Next Article in Journal
Polyvinyl-Pyrrolidone-Coated Silver Nanoparticles—The Colloidal, Chemical, and Biological Consequences of Steric Stabilization under Biorelevant Conditions
Next Article in Special Issue
The Heparanase Regulatory Network in Health and Disease
Previous Article in Journal
NMDARs Containing NR2B Subunit Do Not Contribute to the LTP Form of Hippocampal Plasticity: In Vivo Pharmacological Evidence in Rats
Previous Article in Special Issue
Thrombospondin-1 CD47 Signalling: From Mechanisms to Medicine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Therapeutic Approaches Targeting Proteostasis in Kidney Disease and Fibrosis

1
Graduate Institute of Toxicology, College of Medicine, National Taiwan University, Taipei 100233, Taiwan
2
Department of Physiology of Visceral Function and Body Fluid, Nagasaki University Graduate School of Biomedical Sciences, Nagasaki 852-8523, Japan
3
Department of Integrated Diagnostics & Therapeutics, National Taiwan University Hospital, Taipei 100225, Taiwan
4
Center for Biotechnology, National Taiwan University, Taipei 10672, Taiwan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(16), 8674; https://doi.org/10.3390/ijms22168674
Submission received: 13 July 2021 / Revised: 9 August 2021 / Accepted: 10 August 2021 / Published: 12 August 2021
(This article belongs to the Special Issue Proteolysis of Extracellular Matrix in Human Disease 2.0)

Abstract

:
Pathological insults usually disturb the folding capacity of cellular proteins and lead to the accumulation of misfolded proteins in the endoplasmic reticulum (ER), which leads to so-called “ER stress”. Increasing evidence indicates that ER stress acts as a trigger factor for the development and progression of many kidney diseases. The unfolded protein responses (UPRs), a set of molecular signals that resume proteostasis under ER stress, are thought to restore the adaptive process in chronic kidney disease (CKD) and renal fibrosis. Furthermore, the idea of targeting UPRs for CKD treatment has been well discussed in the past decade. This review summarizes the up-to-date literature regarding studies on the relationship between the UPRs, systemic fibrosis, and renal diseases. We also address the potential therapeutic possibilities of renal diseases based on the modulation of UPRs and ER proteostasis. Finally, we list some of the current UPR modulators and their therapeutic potentials.

1. Introduction

Kidney diseases have recently received considerable attention because the renal function is vulnerable to pathogenic insults, including inflammation, hypoxia, hypertension, and aging. Once acute kidney injury (AKI) occurs, the event frequently leads to the progression of chronic kidney disease (CKD) despite a transient recovery from AKI [1]. Renal tubulointerstitial fibrosis is a common pathway of advanced CKD, which is associated with vasoconstriction, capillary obliteration caused by fibrotic expansion, and finally the formation of a hypoxic microenvironment that worsens renal function [2,3].
There is accumulating evidence indicating that the disruption of endoplasmic reticulum (ER) homeostasis is involved in various pathological processes, including cancer, metabolic diseases, diabetes, neurodegenerative disorders, and liver dysfunction [4]. The unfolded protein responses (UPRs) signaling activated by the ER stress participates in the progression of AKI and CKD [5]. In addition, ER stress is also involved in the progression of organ fibrosis, including those of the kidney, liver, and lung [6]. Because of the limited therapeutic options for the retarding of CKD progression, the modulation of UPRs signaling has become an attractive target for drug discovery [7]. In this mini-review, we discuss the role of UPRs in renal diseases and renal fibrosis, highlight the therapeutic potentials of the modulation of UPRs and the proteostasis of ER, and, in particular, emphasize the role of inositol-requiring protein 1-X-box-binding protein 1 (IRE1-XBP1) signaling.

2. Fundamental Roles of ER Stress and Unfolded Protein Responses

The ER is the organelle where most secretory and transmembrane proteins are synthesized, modified, and folded into their correct conformations. Since the ER plays such an essential role in the maintaining of proteostasis, it must optimally control the quality of protein folding [8]. However, the folding capacity of the ER is susceptible to environmental stress, such as inflammation, oxidative stress, infection, or the deprivation of nutrients, which leads to the abundant accumulation of misfolded or unfolded protein in the ER, which is hence termed the “ER stress” [9]. Restoring the proteostasis of the ER will initiate the signaling of UPRs, thereby reducing protein load, increasing folding ability, or triggering cell apoptosis for as long as overwhelming ER stress persists [8]. Three ER transmembrane proteins, including IRE1, pancreatic eukaryotic translation initiation factor 2-alpha kinase (PKR-like protein kinase, PERK), and activating transcription factor 6 (ATF6), mediate the transduction of UPRs signaling. In normal proteostasis, the immunoglobulin heavy-chain binding protein (BiP) will conjugate with these three transmembrane proteins, and no UPRs cascade will be initiated. When ER stress occurs, BiP acts as an endogenous chaperon that binds to misfolded proteins, and therefore, detaches from the UPRs initiators, resulting in the activation of UPRs signaling [10].
Upon ER stress, IRE1 is activated after the dissociation of BiP, which leads to the conformational change of IRE1. This promotes its dimerization and trans-autophosphorylation, followed by activation of the kinase and RNase domains [11,12]. XBP1 RNA is spliced by unconventional endonuclease, causing a frameshift and generating spliced XBP1 (XBP1s), which possesses a completely different function from its intact form (XBP1u) [13]. XBP1s acts as a transcription activator, upregulating genes that encode ER-associated degradation (ERAD) and chaperone [14,15]. However, the overwhelming level of ER stress induces the activation of the kinase domain, and then recruits tumor necrosis factor receptor (TNFR)-associated factor-2 (TRAF2), before activating the signaling of c-Jun amino-terminal kinases (JNK), which are pro-apoptosis mediators [16]. In addition, hyper-activation of IRE1α is reported to activate regulated IRE1α-dependent degradation (RIDD) and digest a subset of mRNAs that encode secretory proteins [17,18].
The manner of activation of PERK is similar to the way in which IRE is activated. After dimerization and trans-autophosphorylation, PERK suppresses global protein translation by phosphorylating the α subunit of eukaryotic translation initiation factor-2 (eIF2α) [19]. However, ATF4, a critical transcriptional regulator in the PERK-eIF2α pathway, is upregulated by ribosomal skipping [20]. Following the activation of downstream stress-induced redox proteins, C/EBP-homologous protein (CHOP) [21], which is well recognized as a pro-apoptotic protein that mediates UPR-related cell death [22], directly activates both growth arrest and DNA damage-inducible protein-34 (GADD34) [23], the protein that dephosphorylates the phospho-eIF2α in cells under ER stress and helps cells to recover from translational inhibition [23].
Activated ATF6 translocates from the ER to the Golgi apparatus, and is cleaved by site-1 and site-2 proteases in the Golgi apparatus [24,25], thereby generating the cleaved fragment of ATF6 (cATF6), and then cATF6 enters the nucleus to activate the genes that enhance protein folding, including BiP, GRP94, calreticulin, protein disulfide isomerase (PDI), and XBP1 [26,27,28,29].
The three-branch axis orchestrates the process of UPRs, and their interdependency in terms of regulation is well documented. For instance, the downstream target genes of ATF6 can be compensated by XBP1 during acute silence of ATF6 [30]. Inhibition of PERK leads to the compensatory activation of XBP1s, and the inhibition of IRE1α contributes to the sustained activation of PERK and CHOP [31]. In addition, hyperactivation of IRE1α is found in the XBP1 deletion hepatocyte [32]. Recent evidence demonstrated that cell fate is determined by elaborate ER compensation during ER stress.

3. Homeostatic Role of Proteolysis through Adaptive UPRs Activation in Disease Progression

ERAD plays a pivotal role in the removal of misfolded protein to maintain the homeostasis and cell survival of ER. The process of ERAD is initiated by unfolded substrate recruitment, assisted by chaperones such as Bip and the ER degradation-enhancing alpha-mannosidase-like protein (EDEM) protein family. The tagged protein is eventually degraded by the 26S proteasome and resolves ER stress [33,34]. However, it may shut down the capacity of ERAD and activate apoptotic ER stress during the overloading of misfolded/unfolded proteins, as well as during oxidative stress or heat shock states [35]. Various ERAD-deficient mouse models developed organ dysfunction, including enteritis, obesity, and glucose intolerance [36,37,38]. In addition, ERAD inhibition was identified as a therapeutic target in cancer treatment [39,40]. Furthermore, the enhancement of ERAD through the overexpression of XBP1s reduced amyloid β-peptide accumulation in an Alzheimer’s disease model [41,42].

4. ER Stress-Mediated Autophagy and Proteostasis

There is strong evidence supporting the crosstalk between ER stress and autophagy [43,44,45]. Autophagy is triggered by a mechanistic rapamycin (mTOR) inhibition target and involves sequential steps, including the initiation of phagophore, which begins with Unc-51-like autophagy activating the complex formation and nucleation of kinase 1 (ULK1). Then, the autophagosome membrane elongation is assisted by the conjugation of autophagy-related protein (ATG) to LC3 phosphatidylethanolamine (PE) and the lysosome, and finally, proteolytic degradation is initiated [46].
The interplay between ER stress and autophagy has been frequently mentioned. For example, the accumulation of cytosol calcium from ER will activate UPRs, followed by the inhibition of mTOR and the induction of autophagy [47]. Margariti et al. and Ogata et al. demonstrated that autophagy transcriptionally induced the activation of beclin-1 via the IRE1/XBP1s and IRE1/JNK axis [48,49]. B’chir et al. elegantly showed that induction of PERK/eIF2α/ATF4 axis is essential for ATGs genes expression [50]. Furthermore, Fang et al. and Qi et al. showed that the chemical chaperone 4-PBA and TUDCA can attenuate STZ and obesity-induced diabetic nephropathy, extracellular matrix deposition, and autophagy in an ER-stress-dependent manner [51,52]. All of the above-mentioned reports support the connection between ER stress and autophagy.

5. UPRs and Systemic Fibrosis Progression

Many studies proposed the relationship between UPRs and organ fibrosis [6,53]. Familial interstitial pneumonia (FIP), a class of interstitial pneumonitis that may be caused by the genetic mutation in surfactant protein C (SPC) [54], suggests a potential connection between organ fibrosis and ER stress. SPC is secreted by type II alveolar epithelial cells in order to maintain alveolar distensibility. In vitro studies revealed that the mutation of SPC in the carboxyl domain leads to its accumulation, in a misfolded form, in the ER lumen [55]; furthermore, tissue samples from FIP patients with SPC mutation showed prominent expression of BiP, and XBP1 expression co-localized with fibrotic areas [56].
In murine models of cardiac fibrosis, subcutaneous injection of isoproterenol and angiotensin II-induced fibrosis activated UPR signaling and upregulated the pro-apoptotic expression of CHOP. In these models, the severity of fibrosis was attenuated through the administration of chemical chaperone 4-phenylbutyric acid (4-PBA) [57,58].

