tethering can regulate the ER stress and UPR pathways (Figure 2), including mitochondrial fusion protein mitofusin 2 (MFN2), inositol triphosphate receptor (IP3R), voltage-dependent anion-selective channel protein 1 (VDAC1), GRP75, vesicle-associated membrane protein-associated protein B (VAPB), protein tyrosine phosphatase interacting protein 51 (PTPIP51), mitochondrial fission 1 (FIS1), and B-cell receptor-associated protein 31 (BAP31) [119]. At the contact sites, calcium is released through the ER channels to the mitochondrial membrane, and this requires VDAC1 interaction with IP3R [3]. Other mediators that facilitate the interactions at the contact sites include the ER-mitochondria encounter structures (ERMES) complex identified in yeast and MFN2 in mammalian cells [120], which regulate the shape of ER-controlling ER-mitochondrial contacts. In addition to the accumulation of unfolded proteins, metabolite imbalance can activate PERK in the lumen of the ER, and sustained activation of PERK leads to mitochondrial dysfunction [35, 121]. Furthermore, PERK kinases communicate with mitochondria to regulate oxidative stress at the MAMs [1, 122]. As the second tier of ER-mitochondrial tethering, PERK is MAM-rich components that interact with multifunctional MAM tethering proteins (the first tier) and integrally regulate the exchange of lipids, ROS, and metabolites such as calcium at the contact sites [123]. To maintain ER-mitochondrial connections, PERK interacts with MFN2 [122]. This structural effect on MAMs appears to require the cytosolic kinase domain of PERK, as it does not occur because of PERK downstream signaling of eIF2 [122]. A new report showed that PERK recruits E-Syt1 at ER-mitochondrial contacts for mitochondrial lipid transport, respiration [124], and mitochondrial membrane biogenesis [125]. In addition to being a tethering protein, MFN2 regulates PERK functions involved in mitochondrial fusion, inflammation, and the effective uptake of calcium at the ER-mitochondrial contact sites [126–128]. Thus, PERK functions as a crucial UPR sensor to mediate the crosstalk between the ER and mitochondria [129]. MAM dysfunctions may be involved in the pathophysiology of a variety of diseases, including metabolic and neurodegenerative disorders [130, 131].
Figure 2. ER-mitochondria and ER-plasma membrane contact sites derive their close apposition from tethering proteins and mediate inter-organelle signaling pathways.

Several pairs of integral membrane proteins located on the mitochondria, ER, and plasma membrane that are important for the physical tethering of the organelles were identified, reflecting and supporting the diverse functions of the MAMs and PAMs. (1) IP3R-GRP75-VDAC1 ternary binding complex: IP3R at the ER interacts with the mitochondrial channel VDAC1 owing to the chaperone GRP75, which interacts directly with the two proteins. At the ER, SIGMAR1 is chaperoning IP3R. This complex is essential for the efficient calcium transfer from the ER to the mitochondrial matrix through the mitochondrial Ca2+ uniporter (MCU) in the inner mitochondrial membrane. (2) VAPB-PTPIP51 tether: VAPB located at the ER membrane interacts with PTPIP51 located at the mitochondrial outer membrane. (3) BAP31-FIS1 tether: ER-resident protein BAP31 interacts with the mitochondrial fission protein FIS1. (4) MFN1/2 tether: in addition to its predominant location at the outer mitochondrial membrane, MFN2 can be found in the ER and interact with mitochondrial MFN-1/2. (5) PERK can also interact with mitochondrial MFN2. ER-PM contact sites have emerged as key regulators of intracellular calcium dynamics. Stromal-interacting molecule 1 (STIM1) molecule mainly exists as a dimer on the ER membrane that regulates the formation of ER-PM contact sites in response to calcium depletion in the ER lumen. Calcium depletion from the ER store induces dimerization or oligomerization of STIM1. The STIM1 dimer in the active state may directly couple to and activate the Orai1 channel on the PM, followed by replenishing calcium levels in the ER lumen. Created with BioRender.com
In addition, the contact sites between the ER and plasma membrane play a significant role in calcium exchange. The stromal-interacting molecule (STIM) proteins can sense a reduction in calcium within the ER. This leads to a conformational change in the STIM proteins, which then relocate tubular structures within the ER towards the ER-plasma membrane junctions. Ultimately, this results in the direct activation of the calcium release-activated calcium modulator (Orai), the pore-forming component of the calcium release-activated calcium (CRAC) channel. This activation initiates the channel opening and the subsequent calcium influx [132]. Furthermore, ER-plasma membrane contact sites play a crucial role in the metabolism of phosphatidylinositol, particularly in the regulation of phosphatidylinositol 4-phosphate (PI4P). This process is regulated by the oxysterol-binding homology (Osh) protein family and the ER membrane integral proteins, such as vesicle-associated membrane protein-associated protein (VAP), following Sac1-like phosphatidylinositide phosphatase (SACM1L) activation [3]. Interestingly, PERK has been shown to interact with filamin-A and F-actin remodeling to coordinate the establishment of ER-plasma membrane contact sites [133].
Furthermore, recent studies have identified the role of IRE1 in ER-mitochondrial communication by controlling ER-mitochondrial dynamics, mitochondrial calcium uptake, and cell survivability. IRE1 acts as a scaffold to stabilize IP3R at MAMS [134]. All these indicate a complex interrelation of IRE1 and PERK, two membrane-bound ER stress sensors, in controlling ER-mitochondrial communication, which suggests that PERK functions via ER-mitochondrial tethering are a new branch of PERK pathways in addition to cytosol kinase activation and transcription activation via ATF4. In addition, ER is not only a network hub that physically connects membrane organelles but also plays a fundamental role in the regulation of membraneless organelles [135].
ER stress pathways may facilitate the dynamics and formation of cytoplasmic and nuclear membraneless organelles.
Membraneless organelles are relatively new cellular structures that are not surrounded by a sealed phospholipid membrane and are formed through phase separation in either a liquid-liquid or liquid-solid manner [136]. Liquid-liquid phase separation can result in the production of stable liquid droplets within another liquid; however, liquid-solid phase separation can result in gel-like stress granules and solid and crystalline structures [137]. Membraneless organelles can be found in both the nucleus, such as the nucleolus, Cajal bodies (sub-nuclear structures often associated with the nucleolus), nuclear stress bodies, nuclear speckles, interchromatin granule clusters, paraspeckles, and transcription histone locus bodies, and in the cytoplasm, such as the centrosome, P-granules, pyrenoid, and processing-bodies (P-bodies) [138, 139] (Figure 3). Most constitutive membraneless organelles are regulated according to the cell cycle phase [140]. Membraneless organelles have crucial roles in cell physiology, but when they develop as a result of the production of mutant proteins, they might develop pathogenic characteristics such as amyloids [141]. Cellular stress leads to the formation of reversible membraneless organelles that are known as stress assemblies, among which the best studied are the P-bodies and stress granules [137, 142]. Environmental stressors trigger cellular signaling, resulting in the formation of stress granules. Stress granules can be formed in the nucleus or in the cytoplasm [143]. ER stress inhibits mRNA translation initiation and polysome disassembly, leading to the accumulation of untranslated 80S ribosome-free mRNA in the cytoplasm, which binds to RNA-binding proteins [144] and combines into membraneless foci called reversible stress granules [145]. It contains polyadenylated mRNAs, eukaryotic translation initiation factors such as eIF2A and eIF3, 40S ribosomes, and RNA-binding proteins such as Tia1 cytotoxic granule-associated RNA-binding protein and G3BP1 stress granule assembly factor 1, which are the main drivers for stress granule formation.
Figure 3. ER stress pathways may facilitate the dynamics and formation of cytoplasmic and nuclear membraneless organelles.

The membrane-bound organelles in the cells, including mitochondria, Golgi, nucleus, lysosomes, and endosomes, are labeled in green. A growing number of membraneless organelles have been identified in the cytoplasm and nucleus. The cytosolic membraneless organelles such as P-bodies, U bodies, P-granules, centrosomes, and pyrenoid and the nuclear membraneless organelles such as the nucleolus, nuclear speckles, paraspeckles, cajal bodies, and promyelocytic leukemia (PML) bodies are labeled in red. Under cellular stress, the formation of stress granules and enlarged P-bodies is induced. Created with BioRender.com
The heterogeneity in the development and composition of the stress granule pathway is depending on the cell type, specific stressors, and the signaling pathways that are activated. Stress granule formation can triggered by ER stress, oxidative stress, heat shock, hypoxia, malnutrition, the presence of translation-blocking medications, viral infection, knockdown of particular translation initiation factors, and overexpression of particular RNA-binding proteins [146]. There are two types of stress granules: canonical and non-canonical stress granules (Figure 4) [142]. The canonical stress granules can be triggered by a variety of cell-damaging stimuli, including ER stress (heat chock and thapsigargin) [147], oxidative stress (sodium arsenite) [148], nutrient deprivation, and proteasome inhibition (MG132) [149]. Particular stress conditions are detected by specific kinases, including heme-regulated initiation factor 2α kinase (HRI, or eIF2α kinase 1 (eIF2αK1)), protein kinase RNA-activated (PKR, eIF2αK2), PERK (EIF2AK3), and general control non-derepressible 2 (GCN2, or eIF2α kinase 4 (eIF2αK4)) [150], which then become activated and phosphorylate eIF2α. The phosphorylation of eIF2α resulted in the accumulation of several types of RNA-binding proteins, such as poly(A) binding protein cytoplasmic 1 (PAB1), ras GTPase-activating protein-binding protein 1 (G3BP1), Caprin family member 2, tristetraprolin (TTP), T-cell intracellular antigen-1 (TIA-1), fragile X messenger ribonucleoprotein 1 (FMR1), AU-rich element binding protein (HuR), TAR DNA binding protein (TDP-43), as well as polyadenylated mRNA, and specific RNAs, including long RNA, arginines and glycines (RG) rich motifs (induced by heat shock) [151], and adenylate/uridylate (AU)–rich motifs (induced by ER stress) [152]. A non-canonical stress granule pathway is independent of eIF2α phosphorylation [153]. It is dependent on how pateamine A (PatA) and hippuristanol (Hipp) inhibits eIF4A activity (an RNA helicase needed for the ribosome recruitment phase of translation initiation). This inhibits ribosome activity and stops translation initiation by preventing the formation of the 48S initiation complex [153]. It also formed upon exposure to ultraviolet (UV) and osmotic chock. This type of stress granule contains RNA-binding proteins and non-polyadenylated mRNAs. In addition to their active role in the stress response, stress granules participate in mRNA triage and expression regulation, cell signaling and apoptosis, and inhibition of viral replication. Taken together, this indicates that the induction of ER stress and PERK kinase pathway activation plays a significant role in inducing canonical stress granule formation.
Figure 4. Membrane organelles and membraneless organelles are both integrated to mediate cellular responses to stress. Stress granules are multimolecular cytoplasmic foci that assemble as part of the cellular response to stress in two different pathways.

Canonical stress granules are induced by activation and phosphorylation of eukaryotic initiation factor 2 alpha (eIF2α) by four kinases, including general control non-derepressible-2 (GCN2, or eIF2α kinase 4 (EIF2AK4)); pancreatic eIF2α kinase (PKR-like ER kinase (PERK) or eIF2α kinase 3 (EIF2AK3)); protein kinase R (PKR), and haem-regulated inhibitor (HRI); or eIF2α kinase 1 (EIF2AK1). Each kinase contains specific regulatory regions that detect and recognize different stress stimuli, including viral infection, oxidative stress, ER stress, starvation, and arsenite. It is mainly comprised of eukaryotic translation initiation factor 2 subunit alpha (eIF2α), polyadenylated (poly-A) mRNA molecules, long RNAs, AU and RG motifs, and a vast array of proteins, including RNA-binding proteins (RBPs) such as ras GTPase-activating protein-binding protein 1 (G3BP1), T-cell intracellular antigen-1 (TIA-1) and TIA1-related protein (TIAR), polyadenylate-binding protein (PABP), AU-rich element binding protein (HuR), ataxin-2, fragile X mental retardation protein (FMRP), and tristetraprolin (TTP). The non-canonical stress granules are dependent on the inhibition of the eukaryotic initiation factor 4 alpha (eIF4α) by pateamine A (PatA) and hippuristanol (Hipp), which prevents the formation of the 48S initiation complex, inhibits ribosome activity, and arrests translation initiation. These events lead to SG formation independently of eIF2α phosphorylation. Created with BioRender.com.
Regulation of protein translation in the ER stress response.
Protein synthesis is an essential process in cells and responds to dynamic cellular and environmental changes. Translational changes occur more quickly and are better able to adapt to changing environments than transcriptional changes [154]. The ER is the main site for folding and posttranslational modification of secreted and integral membrane proteins. Regulation of protein translation occurs mostly at the level of initiation, which is the rate-limiting step in protein synthesis. Protein translation depends on the formation of the 43S preinitiation complex, a 40S ribosomal subunit, and the translation initiation factors, including eukaryotic translation initiation factor 1 (eIF1), eIF1A, and eIF3. Two distinct pathways are defined in protein translational processes, including (i) the regulation of ternary complex formation by eIF2 phosphorylation at its conserved residue, serine 51 of its α subunit (eIF2α) and suppression of the polypeptide biosynthesis, and (ii) the regulation of the eIF4F complex via the mammalian target of rapamycin complex 1 (mTORC1) [155, 156]. One of the most important regulatory functions of PERK is its role as a regulator of protein translation. Once PERK is activated by ER stress, it phosphorylates eIF2α. eIF2α regulates the binding of the methionyl tRNA to the small ribosomal subunit. Phosphorylation of eIF2α prevents the recycling of eIF2α to its active GTP-bound form and inhibits the delivery of the initiator Met-tRNAi to the translation initiation complex, thereby reducing the rate of general protein translation (global translation inhibition) and affecting the translation of both cytoplasmic and ER client proteins [157]. The formation of stress granules is dependent on eIF2α phosphorylation under certain conditions [158]. Meanwhile, PERK activation and eIF2α phosphorylation result in the translational upregulation of selected proteins, including ATF4 (PERK downstream transcription factor), because of the upstream short open reading frame (uORF) [35, 159–161]. PERK-dependent reduction of protein translation restricts the trafficking of nascent proteins into the ER lumen, thereby reducing potential chaperone loading and allowing chaperones to clear misfolded aggregates. In response to ER stress, PERK knockout cells are hypersensitive to the lethal doses of toxins and lose their ability to regulate protein translation associated with hyperactivation of the IRE1 pathway, implying that these cells experience more ER stress [44] and that PERK acts upstream of the IRE1 pathway. PERK knockout animals develop progressive destruction of multiple cell types normally responsible for high protein secretion. The most notable destructions are the endocrine and exocrine cells of the pancreas and the collagen I-secreting osteoblasts, which lead to diabetes mellitus and severe bone defects [162, 163].
ER stress sensor PERK activation promotes trained immunity, an implication for disease interaction.
Trained immunity is quit novel concept. DAMP receptors signaling can modulate immunity. Innate immune cells develop an enhanced long-term immune response and inflammatory phenotype after brief exposure to endogenous or exogenous DAMPs. As a result, an inflammatory response is induced and greatly enhanced after exposure to the second challenge, and this phenomenon is known as “innate immune memory or trained immunity” [15, 58, 60–63, 90, 108, 164]. ER stress profoundly affects the innate and adaptive immune responses and is associated with autoimmune and inflammatory diseases, including diabetes mellitus and atherosclerosis [60, 61, 64, 165]. Upon activation, the ER stress response activates immune signaling pathways, including signal transducer and activator of transcription (STAT), c-Jun N-terminal kinases (JNKs), and NF-κB, to increase inflammatory cytokine release and affect the activation and function/dysfunction of the innate immune cells [64]. It has been demonstrated that trained immunity is triggered by the activation of classical DAMP receptors like TLR by TLR agonists. This process involves metabolic reprogramming and epigenetic alterations, which lead to a significant enhancement of antimicrobial functions [15, 58–60, 62, 90, 108]. During respiratory syncytial virus (RSV) infection, ER stress was induced and PERK was activated in dendritic cells, resulting in an altered innate immune response phenotype and promoting a pathogenic response via transcriptional regulation of key innate immune cytokines, enhancement of eIF2α-phosphorylation signaling, and induction of CD4+ T cell recruitment associated with interleukin-13 (IL-13) and IL-17 production [166, 167]. Furthermore, we recently reported that priming human aortic endothelial cells with β-glucan followed by TMAO binds to its conditional DAMP receptor, PERK, enhances endothelial cell activation, and increases tumor necrosis factor-α (TNF-α), a classical readout for trained immunity [15]. However, the inhibition of PERK activation resulted in reduced TMAO-enhanced trained immunity. Chronic kidney disease (CKD) is characterized by marked increases in proinflammatory cytokine levels, particularly TNF-α and IL-6 [168], which are inversely correlated with decreased glomerular filtration rate (GFR) [169, 170]. Monocytes and neutrophils from CKD patients showed an exaggerated response to LPS stimulation [171]. This may be due to the uremic environment that induces increased expression of TLR2 and 4 [168, 172, 173]. However, persistent activation of monocytes during CKD may also be a direct consequence of uremic toxins [6, 48]. Our latest findings indicate that the PERK pathway and ER stress are responsible for mediating the angiotensin II-accelerated innate immune response, as well as the differences in disease susceptibilities between the thoracic and abdominal aortas [174]. In monocytes, indoxyl sulfate binds to and activates its conditional DAMP receptor, aryl hydrocarbon receptors [175], thereby promoting the production of proinflammatory cytokines [176], which provoke endothelial damage and increase the risk of cardiovascular complications [177, 178].
ER stress sensors are critical regulators for metabolic reprogramming.
The energetic metabolism reprogramming by ER stress usually interferes with mitochondrial respiration to limit the production of adenosine triphosphate (ATP), increase the production of ROS, and switch mitochondrial energetics to glycolysis, otherwise called the “Warburg effect” if it happens under normoxia, which favors cell transformation and cancer development [179]. The protein disulfide isomerase family A member 2 (PDIA2), an ER stress protein, translocates into mitochondria and engages in interaction with components of the electron transport chain, inhibiting mitochondrial respiration and switching energy metabolism to hyperglycolysis in the tumor microenvironment [180]. Furthermore, new research has shown a direct link between PERK-mediated control on insulin resistance [47]. This work is an example of innovative regulation of metabolic circuits through the PERK-ATF4-phosphoserine aminotransferase 1 (PSAT1) axis. At physiologically relevant concentrations, the gut microbiota-generated TMAO binds to PERK and specifically stimulates the PERK-mediated activation of the transcription factor FoxO1, which is a major driver of metabolic diseases. TMAO is increased by insulin resistance and is linked to a number of human metabolic disorders. However, at pathological concentrations, TMAO binds to and activates PERK, which in turn promotes hyperglycemia. In addition, inhibition of flavin-containing monooxygenase 3 (FMO3), a TMAO synthesizing enzyme, reduces PERK activation and insulin resistance in vivo, suggesting a potential target for a new therapeutic intervention for metabolic syndrome. Furthermore, PERK signaling regulates mitochondrial autophagy in immune cells via mitochondrial reprogramming and epigenetic changes [181]. In addition to the promotion of trained immunity and endothelial cell activation, as we reported [15], the PERK arm of UPR signaling enhances macrophage metabolic functions and promotes an immunosuppressive M2 macrophage phenotype [182, 183]. However, PERK deficiency suppresses mitochondrial respiration (oxygen consumption rate), decreased ATP production [113], reduced extracellular acidification rate (ECAR) and the rate of glycolysis [166]. In addition, PERK deficiency significantly downregulates the expression of key genes involved in lipid metabolism, glutamine metabolism, and amino acid metabolism, including peroxisome proliferator-activated receptor γ (PPAR-γ) [184], lysosomal acid lipase (LIPA), and PPAR-γ coactivator-1β (PGC-1β), resulting in a significant decrease in lipid intake, lipolysis, and energy intake in M2 macrophages [185]. Furthermore, stimulation of bone marrow-derived macrophages (BMDMs) isolated from mice with IL-4 (M2 macrophages) increases PERK activation [186]. However, future work needs to determine how PERK-promoted M2 macrophage phenotype is correlated with its trained immunity promotion.
ER stress sensor signaling pathway promotes epigenetic modifications.
Epigenetic changes [55] have been associated with the progression of kidney diseases. During the progression of CKD, several uremic toxins accumulated can stimulate the production of ROS, resulting in epigenetic alterations and further promoting the progression of CKD [187]. We recently reported that ER stress and mitochondrial genes are co-upregulated in CKD environments [15] and that the uremic toxin TMAO binds to and activates the ER stress protein PERK [15, 47]. PERK signaling can regulate mitochondrial autophagy by inducing mitochondrial reprogramming and distinct epigenetic modifications in T cells [181]. To meet cellular energy needs, PERK signaling promotes mitochondrial respirations and signaling through ATF4, which controls PSAT1 activity, mediates the pathway for serine biosynthesis, and results in enhanced mitochondrial function and α-ketoglutarate production required for JMJD3-dependent histone demethylation in M2 macrophages. This finding indicates PERK-directed rewiring of metabolic reprogramming and epigenetic alteration through PERK-ATF4-PSAT1. Furthermore, PERK has been shown to modulate histone 3 lysine 27 (H3K27) methylation in macrophages [185, 186]. Histone H4 acetylation is increased by the induction of the ER chaperone GRP78, which binds to unfolded proteins and promotes the activity of ER stress transducers [188]. In aortic stenosis, the levels of histone deacetylase-6 (HDAC6) are significantly reduced, which is associated with ER stress-mediated osteogenesis and calcification in aortic valve tissue [189]. However, the pharmacological inhibition of HDAC6 significantly reduced apoptosis by inhibiting ER stress in tubular epithelial cells in an acute kidney injury mouse model [190].
ER stress sensor signaling pathways increase inflammation and promote tumor progression.
A crucial characteristic of tumors is the uncontrolled growth of the transformed cells. As a result of poor vascularization, tumors are constantly challenged with a restricted supply of nutrients and oxygen [191]. Tumor growth is linked to increased ER stress inducers such as glucose and oxygen restriction, as well as increased ROS production [192, 193], hence targeting ER stress and UPR provides a new therapeutic target for tumors [194]. UPR activation has been reported in a number of human cancers as well as in various cellular and animal cancer models [195]. In tumors, the upregulation of UPR markers is commonly observed, indicating the development of ER stress, and all UPR signaling branches contribute to tumor growth, angiogenesis, and immune evasion. The high baseline levels of UPR activation in tumor cells give a survival advantage, but it also maintains tumor cells on a tight threshold of survival-death transition. Tumor cells likely require optimum UPR machinery for survival, and thus suppressing oncogenesis by blocking the UPR response or boosting ER stress levels may be an effective way to suppress oncogenesis. Substantial evidence suggests that ER stress pathways and UPR signaling are involved in the activation of different classical inflammatory processes such as NF-κB activation and cytokine and chemokine production within and surrounding the tumor microenvironment [196, 197]. The ER stress response in myeloid cells can be transmitted cell-to-cell, promoting macrophage activation and triggering proinflammatory responses in the tumor microenvironment [198]. Activation of the UPR in tumor dendritic cells results in impaired immune function associated with impaired lipid metabolism and a reduction of T cell anti-tumor immunity. Recent studies have shown that tumor-infiltrating dendritic cells exhibit a high level of spliced XBP1, which promotes the progression of primary and metastatic ovarian cancer [199]. The IRE1α/XBP1 signaling pathway is implicated in the development and progression of various tumor types, including triple-negative breast cancer (TNBC) [200], prostate cancer, and pancreatic cancer [201]. These tumor-promoting effects appear to extend beyond increased protein folding capacity and involve immunomodulation [202]. Treatment of breast cancer cells with the IRE1 RNase inhibitor MKC8866 decreased the expression of numerous immunomodulators, including IL-6, IL-8, CXCL1, TGFβ2, and GM-CSF.
The role of the PERK pathway in promoting cell survival has been established since its discovery [44]. During tumor initiation and expansion, PERK-dependent signaling is utilized to maintain redox homeostasis, thereby facilitating tumor growth [203]. Previous research has demonstrated that the PERK pathway is substantially activated in multiple myeloma, probably contributing to the development of treatment resistance in these patients. The PERK inhibitor GSK2606414 is very effective against myeloma cells. These results imply that the PERK pathway may be a possible therapeutic target for the treatment of multiple myeloma patients since they were more prominent when the inhibitor was combined with an anti-myeloma medication, such as the proteasome inhibitor bortezomib [204, 205]. The PERK-eIF2α-ATF4 pathway has been demonstrated to be a crucial factor in the ability of cancer cells to adjust to hypoxic stress both in vivo and in vitro [206]. Inhibition of PERK kinase function in human cancer cell lines has demonstrated an increase in apoptosis under hypoxia in vitro and impaired tumor growth in vivo [207]. Additionally, tumors derived from transformed PERK-deficient fibroblasts tend to exhibit significantly reduced tumor growth, and it was determined that the impaired angiogenesis and the PERK-deficient cells’ sensitivity to the resulting hypoxic environment were the causes of the decreased tumor growth [207, 208]. There has been further proof shown that PERK plays a role in promoting tumor growth in genetically engineered mouse mammary tumor virus (MMTV)-c-Neu oncogene transgenic mice crossed with PERK knockout mice, where no delay is found in tumor development but rather a significant defect in tumor progression and a significantly reduced rate of metastatic spread are identified as a result of extensive DNA damage caused by increased ROS accumulation [203]. Transgenic mice that expressed the SV40 T-antigen in the pancreatic insulin-secreting beta cells showed that insulinoma growth and angiogenesis were significantly reduced in PERK knockout mice compared with wild-type controls [209]. It has been noted that imatinib mesylate (STI571)-resistant BCR/ABL (a fused oncogene formed from the ABL gene from chromosome 9 fused to the BCR gene on chromosome 22) leukemia cell lines exhibit increased PERK activation and upregulation of the eIF2 pathway; therefore, it has been hypothesized that PERK inhibition may make chronic myeloid leukemia cells more responsive to therapy [210–213]. There is evidence that PERK signaling has effects on the tumor microenvironment in relation to tumor progression. For instance, proinflammatory cytokine production is associated with PERK activation [214, 215]. Thus, the PERK pathway seems to drive carcinogenesis by triggering prosurvival pathways that assist cancer cells in adjusting to the hostile tumor microenvironment. It is believed that the typical role of PERK is to protect secretory cells against ER stress. Pharmacological or hereditary suppression of PERK decelerates the growth of tumor xenografts in mice. Conventional PERK knockout mice exhibit diabetes as a result of pancreatic islet cell death as well as significant developmental abnormalities and growth retardation. The characteristics are similar to those observed in patients with Wolcott-Rallison syndrome who possess inactivating germline mutations in the PERK gene [162, 163]. Also, embryonic PERK deletion results in pancreatic failure, and glucose homeostasis ultimately results in prenatal death [162, 216]. In addition, conditional PERK deletion with a tamoxifen-inducible CRE enzyme showed that PERK deletion leads to the destruction of pancreatic tissue [217]. Although certain types of cancer and metabolic disorders exhibit excessive PERK activation, genetic ablation of PERK results in pancreatic inflammation and irregular glucose levels. This implies that a certain level of PERK function is necessary for maintaining pancreatic health. It could be feasible that a reduction in PERK activity, instead of complete inhibition, may offer therapeutic advantages while maintaining the regular functioning of the pancreas.