6. UPRs in Renal Disease and Fibrosis

Renal fibrosis is the final common pathway of CKD and end-stage renal disease (ESRD) [59], which results from the loss of parenchymal due to the occurrence of natural senescence, diabetes, or acute kidney insults. It has been well documented that AKI is recognized as a cause of long-term risk of CKD or ESRD and maladaptive repair of kidney injury, leading to renal fibrosis and the transition from AKI to CKD [60,61]. Myofibroblasts play a critical role in the inducing of excessive deposition of the extracellular matrix, which contributes to renal fibrosis. Myofibroblasts can originate from the activation of renal interstitial fibroblasts, perivascular fibroblasts, pericytes, and bone marrow-derived mesenchymal cells, as well as the transition of endothelial cells or tubular epithelial cells [62,63,64,65,66]. Recent studies revealed that perivascular fibroblasts and pericytes, but not injured tubular epithelial cells, transdifferentiate into myofibroblasts and contribute to fibrosis in renal fibrosis animal models. However, an emerging concept is that kidney damage caused by AKI or unresolved injuries leads to prolonged cell arrest in the cell cycle G2/M phase and leads to the appearance of profibrotic and proinflammatory features in tubule cells. Profibrogenic growth factors and inflammatory cytokines secreted from injured tubules can stimulate the proliferation of fibroblasts and the production of extracellular matrix, and eventually contribute to progressive renal fibrosis. In addition, many publications also revealed that UPRs signaling is involved in the progression of renal diseases, which is triggered by hypoxia, oxidative stress, inflammation, high glucose, and functional genetic deficiency of the glomerular protein [67].

6.1. Disturbance of UPR Contributes to AKI-to-CKD Transition

Renal ischemia-reperfusion injury is a common cause of AKI that results in hypoxia, and ER stress is well recognized as the initial response to ischemia-reperfusion injury [68,69]. A calcitonin/calcitonin gene-related peptide, namely intermedin, ameliorates renal ischemia-reperfusion injury by inhibiting ER stress-mediated apoptotic signaling, for example, the expression of CHOP and caspase-12 [70]. The unilateral ureteral obstruction (UUO) model activated all three effectors of the UPR signaling during the development of renal fibrosis. The pro-apoptotic signals, such as CHOP, caspase-12, JNK, and Bax, were also increased [65]. Furthermore, Fan and Xiao et al. showed that the dysregulation of UPRs was correlated to the severity of the progression from AKI to CKD in humans, with upregulated expression of Bip, p-PERK, and CHOP and reduced expression of XBP1s in patients with progressive AKI renal biopsy [71]. Jao et al. revealed that the dysregulation of UPRs induced the accumulation of lipids, as well as renal fibrosis [72]. The renal fibrosis induced in the UIRI mice model coincides with the accumulation of lipids and the activation of ATF6 in tubular epithelial cells. Furthermore, ATF6 knockout mice demonstrated less tubulointerstitial fibrosis and lipid accumulation through the activation of PPARα.
The production of reactive oxygen species (ROS) during renal ischemia-reperfusion injury also causes the pathogenesis of CKD progression [73]. Antioxidant therapy for CKD patients showed significant benefits, including attenuation of the risk of ESRD development, reduced serum creatinine levels, and improved creatinine clearance [74]. The production of ROS interferes with cellular redox-dependent metabolism and protein-folding capacity, resulting in the accumulation of misfolded proteins in the ER [75]. The antioxidative effects of UPRs were also found during the stimulation of cells against ROS [76]. Nuclear factor E2-related factor 2 (Nrf2) is an antioxidative transcription regulator that resists oxidative stress through the activation of antioxidative genes, such as catalase, heme oxygenase-1 (HO-1), and superoxide dismutase [77]. In terms of the connection between Nrf2 and three branches of UPRs initiators, it was demonstrated that Nrf2 is the downstream target of the ATF6 [78], IRE1/JNK [76], and PERK pathways [79]. The PERK-Nrf2 pathway, for example, plays an essential role in the maintenance of redox homeostasis, as shown by the fact that a deficiency of PERK leads to the accumulation of ROS in cells [80]. Cadmium induces kidney injury through the generation of ROS and leads to ER stress-mediated apoptosis [81]. This evidence suggests that ER stress is an important pathogenic mediator of renal diseases.

6.2. Dysregulation of UPR Mediates Renal Fibrosis in Diabetic Nephropathy and Podocyte Defect Mice Model

Diabetic nephropathy (DN) accounts for up to 40% of incident ESRD [82]. In the streptozotocin (STZ)-induced DN model, the upregulation of BiP and CHOP, and the activation of PERK signaling, are observed in 22-month-old mice with DN accompanied by tubulointerstitial fibrosis and extensive inflammatory cell infiltration [83]. The renal renin-angiotensin system (RAS) is well known for governing blood pressure and the homeostasis of target organs. In the cascade of RAS, angiotensin II is the key regulator that contributes to vasoconstriction. In addition, the proinflammatory and profibrogenic effects of angiotensin II are widely discussed in the progression of CKD [84]. Therefore, angiotensin-converting enzyme inhibitor (ACEI) or angiotensin receptor blocker (ARB) are the standard treatments for hypertension or heart failure patients. Evidence showed that ACEIs reduce the apoptosis of renal tubular cells and suppress the signaling of UPRs, as demonstrated by, for example, the activation of phospho-eIF2α and phospho-PERK in STZ-induced diabetic rat models [85]. Human kidney biopsy samples demonstrated that diabetic nephropathy had higher expression of BiP, XBP1, and CHOP, which is consistent with in vitro findings. In addition, albumin and high glucose administration induced the activation of ER stress in human and rodent renal tubules [86]. Furthermore, the deletion of podocyte-specific IRE1a and XBP1s in mice led to more severe cases of albuminuria, glomerular basement membrane thickening, and ER stress induction [87,88].

7. Therapeutic Strategies: Targeting the IRE1-XBP1 and PERK-eIF2α Axis

7.1. IRE1-XBP1 Axis of UPRs

The XBP1 arm of the UPRs is generally recognized as cytoprotective. As mentioned above, the three UPRs branches are all activated in the UUO model [65]. In this model, XBP1s has been downregulated during the development of renal fibrosis, which ARB could reverse. Somlo and colleagues reported that site-specific deficiency of XBP1 in podocytes resulted in severe albuminuria, glomerulosclerosis, and kidney fibrosis in a Sec63 and XBP1 double knockout model [89,90]. In this first part of the study, the authors demonstrated the accumulation of unfolded proteins without proteinuria or pathological features in podocyte-specific Sec63 or XBP1s single knockout mice. However, podocyte-specific Sec63 and XBP1s double-knockout mice developed defects in the integrity of the glomerular filtration barrier and progressive tubulointerstitial fibrosis in association with the loss of podocytes, which occurred through the activation of the JNK-apoptotic pathway in 2-month-old mice. Moreover, the re-expression of XBP1s in vivo completely rescues chronic tubulointerstitial kidney injury in XBP1 and Sec63 double knockout mice. Madhusudhan et al. also revealed an essential role of XBP1 in DN. They demonstrated that sXBP1 lies downstream of insulin signaling, and attenuates insulin signaling in podocytes through the genetic ablation of the insulin receptor or the regulatory subunits phosphatidylinositol 3-kinase (PI3K) p85α or p85β, which impair sXBP1 nuclear translocation and exacerbate DN [87] (Figure 1).
Oxidative stress is also a potential contributor to renal fibrosis [91]. Liu and colleagues showed that mouse embryonic fibroblasts with a genetic deficiency of XBP1 are more sensitive to H2O2-induced apoptosis. The anti-oxidative properties of XBP1 resulted from the transcriptional upregulation of catalase that occurred through binding to its promoter region [92]. Another group also showed that XBP1 attenuated disturbed flow-induced oxidative stress through the upregulation of heme oxygenase 1 (HO-1) in endothelial cells [93]. Angiogenin (ANG) is a ribonuclease that promotes the adaptation of tissue to injury. By promoting the cleavage of tRNA, ANG plays a physiologically relevant, ER stress-mediated, adaptive role in the translational control of kidney injuries, in a IRE1-XBP1-dependent manner [94,95] (Figure 1). Inflammation also plays a critical role in the development of renal fibrosis. In a recent study, ER stress preconditioning could attenuate LDL-induced inflammation in human mesangial cells. Its underlying mechanism is mainly through the induction of XBP1 followed by the blocking of the IRE1α/IKK/NF-κB-mediated inflammatory response [94]. Collectively, modulation of the IRE1-XBP1 axis might be a favorable therapeutic target among ER stress-mediated injuries.

7.1.1. IRE1-XBP1-Mediated Inflammasome in Renal Fibrosis

Kidney inflammation is considered to be a pathogenic factor that triggers CKD progression [96]. Prolonged inflammation resulting from mononuclear infiltration or the deposition of immunogens leads to the recruitment of fibrocytes, and consequently, renal fibrosis [97]. NOD-, LRR- and pyrin domain-containing protein 3 (NLRP3), a protein oligomerization that forms inflammasomes, is involved in the cleavage of pro-IL-1β and pro-IL-18, which contribute to renal inflammation and fibrosis in various non-diabetic kidney diseases [98,99,100,101]. NLRP3 knockout ameliorates tubular injury, inflammation, and fibrosis in the UUO kidney. Notably, the hyperactivation of IRE1α results in the degradation of miR-17 following the activation of the NLRP3 inflammasomes. By blocking the activation of XBP1s by STF-083010, the RNA endonuclease inhibitor attenuates the secretion of IL-1β during treatment with thapsigargin [102]. Therefore, the ER stress-mediated activation of NLRP3 inflammasome can be specifically attenuated by the activation of IRE1-XBP1s.

7.1.2. Application of Small Molecular Compounds Targeting the IRE1-XBP1 Pathway

Kinase-Inhibiting RNase Attenuators (KIRA)

Hyperactivation of IRE1 may trigger the RIDD and lead to cell apoptosis. Recently, Dr. FR Papa has developed a KIRA that allosterically inhibits IRE1α kinase activity by breaking oligomers. KIRA prevents hyper-oligomerization of IRE1 and maintains the RNA endonuclease of IRE1, which allows XBP1 to be spliced properly. It successfully protected pancreatic β cells and retinal cells from ER stress-induced injury in vivo and in vitro [103]. It might be a vital molecule under research in the future; however, more tests are needed in different disease models.

Quercetin

Quercetin is one of the most abundant flavonoids with antioxidative properties in the human diet. Recent studies demonstrated its ability in the activation of IRE1 RNase and the inhibition of IRE1-JNK signaling [104,105,106]. Previous research studies showed that quercetin has a renoprotection effect in many types of kidney injuries, including DN [107], cadmium toxicity [108], and UUO kidney (Figure 1) [109]. Furthermore, in vitro investigation showed that the anti-fibrotic effect of quercetin is mediated through the attenuation of TGF-β1-induced collagen I and α-SMA expression in normal rat kidney fibroblast cells (NRK-49F) [110] and UUO kidney [111]. In addition, another research study demonstrated that quercetin protects glomerular endothelial cells from asymmetric dimethylarginine (ADMA)-induced apoptosis through PERK- and IRE1-associated and TGF-β-enhanced pathways [112] (Table 1).

XBP1 Agonists

Although the hyperactivation of IRE1-induced renal inflammation and fibrosis were found, the therapeutic effect of the IRE1 RNase targeted molecule, XBP1s, was proved by Lian Qiu et al. in a mouse model of ulcerative colitis (UC) or inflammatory bowel disease (IBD). IBD is a chronic intestinal inflammatory disease that revealed hypomorphic variants of XBP1 as a susceptibility factor that leads to IBD [113]. The XBP1 agonists, HLJ2 and (±)-8-ADC, are monomeric compounds that are extracted and modified from Ranunculaceae and Papaveraceae plant families [114]. These XBP1 agonists exert an anti-inflammatory effect toward dextran sulfate sodium-induced colitis through the inhibition of NF-κB. Furthermore, HLJ2 demonstrated an anti-epithelial mesenchymal transition (EMT) effect under TGF-β1 stimulation [115,116]. These findings suggest the promising therapeutic value of XBP1 agonists in inflammatory disease and organ fibrosis (Table 1).

7.2. PERK-eIF2α Axis

7.2.1. Application of Salubrinal

Salubrinal, a selective inhibitor of GADD34-phosphatase-1 (PP1), which prevents dephosphorylation of eIF2α, was shown to protect cells from ER stress-induced apoptosis [117]. Cadmium increased the generation of ROS with subsequent induction of ER stress in a cultured renal proximal tubular cell line [118,119]. The administration of Salubrinal attenuated the cadmium-induced expression of BiP and CHOP, and the activation of cell death signaling of JNK, in HK-2 cells [120]. Salubrinal also ameliorates podocyte damage caused by hyperglycemia, as well as damage induced by other xenotoxicant agents such as arsenic, paraquat, cyclosporine, and cisplatin [51,121]. However, our previous animal study of the effects of salubrinal on cisplatin-induced renal cell damage showed a different conclusion from that obtained through in vitro studies. We found that salubrinal enhanced cisplatin-induced nephrotoxicity in mice, and this was accompanied by the activation of ATF4 and CHOP, and the cleavage of caspases-12, 9, and 3 [64]. Since salubrinal itself did not trigger renal cell injury in mice, the acceptable reason for this discrepancy in the results might be the hyperactivation of eIF2α and the enhanced cisplatin-triggered oxidative stress in the kidney.
More recently, salubrinal was shown to prevent the suppression of HK-2 cell proliferation by indoxyl sulfate [122]. In addition to salubrinal, other PERK signaling modulators including Guanabenz and GSK2606414 were investigated. Guanabenz, the α2-adrenergic receptor agonist that was originally used to treat hypertension, could prevent the dephosphorylation of eIF2α by GADD34 through competitive binding to PP1c [123]. GSK2606414 inhibits the activation of PERK by binding to the active site of the PERK kinase domain, thereby repressing tumor growth in several mouse xenograft models [124] (Table 1).

7.2.2. CHOP Is the Most Potential Therapeutic Target

It is well recognized that CHOP expression is involved in many diseases, and the modulation of CHOP expression might be the most potent target to retard the progression of renal disease and fibrosis. In the ischemia-reperfusion injury model, CHOP-deficiency significantly reduced serum creatinine and BUN levels compared to wild-type mice. Histological scores also showed decreased renal tubule dilation, tubular cell death, and cast formation in CHOP-deficient mice. This effect may be mediated by the regulation of pro-apoptosis, inflammation, and ROS-related signals, including cyclo-oxygenase-2 (COX-2), Caspase-3, and Caspase-8 [66]. CHOP-deficiency also attenuates renal fibrosis in the UUO model through a reduction in inflammatory infiltration, collagen deposition, and interstitial fibrotic area, and a repression of the expression of fibrotic markers including fibronectin, collagen, and α-SMA. The underlying mechanism may be mediated through the attenuation of CHOP-mediated inflammatory factors including IL-1β and TGF-β1 production, as well as PI3K/Akt activity [125]. These findings are consistent with the liver fibrotic murine model, in which CHOP deficiency attenuates hepatic fibrosis, inflammatory gene expression, and oncogenesis [126].
In terms of therapeutic approaches, one notable study on malignant cells states that the PERK-CHOP pro-apoptotic pathway is the barrier to malignancy. At the same time, the deletion of CHOP increases tumor incidence, and the molecular chaperone p58IPK selectively attenuates PERK-CHOP-mediated apoptosis [127]. Since the repression of CHOP expression leads to some beneficial effects, the modulation of CHOP expression might offer potential in dealing with ER stress-mediated kidney injury. In contrast, CHOP deficiency results in more severe kidney injury in the LPS-induced acute kidney injury model [128]. LPS injection induces the activation of UPRs, including XBP1s, GRP78, GRP94, and CHOP. However, CHOP deficiency leads to higher BUN levels, albuminuria, and apoptotic cell death compared to WT mice under LPS treatment. The authors suggest that the mechanism is independent of the modulation of UPR singling; instead, it is mainly caused by changes in the inflammatory response of the downstream target C/EBP family proteins. CHOP might play an anti-inflammation role in LPS-induced kidney injury. Additionally, we should carefully discuss the role of CHOP in various kidney diseases, especially in kidney damage that is mainly caused by the inflammatory response.

8. Other Beneficial Effects in Moderating the ER Protein Homeostasis

8.1. Pre-Conditioning ER Stress

ER stress pre-conditioning showed protective effects in many studies, both in vitro and in vivo. In 1999, Bush and colleagues revealed that ER stress pre-conditioning protects ATP depletion-induced cell damage [129], and that pretreatment with tunicamycin could protect mice from acute ischemic injury [130]. This protective effect could be seen in other in vitro disease models, such as prolonged ATP depletion or oxidative stress-induced cardiomyocytes injury [131]. The underlining mechanisms may involve the prevention of increased intracellular Ca2+ concentration and ERK activation, and the decreasing of JNK activation [132], or they may be caused by the preserving of intercellular junctions, cytoarchitecture, and cell-substratum interactions in ATP-depleted epithelial cells. In addition, the administration of the ER stress inducers, tunicamycin or thapsigargin, significantly reduced mesangial proliferation and the adhesion of Bowman’s capsule to the glomerular tuft and proteinuria in a rat glomerulonephritis model [62].

8.2. AMPK Activation

AMP-activated protein kinase (AMPK) is a serine/threonine kinase that is thought to be a metabolism modulator that controls cellular and whole-body energy balance [133]. Recent studies showed that AMPK activation might protect renal cells from ER stress-induced injury. Administration of metformin, an AMPK activator, protects renal tubular cells against albumin loading-induced ER stress via inhibition of ROS through the induction of thioredoxin, an endogenous antioxidative molecule, and the suppression of BiP expression in an albumin-overloaded rat model [134]. Furthermore, AMPK suppressed tunicamycin, thapsigargin and/or TGF-β, angiotensin II, aldosterone, and high-glucose-induced ER stress in tubular epithelial cells [135]. Metformin also suppressed ER stress and fibrosis in both tunicamycin-induced AKI and UUO mouse models [135].

8.3. Roles of Chemical Chaperone 4-PBA and TUDCA

Both 4-phenylbutyrate (4-PBA) and TUDCA are small molecular weight chemical chaperones. The FDA approved 4-PBA for children with urea cycle disorders. They were also shown to have therapeutic potential for neurodegenerative diseases [136] and several types of cancers [137,138,139]. Both compounds assist in the adequate folding of protein in order to relieve ER stress-associated diseases [69,140,141], including renal fibrosis and diabetic nephropathy. In STZ-induced diabetes nephropathy, 4-PBA and TUDCA reduced the excretion of albuminuria and the expression levels of BiP, ATF6, PERK, JNK, and CHOP, and the inflammatory mediators. The chemical chaperons also restored the adaptive UPRs molecule XBP1s [52,87]. Furthermore, the chemical chaperones exerted an anti-fibrotic effect in post-injured UIRI or UUO mice models. Notably, recent research found that ER stress is involved in angiotensin II-induced NLRP3 inflammasome activation, and pretreatment with 4-PBA could reduce the expression of both NLRP3 and inflammatory cytokines, suggesting that 4-PBA serves as a potential agent to reduce renal fibrosis [142] (Table 1).
Table 1. Summarized the therapeutic effect of UPR modulators.
Table 1. Summarized the therapeutic effect of UPR modulators.
ChemicalMechanism of ActionAnimal ModelTherapeutic EffectRef.
4-PBAunfolded protein foldingUIRI, UUO, Dahl salt-sensitive rat, STZ-DNYes[57,69,140,141]
GSK2606414PERK inhibitorPrion infected miceNot clear[143]
GuanabenzeIF2α phosphatase inhibitorN/ANot clear[7,123]
KIRA6IRE1 RNase inhibitorAkita MouseYes[103]
SalubrinaleIF2α phosphatase inhibitorCisplatin, Cadmium, Arsenic, Paraquat, CyclosporineNot clear[64,120,121]
TUDCAunfolded protein foldingUIRI, STZ-induced DNYes[87,140]
HLJ2XBP1s agonistUCYes[115,144]
(±)-8-ADCXBP1s agonistUCYes[116]
QuercetinIRE1 RNase activatorUUO, DN, Cadmium, ADAMYes[104,105,106,107,108,109,110,111,112]

9. Conclusions

A number of therapeutic strategies related to UPR signaling were investigated in recent years; however, most focused on CNS disease, cancer, or diabetes [126,145]. In this article, we highlighted the findings of recent publications on the use of UPRs signals to target renal diseases or fibrosis progression either directly or indirectly, as well as some UPR modulators. We suggest that the modulation of phosphorylation-dependent PERK and IRE1 activities might be a promising direction in future pharmaceutical investigations [146]. As for the UPR modulator, there is lack of prominent reports, but based on the mechanisms of pathophysiology, it might still be a potential therapeutic target. Nevertheless, UPR modulators play the role of a double-edged sword, as they might induce either the systemic activation or inhibition of UPRs signaling, and be accompanied by unpredictable cellular stress. However, it is worth noting that the activation of adaptive UPR molecules, especially XBP1s in DN, podocyte-defective, or inflammatory models, is a promising therapeutic approach. Furthermore, the elimination of unfolded proteins or the enhancement of the ERAD function is also beneficial for kidney injury. In contrast, the role of the apoptotic UPR molecule CHOP is still not conclusive. For clinical applications of UPR modulators in renal fibrosis, some candidates, such as 4-PBA and TUDCA, are involved in ongoing DN clinical trials. Still, most of the compounds were only confirmed in vivo or in vitro. Therefore, more studies are required to confirm the beneficial effects of the modulation of UPRs on renal diseases.