In addition, increased levels of ER chaperones, such as GRP78 and GRP94, which are classical UPR markers, are frequently seen in solid tumors. GRP78 has been linked to carcinogenesis [218]. GRP78 is highly expressed in a variety of tumors, presumably due to the ER stress triggered by the hypoxic and nutrient-deficient tumor microenvironment, and its expression correlates to tumor growth, invasion, and tumor metastases [218]. Previous studies have shown that the downregulation of GRP78 in fibrosarcoma cells impairs their ability to form tumors when transplanted into mice [219].
Targeting and inhibiting the ER stress pathways via chemical compounds has been demonstrated to be effective against multiple tumors in preclinical and clinical studies [220, 221] (Table 1). Therefore, the discovery of small-molecule inhibitors that target ER stress sensors, specifically PERK, will provide an attractive therapeutic approach for anticancer and antiangiogenic activity. Inhibiting the integrated stress response (ISR) by using ISRIB reduced tumor size in transgenic mice and patient-derived xenograft models of metastatic disease [222]. Recently, small-molecule inhibitors have entered clinical trials for various tumors, including clear cell renal cell carcinoma and multiple myeloma (https://clinicaltrials.gov/ct2/show/NCT04834778; https://clinicaltrials.gov/ct2/show/NCT05027594). The first identified small-molecule inhibitor of PERK is GSK2606414 [223]. GSK2606414 is an inhibitor that competes with ATP and is highly selective for PERK. Another compound, GSK2656157, was developed for preclinical studies. In various human tumor xenograft models, GSK2656157 was found to efficiently reduce cancer growth [224, 225]. In conclusion, depending on the significant role of PERK in promoting inflammation, tumor progression, angiogenesis, and tumor metastases, targeting the PERK signaling pathway presents a novel therapeutic strategy for cancers, particularly those of secretory function such as neuroendocrine tumors, adenocarcinoma, and multiple myeloma, as well as those with more significant hypoxia such as pancreatic adenocarcinoma and glioblastoma.
Table 1. New therapeutics targeting endoplasmic reticulum stress signaling pathways are in preclinical and clinical research to treat cancer and potentially other diseases in the future.
Emerging evidence suggests that agents affecting ER stress and UPR could be exploited as promising anticancer drug candidates. Several compounds that target ER stress and UPR have been confirmed in preclinical studies and clinical trials for the treatment of various cancer types, and some drugs have been approved by the Food and Drug Administration (FDA). Agents that either disrupt the cytoprotective function of the altered UPR of cancer cells or cause severe ER stress and cell death might be utilized alone or in conjunction with traditional anticancer therapies.
| Experiments | Molecules | Targets | Mechanism | Tumor type | Phase | Pharmacological effects | PMID |
|---|---|---|---|---|---|---|---|
| Preclinical | MAL3–101 | HSP70 | Inhibition of molecular chaperones | Multiple myeloma | Preclinical | Triggers tumor-specific apoptosis and cell death | 21977030 |
| MG-132 | 26S proteasome complex | Inhibition of proteasomes activity | Glioblastoma | Preclinical | Inhibits tumor cells growth Induces tumor cell apoptosis | 22959274 22897979 |
|
| Toyocamycin | IRE1 | Inhibition of XBP1 splicing | Leukemia | Preclinical | Inhibits tumor cells | 18653996 22538852 |
|
| 4μ8C | IRE1 | Inhibition of XBP1 mRNA splicing | Multiple myeloma | Preclinical | Inhibits tumor growth Enhances the efficacy of chemotherapy to tumor cells | 22315414 24821775 |
|
| MKC-3946 | IRE1 | Inhibition of XBP1 mRNA splicing | Multiple myeloma | Preclinical | Inhibits tumor growth Induces tumor cell apoptosis Enhances the efficacy of bortezomib or 17-AAG to tumor cells | 14559994 25069067 22538852 |
|
| STF-083010 | IRE1 | Inhibition of XBP1 mRNA splicing | Multiple myeloma | Preclinical | Inhibits tumor growth Induces tumor cell apoptosis | 21081713 19609461 |
|
| GSK2656157 | PERK | Inhibition of PERK activity | Pancreatic cancer Multiple myeloma | Preclinical | Inhibits tumor growth and angiogenesis | 23333938 | |
| GSK2606414 | PERK | Inhibition of PERK activity | Multiple myeloma | Preclinical | Inhibits tumor growth | 33028016 | |
| ISRIB | eIF2α | Inhibition of eIF2α phosphorylation | Chronic myeloid leukemia | Preclinical | Sensitizes CML cells to imatinib Eliminates Leukemia cells | 36460969 | |
| Ceapin-A7 | ATF6 | Inhibition of ATF6 activity | Preclinical | Sensitizes tumor cells to ER stressor | 27435960 | ||
| CB-5083 | P97 (a key regulator of ERAD) | Activation of UPR | Solid tumors | Preclinical | Induces tumor cell apoptosis | 26555175 | |
| Clinical trials | Sunitinib | Multitargeted tyrosine kinase | Inhibition of IRE1 activity | Renal cell carcinoma | Phase III; FDA approved for renal cell carcinoma | Increases patient survival rate after nephrectomy | 27718781 |
| Sorafenib | Multitargeted tyrosine kinase | Activation of UPR | Hepatocellular carcinoma Desmoid tumors | Phase II and Phase III; FDA approved for hepatocellular carcinoma and renal carcinoma | Inhibits tumor-cell proliferation and tumor angiogenesis Increases the rate of apoptosis | 18650514 30575484 |
|
| Delanzomib CEP-18770 | 26S proteasome | Inhibition of proteasome activity | multiple myeloma | Phase I/II clinical trials | 55% of patients had stable disease and 9% had a partial response | 28140719 | |
| Lobaplatin | PERK | Inhibition of PERK-eIF2α-ATF4-CHOP pathway | breast cancer Hepatic cancer | Phase II clinical trial NCT03210389 | Reduce tumor recurrence and metastasis | 29483583 | |
| Bortezomib PS-341 | 26S proteasome | Inhibition of proteasome activity | Multiple myeloma Cell lymphoma | Phase II, III VISTA trial; FDA approved for multiple myeloma | Increases survival rate Increases response to Dexamethasone | 20368561 14657528 15953001 15953004 12826635 |
ER stress sensors mediate inflammation in cardiovascular diseases.
Cardiovascular diseases (CVD), including atherosclerosis, ischemic heart disease, and heart failure, are the main causes of morbidity and death in CKD patients. Proper cardiovascular function is associated with ER homeostasis, and ER stress is a factor in the development of many different types of CVDs. There is mounting evidence that ER stress contributes significantly to the development and progression of atherosclerotic cardiovascular diseases (CVDs) by being linked to inflammatory signaling pathways via a variety of mechanisms [226, 227]. Specifically, prolonged ER stress activates the three ER transmembrane stress sensors, PERK, ATF6, and IRE1, to initiate UPR signaling that induces oxidative stress and specific inflammatory responses associated with CVD, particularly in ECs and macrophages [228, 229]. Human aortic endothelial cells that have experienced ER stress and cellular UPR activation produce more inflammatory molecules such as IL-8, IL-6, monocyte chemoattractant protein 1 (MCP1), TNF-α, and the chemokine CXC motif ligand 3 (CXCL3), which are blocked by gene silencing of ATF4 and/or XBP1 [230], resulting in vulnerable atherosclerosis plaques [231]. A recent report showed that ER stress plays a role in cardiac remodeling in hypertensive animals [232]. Additionally, it was demonstrated that in rats receiving aldosterone salt treatment, inhibiting ER stress reduces cardiovascular remodeling [233]. Increased PERK and IRE1 levels, as well as accumulated ROS, activate and augment the inflammatory response in atherosclerosis [74]. PERK-mediated translation inhibition results in phosphorylation of the inhibitor of nuclear factor-κB (IκB). Then NF-κB is released and translocated to the nucleus, where it promotes the activation and expression of genes involved in downstream inflammatory pathways, such as those that encode cytokines like TNF-α and IL-1 [227, 234]. Oxidized low-density lipoprotein (ox-LDL) is a key mediator of endothelial dysfunction and the initiation and progression of atherosclerosis [235]. Studies have shown that ox-LDL induces expression of PERK and phosphorylation of eIF2α in human umbilical vein endothelial cells (HUVECs) [236]. We recently showed that PERK activation increases endothelial cell activation, trained immunity, and vascular inflammation in human aortic endothelial cells [15]. In addition, PERK deletion in vascular smooth muscle cells (VSMCs) has been shown to prevent up to 80% of plaque formation in hyperlipidemic mice while also inhibiting VSMC phenotypic modulation [237] and migration [72, 238].
Given the importance of ER stress in the pathophysiology of CVD, strategies targeting ER stress, particularly the dysfunctional UPR, are emerging as potential therapeutic pathways for disease intervention. To prevent and treat CVD, selectively and effectively targeting ER stress in an organelle- or organ-specific manner, such as the mitochondria, holds promise.
ER stress sensor signaling pathways and neurodegenerative diseases.
In contrast to the majority of mammalian cell types, neurons have a limited ability to regenerate. Consequently, the loss of neurons due to cell death leads to neurodegeneration, a loss of neuronal function in the tissues of the central nervous system. The UPR increases the production and release of proinflammatory mediators such as NF-κB, which regulates the expression of the nuclear proto-oncogene SET. SET is directly implicated in the pathogenesis of neurodegenerative diseases. Among the most well-known neurodegenerative diseases are Alzheimer’s disease, Parkinson’s disease, and Huntington’s disease [239]. Alzheimer’s disease is characterized by hyperphosphorylation of the Tau protein (pTau), which destabilizes the microtubules in neurons, leading to the formation of intracellular neurofibrillary tangles (NFTs) and extracellular plaques due to the accumulation of Amyloid-β (Aβ) peptides as a result of mutations in the amyloid precursor protein. Consequently, the accumulation of pTau, NFTs, and Aβ-oligomers increases protein load in the ER, which triggers ER stress and initiates the UPR [240–242]. In models of Alzheimer’s disease, there has been an observed increase in the expression of heat shock protein family A (HSP70) member 5 (BiP), p-PERK, p-IRE1α, P-eIF2α, and ATF4. The activation of the apoptosis signal-regulating kinase 1 (ASK1) branch of the IRE1α pathway may also result from ER stress and oxidative stress caused by Aβ/pTau [243], suggesting that ASK1 might be a new therapeutic target to prevent or treat Alzheimer’s disease. Furthermore, prolonged PERK activation has been demonstrated to impact memory and promote neurodegeneration by impairing protein synthesis [76]. In Parkinson’s disease, the accumulation of mutant α-synuclein (lewy body dementia, abnormal deposits of a-synuclein in the brain) in the dopaminergic neurons results in the activation of the UPR and PERK pathways through increased levels of p-PERK and p-eIF2α [244, 245]. In Huntington’s disease, the accumulation of mutant huntingtin (mHtt) causes ER stress and upregulation of the IRE1, ATF6, and PERK pathways [246]. Although ER stress is a common characteristic of neurodegenerative disorders, it triggers all three pathways of UPR. Yet, it is evident that the PERK pathway plays the main role in the development and progression of neurodegenerative disorders. Therefore, the use of small-molecule drugs to modulate the PERK pathway shows great potential as a therapeutic strategy for various neurodegenerative disorders.
ER stress and UPR in kidney diseases
Healthcare professionals face several difficulties while treating patients with acute kidney injury (AKI) and chronic kidney disease (CKD). AKI is characterized by the rapid loss of renal function and represents an independent risk factor for CKD and end-stage renal disease (ESRD) [247]. CKD is a cardiometabolic disorder characterized by vascular calcification and the accumulation of toxic metabolites, ultimately leading to decreased renal function. Understanding the underlying causes of CKD can help reduce renal dysfunction and associated complications, such as cardiovascular abnormalities. The discovery of novel pharmacological targets for kidney disease is necessary for the development of more effective therapies. ER stress contributes to the development and progression of a wide variety of kidney diseases, including AKI, diabetic nephropathy, nephrotoxicity, and renal fibrosis [75]. The sustained UPR activation leads to the upregulation of inflammatory cytokines and fibrotic genes, which leads to progressive tissue injury and fibrosis. Increased expression of phosphorylated PERK, phosphorylated IRE1α, ATF6, XBP1, and CHOP was observed in the kidneys of diabetic db/db (leptin receptor mutation) mice and renal ischemia/reperfusion (I/R) mouse models. PERK-ATF4-CHOP and IRE1-XBP1 signaling pathways are activated in ischemic acute kidney injury and mediate immunological responses, autophagy, inflammation, apoptosis, and renal fibrosis [248–251] and inhibition of this pathway by chemical chaperones can mitigate renal dysfunction and fibrosis [252]. ATF6 knockout animals showed less lipid deposition and tubulointerstitial kidney fibrosis through decreased expression of proliferator-activated receptor alpha (PPARα), which causes mitochondrial malfunction, oxidative stress, inflammation, apoptotic cell death, and an increase in pro-fibrotic genes [250]. Modulation of CHOP expression may be the most effective target to slow the progression of kidney disease and fibrosis. CHOP deficiency attenuates renal fibrosis and kidney renal diseases [253]. Significant advancements have been achieved in treatments that target ER stress to attenuate the progression of renal diseases. Chemical chaperones approved by the Food and Drug Administration (FDA), such as ursodeoxycholic acid (TUDCA) and 4-phenylbutyric acid (4-PBA), are classic ER stress suppressors that improve insulin sensitivity in metabolic disorders associated with diabetes, restore glucose tolerance, and enhance ER protein folding capacity [254, 255].
Conclusion
The ER is a crucial organelle for maintaining protein homeostasis. Misfolded protein accumulations occur in the ER as a result of physiological or pathological stresses disrupting ER equilibrium. ER stress triggers the UPR, which activates the compensatory protection mechanism as the body’s adaptive and defensive reaction to damaging stimuli. ER stress can be triggered by various stressors that affect homeostatic cellular functions, and prolonged ER stress has been identified as a pathogenic mechanism in the development and progression of a large number of diseases. Emerging evidence indicates that ER stress is implicated in inflammatory responses and plays rather fundamental roles in the onset and progression of cardiovascular diseases [227], cancer [192], kidney diseases, and neurodegenerative diseases [242]. In response to ER stress, UPR is activated under the mediation of three ER membrane-associated proteins: PERK, IRE1, and ATF6 [227]. Pharmacologically, there are two ways to target ER stress: either directly by modifying the misfolded protein accumulation within the ER, or indirectly by controlling UPR signaling via ER stress sensors or the enzymes that mediate their downstream effects. Targeted inhibition of PERK/eIF2, ATF6, and IRE1, the three primary UPR branches, can reduce ER stress and, as a result, exert a protective effect.
In this review, we summarized the ER stress new sensor functions, particularly PERK as a conditional DAMP receptor, and its roles in organelle crosstalk regulation, regulation of protein translation, regulation of membraneless stress assemblies’ formation, as well as modulation of the innate immune response, regulation of metabolic reprogramming, epigenetic modification, and trained immunity. We also summarized the roles of ER stress in several diseases, including cardiovascular diseases, cancers, neurodegenerative diseases, and kidney diseases. We proposed that, in addition to the pattern recognition receptors (PRRs), such as Toll-like receptors (TLRs) and inflammasomes, the disease risk factors, DAMPs, and PAMPs can be sensed by protein folding as well as the three ER stress molecules, including IRE1, PERK, and ATF6. It has been well characterized that PRR signaling pathways are the initiators of cascades that eventually lead to the amplification of the inflammatory response.
Our new working model (Figure 4) proposes that the ER has two new categorized sensor types: protein folding and ER stress regulators as DAMP receptors (PERK). Under disease conditions, disease risk factors, DAMPs, and PAMPs can be recognized by pattern recognition receptors (PRRs), such as Toll-like receptors (TLRs), and inflammasome pathways to enhance inflammatory responses. DAMPs and PAMPs can also be sensed by protein folding and the three ER stress molecules, including IRE1, PERK, and ATF6. A fundamental factor in the development of many pathologies is the loss of protein folding homeostasis, and ER stress plays a significant role in this process. The activation of the ER stress pathways results in impaired calcium homeostasis, increased mitochondrial ROS production, and oxidative stress. The toxic accumulation of ROS within the mitochondria and ER and sustained ER stress trigger inflammatory responses. Dysfunctional UPR pathways have been linked to a variety of chronic diseases, such as inflammatory disease, cardiovascular disease, metabolic disorders, neurodegenerative disorders, cancer, and others. An improved understanding of the molecular mechanisms underlying the ER stress pathways may lead to the discovery of highly promising therapeutic targets for chronic inflammatory diseases, cardiovascular diseases, immune disorders, kidney diseases, neurodegenerative diseases, diabetes, and cancers. Given the detrimental effects of ER stress, the development of novel drugs that can specifically target ER stress in a cell- and disease-specific manner is critically needed. In summary, the advancement and knowledge gained would also support the treatment of diseases linked to ER stress.
Supplementary Material
Figure 5. Our new working model.

ER has two new categorized sensor types: protein folding and ER stress regulators as DAMP receptors (PERK). Under disease conditions, disease risk factors, DAMPs, and PAMPs from protein-rich food and sea foods can be recognized by pattern recognition receptors (PRRs), such as Toll-like receptors (TLRs), and inflammasome pathways to enhance inflammatory responses. DAMPs and PAMPs can also be sensed by protein folding and the three ER stress molecules, including IRE1, PERK, and ATF6. Loss of protein folding homeostasis is central to many diseases, and ER stress is a major contributor to the development of these pathologies. The activation of the ER stress pathways results in impaired calcium homeostasis, increased mitochondrial ROS production, and oxidative stress. The toxic accumulation of ROS within the mitochondria and ER and sustained ER stress trigger inflammatory responses. Dysfunctional UPR pathways have been linked to a variety of chronic diseases, such as neurodegenerative diseases, metabolic disorders, cancer, inflammatory disease, diabetes mellitus, cardiovascular disease, and others. Created with BioRender.com
Source of Funding
Our research activities are supported by grants from the National Institutes of Health (HL163570, HL138749, HL147565, HL130233, DK104116, DK113775, R01-DK098511, R01-DK121227, R01-DK132888, and K08-HL151747), and AHA fellowship (916828 to KX). The content in this article is solely the responsibility of the authors and does not necessarily represent the official views of the NIH.
Abbreviations:
- AIM2
absent in melanoma 2
- AKI
acute kidney injury
- ATF4
activating transcription factor 4
- ATF6
activating transcription factor 6
- AKT1/2/3
protein kinase B
- ATP
adenosine triphosphate
- BAP31
B-cell receptor-associated protein 31
- BiP
binding immunoglobulin protein
- CHOP
C/EBP-homologous protein
- C/EBP
CCAAT-enhancer binding protein
- CKD
chronic kidney disease
- CRAC
calcium release-activated calcium
- CREB
cAMP response element binding protein kinase
- CVD
cardiovascular diseases
- CXCL3
chemokine CXC motif ligand 3
- DAG
diacylglycerol
- DAMPs
danger-associated molecular patterns
- ECAR
extracellular acidification rate
- eIF2AK3
eukaryotic translation initiation factor-2α kinase 3
- EGFR
epidermal growth factor receptor
- ER
endoplasmic reticulum
- ERAD
ER-associated degradation
- ERK1/2MAPK
extracellular signal-regulated kinase
- ERMES
ER-mitochondria encounter structures
- ESRD
end-stage renal disease
- FADD
FAS-associated death domain protein
- FIS1
mitochondrial fission 1 protein
- FMO3
flavin-containing monooxygenase 3
- FMR1
fragile X messenger ribonucleoprotein 1
- FOXO
forkhead box O protein
- G3BP1
GTPase-activating protein-binding protein 1
- GCN2
general control non-derepressible 2
- GFR
glomerular filtration rate
- GPCRs
G-protein-coupled receptors
- GRP75
glucose regulatory protein 75
- GRP78
glucose-related protein 78
- GSK-3
glycogen synthase-3
- HAMPs
homeostasis-associated molecular patterns
- HDAC6
histone deacetylase-6
- H3K27
histone 3 lysine 27
- HMGB1
high-mobility group box 1
- HMG-CoA
3-hydroxy-3-methylglutaryl coenzyme-A
- HRI
heme-regulated initiation factor 2α kinase
- HSP
heat shock protein
- IL-13
interleukin-13
- InsP3
inositol-1,4,5-trisphosphate
- IP3R
inositol triphosphate receptor
- IRE1
inositol-requiring enzyme-1
- ISR
integrated stress response
- JNK1/2/3MAPK
c-Jun N-terminal kinase
- LPS
lipopolysaccharides
- MAM
mitochondria-associated ER membranes
- MCP1
monocyte chemoattractant protein 1
- MFN1
mitochondrial fusion protein 1
- MFN2
mitofusin 2
- MSK1/2
mitogen- and stress-activated kinases
- mTOR
mammalian target of rapamycin
- NFTs
neurofibrillary tangles
- NLRs
Nucleotide-binding oligomerization domain (NOD)-like receptors
- Nrf2
nuclear factor erythroid-derived 2-like 2
- ORFs
open reading frames
- Osh
oxysterol-binding homology
- ox-LDL
Oxidized low-density lipoprotein
- PAB1
poly(A) binding protein cytoplasmic 1
- PAMs
plasma membrane-associated ER membranes
- PAMPs
pathogen-associated molecular patterns
- PDI
protein disulfide isomerases
- PDGFRβ
platelet-derived growth factor receptor β
- PERK
protein kinase R-like endoplasmic reticulum kinase
- P38αMAPK
p38 mitogen-activated protein kinase (MAPK)
- PI4P
phosphatidylinositol 4-phosphate
- PKR
protein kinase RNA-activated
- PPAR-γ
peroxisome proliferator-activated receptor γ
- PM
plasma membrane
- PRKAA1
protein kinase AMP-activated catalytic subunit alpha 1
- PRRs
pattern recognition receptors
- PTPIP51
protein tyrosine phosphatase interacting protein 51
- RAGE
advanced glycation end products
- RIG-I
retinoid acid-inducible gene I
- RIPK1
receptor-interacting serine/threonine protein kinase 1
- ROS
reactive oxygen species
- RYRs
ryanodine receptors
- SACM1L
Sac1-like phosphatidylinositide phosphatase
- SAMPs
stress-associated molecular patterns
- SERCA
sarco-ER calcium-ATPase
- S1P
site-1 protease
- S2P
site-2 protease
- STAT
signal transducer and activator of transcription
- STIM
stromal interacting molecule
- TIA-1
T-cell intracellular antigen-1
- TLR4
toll-like receptor 4
- TMAO
trimethylamine-N-oxide
- TNF-α
tumor necrosis factor-α
- TRAF2
the tumor necrosis factor receptor-associated factor 2
- TRAIL
tumor necrosis factor-related apoptosis-inducing ligand
- TTP
tristetraprolin
- UPR
unfolded protein response
- VAPB
vesicle-associated membrane protein-associated protein B
- VDAC1
voltage-dependent anion-selective channel protein 1
- VSMCs
vascular smooth muscle cells
- XBP1
X box-binding protein 1
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Conflict of Interest statement
The authors have declared that no conflict of interest exists.