Author Contributions

J.-H.C., C.-H.W. and C.-K.C. contributed to the manuscript’s design and writing. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by grants from the Taiwan Ministry of Science and Technology (MOST-104-2314-B-002-126-MY3, MOST-107-2314-B-002-027-MY3 and MOST- 110-2314-B-002 -130) and grants from the National Taiwan University Hospital (NTUH-106-S3574 and NTUH-107-S3826, NTUH.108-P02, NTUH-110-S5063).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank the staff of the Second Core Lab, Department of Medical Research, National Taiwan University Hospital for technical support during the study.

Conflicts of Interest

The authors declare no conflict of interest in relation to this review.

Abbreviations

4-PBA4-phenylbutyric acid
ACEIAngiotensin-converting enzyme inhibitor
AKIAcute kidney injury
AMPKAMP-activated protein kinase
ANGAngiogenin
ATF4Activating transcription factor 4
ATF6Activating transcription factor 6
BiPBinding immunoglobulin protein
CHOPCAAT/enhancer-binding protein (C/EBP) homologous protein
CKDChronic kidney disease
COX-2Cyclo-oxygenase-2
CREBcAMP response element-binding protein
DNDiabetic nephropathy
EDEMER degradation-enhancing alpha-mannosidase-like protein
eIF2αEukaryotic translation initiation factor 2α
EMTEpithelial-mesenchymal transition
EREndoplasmic reticulum
ERADER-associated protein degradation
FIPFamilial interstitial pneumonia
GADD34Growth arrest and DNA-damage-inducible 34
GRP78Glucose-regulated protein 78
HDAC3Histone deacetylase 3
IBDInflammatory bowel disease
IRE1αInositol-requiring enzyme 1 α
JNKc-Jun amino-terminal kinases
KIRAKinase-Inhibiting RNase Attenuators
NLRP3NOD-, LRR- and pyrin domain-containing protein 3
PDIProtein disulfide isomerase
PP1Phosphatase-1
RIDDRegulated IRE1-dependent decay
ROSReactive oxygen species
SPCSurfactant protein C
STZStreptozotocin
TNRFTumor necrosis factor receptor
TRAF2Tumor necrosis factor receptor-associated factor 2
TUDCATauroursodeoxycholic acid
UCUlcerative colitis
UPRUnfolded protein response
UUOUnilateral ureteral obstruction
XBP1X-box binding protein 1
XBP1sSpliced isoform of XBP1
XBP1uUnspliced isoform of XBP1

References

  1. Wald, R.; Quinn, R.R.; Luo, J.; Li, P.; Scales, D.C.; Mamdani, M.M.; Ray, J.G. Chronic Dialysis and Death Among Survivors of Acute Kidney Injury Requiring Dialysis. JAMA 2009, 302, 1179–1185. [Google Scholar] [CrossRef] [Green Version]
  2. Risdon, R.A.; Sloper, J.C.; De Wardener, H.E. Relationship between Renal Function and Histological Changes Found in Renal-Biopsy Specimens from Patients with Persistent Glomerular Nephritis. Lancet 1968, 2, 363–366. [Google Scholar] [CrossRef]
  3. Clements, M.E.; Chaber, C.J.; Ledbetter, S.R.; Zuk, A. Increased Cellular Senescence and Vascular Rarefaction Exacerbate the Progression of Kidney Fibrosis in Aged Mice Following Transient Ischemic Injury. PLoS ONE 2013, 8, e70464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Wang, S.; Kaufman, R.J. The impact of the unfolded protein response on human disease. J. Cell Biol. 2012, 197, 857–867. [Google Scholar] [CrossRef] [Green Version]
  5. Inagi, R.; Ishimoto, Y.; Nangaku, M. Proteostasis in endoplasmic reticulum—New mechanisms in kidney disease. Nat. Rev. Nephrol. 2014, 10, 369–378. [Google Scholar] [CrossRef]
  6. Tanjore, H.; Lawson, W.E.; Blackwell, T.S. Endoplasmic reticulum stress as a pro-fibrotic stimulus. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 1832, 1832, 940–947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Hetz, C.; Chevet, E.; Harding, H. Targeting the unfolded protein response in disease. Nat. Rev. Drug Discov. 2013, 12, 703–719. [Google Scholar] [CrossRef] [PubMed]
  8. Ron, D.; Walter, P. Signal integration in the endoplasmic reticulum unfolded protein response. Nat. Rev. Mol. Cell Biol. 2007, 8, 519–529. [Google Scholar] [CrossRef]
  9. Wang, M.; Kaufman, R.J. The impact of the endoplasmic reticulum protein-folding environment on cancer development. Nat. Rev. Cancer 2014, 14, 581–597. [Google Scholar] [CrossRef] [PubMed]
  10. Bertolotti, A.; Zhang, Y.; Hendershot, L.M.; Harding, H.; Ron, D. Dynamic interaction of BiP and ER stress transducers in the unfolded-protein response. Nat. Cell Biol. 2000, 2, 326–332. [Google Scholar] [CrossRef]
  11. Liu, C.Y.; Schröder, M.; Kaufman, R.J. Ligand-independent Dimerization Activates the Stress Response Kinases IRE1 and PERK in the Lumen of the Endoplasmic Reticulum. J. Biol. Chem. 2000, 275, 24881–24885. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Han, D.; Lerner, A.G.; Walle, L.V.; Upton, J.-P.; Xu, W.; Hagen, A.; Backes, B.J.; Oakes, S.A.; Papa, F.R. IRE1α Kinase Activation Modes Control Alternate Endoribonuclease Outputs to Determine Divergent Cell Fates. Cell 2009, 138, 562–575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Yoshida, H.; Matsui, T.; Yamamoto, A.; Okada, T.; Mori, K. XBP1 mRNA Is Induced by ATF6 and Spliced by IRE1 in Response to ER Stress to Produce a Highly Active Transcription Factor. Cell 2001, 107, 881–891. [Google Scholar] [CrossRef] [Green Version]
  14. Yoshida, H.; Matsui, T.; Hosokawa, N.; Kaufman, R.J.; Nagata, K.; Mori, K. A Time-Dependent Phase Shift in the Mammalian Unfolded Protein Response. Dev. Cell 2003, 4, 265–271. [Google Scholar] [CrossRef] [Green Version]
  15. Wakabayashi, S.; Yoshida, H. The essential biology of the endoplasmic reticulum stress response for structural and computational biologists. Comput. Struct. Biotechnol. J. 2013, 6, e201303010. [Google Scholar] [CrossRef] [Green Version]
  16. Urano, F.; Wang, X.; Bertolotti, A.; Zhang, Y.; Chung, P.; Harding, H.P.; Ron, D. Coupling of Stress in the ER to Activation of JNK Protein Kinases by Transmembrane Protein Kinase IRE. Science 2000, 287, 664–666. [Google Scholar] [CrossRef] [Green Version]
  17. Hollien, J.; Weissman, J.S. Decay of Endoplasmic Reticulum-Localized mRNAs During the Unfolded Protein Response. Science 2006, 313, 104–107. [Google Scholar] [CrossRef] [Green Version]
  18. Tam, A.B.; Koong, A.; Niwa, M. Ire1 Has Distinct Catalytic Mechanisms for XBP1/HAC1 Splicing and RIDD. Cell Rep. 2014, 9, 850–858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Wek, R.C.; Cavener, D.R. Translational control and the unfolded protein response. Antioxid. Redox Signal. 2007, 9, 2357–2372. [Google Scholar] [CrossRef] [PubMed]
  20. Vattem, K.M.; Wek, R.C. Reinitiation involving upstream ORFs regulates ATF4 mRNA translation in mammalian cells. Proc. Natl. Acad. Sci. USA 2004, 101, 11269–11274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Ma, Y.; Brewer, J.W.; Diehl, J.A.; Hendershot, L.M. Two Distinct Stress Signaling Pathways Converge Upon the CHOP Promoter During the Mammalian Unfolded Protein Response. J. Mol. Biol. 2002, 318, 1351–1365. [Google Scholar] [CrossRef]
  22. McCullough, K.D.; Martindale, J.L.; Klotz, L.-O.; Aw, T.-Y.; Holbrook, N.J. Gadd153 Sensitizes Cells to Endoplasmic Reticulum Stress by Down-Regulating Bcl2 and Perturbing the Cellular Redox State. Mol. Cell. Biol. 2001, 21, 1249–1259. [Google Scholar] [CrossRef] [Green Version]
  23. Marciniak, S.; Yun, C.Y.; Oyadomari, S.; Novoa, I.; Zhang, Y.; Jungreis, R.; Nagata, K.; Harding, H.; Ron, D. CHOP induces death by promoting protein synthesis and oxidation in the stressed endoplasmic reticulum. Genes Dev. 2004, 18, 3066–3077. [Google Scholar] [CrossRef] [Green Version]
  24. Hong, M.; Li, M.; Mao, C.; Lee, A.S. Endoplasmic reticulum stress triggers an acute proteasome-dependent degradation of ATF. J. Cell. Biochem. 2004, 92, 723–732. [Google Scholar] [CrossRef] [PubMed]
  25. Shen, J.; Prywes, R. Dependence of Site-2 Protease Cleavage of ATF6 on Prior Site-1 Protease Digestion Is Determined by the Size of the Luminal Domain of ATF. J. Biol. Chem. 2004, 279, 43046–43051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Lee, K.; Tirasophon, W.; Shen, X.; Michalak, M.; Prywes, R.; Okada, T.; Yoshida, H.; Mori, K.; Kaufman, R.J. IRE1-mediated unconventional mRNA splicing and S2P-mediated ATF6 cleavage merge to regulate XBP1 in signaling the unfolded protein response. Genes Dev. 2002, 16, 452–466. [Google Scholar] [CrossRef] [Green Version]
  27. Okada, T.; Yoshida, H.; Akazawa, R.; Negishi, M.; Mori, K. Distinct roles of activating transcription factor 6 (ATF6) and double-stranded RNA-activated protein kinase-like endoplasmic reticulum kinase (PERK) in transcription during the mammalian unfolded protein response. Biochem. J. 2002, 366, 585–594. [Google Scholar] [CrossRef]
  28. Yamamoto, K.; Sato, T.; Matsui, T.; Sato, M.; Okada, T.; Yoshida, H.; Harada, A.; Mori, K. Transcriptional Induction of Mammalian ER Quality Control Proteins Is Mediated by Single or Combined Action of ATF6α and XBP. Dev. Cell 2007, 13, 365–376. [Google Scholar] [CrossRef] [Green Version]
  29. Galindo, I.; Hernaez, B.; Muñoz-Moreno, R.; Cuesta-Geijo, M.A.; Dalmau-Mena, I.; Alonso, C. The ATF6 branch of unfolded protein response and apoptosis are activated to promote African swine fever virus infection. Cell Death Dis. 2012, 3, e341. [Google Scholar] [CrossRef] [Green Version]
  30. Lee, A.-H.; Iwakoshi, N.N.; Glimcher, L.H. XBP-1 Regulates a Subset of Endoplasmic Reticulum Resident Chaperone Genes in the Unfolded Protein Response. Mol. Cell. Biol. 2003, 23, 7448–7459. [Google Scholar] [CrossRef] [Green Version]
  31. Moore, P.C.; Qi, J.Y.; Thamsen, M.; Ghosh, R.; Peng, J.; Gliedt, M.J.; Meza-Acevedo, R.; Warren, R.E.; Hiniker, A.; Kim, G.E.; et al. Parallel Signaling through IRE1α and PERK Regulates Pancreatic Neuroendocrine Tumor Growth and Survival. Cancer Res. 2019, 79, 6190–6203. [Google Scholar] [CrossRef] [Green Version]
  32. Duwaerts, C.C.; Siao, K.; Soon, R.K.; Her, C.; Iwawaki, T.; Kohno, K.; Mattis, A.N.; Maher, J.J. Hepatocyte-specific deletion of XBP1 sensitizes mice to liver injury through hyperactivation of IRE1α. Cell Death Differ. 2021, 28, 1455–1465. [Google Scholar] [CrossRef]
  33. Meusser, B.; Hirsch, C.R.; Jarosch, E.; Sommer, T. ERAD: The long road to destruction. Nat. Cell Biol. 2005, 7, 766–772. [Google Scholar] [CrossRef] [PubMed]
  34. Vembar, S.S.; Brodsky, J.L. One step at a time: Endoplasmic reticulum-associated degradation. Nat. Rev. Mol. Cell Biol. 2008, 9, 944–957. [Google Scholar] [CrossRef] [PubMed]
  35. Kelly, S.M.; Vanslyke, J.K.; Musil, L.S. Regulation of Ubiquitin-Proteasome System–mediated Degradation by Cytosolic Stress. Mol. Biol. Cell 2007, 18, 4279–4291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Sun, S.; Lourie, R.; Cohen, S.B.; Ji, Y.; Goodrich, J.K.; Poole, A.C.; Ley, R.E.; Denkers, E.Y.; McGuckin, M.A.; Long, Q.; et al. Epithelial Sel1L is required for the maintenance of intestinal homeostasis. Mol. Biol. Cell 2016, 27, 483–490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Hu, Y.; Gao, Y.; Zhang, M.; Deng, K.-Y.; Singh, R.; Tian, Q.; Gong, Y.; Pan, Z.; Liu, Q.; Boisclair, Y.R.; et al. Endoplasmic Reticulum–Associated Degradation (ERAD) Has a Critical Role in Supporting Glucose-Stimulated Insulin Secretion in Pancreatic β-Cells. Diabetes 2019, 68, 733–746. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Fujita, H.; Yagishita, N.; Aratani, S.; Saito-Fujita, T.; Morota, S.; Yamano, Y.; Hansson, M.; Inazu, M.; Kokuba, H.; Sudo, K.; et al. The E3 ligase synoviolin controls body weight and mitochondrial biogenesis through negative regulation of PGC -1β. EMBO J. 2015, 34, 1042–1055. [Google Scholar] [CrossRef] [Green Version]
  39. Anderson, D.J.; Le Moigne, R.; Djakovic, S.; Kumar, B.; Rice, J.; Wong, S.; Wang, J.; Yao, B.; Valle, E.; Von Soly, S.K.; et al. Targeting the AAA ATPase p97 as an Approach to Treat Cancer through Disruption of Protein Homeostasis. Cancer Cell 2015, 28, 653–665. [Google Scholar] [CrossRef] [Green Version]
  40. Le Moigne, R.; Aftab, B.T.; Djakovic, S.; Dhimolea, E.; Valle, E.; Murnane, M.; King, E.M.; Soriano, F.; Menon, M.-K.; Wu, Z.Y.; et al. The p97 Inhibitor CB-5083 Is a Unique Disrupter of Protein Homeostasis in Models of Multiple Myeloma. Mol. Cancer Ther. 2017, 16, 2375–2386. [Google Scholar] [CrossRef] [Green Version]