Declarations of interest: none
References
- 1.Liu M, et al. , Organelle Crosstalk Regulators Are Regulated in Diseases, Tumors, and Regulatory T Cells: Novel Classification of Organelle Crosstalk Regulators. Front Cardiovasc Med, 2021. 8: p. 713170. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Diaz P, et al. , Perspectives on Organelle Interaction, Protein Dysregulation, and Cancer Disease. Front Cell Dev Biol, 2021. 9: p. 613336. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.English AR and Voeltz GK, Endoplasmic reticulum structure and interconnections with other organelles. Cold Spring Harb Perspect Biol, 2013. 5(4): p. a013227. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Berridge MJ, The endoplasmic reticulum: a multifunctional signaling organelle. Cell Calcium, 2002. 32(5–6): p. 235–49. [DOI] [PubMed] [Google Scholar]
- 5.Lu Y, et al. , Aorta in Pathologies May Function as an Immune Organ by Upregulating Secretomes for Immune and Vascular Cell Activation, Differentiation and Trans-Differentiation-Early Secretomes may Serve as Drivers for Trained Immunity. Front Immunol, 2022. 13: p. 858256. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Zhang R, et al. , End-stage renal disease is different from chronic kidney disease in upregulating ROS-modulated proinflammatory secretome in PBMCs - A novel multiple-hit model for disease progression. Redox Biol, 2020. 34: p. 101460. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Ni D, et al. , Canonical Secretomes, Innate Immune Caspase-1-, 4/11-Gasdermin D Non-Canonical Secretomes and Exosomes May Contribute to Maintain Treg-Ness for Treg Immunosuppression, Tissue Repair and Modulate Anti-Tumor Immunity via ROS Pathways. Front Immunol, 2021. 12: p. 678201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Moncan M, et al. , Regulation of lipid metabolism by the unfolded protein response. J Cell Mol Med, 2021. 25(3): p. 1359–1370. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Schönthal AH, Endoplasmic reticulum stress: its role in disease and novel prospects for therapy. Scientifica (Cairo), 2012. 2012: p. 857516. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Kanapin A, et al. , Mouse proteome analysis. Genome Res, 2003. 13(6b): p. 1335–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Sun Y, et al. , ROS systems are a new integrated network for sensing homeostasis and alarming stresses in organelle metabolic processes. Redox Biol, 2020. 37: p. 101696. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Schröder M, Endoplasmic reticulum stress responses. Cell Mol Life Sci, 2008. 65(6): p. 862–94. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Loi M and Molinari M, Mechanistic insights in recov-ER-phagy: micro-ER-phagy to recover from stress. Autophagy, 2020. 16(2): p. 385–386. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Schröder M, The unfolded protein response. Mol Biotechnol, 2006. 34(2): p. 279–90. [DOI] [PubMed] [Google Scholar]
- 15.Saaoud F, et al. , Aorta- and liver-generated TMAO enhances trained immunity for increased inflammation via ER stress/mitochondrial ROS/glycolysis pathways. JCI Insight, 2023. 8(1). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Shi Y, et al. , Identification and characterization of pancreatic eukaryotic initiation factor 2 α-subunit kinase, PEK, involved in translational control. Molecular and cellular biology, 1998. 18(12): p. 7499–7509. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Li M, et al. , ATF6 as a transcription activator of the endoplasmic reticulum stress element: thapsigargin stress-induced changes and synergistic interactions with NF-Y and YY1. Molecular and cellular biology, 2000. 20(14): p. 5096–5106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Bertolotti A, et al. , Dynamic interaction of BiP and ER stress transducers in the unfolded-protein response. Nat Cell Biol, 2000. 2(6): p. 326–32. [DOI] [PubMed] [Google Scholar]
- 19.Schwarz DS and Blower MD, The endoplasmic reticulum: structure, function and response to cellular signaling. Cell Mol Life Sci, 2016. 73(1): p. 79–94. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Song S, et al. , Crosstalk of autophagy and apoptosis: Involvement of the dual role of autophagy under ER stress. J Cell Physiol, 2017. 232(11): p. 2977–2984. [DOI] [PubMed] [Google Scholar]
- 21.Koumenis C, et al. , Regulation of protein synthesis by hypoxia via activation of the endoplasmic reticulum kinase PERK and phosphorylation of the translation initiation factor eIF2alpha. Mol Cell Biol, 2002. 22(21): p. 7405–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.de la Cadena SG, et al. , Glucose deprivation induces reticulum stress by the PERK pathway and caspase-7- and calpain-mediated caspase-12 activation. Apoptosis, 2014. 19(3): p. 414–27. [DOI] [PubMed] [Google Scholar]
- 23.Kozutsumi Y, et al. , The presence of malfolded proteins in the endoplasmic reticulum signals the induction of glucose-regulated proteins. Nature, 1988. 332(6163): p. 462–4. [DOI] [PubMed] [Google Scholar]
- 24.Pineau L, et al. , Lipid-induced ER stress: synergistic effects of sterols and saturated fatty acids. Traffic, 2009. 10(6): p. 673–90. [DOI] [PubMed] [Google Scholar]
- 25.Karaskov E, et al. , Chronic palmitate but not oleate exposure induces endoplasmic reticulum stress, which may contribute to INS-1 pancreatic beta-cell apoptosis. Endocrinology, 2006. 147(7): p. 3398–407. [DOI] [PubMed] [Google Scholar]
- 26.Wang S, et al. , Ginkgolide K protects the heart against endoplasmic reticulum stress injury by activating the inositol-requiring enzyme 1α/X box-binding protein-1 pathway. Br J Pharmacol, 2016. 173(15): p. 2402–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Olzmann JA, Kopito RR, and Christianson JC, The mammalian endoplasmic reticulum-associated degradation system. Cold Spring Harb Perspect Biol, 2013. 5(9). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Stevenson J, Huang EY, and Olzmann JA, Endoplasmic Reticulum-Associated Degradation and Lipid Homeostasis. Annu Rev Nutr, 2016. 36: p. 511–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Sha X, et al. , Interleukin-35 Inhibits Endothelial Cell Activation by Suppressing MAPK-AP-1 Pathway. J Biol Chem, 2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Huang X, et al. , Identification of novel pretranslational regulatory mechanisms for NF-kappaB activation. J Biol Chem, 2013. 288(22): p. 15628–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Xiong Z, et al. , Expression of TCTP antisense in CD25(high) regulatory T cells aggravates cuff-injured vascular inflammation. Atherosclerosis, 2009. 203(2): p. 401–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Haze K, et al. , Mammalian transcription factor ATF6 is synthesized as a transmembrane protein and activated by proteolysis in response to endoplasmic reticulum stress. Molecular biology of the cell, 1999. 10(11): p. 3787–3799. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Ye J, et al. , ER stress induces cleavage of membrane-bound ATF6 by the same proteases that process SREBPs. Molecular cell, 2000. 6(6): p. 1355–1364. [DOI] [PubMed] [Google Scholar]
- 34.Almanza A, et al. , Endoplasmic reticulum stress signalling - from basic mechanisms to clinical applications. Febs j, 2019. 286(2): p. 241–278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Harding HP, Zhang Y, and Ron D, Protein translation and folding are coupled by an endoplasmic-reticulum-resident kinase. Nature, 1999. 397(6716): p. 271–4. [DOI] [PubMed] [Google Scholar]
- 36.Cullinan SB, et al. , Nrf2 is a direct PERK substrate and effector of PERK-dependent cell survival. Molecular and cellular biology, 2003. 23(20): p. 7198–7209. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Walter P and Ron D, The unfolded protein response: from stress pathway to homeostatic regulation. Science, 2011. 334(6059): p. 1081–6. [DOI] [PubMed] [Google Scholar]
- 38.Wek RC and Cavener DR, Translational control and the unfolded protein response. Antioxid Redox Signal, 2007. 9(12): p. 2357–71. [DOI] [PubMed] [Google Scholar]
- 39.Liu Z, et al. , Protein kinase R-like ER kinase and its role in endoplasmic reticulum stress-decided cell fate. Cell Death Dis, 2015. 6(7): p. e1822. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Yao Y, et al. , A non-canonical pathway regulates ER stress signaling and blocks ER stress-induced apoptosis and heart failure. Nature communications, 2017. 8(1): p. 133. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Szegezdi E, et al. , Mediators of endoplasmic reticulum stress‐induced apoptosis. EMBO reports, 2006. 7(9): p. 880–885. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Bobrovnikova-Marjon E, et al. , PERK utilizes intrinsic lipid kinase activity to generate phosphatidic acid, mediate Akt activation, and promote adipocyte differentiation. Mol Cell Biol, 2012. 32(12): p. 2268–78. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Zhang W, et al. , ER stress potentiates insulin resistance through PERK-mediated FOXO phosphorylation. Genes Dev, 2013. 27(4): p. 441–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Harding HP, et al. , Perk is essential for translational regulation and cell survival during the unfolded protein response. Mol Cell, 2000. 5(5): p. 897–904. [DOI] [PubMed] [Google Scholar]
- 45.Kapetanaki S, et al. , The Fibrotic Effects of TMAO on Human Renal Fibroblasts Is Mediated by NLRP3, Caspase-1 and the PERK/Akt/mTOR Pathway. Int J Mol Sci, 2021. 22(21). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Chan MM, et al. , The Microbial Metabolite Trimethylamine N-Oxide Links Vascular Dysfunctions and the Autoimmune Disease Rheumatoid Arthritis. Nutrients, 2019. 11(8). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Chen S, et al. , Trimethylamine N-Oxide Binds and Activates PERK to Promote Metabolic Dysfunction. Cell Metab, 2019. 30(6): p. 1141–1151.e5. [DOI] [PubMed] [Google Scholar]
- 48.Sun Y, et al. , Uremic toxins are conditional danger- or homeostasis-associated molecular patterns. Front Biosci (Landmark Ed), 2018. 23(2): p. 348–387. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Kraskiewicz H and FitzGerald U, InterfERing with endoplasmic reticulum stress. Trends in pharmacological sciences, 2012. 33(2): p. 53–63. [DOI] [PubMed] [Google Scholar]
- 50.Healy SJ, et al. , Targeting the endoplasmic reticulum-stress response as an anticancer strategy. European journal of pharmacology, 2009. 625(1–3): p. 234–246. [DOI] [PubMed] [Google Scholar]
- 51.Cojocari D, et al. , New small molecule inhibitors of UPR activation demonstrate that PERK, but not IRE1α signaling is essential for promoting adaptation and survival to hypoxia. Radiotherapy and Oncology, 2013. 108(3): p. 541–547. [DOI] [PubMed] [Google Scholar]
- 52.Mendez JM, et al. , Activation of the Endoplasmic Reticulum Stress Response Impacts the NOD1 Signaling Pathway. Infect Immun, 2019. 87(8). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Wang X, et al. , Lysophospholipid Receptors, as Novel Conditional Danger Receptors and Homeostatic Receptors Modulate Inflammation-Novel Paradigm and Therapeutic Potential. J Cardiovasc Transl Res, 2016. 9(4): p. 343–59. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Shao Y, et al. , Lysophospholipids and Their Receptors Serve as Conditional DAMPs and DAMP Receptors in Tissue Oxidative and Inflammatory Injury. Antioxid Redox Signal, 2018. 28(10): p. 973–986. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Shao Y, et al. , Metabolic Diseases Downregulate the Majority of Histone Modification Enzymes, Making a Few Upregulated Enzymes Novel Therapeutic Targets--”Sand Out and Gold Stays”. J Cardiovasc Transl Res, 2016. 9(1): p. 49–66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.So JS, Roles of Endoplasmic Reticulum Stress in Immune Responses. Mol Cells, 2018. 41(8): p. 705–716. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Martins AS, et al. , The Unfolded Protein Response in Homeostasis and Modulation of Mammalian Immune Cells. Int Rev Immunol, 2016. 35(6): p. 457–476. [DOI] [PubMed] [Google Scholar]