  41. Cui, H.; Deng, M.; Zhang, Y.; Yin, F.; Liu, J. Geniposide Increases Unfolded Protein Response-Mediating HRD1 Expression to Accelerate APP Degradation in Primary Cortical Neurons. Neurochem. Res. 2018, 43, 669–680. [Google Scholar] [CrossRef] [PubMed]
  42. Kaneko, M.; Koike, H.; Saito, R.; Kitamura, Y.; Okuma, Y.; Nomura, Y. Loss of HRD1-Mediated Protein Degradation Causes Amyloid Precursor Protein Accumulation and Amyloid- Generation. J. Neurosci. 2010, 30, 3924–3932. [Google Scholar] [CrossRef]
  43. Yan, M.-M.; Ni, J.-D.; Song, D.; Ding, M.; Huang, J. Interplay between unfolded protein response and autophagy promotes tumor drug resistance. Oncol. Lett. 2015, 10, 1959–1969. [Google Scholar] [CrossRef] [Green Version]
  44. Cybulsky, A.V. Endoplasmic reticulum stress, the unfolded protein response and autophagy in kidney diseases. Nat. Rev. Nephrol. 2017, 13, 681–696. [Google Scholar] [CrossRef]
  45. Senft, D.; Ronai, Z.A. UPR, autophagy, and mitochondria crosstalk underlies the ER stress response. Trends Biochem. Sci. 2015, 40, 141–148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Mizushima, N. Autophagy: Process and function. Genes Dev. 2007, 21, 2861–2873. [Google Scholar] [CrossRef] [Green Version]
  47. Hoyerhansen, M.; Jaattela, M. Connecting endoplasmic reticulum stress to autophagy by unfolded protein response and calcium. Cell Death Differ. 2007, 14, 1576–1582. [Google Scholar] [CrossRef] [PubMed]
  48. Margariti, A.; Li, H.; Chen, T.; Martin, D.; Vizcay-Barrena, G.; Alam, S.; Karamariti, E.; Xiao, Q.; Zampetaki, A.; Zhang, Z.; et al. XBP1 mRNA Splicing Triggers an Autophagic Response in Endothelial Cells through BECLIN-1 Transcriptional Activation. J. Biol. Chem. 2013, 288, 859–872. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Ogata, M.; Hino, S.-I.; Saito, A.; Morikawa, K.; Kondo, S.; Kanemoto, S.; Murakami, T.; Taniguchi, M.; Tanii, I.; Yoshinaga, K.; et al. Autophagy Is Activated for Cell Survival after Endoplasmic ReticulumStress. Mol. Cell. Biol. 2006, 26, 9220–9231. [Google Scholar] [CrossRef] [Green Version]
  50. B’Chir, W.; Maurin, A.-C.; Carraro, V.; Averous, J.; Jousse, C.; Muranishi, Y.; Parry, L.; Stepien, G.; Fafournoux, P.; Bruhat, A. The eIF2α/ATF4 pathway is essential for stress-induced autophagy gene expression. Nucleic Acids Res. 2013, 41, 7683–7699. [Google Scholar] [CrossRef] [Green Version]
  51. Fang, L.; Zhou, Y.; Cao, H.; Wen, P.; Jiang, L.; He, W.; Dai, C.; Yang, J. Autophagy Attenuates Diabetic Glomerular Damage through Protection of Hyperglycemia-Induced Podocyte Injury. PLoS ONE 2013, 8, e60546. [Google Scholar] [CrossRef]
  52. Qi, W.; Mu, J.; Luo, Z.-F.; Zeng, W.; Guo, Y.-H.; Pang, Q.; Ye, Z.-L.; Liu, L.; Yuan, F.-H.; Feng, B. Attenuation of diabetic nephropathy in diabetes rats induced by streptozotocin by regulating the endoplasmic reticulum stress inflammatory response. Metabolism 2011, 60, 594–603. [Google Scholar] [CrossRef]
  53. Lee, T.-H.; Yeh, C.-F.; Lee, Y.-T.; Shih, Y.-C.; Chen, Y.-T.; Hung, C.-T.; You, M.-Y.; Wu, P.-C.; Shentu, T.-P.; Huang, R.-T.; et al. Fibroblast-enriched endoplasmic reticulum protein TXNDC5 promotes pulmonary fibrosis by augmenting TGFβ signaling through TGFBR1 stabilization. Nat. Commun. 2020, 11, 4254. [Google Scholar] [CrossRef] [PubMed]
  54. Nogee, L.M.; Dunbar, A.E.; Wert, S.E.; Askin, F.; Hamvas, A.; Whitsett, J.A. A Mutation in the Surfactant Protein C Gene Associated with Familial Interstitial Lung Disease. N. Engl. J. Med. 2001, 344, 573–579. [Google Scholar] [CrossRef] [PubMed]
  55. Mulugeta, S.; Nguyen, V.; Russo, S.J.; Muniswamy, M.; Beers, M.F. A Surfactant Protein C Precursor Protein BRICHOS Domain Mutation Causes Endoplasmic Reticulum Stress, Proteasome Dysfunction, and Caspase 3 Activation. Am. J. Respir. Cell Mol. Biol. 2005, 32, 521–530. [Google Scholar] [CrossRef]
  56. Lawson, W.E.; Crossno, P.F.; Polosukhin, V.V.; Roldan, J.; Cheng, D.-S.; Lane, K.B.; Blackwell, T.R.; Xu, C.; Markin, C.; Ware, L.B.; et al. Endoplasmic reticulum stress in alveolar epithelial cells is prominent in IPF: Association with altered surfactant protein processing and herpesvirus infection. Am. J. Physiol. Lung Cell. Mol. Physiol. 2008, 294, L1119–L1126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Ayala, P.; Montenegro, J.; Vivar, R.; Letelier, A.; Urroz, P.A.; Copaja, M.; Pivet, D.; Humeres, C.; Troncoso, R.; Vicencio, J.M.; et al. Attenuation of endoplasmic reticulum stress using the chemical chaperone 4-phenylbutyric acid prevents cardiac fibrosis induced by isoproterenol. Exp. Mol. Pathol. 2012, 92, 97–104. [Google Scholar] [CrossRef]
  58. Kassan, M.; Galán, M.; Partyka, M.; Saifudeen, Z.; Henrion, D.; Trebak, M.; Matrougui, K. Endoplasmic Reticulum Stress Is Involved in Cardiac Damage and Vascular Endothelial Dysfunction in Hypertensive Mice. Arter. Thromb. Vasc. Biol. 2012, 32, 1652–1661. [Google Scholar] [CrossRef] [Green Version]
  59. Nangaku, M. Mechanisms of Tubulointerstitial Injury in the Kidney: Final Common Pathways to End-stage Renal Failure. Intern. Med. 2004, 43, 9–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Doyle, J.; Forni, L.G. Acute kidney injury: Short-term and long-term effects. Crit. Care 2016, 20, 1–7. [Google Scholar] [CrossRef] [Green Version]
  61. Coca, S.; Yusuf, B.; Shlipak, M.G.; Garg, A.X.; Parikh, C.R. Long-term Risk of Mortality and Other Adverse Outcomes After Acute Kidney Injury: A Systematic Review and Meta-analysis. Am. J. kidney Dis. 2009, 53, 961–973. [Google Scholar] [CrossRef] [Green Version]
  62. Inagi, R.; Kumagai, T.; Nishi, H.; Kawakami, T.; Miyata, T.; Fujita, T.; Nangaku, M. Preconditioning with Endoplasmic Reticulum Stress Ameliorates Mesangioproliferative Glomerulonephritis. J. Am. Soc. Nephrol. 2008, 19, 915–922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Wu, C.T.; Weng, T.I.; Chen, L.P.; Chiang, C.K.; Liu, S.H. Involvement of caspase-12-dependent apoptotic pathway in ionic radiocontrast urografin-induced renal tubular cell injury. Toxicol. Appl. Pharmacol. 2013, 266, 167–175. [Google Scholar] [CrossRef]
  64. Wu, C.T.; Sheu, M.-L.; Tsai, K.-S.; Chiang, C.K.; Liu, S.H. Salubrinal, an eIF2α dephosphorylation inhibitor, enhances cisplatin-induced oxidative stress and nephrotoxicity in a mouse model. Free. Radic. Biol. Med. 2011, 51, 671–680. [Google Scholar] [CrossRef] [PubMed]
  65. Chiang, C.-K.; Hsu, S.-P.; Wu, C.-T.; Huang, J.-W.; Cheng, H.-T.; Chang, Y.-W.; Hung, K.-Y.; Wu, K.-D.; Liu, S.-H. Endoplasmic Reticulum Stress Implicated in the Development of Renal Fibrosis. Mol. Med. 2011, 17, 1295–1305. [Google Scholar] [CrossRef] [PubMed]
  66. Chen, B.L.; Sheu, M.-L.; Tsai, K.-S.; Lan, K.C.; Guan, S.S.; Wu, C.T.; Chen, L.P.; Hung, K.-Y.; Huang, J.-W.; Chiang, C.-K.; et al. CCAAT-Enhancer-Binding Protein Homologous Protein Deficiency Attenuates Oxidative Stress and Renal Ischemia-Reperfusion Injury. Antioxid. Redox Signal. 2015, 23, 1233–1245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Inoue, T.; Maekawa, H.; Inagi, R. Organelle crosstalk in the kidney. Kidney Int. 2019, 95, 1318–1325. [Google Scholar] [CrossRef] [Green Version]
  68. Liu, X.-H.; Zhang, Z.-Y.; Sun, S.; Wu, X.-D. Ischemic postconditioning protects myocardium from ischemia/reperfusion injury through attenuating endoplasmic reticulum stress. Shock 2008, 30, 422–427. [Google Scholar] [CrossRef] [PubMed]
  69. Shu, S.; Zhu, J.; Liu, Z.; Tang, C.; Cai, J.; Dong, Z. Endoplasmic reticulum stress is activated in post-ischemic kidneys to promote chronic kidney disease. EBioMedicine 2018, 37, 269–280. [Google Scholar] [CrossRef] [Green Version]
  70. Wang, Y.; Tian, J.; Qiao, X.; Su, X.; Mi, Y.; Zhang, R.; Li, R. Intermedin protects against renal ischemia-reperfusion injury by inhibiting endoplasmic reticulum stress. BMC Nephrol. 2015, 16, 169. [Google Scholar] [CrossRef] [Green Version]
  71. Fan, Y.; Xiao, W.; Lee, K.; Salem, F.; Wen, J.; He, L.; Zhang, J.; Fei, Y.; Cheng, D.; Bao, H.; et al. Inhibition of Reticulon-1A–Mediated Endoplasmic Reticulum Stress in Early AKI Attenuates Renal Fibrosis Development. J. Am. Soc. Nephrol. 2017, 28, 2007–2021. [Google Scholar] [CrossRef] [Green Version]
  72. Jao, T.-M.; Nangaku, M.; Wu, C.-H.; Sugahara, M.; Saito, H.; Maekawa, H.; Ishimoto, Y.; Aoe, M.; Inoue, T.; Tanaka, T.; et al. ATF6α downregulation of PPARα promotes lipotoxicity-induced tubulointerstitial fibrosis. Kidney Int. 2019, 95, 577–589. [Google Scholar] [CrossRef] [Green Version]
  73. Kaushal, G.P.; Chandrashekar, K.; Juncos, L.A. Molecular Interactions Between Reactive Oxygen Species and Autophagy in Kidney Disease. Int. J. Mol. Sci. 2019, 20, 3791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Jun, M. Antioxidants for chronic kidney disease. Nephrology 2013, 18, 576–578. [Google Scholar] [CrossRef] [PubMed]
  75. Koritzinsky, M.; Levitin, F.; Beucken, J.V.D.; Rumantir, R.A.; Harding, N.J.; Chu, K.C.; Boutros, P.; Braakman, I.; Wouters, B. Two phases of disulfide bond formation have differing requirements for oxygen. J. Cell Biol. 2013, 203, 615–627. [Google Scholar] [CrossRef] [Green Version]
  76. Pajares, M.; Cuadrado, A.; Rojo, A.I. Modulation of proteostasis by transcription factor NRF2 and impact in neurodegenerative diseases. Redox Biol. 2017, 11, 543–553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Chen, J.; Zhang, Z.; Cai, L. Diabetic Cardiomyopathy and Its Prevention by Nrf2: Current Status. Diabetes Metab. J. 2014, 38, 337–345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Jin, J.-K.; Blackwood, E.A.; Azizi, K.; Thuerauf, D.J.; Fahem, A.G.; Hofmann, C.; Kaufman, R.J.; Doroudgar, S.; Glembotski, C.C. ATF6 Decreases Myocardial Ischemia/Reperfusion Damage and Links ER Stress and Oxidative Stress Signaling Pathways in the Heart. Circ. Res. 2017, 120, 862–875. [Google Scholar] [CrossRef] [Green Version]
  79. Cullinan, S.B.; Zhang, D.; Hannink, M.; Arvisais, E.; Kaufman, R.J.; Diehl, J.A. Nrf2 Is a Direct PERK Substrate and Effector of PERK-Dependent Cell Survival. Mol. Cell. Biol. 2003, 23, 7198–7209. [Google Scholar] [CrossRef] [Green Version]
  80. Cullinan, S.B.; Diehl, J.A. Coordination of ER and oxidative stress signaling: The PERK/Nrf2 signaling pathway. Int. J. Biochem. Cell Biol. 2006, 38, 317–332. [Google Scholar] [CrossRef]
  81. Yokouchi, M.; Hiramatsu, N.; Hayakawa, K.; Okamura, M.; Du, S.; Kasai, A.; Takano, Y.; Shitamura, A.; Shimada, T.; Yao, J.; et al. Involvement of Selective Reactive Oxygen Species Upstream of Proapoptotic Branches of Unfolded Protein Response. J. Biol. Chem. 2008, 283, 4252–4260. [Google Scholar] [CrossRef] [Green Version]
  82. Molitch, M.E.; DeFronzo, R.A.; Franz, M.J.; Keane, W.F. American Diabetes Association Nephropathy in Diabetes. Diabetes Care 2004, 27, S79–S83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Wu, J.; Zhang, R.; Torreggiani, M.; Ting, A.; Xiong, H.; Striker, G.E.; Vlassara, H.; Zheng, F. Induction of Diabetes in Aged C57B6 Mice Results in Severe Nephropathy: An Association with Oxidative Stress, Endoplasmic Reticulum Stress, and Inflammation. Am. J. Pathol. 2010, 176, 2163–2176. [Google Scholar] [CrossRef]
  84. Mezzano, S.A.; Ruiz-Ortega, M.; Egido, J. Angiotensin II and Renal Fibrosis. Hypertension 2001, 38, 635–638. [Google Scholar] [CrossRef] [Green Version]
  85. Sun, H.; Sun, L.; Li, Y.; Shao, M.; Cheng, X.; Ge, N.; Lu, J.; Li, S. ACE-inhibitor Suppresses the Apoptosis Induced by Endoplasmic Reticulum Stress in Renal Tubular in Experimental Diabetic Rats. Exp. Clin. Endocrinol. Diabetes 2009, 117, 336–344. [Google Scholar] [CrossRef]
  86. Lindenmeyer, M.T.; Rastaldi, M.P.; Ikehata, M.; Neusser, M.A.; Kretzler, M.; Cohen, C.D.; Schlöndorff, D. Proteinuria and Hyperglycemia Induce Endoplasmic Reticulum Stress. J. Am. Soc. Nephrol. 2008, 19, 2225–2236. [Google Scholar] [CrossRef] [Green Version]
  87. Madhusudhan, T.; Wang, H.; Dong, W.; Ghosh, S.; Bock, F.; Thangapandi, V.R.; Ranjan, S.; Wolter, J.; Kohli, S.; Shahzad, K.; et al. Defective podocyte insulin signalling through p85-XBP1 promotes ATF6-dependent maladaptive ER-stress response in diabetic nephropathy. Nat. Commun. 2015, 6, 6496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Navarro-Betancourt, J.R.; Papillon, J.; Guillemette, J.; Iwawaki, T.; Chung, C.-F.; Cybulsky, A.V. Role of IRE1α in podocyte proteostasis and mitochondrial health. Cell Death Discov. 2020, 6, 1–15. [Google Scholar] [CrossRef] [PubMed]
  89. Hassan, H.; Tian, X.; Inoue, K.; Chai, N.; Liu, C.; Soda, K.; Moeckel, G.; Tufro, A.; Lee, A.-H.; Somlo, S.; et al. Essential Role of X-Box Binding Protein-1 during Endoplasmic Reticulum Stress in Podocytes. J. Am. Soc. Nephrol. 2015, 27, 1055–1065. [Google Scholar] [CrossRef] [PubMed]
  90. Ishikawa, Y.; Fedeles, S.; Marlier, A.; Zhang, C.; Gallagher, A.-R.; Lee, A.-H.; Somlo, S. Spliced XBP1 Rescues Renal Interstitial Inflammation Due to Loss of Sec63 in Collecting Ducts. J. Am. Soc. Nephrol. 2019, 30, 443–459. [Google Scholar] [CrossRef] [Green Version]
  91. Ozbek, E. Induction of Oxidative Stress in Kidney. Int. J. Nephrol. 2012, 2012, 465897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Liu, Y.; Adachi, M.; Zhao, S.; Hareyama, M.; Koong, A.; Luo, D.; Rando, T.; Imai, K.; Shinomura, Y. Preventing oxidative stress: A new role for XBP. Cell Death Differ. 2009, 16, 847–857. [Google Scholar] [CrossRef]
  93. Martin, D.; Li, Y.; Yang, J.; Wang, G.; Margariti, A.; Jiang, Z.; Yu, H.; Zampetaki, A.; Hu, Y.; Xu, Q.; et al. Unspliced X-box-binding Protein 1 (XBP1) Protects Endothelial Cells from Oxidative Stress through Interaction with Histone Deacetylase. J. Biol. Chem. 2014, 289, 30625–30634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Fu, H.; Feng, J.; Liu, Q.; Sun, F.; Tie, Y.; Zhu, J.; Xing, R.; Sun, Z.; Zheng, X. Stress induces tRNA cleavage by angiogenin in mammalian cells. FEBS Lett. 2008, 583, 437–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Mami, I.; Bouvier, N.; El Karoui, K.; Gallazzini, M.; Rabant, M.; Laurent-Puig, P.; Li, S.; Tharaux, P.-L.; Beaune, P.; Thervet, E.; et al. Angiogenin Mediates Cell-Autonomous Translational Control under Endoplasmic Reticulum Stress and Attenuates Kidney Injury. J. Am. Soc. Nephrol. 2015, 27, 863–876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Kurts, C.; Panzer, U.; Anders, H.-J.; Rees, A.J. The immune system and kidney disease: Basic concepts and clinical implications. Nat. Rev. Immunol. 2013, 13, 738–753. [Google Scholar] [CrossRef]
  97. Tampe, D.; Zeisberg, M. Potential approaches to reverse or repair renal fibrosis. Nat. Rev. Nephrol. 2014, 10, 226–237. [Google Scholar] [CrossRef] [PubMed]
  98. Vesey, D.; Cheung, C.W.; Cuttle, L.; Endre, Z.; Gobé, G.; Johnson, D. Interleukin-1β induces human proximal tubule cell injury, α-smooth muscle actin expression and fibronectin production. Kidney Int. 2002, 62, 31–40. [Google Scholar] [CrossRef]
  99. Turner, C.M.; Arulkumaran, N.; Singer, M.; Unwin, R.J.; Tam, F.W. Is the inflammasome a potential therapeutic target in renal disease? BMC Nephrol. 2014, 15, 21. [Google Scholar] [CrossRef] [Green Version]
  100. Martinon, F.; Mayor, A.; Tschopp, J. The Inflammasomes: Guardians of the Body. Annu. Rev. Immunol. 2009, 27, 229–265. [Google Scholar] [CrossRef] [Green Version]
  101. Vilaysane, A.; Chun, J.; Seamone, M.E.; Wang, W.; Chin, R.; Hirota, S.; Li, Y.; Clark, S.A.; Tschopp, J.; Trpkov, K.; et al. The NLRP3 Inflammasome Promotes Renal Inflammation and Contributes to CKD. J. Am. Soc. Nephrol. 2010, 21, 1732–1744. [Google Scholar] [CrossRef] [Green Version]
  102. Papandreou, I.; Denko, N.C.; Olson, M.; Van Melckebeke, H.; Lust, S.; Tam, A.; Solow-Cordero, D.E.; Bouley, D.M.; Offner, F.; Niwa, M.; et al. Identification of an Ire1alpha endonuclease specific inhibitor with cytotoxic activity against human multiple myeloma. Blood 2011, 117, 1311–1314. [Google Scholar] [CrossRef] [Green Version]
  103. Ghosh, R.; Wang, L.; Wang, E.S.; Perera, B.G.K.; Igbaria, A.; Morita, S.; Prado, K.; Thamsen, M.; Caswell, D.; Macias, H.; et al. Allosteric Inhibition of the IRE1α RNase Preserves Cell Viability and Function during Endoplasmic Reticulum Stress. Cell 2014, 158, 534–548. [Google Scholar] [CrossRef] [Green Version]
  104. De Vries, J.H.; Hollman, P.C.; Meyboom, S.; Buysman, M.N.; Zock, P.L.; van Staveren, W.A.; Katan, M.B. Plasma concentra-tions and urinary excretion of the antioxidant flavonols quercetin and kaempferol as biomarkers for dietary intake. Am. J. Clin. Nutr. 1998, 68, 60–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Wiseman, L.; Zhang, Y.; Lee, K.P.; Harding, H.; Haynes, C.M.; Price, J.; Sicheri, F.; Ron, D. Flavonol Activation Defines an Unanticipated Ligand-Binding Site in the Kinase-RNase Domain of IRE. Mol. Cell 2010, 38, 291–304. [Google Scholar] [CrossRef] [Green Version]
  106. Liu, C.-M.; Zheng, G.H.; Ming, Q.L.; Sun, J.M.; Cheng, C. Protective effect of quercetin on lead-induced oxidative stress and endoplasmic reticulum stress in rat liver via the IRE1/JNK and PI3K/Akt pathway. Free. Radic. Res. 2013, 47, 192–201. [Google Scholar] [CrossRef]
  107. Anjaneyulu, M.; Chopra, K. Quercetin, an anti-oxidant bioflavonoid, attenuates diabetic nephropathy in rats. Clin. Exp. Pharmacol. Physiol. 2004, 31, 244–248. [Google Scholar] [CrossRef] [PubMed]
  108. Morales, A.; Vicente-Sánchez, C.; Sandoval, J.S.; Egido, J.; Mayoral, P.; Arévalo, M.; Fernández-Tagarro, M.; Lopez-Novoa, J.M.; Pérez-Barriocanal, F. Protective effect of quercetin on experimental chronic cadmium nephrotoxicity in rats is based on its antioxidant properties. Food Chem. Toxicol. 2006, 44, 2092–2100. [Google Scholar] [CrossRef] [PubMed]
  109. Jones, E.A.; Shahed, A.; Shoskes, D.A. Modulation of apoptotic and inflammatory genes by bioflavonoids and angiotensin II inhibition in ureteral obstruction. Urololgy 2000, 56, 346–351. [Google Scholar] [CrossRef]
  110. Hu, Q.; Noor, M.; Wong, Y.F.; Hylands, P.J.; Simmonds, M.S.J.; Xu, Q.; Jiang, D.; Hendry, B.M.; Xu, Q. In vitro anti-fibrotic activities of herbal compounds and herbs. Nephrol. Dial. Transplant. 2009, 24, 3033–3041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Yan, Y.; Feng, Y.; Li, W.; Che, J.-P.; Wang, G.-C.; Liu, M.; Zheng, J.-H. Protective effects of quercetin and hyperoside on renal fibrosis in rats with unilateral ureteral obstruction. Exp. Ther. Med. 2014, 8, 727–730. [Google Scholar] [CrossRef] [Green Version]
  112. Guo, W.; Ding, J.; Zhang, A.; Dai, W.; Liu, S.; Diao, Z.; Wang, L.; Han, X.; Liu, W. The Inhibitory Effect of Quercetin on Asymmetric Dimethylarginine-Induced Apoptosis Is Mediated by the Endoplasmic Reticulum Stress Pathway in Glomerular Endothelial Cells. Int. J. Mol. Sci. 2014, 15, 484–503. [Google Scholar] [CrossRef] [Green Version]
  113. Kaser, A.; Lee, A.-H.; Franke, A.; Glickman, J.N.; Zeissig, S.; Tilg, H.; Nieuwenhuis, E.E.; Higgins, D.E.; Schreiber, S.; Glimcher, L.H.; et al. XBP1 Links ER Stress to Intestinal Inflammation and Confers Genetic Risk for Human Inflammatory Bowel Disease. Cell 2008, 134, 743–756. [Google Scholar] [CrossRef] [Green Version]
  114. Wang, W.; Zhang, Q.W.; Ye, W.C.; Wang, Y.T. Isoquinoline alkaloids from the rhizoma of Coptis chinensis. J. Nat. Med. 2007, 5, 348. [Google Scholar]
  115. Zhang, H.; Zhang, Z.; Song, G.; Tang, X.; Song, H.; Deng, A.; Wang, W.; Wu, L.; Qin, H. Development of an XBP1 agonist, HLJ2, as a potential therapeutic agent for ulcerative colitis. Eur. J. Pharm. Sci. 2017, 109, 56–64. [Google Scholar] [CrossRef]