- 58.Lu Y, et al. , Increased acetylation of H3K14 in the genomic regions that encode trained immunity enzymes in lysophosphatidylcholine-activated human aortic endothelial cells - Novel qualification markers for chronic disease risk factors and conditional DAMPs. Redox Biol, 2019. 24: p. 101221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Drummer CIV, et al. , Hyperlipidemia May Synergize with Hypomethylation in Establishing Trained Immunity and Promoting Inflammation in NASH and NAFLD. J Immunol Res, 2021. 2021: p. 3928323. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Drummer C.t., et al. , Trained Immunity and Reactivity of Macrophages and Endothelial Cells. Arterioscler Thromb Vasc Biol, 2021. 41(3): p. 1032–1046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Zhong C, et al. , Trained Immunity: An Underlying Driver of Inflammatory Atherosclerosis. Front Immunol, 2020. 11: p. 284. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Fagenson AM, et al. , Liver Ischemia Reperfusion Injury, Enhanced by Trained Immunity, Is Attenuated in Caspase 1/Caspase 11 Double Gene Knockout Mice. Pathogens, 2020. 9(11). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Saaoud F, et al. , Cigarette smoke modulates inflammation and immunity via ROS-regulated trained immunity and trained tolerance mechanisms. Antioxid Redox Signal, 2022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Di Conza G and Ho PC, ER Stress Responses: An Emerging Modulator for Innate Immunity. Cells, 2020. 9(3). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Preusse C, et al. , Endoplasmic reticulum-stress and unfolded protein response-activation in immune-mediated necrotizing myopathy. Brain Pathol, 2022. 32(6): p. e13084. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Walter F, et al. , Imaging of single cell responses to ER stress indicates that the relative dynamics of IRE1/XBP1 and PERK/ATF4 signalling rather than a switch between signalling branches determine cell survival. Cell Death Differ, 2015. 22(9): p. 1502–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Lee YS, et al. , Ferroptosis-Induced Endoplasmic Reticulum Stress: Cross-talk between Ferroptosis and Apoptosis. Mol Cancer Res, 2018. 16(7): p. 1073–1076. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Sano R and Reed JC, ER stress-induced cell death mechanisms. Biochim Biophys Acta, 2013. 1833(12): p. 3460–3470. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Yao RQ, et al. , Organelle-specific autophagy in inflammatory diseases: a potential therapeutic target underlying the quality control of multiple organelles. Autophagy, 2021. 17(2): p. 385–401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Kaarniranta K, et al. , Autophagy in age-related macular degeneration. Autophagy, 2023. 19(2): p. 388–400. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Soto-Alarcon SA, et al. , Liver Protective Effects of Extra Virgin Olive Oil: Interaction between Its Chemical Composition and the Cell-signaling Pathways Involved in Protection. Endocr Metab Immune Disord Drug Targets, 2018. 18(1): p. 75–84. [DOI] [PubMed] [Google Scholar]
- 72.Chattopadhyay A, et al. , Cholesterol-Induced Phenotypic Modulation of Smooth Muscle Cells to Macrophage/Fibroblast-like Cells Is Driven by an Unfolded Protein Response. Arterioscler Thromb Vasc Biol, 2021. 41(1): p. 302–316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Hong J, et al. , The Role of Endoplasmic Reticulum Stress in Cardiovascular Disease and Exercise. Int J Vasc Med, 2017. 2017: p. 2049217. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Zhou AX and Tabas I, The UPR in atherosclerosis. Seminars in Immunopathology, 2013. 35(3): p. 321–332. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Gallazzini M and Pallet N, Endoplasmic reticulum stress and kidney dysfunction. Biol Cell, 2018. 110(9): p. 205–216. [DOI] [PubMed] [Google Scholar]
- 76.Ohno M, PERK as a hub of multiple pathogenic pathways leading to memory deficits and neurodegeneration in Alzheimer’s disease. Brain Res Bull, 2018. 141: p. 72–78. [DOI] [PubMed] [Google Scholar]
- 77.Ghemrawi R and Khair M, Endoplasmic Reticulum Stress and Unfolded Protein Response in Neurodegenerative Diseases. Int J Mol Sci, 2020. 21(17). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Shacham T, Patel C, and Lederkremer GZ, PERK Pathway and Neurodegenerative Disease: To Inhibit or to Activate? Biomolecules, 2021. 11(3). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Kang Z, et al. , UPR(mt) and coordinated UPR(ER) in type 2 diabetes. Front Cell Dev Biol, 2022. 10: p. 974083. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Rozpedek W, et al. , The Role of the PERK/eIF2alpha/ATF4/CHOP Signaling Pathway in Tumor Progression During Endoplasmic Reticulum Stress. Curr Mol Med, 2016. 16(6): p. 533–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Tian X, et al. , Targeting the Integrated Stress Response in Cancer Therapy. Front Pharmacol, 2021. 12: p. 747837. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Cubillos-Ruiz JR, Bettigole SE, and Glimcher LH, Tumorigenic and Immunosuppressive Effects of Endoplasmic Reticulum Stress in Cancer. Cell, 2017. 168(4): p. 692–706. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Salvagno C, et al. , Decoding endoplasmic reticulum stress signals in cancer cells and antitumor immunity. Trends Cancer, 2022. 8(11): p. 930–943. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Liang D, et al. , The unfolded protein response as regulator of cancer stemness and differentiation: Mechanisms and implications for cancer therapy. Biochem Pharmacol, 2021. 192: p. 114737. [DOI] [PubMed] [Google Scholar]
- 85.Urra H, et al. , Endoplasmic Reticulum Stress and the Hallmarks of Cancer. Trends Cancer, 2016. 2(5): p. 252–262. [DOI] [PubMed] [Google Scholar]
- 86.Yang XF, Yin Y, and Wang H, VASCULAR INFLAMMATION AND ATHEROGENESIS ARE ACTIVATED VIA RECEPTORS FOR PAMPs AND SUPPRESSED BY REGULATORY T CELLS. Drug Discov Today Ther Strateg, 2008. 5(2): p. 125–142. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Yin Y, et al. , Inflammasomes: sensors of metabolic stresses for vascular inflammation. Front Biosci, 2013. 18: p. 638–49. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Virtue A, Wang H, and Yang XF, MicroRNAs and toll-like receptor/interleukin-1 receptor signaling. J Hematol Oncol, 2012. 5: p. 66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Drummer C.t., et al. , Caspase-11 promotes high-fat diet-induced NAFLD by increasing glycolysis, OXPHOS, and pyroptosis in macrophages. Front Immunol, 2023. 14: p. 1113883. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Shao Y, et al. , Endothelial Immunity Trained by Coronavirus Infections, DAMP Stimulations and Regulated by Anti-Oxidant NRF2 May Contribute to Inflammations, Myelopoiesis, COVID-19 Cytokine Storms and Thromboembolism. Front Immunol, 2021. 12: p. 653110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Zindel J and Kubes P, DAMPs, PAMPs, and LAMPs in Immunity and Sterile Inflammation. Annu Rev Pathol, 2020. 15: p. 493–518. [DOI] [PubMed] [Google Scholar]
- 92.Danieli MG, et al. , Alarmins in autoimmune diseases. Autoimmun Rev, 2022. 21(9): p. 103142. [DOI] [PubMed] [Google Scholar]
- 93.Yin Y, et al. , Early hyperlipidemia promotes endothelial activation via a caspase-1-sirtuin 1 pathway. Arterioscler Thromb Vasc Biol, 2015. 35(4): p. 804–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Yin Y, et al. , Inflammasomes are differentially expressed in cardiovascular and other tissues. Int J Immunopathol Pharmacol, 2009. 22(2): p. 311–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Lopez-Pastrana J, et al. , Inhibition of Caspase-1 Activation in Endothelial Cells Improves Angiogenesis: A NOVEL THERAPEUTIC POTENTIAL FOR ISCHEMIA. J Biol Chem, 2015. 290(28): p. 17485–94. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Vénéreau E, Ceriotti C, and Bianchi ME, DAMPs from Cell Death to New Life. Front Immunol, 2015. 6: p. 422. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Gong T, et al. , DAMP-sensing receptors in sterile inflammation and inflammatory diseases. Nat Rev Immunol, 2020. 20(2): p. 95–112. [DOI] [PubMed] [Google Scholar]
- 98.Li X, et al. , IL-35, as a newly proposed homeostasis-associated molecular pattern, plays three major functions including anti-inflammatory initiator, effector, and blocker in cardiovascular diseases. Cytokine, 2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Li X, et al. , Lysophospholipids induce innate immune transdifferentiation of endothelial cells, resulting in prolonged endothelial activation. Journal of Biological Chemistry, 2018: p. jbc.RA118.002752. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Dai J, et al. , Metabolism-associated danger signal-induced immune response and reverse immune checkpoint-activated CD40(+) monocyte differentiation. J Hematol Oncol, 2017. 10(1): p. 141. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Shen W, et al. , Homocysteine-methionine cycle is a metabolic sensor system controlling methylation-regulated pathological signaling. Redox Biol, 2020. 28: p. 101322. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Zeng H, et al. , DNA Checkpoint and Repair Factors Are Nuclear Sensors for Intracellular Organelle Stresses-Inflammations and Cancers Can Have High Genomic Risks. Front Physiol, 2018. 9: p. 516. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Sullivan GP, et al. , TRAIL Receptors Serve as Stress-Associated Molecular Patterns to Promote ER-Stress-Induced Inflammation. Developmental Cell, 2020. 52(6): p. 714–730.e5. [DOI] [PubMed] [Google Scholar]
- 104.Stubbs JR, et al. , Trimethylamine N-Oxide and Cardiovascular Outcomes in Patients with ESKD Receiving Maintenance Hemodialysis. Clin J Am Soc Nephrol, 2019. 14(2): p. 261–267. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Chen ML, et al. , Trimethylamine-N-Oxide Induces Vascular Inflammation by Activating the NLRP3 Inflammasome Through the SIRT3-SOD2-mtROS Signaling Pathway. J Am Heart Assoc, 2017. 6(9). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Ke X, et al. , Gut bacterial metabolites modulate endoplasmic reticulum stress. Genome Biology, 2021. 22(1): p. 292. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Rainbolt TK, Saunders JM, and Wiseman RL, Stress-responsive regulation of mitochondria through the ER unfolded protein response. Trends Endocrinol Metab, 2014. 25(10): p. 528–37. [DOI] [PubMed] [Google Scholar]
- 108.Shao Y, et al. , Vascular Endothelial Cells and Innate Immunity. Arterioscler Thromb Vasc Biol, 2020. 40(6): p. e138–e152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Li X, et al. , Mitochondrial Reactive Oxygen Species Mediate Lysophosphatidylcholine-Induced Endothelial Cell ActivationHighlights. Arteriosclerosis, Thrombosis, and Vascular Biology, 2016. 36(6): p. 1090–1100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Li X, et al. , Mitochondrial ROS, uncoupled from ATP synthesis, determine endothelial activation for both physiological recruitment of patrolling cells and pathological recruitment of inflammatory cells. Can J Physiol Pharmacol, 2017. 95(3): p. 247–252. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Li X, et al. , IL-35 (Interleukin-35) Suppresses Endothelial Cell Activation by Inhibiting Mitochondrial Reactive Oxygen Species-Mediated Site-Specific Acetylation of H3K14 (Histone 3 Lysine 14). Arterioscler Thromb Vasc Biol, 2018. 38(3): p. 599–609. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Li X, et al. , Targeting mitochondrial reactive oxygen species as novel therapy for inflammatory diseases and cancers. J Hematol Oncol, 2013. 6: p. 19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Li X, et al. , Mitochondrial Reactive Oxygen Species Mediate Lysophosphatidylcholine-Induced Endothelial Cell Activation. Arterioscler Thromb Vasc Biol, 2016. 36(6): p. 1090–100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Cheng J, et al. , Mitochondrial Proton Leak Plays a Critical Role in Pathogenesis of Cardiovascular Diseases. Adv Exp Med Biol, 2017. 982: p. 359–370. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Nanayakkara GK, Wang H, and Yang X, Proton leak regulates mitochondrial reactive oxygen species generation in endothelial cell activation and inflammation - A novel concept. Arch Biochem Biophys, 2019. 662: p. 68–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Inigo JR and Chandra D, The mitochondrial unfolded protein response (UPR(mt)): shielding against toxicity to mitochondria in cancer. J Hematol Oncol, 2022. 15(1): p. 98. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Scorrano L, et al. , Coming together to define membrane contact sites. Nat Commun, 2019. 10(1): p. 1287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Simmen T and Herrera-Cruz MS, Plastic mitochondria-endoplasmic reticulum (ER) contacts use chaperones and tethers to mould their structure and signaling. Curr Opin Cell Biol, 2018. 53: p. 61–69. [DOI] [PubMed] [Google Scholar]
- 119.Phillips MJ and Voeltz GK, Structure and function of ER membrane contact sites with other organelles. Nat Rev Mol Cell Biol, 2016. 17(2): p. 69–82. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Friedman JR, et al. , ER tubules mark sites of mitochondrial division. Science, 2011. 334(6054): p. 358–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Volmer R, van der Ploeg K, and Ron D, Membrane lipid saturation activates endoplasmic reticulum unfolded protein response transducers through their transmembrane domains. Proc Natl Acad Sci U S A, 2013. 110(12): p. 4628–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Verfaillie T, et al. , PERK is required at the ER-mitochondrial contact sites to convey apoptosis after ROS-based ER stress. Cell Death Differ, 2012. 19(11): p. 1880–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Muñoz JP, et al. , Mfn2 modulates the UPR and mitochondrial function via repression of PERK. Embo j, 2013. 32(17): p. 2348–61. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Sassano ML, et al. , PERK recruits E-Syt1 at ER-mitochondria contacts for mitochondrial lipid transport and respiration. J Cell Biol, 2023. 222(3). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Leterme S and Michaud M, Mitochondrial membrane biogenesis: A new pathway for lipid transport mediated by PERK/E-Syt1 complex. J Cell Biol, 2023. 222(3). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Tur J, et al. , Mitofusin 2 in Macrophages Links Mitochondrial ROS Production, Cytokine Release, Phagocytosis, Autophagy, and Bactericidal Activity. Cell Rep, 2020. 32(8): p. 108079. [DOI] [PubMed] [Google Scholar]
- 127.Ichinohe T, et al. , Mitochondrial protein mitofusin 2 is required for NLRP3 inflammasome activation after RNA virus infection. Proc Natl Acad Sci U S A, 2013. 110(44): p. 17963–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.de Brito OM and Scorrano L, Mitofusin 2 tethers endoplasmic reticulum to mitochondria. Nature, 2008. 456(7222): p. 605–10. [DOI] [PubMed] [Google Scholar]
- 129.Fan Y and Simmen T, Mechanistic connections between endoplasmic reticulum (ER) redox control and mitochondrial metabolism. Cells, 2019. 8(9): p. 1071. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Tubbs E and Rieusset J, Metabolic signaling functions of ER-mitochondria contact sites: role in metabolic diseases. J Mol Endocrinol, 2017. 58(2): p. R87–r106. [DOI] [PubMed] [Google Scholar]
- 131.Hattori N, et al. , Mitochondrial-Associated Membranes in Parkinson’s Disease. Adv Exp Med Biol, 2017. 997: p. 157–169. [DOI] [PubMed] [Google Scholar]
- 132.Burgoyne T, Patel S, and Eden ER, Calcium signaling at ER membrane contact sites. Biochim Biophys Acta, 2015. 1853(9): p. 2012–7. [DOI] [PubMed] [Google Scholar]
- 133.van Vliet AR, et al. , The ER Stress Sensor PERK Coordinates ER-Plasma Membrane Contact Site Formation through Interaction with Filamin-A and F-Actin Remodeling. Mol Cell, 2017. 65(5): p. 885–899 e6. [DOI] [PubMed] [Google Scholar]
- 134.Carreras-Sureda A, et al. , Non-canonical function of IRE1α determines mitochondria-associated endoplasmic reticulum composition to control calcium transfer and bioenergetics. Nat Cell Biol, 2019. 21(6): p. 755–767. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Ma JH, et al. , Comparative Proteomic Analysis of the Mitochondria-associated ER Membrane (MAM) in a Long-term Type 2 Diabetic Rodent Model. Sci Rep, 2017. 7(1): p. 2062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Gomes E and Shorter J, The molecular language of membraneless organelles. J Biol Chem, 2019. 294(18): p. 7115–7127. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Patel A, et al. , A liquid-to-solid phase transition of the ALS protein FUS accelerated by disease mutation. Cell, 2015. 162(5): p. 1066–1077. [DOI] [PubMed] [Google Scholar]
- 138.Hyman AA, Weber CA, and Jülicher F, Liquid-liquid phase separation in biology. Annu Rev Cell Dev Biol, 2014. 30: p. 39–58. [DOI] [PubMed] [Google Scholar]
- 139.Luo Y, Na Z, and Slavoff SA, P-Bodies: Composition, Properties, and Functions. Biochemistry, 2018. 57(17): p. 2424–2431. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Rai AK, et al. , Kinase-controlled phase transition of membraneless organelles in mitosis. Nature, 2018. 559(7713): p. 211–216. [DOI] [PubMed] [Google Scholar]
- 141.Spannl S, et al. , Biomolecular condensates in neurodegeneration and cancer. Traffic, 2019. 20(12): p. 890–911. [DOI] [PubMed] [Google Scholar]
- 142.van Leeuwen W and Rabouille C, Cellular stress leads to the formation of membraneless stress assemblies in eukaryotic cells. Traffic, 2019. 20(9): p. 623–638. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Denegri M, et al. , Stress-induced nuclear bodies are sites of accumulation of pre-mRNA processing factors. Mol Biol Cell, 2001. 12(11): p. 3502–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Liu M, et al. , 29 m(6)A-RNA Methylation (Epitranscriptomic) Regulators Are Regulated in 41 Diseases including Atherosclerosis and Tumors Potentially via ROS Regulation - 102 Transcriptomic Dataset Analyses. J Immunol Res, 2022. 2022: p. 1433323. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Anderson P and Kedersha N, Stressful initiations. J Cell Sci, 2002. 115(Pt 16): p. 3227–34. [DOI] [PubMed] [Google Scholar]
- 146.Vanderweyde T, et al. , Role of stress granules and RNA-binding proteins in neurodegeneration: a mini-review. Gerontology, 2013. 59(6): p. 524–533. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Wheeler JR, et al. , Distinct stages in stress granule assembly and disassembly. Elife, 2016. 5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.McEwen E, et al. , Heme-regulated inhibitor kinase-mediated phosphorylation of eukaryotic translation initiation factor 2 inhibits translation, induces stress granule formation, and mediates survival upon arsenite exposure. J Biol Chem, 2005. 280(17): p. 16925–33. [DOI] [PubMed] [Google Scholar]
- 149.Mazroui R, et al. , Inhibition of the ubiquitin-proteasome system induces stress granule formation. Mol Biol Cell, 2007. 18(7): p. 2603–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Donnelly N, et al. , The eIF2α kinases: their structures and functions. Cell Mol Life Sci, 2013. 70(19): p. 3493–511. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Thandapani P, et al. , Defining the RGG/RG motif. Mol Cell, 2013. 50(5): p. 613–23. [DOI] [PubMed] [Google Scholar]
- 152.Ripin N, et al. , Molecular basis for AU-rich element recognition and dimerization by the HuR C-terminal RRM. Proc Natl Acad Sci U S A, 2019. 116(8): p. 2935–2944. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Mazroui R, et al. , Inhibition of ribosome recruitment induces stress granule formation independently of eukaryotic initiation factor 2α phosphorylation. Molecular biology of the cell, 2006. 17(10): p. 4212–4219. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Sonenberg N and Hinnebusch AG, Regulation of translation initiation in eukaryotes: mechanisms and biological targets. Cell, 2009. 136(4): p. 731–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Ma XM and Blenis J, Molecular mechanisms of mTOR-mediated translational control. Nat Rev Mol Cell Biol, 2009. 10(5): p. 307–18. [DOI] [PubMed] [Google Scholar]
- 156.Jackson RJ, Hellen CU, and Pestova TV, The mechanism of eukaryotic translation initiation and principles of its regulation. Nat Rev Mol Cell Biol, 2010. 11(2): p. 113–27. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Wek RC, Jiang HY, and Anthony TG, Coping with stress: eIF2 kinases and translational control. Biochem Soc Trans, 2006. 34(Pt 1): p. 7–11. [DOI] [PubMed] [Google Scholar]
- 158.Kedersha NL, et al. , RNA-binding proteins TIA-1 and TIAR link the phosphorylation of eIF-2 alpha to the assembly of mammalian stress granules. J Cell Biol, 1999. 147(7): p. 1431–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Shi Y, et al. , Identification and characterization of pancreatic eukaryotic initiation factor 2 alpha-subunit kinase, PEK, involved in translational control. Mol Cell Biol, 1998. 18(12): p. 7499–509. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Pavio N, et al. , Protein synthesis and endoplasmic reticulum stress can be modulated by the hepatitis C virus envelope protein E2 through the eukaryotic initiation factor 2alpha kinase PERK. J Virol, 2003. 77(6): p. 3578–85. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Yaman I, et al. , The zipper model of translational control: a small upstream ORF is the switch that controls structural remodeling of an mRNA leader. Cell, 2003. 113(4): p. 519–31. [DOI] [PubMed] [Google Scholar]
- 162.Harding HP, et al. , Diabetes mellitus and exocrine pancreatic dysfunction in perk−/− mice reveals a role for translational control in secretory cell survival. Mol Cell, 2001. 7(6): p. 1153–63. [DOI] [PubMed] [Google Scholar]
- 163.Zhang P, et al. , The PERK eukaryotic initiation factor 2 alpha kinase is required for the development of the skeletal system, postnatal growth, and the function and viability of the pancreas. Mol Cell Biol, 2002. 22(11): p. 3864–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Ochando J, et al. , Trained immunity - basic concepts and contributions to immunopathology. Nat Rev Nephrol, 2023. 19(1): p. 23–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Zhang K and Kaufman RJ, From endoplasmic-reticulum stress to the inflammatory response. Nature, 2008. 454(7203): p. 455–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Narayanan S, et al. , ER stress protein PERK promotes inappropriate innate immune responses and pathogenesis during RSV infection. J Leukoc Biol, 2022. 111(2): p. 379–389. [DOI] [PubMed] [Google Scholar]
- 167.Mai J, et al. , Interleukin-17A Promotes Aortic Endothelial Cell Activation via Transcriptionally and Post-translationally Activating p38 MAPK Pathway. J Biol Chem, 2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Grabulosa CC, et al. , Chronic kidney disease induces inflammation by increasing Toll-like receptor-4, cytokine and cathelicidin expression in neutrophils and monocytes. Exp Cell Res, 2018. 365(2): p. 157–162. [DOI] [PubMed] [Google Scholar]
- 169.Castillo-Rodríguez E, et al. , Inflammatory Cytokines as Uremic Toxins: “Ni Son Todos Los Que Estan, Ni Estan Todos Los Que Son”. Toxins (Basel), 2017. 9(4). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170.Amdur RL, et al. , Inflammation and Progression of CKD: The CRIC Study. Clin J Am Soc Nephrol, 2016. 11(9): p. 1546–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171.Sela S, et al. , Primed peripheral polymorphonuclear leukocyte: a culprit underlying chronic low-grade inflammation and systemic oxidative stress in chronic kidney disease. J Am Soc Nephrol, 2005. 16(8): p. 2431–8. [DOI] [PubMed] [Google Scholar]
- 172.Gollapudi P, et al. , Leukocyte toll-like receptor expression in end-stage kidney disease. Am J Nephrol, 2010. 31(3): p. 247–54. [DOI] [PubMed] [Google Scholar]
- 173.Koc M, et al. , Toll-like receptor expression in monocytes in patients with chronic kidney disease and haemodialysis: relation with inflammation. Nephrol Dial Transplant, 2011. 26(3): p. 955–63. [DOI] [PubMed] [Google Scholar]
- 174.Lu Y, et al. , ER stress mediates Angiotensin II-augmented innate immunity memory and facilitates distinct susceptibilities of thoracic from abdominal aorta to aneurysm development. Front Immunol, 2023. 14: p. 1268916. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Gondouin B, et al. , Indolic uremic solutes increase tissue factor production in endothelial cells by the aryl hydrocarbon receptor pathway. Kidney Int, 2013. 84(4): p. 733–44. [DOI] [PubMed] [Google Scholar]
- 176.Kim HY, et al. , Indoxyl sulfate-induced TNF-α is regulated by crosstalk between the aryl hydrocarbon receptor, NF-κB, and SOCS2 in human macrophages. Faseb j, 2019. 33(10): p. 10844–10858. [DOI] [PubMed] [Google Scholar]
- 177.Kim HY, et al. , Indoxyl sulfate (IS)-mediated immune dysfunction provokes endothelial damage in patients with end-stage renal disease (ESRD). Sci Rep, 2017. 7(1): p. 3057. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Zimmermann J, et al. , Inflammation enhances cardiovascular risk and mortality in hemodialysis patients. Kidney Int, 1999. 55(2): p. 648–58. [DOI] [PubMed] [Google Scholar]
- 179.Yu L, et al. , The Glycolytic Switch in Tumors: How Many Players Are Involved? J Cancer, 2017. 8(17): p. 3430–3440. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Tao J, et al. , PDIA2 Bridges Endoplasmic Reticulum Stress and Metabolic Reprogramming During Malignant Transformation of Chronic Colitis. Front Oncol, 2022. 12: p. 836087. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Chakraborty P, et al. , Carbon Monoxide Activates PERK-Regulated Autophagy to Induce Immunometabolic Reprogramming and Boost Antitumor T-cell Function. Cancer Res, 2022. 82(10): p. 1969–1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182.Lai B, et al. , Twenty Novel Disease Group-Specific and 12 New Shared Macrophage Pathways in Eight Groups of 34 Diseases Including 24 Inflammatory Organ Diseases and 10 Types of Tumors. Front Immunol, 2019. 10: p. 2612. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Johnson C, et al. , A Novel Subset of CD95(+) Pro-Inflammatory Macrophages Overcome miR155 Deficiency and May Serve as a Switch From Metabolically Healthy Obesity to Metabolically Unhealthy Obesity. Front Immunol, 2020. 11: p. 619951. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 184.Virtue A, et al. , MicroRNA-155 Deficiency Leads to Decreased Atherosclerosis, Increased White Adipose Tissue Obesity, and Non-alcoholic Fatty Liver Disease: A NOVEL MOUSE MODEL OF OBESITY PARADOX. J Biol Chem, 2017. 292(4): p. 1267–1287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Raines LN, et al. , PERK is a critical metabolic hub for immunosuppressive function in macrophages. Nat Immunol, 2022. 23(3): p. 431–445. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186.Pratap UP and Vadlamudi RK, PERK promotes immunosuppressive M2 macrophage phenotype by metabolic reprogramming and epigenetic modifications through the PERK-ATF4-PSAT1 axis. Immunometabolism (Cobham), 2022. 4(3): p. e00007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Young GH and Wu VC, KLOTHO methylation is linked to uremic toxins and chronic kidney disease. Kidney Int, 2012. 81(7): p. 611–2. [DOI] [PubMed] [Google Scholar]
- 188.Baumeister P, et al. , Endoplasmic reticulum stress induction of the Grp78/BiP promoter: activating mechanisms mediated by YY1 and its interactive chromatin modifiers. Mol Cell Biol, 2005. 25(11): p. 4529–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Fu Z, et al. , Histone deacetylase 6 reduction promotes aortic valve calcification via an endoplasmic reticulum stress-mediated osteogenic pathway. J Thorac Cardiovasc Surg, 2019. 158(2): p. 408–417.e2. [DOI] [PubMed] [Google Scholar]
- 190.Feng Y, et al. , Selective Histone Deacetylase 6 Inhibitor 23BB Alleviated Rhabdomyolysis-Induced Acute Kidney Injury by Regulating Endoplasmic Reticulum Stress and Apoptosis. Front Pharmacol, 2018. 9: p. 274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Hanahan D and Weinberg RA, The Hallmarks of Cancer. Cell, 2000. 100(1): p. 57–70. [DOI] [PubMed] [Google Scholar]
- 192.Romero-Ramirez L, et al. , XBP1 is essential for survival under hypoxic conditions and is required for tumor growth. Cancer Res, 2004. 64(17): p. 5943–7. [DOI] [PubMed] [Google Scholar]
- 193.Ackerman D and Simon MC, Hypoxia, lipids, and cancer: surviving the harsh tumor microenvironment. Trends Cell Biol, 2014. 24(8): p. 472–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Chen X and Cubillos-Ruiz JR, Endoplasmic reticulum stress signals in the tumour and its microenvironment. Nat Rev Cancer, 2021. 21(2): p. 71–88. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195.Clarke HJ, et al. , Endoplasmic reticulum stress in malignancy. Cancer cell, 2014. 25(5): p. 563–573. [DOI] [PubMed] [Google Scholar]
- 196.Li Y, et al. , Free cholesterol-loaded macrophages are an abundant source of tumor necrosis factor-α and interleukin-6: model of NF-κB-and map kinase-dependent inflammation in advanced atherosclerosis. Journal of Biological Chemistry, 2005. 280(23): p. 21763–21772. [DOI] [PubMed] [Google Scholar]
- 197.Verfaillie T, Garg AD, and Agostinis P, Targeting ER stress induced apoptosis and inflammation in cancer. Cancer Letters, 2013. 332(2): p. 249–264. [DOI] [PubMed] [Google Scholar]
- 198.Mahadevan NR, et al. , Transmission of endoplasmic reticulum stress and pro-inflammation from tumor cells to myeloid cells. Proceedings of the National Academy of Sciences, 2011. 108(16): p. 6561–6566. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199.Cubillos-Ruiz JR, et al. , ER stress sensor XBP1 controls anti-tumor immunity by disrupting dendritic cell homeostasis. Cell, 2015. 161(7): p. 1527–1538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Chen X, et al. , XBP1 promotes triple-negative breast cancer by controlling the HIF1α pathway. Nature, 2014. 508(7494): p. 103–107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201.Sheng X, et al. , IRE1α-XBP1s pathway promotes prostate cancer by activating c-MYC signaling. Nat Commun, 2019. 10(1): p. 323. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Logue SE, et al. , Inhibition of IRE1 RNase activity modulates the tumor cell secretome and enhances response to chemotherapy. Nat Commun, 2018. 9(1): p. 3267. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Bobrovnikova-Marjon E, et al. , PERK promotes cancer cell proliferation and tumor growth by limiting oxidative DNA damage. Oncogene, 2010. 29(27): p. 3881–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.White-Gilbertson S, Hua Y, and Liu B, The role of endoplasmic reticulum stress in maintaining and targeting multiple myeloma: a double-edged sword of adaptation and apoptosis. Front Genet, 2013. 4: p. 109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Bagratuni T, et al. , Characterization of a PERK Kinase Inhibitor with Anti-Myeloma Activity. Cancers (Basel), 2020. 12(10). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206.Fels DR and Koumenis C, The PERK/eIF2α/ATF4 module of the UPR in hypoxia resistance and tumor growth. Cancer Biology & Therapy, 2006. 5(7): p. 723–728. [DOI] [PubMed] [Google Scholar]
- 207.Bi M, et al. , ER stress-regulated translation increases tolerance to extreme hypoxia and promotes tumor growth. Embo j, 2005. 24(19): p. 3470–81. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208.Blais JD, et al. , Perk-dependent translational regulation promotes tumor cell adaptation and angiogenesis in response to hypoxic stress. Molecular and cellular biology, 2006. 26(24): p. 9517–9532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209.Gupta S, McGrath B, and Cavener DR, PERK regulates the proliferation and development of insulin-secreting beta-cell tumors in the endocrine pancreas of mice. PloS one, 2009. 4(11): p. e8008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210.Kusio-Kobialka M, et al. , The PERK-eIF2α phosphorylation arm is a pro-survival pathway of BCR-ABL signaling and confers resistance to imatinib treatment in chronic myeloid leukemia cells. Cell Cycle, 2012. 11(21): p. 4069–78. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211.Yang XF, et al. , CML28 is a broadly immunogenic antigen, which is overexpressed in tumor cells. Cancer Res, 2002. 62(19): p. 5517–5522. [PMC free article] [PubMed] [Google Scholar]
- 212.Yang XF, et al. , CML66, a broadly immunogenic tumor antigen, elicits a humoral immune response associated with remission of chronic myelogenous leukemia. Proc Natl Acad Sci U S A, 2001. 98(13): p. 7492–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213.Yan Y, et al. , A novel mechanism of alternative promoter and splicing regulates the epitope generation of tumor antigen CML66-L. J Immunol, 2004. 172(1): p. 651–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214.Meares GP, et al. , PERK-dependent activation of JAK1 and STAT3 contributes to endoplasmic reticulum stress-induced inflammation. Mol Cell Biol, 2014. 34(20): p. 3911–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215.Garg AD, et al. , ER stress-induced inflammation: does it aid or impede disease progression? Trends Mol Med, 2012. 18(10): p. 589–98. [DOI] [PubMed] [Google Scholar]
- 216.Zhang W, et al. , PERK EIF2AK3 control of pancreatic beta cell differentiation and proliferation is required for postnatal glucose homeostasis. Cell Metab, 2006. 4(6): p. 491–7. [DOI] [PubMed] [Google Scholar]
- 217.Gao Y, et al. , PERK is required in the adult pancreas and is essential for maintenance of glucose homeostasis. Mol Cell Biol, 2012. 32(24): p. 5129–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218.Lee AS, GRP78 induction in cancer: therapeutic and prognostic implications. Cancer research, 2007. 67(8): p. 3496–3499. [DOI] [PubMed] [Google Scholar]
- 219.Jamora C, Dennert G, and Lee AS, Inhibition of tumor progression by suppression of stress protein GRP78/BiP induction in fibrosarcoma B/C10ME. Proceedings of the National Academy of Sciences, 1996. 93(15): p. 7690–7694. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 220.Pavlović N and Heindryckx F, Targeting ER stress in the hepatic tumor microenvironment. Febs j, 2022. 289(22): p. 7163–7176. [DOI] [PubMed] [Google Scholar]
- 221.Liu Y, Tan Z, and Yang Y, Negative feedback and modern anti-cancer strategies targeting the ER stress response. FEBS Lett, 2020. 594(24): p. 4247–4265. [DOI] [PubMed] [Google Scholar]
- 222.Rabouw HH, et al. , Small molecule ISRIB suppresses the integrated stress response within a defined window of activation. Proc Natl Acad Sci U S A, 2019. 116(6): p. 2097–2102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223.Axten JM, et al. , Discovery of 7-methyl-5-(1-{[3-(trifluoromethyl)phenyl]acetyl}−2,3-dihydro-1H-indol-5-yl)-7H-pyrrolo[2,3-d]pyrimidin-4-amine (GSK2606414), a potent and selective first-in-class inhibitor of protein kinase R (PKR)-like endoplasmic reticulum kinase (PERK). J Med Chem, 2012. 55(16): p. 7193–207. [DOI] [PubMed] [Google Scholar]
- 224.Axten JM, et al. , Discovery of GSK2656157: An Optimized PERK Inhibitor Selected for Preclinical Development. ACS Med Chem Lett, 2013. 4(10): p. 964–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 225.Atkins C, et al. , Characterization of a novel PERK kinase inhibitor with antitumor and antiangiogenic activity. Cancer Res, 2013. 73(6): p. 1993–2002. [DOI] [PubMed] [Google Scholar]
- 226.Tabas I, The role of endoplasmic reticulum stress in the progression of atherosclerosis. Circ Res, 2010. 107(7): p. 839–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227.Hotamisligil GS, Endoplasmic reticulum stress and the inflammatory basis of metabolic disease. Cell, 2010. 140(6): p. 900–917. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228.Bhandary B, et al. , An Involvement of Oxidative Stress in Endoplasmic Reticulum Stress and Its Associated Diseases. International Journal of Molecular Sciences, 2013. 14(1): p. 434–456. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 229.Kassan M, et al. , Endoplasmic reticulum stress is involved in cardiac damage and vascular endothelial dysfunction in hypertensive mice. Arterioscler Thromb Vasc Biol, 2012. 32(7): p. 1652–61. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230.Gargalovic PS, et al. , The unfolded protein response is an important regulator of inflammatory genes in endothelial cells. Arterioscler Thromb Vasc Biol, 2006. 26(11): p. 2490–6. [DOI] [PubMed] [Google Scholar]
- 231.Li Y, et al. , Free Cholesterol-loaded Macrophages Are an Abundant Source of Tumor Necrosis Factor-α and Interleukin-6: MODEL OF NF-κB- AND MAP KINASE-DEPENDENT INFLAMMATION IN ADVANCED ATHEROSCLEROSIS*. Journal of Biological Chemistry, 2005. 280(23): p. 21763–21772. [DOI] [PubMed] [Google Scholar]
- 232.Isodono K, et al. , PARM-1 is an endoplasmic reticulum molecule involved in endoplasmic reticulum stress-induced apoptosis in rat cardiac myocytes. PLoS One, 2010. 5(3): p. e9746. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233.Arias-Loza P-A, et al. , Both Estrogen Receptor Subtypes, α and β, Attenuate Cardiovascular Remodeling in Aldosterone Salt–Treated Rats. Hypertension, 2007. 50(2): p. 432–438. [DOI] [PubMed] [Google Scholar]
- 234.Deng J, et al. , Translational repression mediates activation of nuclear factor kappa B by phosphorylated translation initiation factor 2. Molecular and cellular biology, 2004. 24(23): p. 10161–10168. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235.Bonomini F, et al. , Atherosclerosis and oxidative stress. Histol Histopathol, 2008. 23(3): p. 381–90. [DOI] [PubMed] [Google Scholar]
- 236.Kang −TY, et al. , - Role of PERK/eIF2α/CHOP Endoplasmic Reticulum Stress Pathway in Oxidized Low-density Lipoprotein Mediated Induction of Endothelial Apoptosis. - Biomedical and Environmental Sciences, 2016. - 29(- 12): p. - 868. [DOI] [PubMed] [Google Scholar]
- 237.Monroy MA, et al. , Chronic kidney disease alters vascular smooth muscle cell phenotype. Front Biosci (Landmark Ed), 2015. 20: p. 784–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 238.Chattopadhyay A, et al. , Preventing Cholesterol-Induced Perk (Protein Kinase RNA-Like Endoplasmic Reticulum Kinase) Signaling in Smooth Muscle Cells Blocks Atherosclerotic Plaque Formation. Arterioscler Thromb Vasc Biol, 2022. 42(8): p. 1005–1022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 239.Andreone BJ, Larhammar M, and Lewcock JW, Cell Death and Neurodegeneration. Cold Spring Harb Perspect Biol, 2020. 12(2). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 240.Guo T, et al. , Molecular and cellular mechanisms underlying the pathogenesis of Alzheimer’s disease. Mol Neurodegener, 2020. 15(1): p. 40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 241.Abisambra JF, et al. , Tau accumulation activates the unfolded protein response by impairing endoplasmic reticulum-associated degradation. J Neurosci, 2013. 33(22): p. 9498–507. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 242.Hoozemans JJ, et al. , The unfolded protein response is activated in Alzheimer’s disease. Acta Neuropathol, 2005. 110(2): p. 165–72. [DOI] [PubMed] [Google Scholar]
- 243.Song J, et al. , Apoptosis signal regulating kinase 1 (ASK1): potential as a therapeutic target for Alzheimer’s disease. Int J Mol Sci, 2014. 15(2): p. 2119–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 244.Hoozemans JJ, et al. , Activation of the unfolded protein response in Parkinson’s disease. Biochem Biophys Res Commun, 2007. 354(3): p. 707–11. [DOI] [PubMed] [Google Scholar]
- 245.Ryu EJ, et al. , Endoplasmic reticulum stress and the unfolded protein response in cellular models of Parkinson’s disease. J Neurosci, 2002. 22(24): p. 10690–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246.Shacham T, Sharma N, and Lederkremer GZ, Protein Misfolding and ER Stress in Huntington’s Disease. Front Mol Biosci, 2019. 6: p. 20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 247.Fortrie G, The Aftermath of Acute Kidney Injury. 2019. [DOI] [PMC free article] [PubMed]
- 248.Inagi R, Endoplasmic reticulum stress in the kidney as a novel mediator of kidney injury. Nephron Exp Nephrol, 2009. 112(1): p. e1–9. [DOI] [PubMed] [Google Scholar]
- 249.Cybulsky AV, Endoplasmic reticulum stress, the unfolded protein response and autophagy in kidney diseases. Nat Rev Nephrol, 2017. 13(11): p. 681–696. [DOI] [PubMed] [Google Scholar]
- 250.Ricciardi CA and Gnudi L, The endoplasmic reticulum stress and the unfolded protein response in kidney disease: Implications for vascular growth factors. J Cell Mol Med, 2020. 24(22): p. 12910–12919. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 251.Lindenmeyer MT, et al. , Proteinuria and hyperglycemia induce endoplasmic reticulum stress. J Am Soc Nephrol, 2008. 19(11): p. 2225–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252.Shu S, et al. , Endoplasmic reticulum stress is activated in post-ischemic kidneys to promote chronic kidney disease. EBioMedicine, 2018. 37: p. 269–280. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253.Zhang M, et al. , Chop deficiency prevents UUO-induced renal fibrosis by attenuating fibrotic signals originated from Hmgb1/TLR4/NFκB/IL-1β signaling. Cell Death Dis, 2015. 6(8): p. e1847. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 254.Cao AL, et al. , Ursodeoxycholic acid and 4-phenylbutyrate prevent endoplasmic reticulum stress-induced podocyte apoptosis in diabetic nephropathy. Lab Invest, 2016. 96(6): p. 610–22. [DOI] [PubMed] [Google Scholar]
- 255.Maekawa H and Inagi R, Stress Signal Network between Hypoxia and ER Stress in Chronic Kidney Disease. Front Physiol, 2017. 8: p. 74. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