  116. Zhang, H.; Song, G.; Zhang, Z.; Song, H.; Tang, X.; Deng, A.; Wang, W.; Wu, L.; Qin, H. Colitis Is Effectively Ameliorated by (±)-8-Acetonyl-dihydrocoptisine via the XBP1-NF-κB Pathway. Front. Pharmacol. 2017, 8, 619. [Google Scholar] [CrossRef] [PubMed]
  117. Long, K.; Boyce, M.; Lin, H.; Yuan, J.; Ma, D. Structure–activity relationship studies of salubrinal lead to its active biotinylated derivative. Bioorganic Med. Chem. Lett. 2005, 15, 3849–3852. [Google Scholar] [CrossRef] [PubMed]
  118. Jacquillet, G.; Barbier, O.; Cougnon, M.; Tauc, M.; Namorado, M.C.; Martin, D.; Reyes, J.L.; Poujeol, P. Zinc protects renal function during cadmium intoxication in the rat. Am. J. Physiol. Ren. Physiol. 2006, 290, F127–F137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Thijssen, S.; Cuypers, A.; Maringwa, J.; Smeets, K.; Horemans, N.; Lambrichts, I.; Van Kerkhove, E. Low cadmium exposure triggers a biphasic oxidative stress response in mice kidneys. Toxicology 2007, 236, 29–41. [Google Scholar] [CrossRef] [PubMed]
  120. Komoike, Y.; Inamura, H.; Matsuoka, M. Effects of salubrinal on cadmium-induced apoptosis in HK-2 human renal proximal tubular cells. Arch. Toxicol. 2011, 86, 37–44. [Google Scholar] [CrossRef]
  121. Matsuoka, M.; Komoike, Y. Experimental Evidence Shows Salubrinal, an eIF2α Dephosphorylation Inhibitor, Reduces Xenotoxicant-Induced Cellular Damage. Int. J. Mol. Sci. 2015, 16, 16275–16287. [Google Scholar] [CrossRef]
  122. Jeon, B.J.; Yang, H.M.; Lyu, Y.S.; Pae, H.O.; Ju, S.M.; Jeon, B.H. Apigenin inhibits indoxyl sulfate-induced endoplasmic retic-ulum stress and anti-proliferative pathways, CHOP and IL-6/p21, in human renal proximal tubular cells. Eur. Rev. Med. Pharmacol. Sci. 2015, 19, 2303–2310. [Google Scholar]
  123. Tsaytler, P.; Harding, H.P.; Ron, D.; Bertolotti, A. Selective Inhibition of a Regulatory Subunit of Protein Phosphatase 1 Restores Proteostasis. Science 2011, 332, 91–94. [Google Scholar] [CrossRef] [PubMed]
  124. Atkins, C.; Liu, Q.; Minthorn, E.; Zhang, S.-Y.; Figueroa, D.J.; Moss, K.; Stanley, T.B.; Sanders, B.; Goetz, A.; Gaul, N.; et al. Characterization of a Novel PERK Kinase Inhibitor with Antitumor and Antiangiogenic Activity. Cancer Res. 2013, 73, 1993–2002. [Google Scholar] [CrossRef] [Green Version]
  125. Zhang, M.; Guo, Y.; Fu, H.; Hu, S.; Pan, J.; Wang, Y.; Cheng, J.; Song, J.; Yu, Q.; Zhang, S.; 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, e1847. [Google Scholar] [CrossRef] [Green Version]
  126. DeZwaan-McCabe, D.; Riordan, J.; Arensdorf, A.M.; Icardi, M.; Dupuy, A.; Rutkowski, D.T. The Stress-Regulated Transcription Factor CHOP Promotes Hepatic Inflammatory Gene Expression, Fibrosis, and Oncogenesis. PLoS Genet. 2013, 9, e1003937. [Google Scholar] [CrossRef] [Green Version]
  127. Huber, A.-L.; Lebeau, J.; Guillaumot, P.; Petrilli, V.; Malek, M.; Chilloux, J.; Fauvet, F.; Payen, L.; Kfoury, A.; Renno, T.; et al. p58IPK-Mediated Attenuation of the Proapoptotic PERK-CHOP Pathway Allows Malignant Progression upon Low Glucose. Mol. Cell 2013, 49, 1049–1059. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Esposito, V.; Grosjean, F.; Tan, J.; Huang, L.; Zhu, L.; Chen, J.; Xiong, H.; Striker, G.E.; Zheng, F. CHOP deficiency results in elevated lipopolysaccharide-induced inflammation and kidney injury. Am. J. Physiol. Ren. Physiol. 2013, 304, F440–F450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Bush, K.T.; George, S.K.; Zhang, P.L.; Nigam, S.K. Pretreatment with inducers of ER molecular chaperones protects epithelial cells subjected to ATP depletion. Am. J. Physiol. Content 1999, 277, F211–F218. [Google Scholar] [CrossRef]
  130. Prachasilchai, W.; Sonoda, H.; Yokota-Ikeda, N.; Oshikawa, S.; Aikawa, C.; Uchida, K.; Ito, K.; Kudo, T.; Imaizumi, K.; Ikeda, M. A protective role of unfolded protein response in mouse ischemic acute kidney injury. Eur. J. Pharmacol. 2008, 592, 138–145. [Google Scholar] [CrossRef] [PubMed]
  131. Zhang, P.L.; Lun, M.; Teng, J.; Huang, J.; Blasick, T.M.; Yin, L.; Herrera, G.A.; Cheung, J.Y. Preinduced molecular chaperones in the endoplasmic reticulum protect cardiomyocytes from lethal injury. Ann. Clin. Lab. Sci. 2004, 34, 449–457. [Google Scholar] [PubMed]
  132. Hung, C.-C.; Ichimura, T.; Stevens, J.L.; Bonventre, J.V. Protection of Renal Epithelial Cells against Oxidative Injury by Endoplasmic Reticulum Stress Preconditioning Is Mediated by ERK1/2 Activation. J. Biol. Chem. 2003, 278, 29317–29326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Steinberg, G.R.; Kemp, B. AMPK in Health and Disease. Physiol. Rev. 2009, 89, 1025–1078. [Google Scholar] [CrossRef] [PubMed]
  134. Lee, E.K.; Jeong, J.U.; Chang, J.W.; Yang, W.S.; Kim, S.B.; Kil Park, S.; Park, J.S.; Lee, S.K. Activation of AMP-Activated Protein Kinase Inhibits Albumin-Induced Endoplasmic Reticulum Stress and Apoptosis through Inhibition of Reactive Oxygen Species. Nephron Exp. Nephrol. 2012, 121, e38–e48. [Google Scholar] [CrossRef] [PubMed]
  135. Kim, H.; Moon, S.Y.; Kim, J.-S.; Baek, C.H.; Kim, M.; Min, J.Y.; Lee, S.K. Activation of AMP-activated protein kinase inhibits ER stress and renal fibrosis. Am. J. Physiol. Ren. Physiol. 2015, 308, F226–F236. [Google Scholar] [CrossRef]
  136. Mimori, S.; Okuma, Y.; Kaneko, M.; Kawada, K.; Hosoi, T.; Ozawa, K.; Nomura, Y.; Hamana, H. Protective Effects of 4-Phenylbutyrate Derivatives on the Neuronal Cell Death and Endoplasmic Reticulum Stress. Biol. Pharm. Bull. 2012, 35, 84–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Carducci, M.A.; Nelson, J.B.; Chan-Tack, K.M.; Ayyagari, S.R.; Sweatt, W.H.; Campbell, P.; Nelson, W.G.; Simons, J.W. Phenylbutyrate induces apoptosis in human prostate cancer and is more potent than phenylacetate. Clin. Cancer Res. 1996, 2, 379–387. [Google Scholar]
  138. Dyer, E.S.; Paulsen, M.T.; Markwart, S.M.; Goh, M.; Livant, D.L.; Ljungman, M. Phenylbutyrate inhibits the invasive properties of prostate and breast cancer cell lines in the sea urchin embryo basement membrane invasion assay. Int. J. Cancer 2002, 101, 496–499. [Google Scholar] [CrossRef] [Green Version]
  139. Cubillos-Ruiz, J.R.; Mohamed, E.; Rodriguez, P.C. Unfolding anti-tumor immunity: ER stress responses sculpt tolerogenic myeloid cells in cancer. J. Immunother. Cancer 2017, 5, 5. [Google Scholar] [CrossRef] [Green Version]
  140. Özcan, U.; Yilmaz, E.; Ozcan, L.; Furuhashi, M.; Vaillancourt, E.; Smith, R.O.; Görgün, C.Z.; Hotamisligil, G.S. Chemical Chaperones Reduce ER Stress and Restore Glucose Homeostasis in a Mouse Model of Type 2 Diabetes. Science 2006, 313, 1137–1140. [Google Scholar] [CrossRef] [Green Version]
  141. Liu, S.-H.; Yang, C.-C.; Chan, D.-C.; Wu, C.-T.; Chen, L.-P.; Huang, J.-W.; Hung, K.-Y.; Chiang, C.-K. Chemical chaperon 4-phenylbutyrate protects against the endoplasmic reticulum stress-mediated renal fibrosis in vivo and in vitro. Oncotarget 2016, 7, 22116–22127. [Google Scholar] [CrossRef] [Green Version]
  142. Wang, J.; Wen, Y.; Lv, L.-L.; Liu, H.; Tang, R.-N.; Ma, K.-L.; Liu, B.-C. Involvement of endoplasmic reticulum stress in angiotensin II-induced NLRP3 inflammasome activation in human renal proximal tubular cells in vitro. Acta Pharmacol. Sin. 2015, 36, 821–830. [Google Scholar] [CrossRef] [Green Version]
  143. Moreno, J.A.; Halliday, M.; Molloy, C.; Radford, H.; Verity, N.; Axten, J.M.; Ortori, C.A.; Willis, A.E.; Fischer, P.M.; Barrett, D.A.; et al. Oral Treatment Targeting the Unfolded Protein Response Prevents Neurodegeneration and Clinical Disease in Prion-Infected Mice. Sci. Transl. Med. 2013, 5, 206ra138. [Google Scholar] [CrossRef] [PubMed]
  144. Song, H.; Tang, X.; Li, X.; Wang, Y.; Deng, A.; Wang, W.; Zhang, H.; Qin, H.; Wu, L. HLJ2 Effectively Ameliorates Colitis-Associated Cancer via Inhibition of NF-κB and Epithelial–Mesenchymal Transition. Drug Des. Dev. Ther. 2020, 14, 4291–4302. [Google Scholar] [CrossRef]
  145. Kim, I.; Xu, W.; Reed, J.C. Cell death and endoplasmic reticulum stress: Disease relevance and therapeutic opportunities. Nat. Rev. Drug Discov. 2008, 7, 1013–1030. [Google Scholar] [CrossRef] [PubMed]
  146. Maly, D.J.; Papa, F.R. Druggable sensors of the unfolded protein response. Nat. Chem. Biol. 2014, 10, 892–901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The IRE1-XBP1-related UPRs pathways played pathological and therapeutic roles in kidney disease. (a) Unfolded proteins bind to BiP and allow the activation of IRE1 to occur through its dimerization and trans-autophosphorylation. IRE1 phosphorylation as indicated p in the yellow cycle. Activated IRE1 splices the mRNA of XBP1u to form spliced XBP1 (XBP1s) mRNA, which encodes XBP1s protein then translocates to the nucleus, inducing genes that are thought to be generally cytoprotective, such as chaperones, trafficking associators, ERAD-related proteins. (b) Evidence shows that XBP1 knockout makes cells sensitive to oxidative stress, an important insult to kidney injury. Combining the XBP1 knockout model with genetically induced ER stress in the kidney leads to severe albuminuria, glomerulosclerosis, and renal failure. (c) XBP1s could directly induce angiogenin (ANG) expression in renal epithelial cells, which has been proven to protect cells and attenuate adverse effects caused by ER stress. (d) Severe ER stress triggers cell death through the IRE1-TRAF2-JNK pathway in many kinds of kidney diseases. Quercetin, a natural flavonoid, has been demonstrated to inhibit the IRE1-TRAF2-JNK pathway in unilateral ureteral obstruction (UUO), diabetic nephropathy, and cadmium-induced kidney injury. It might be a strong therapeutic option for the targeting of UPRs.
Figure 1. The IRE1-XBP1-related UPRs pathways played pathological and therapeutic roles in kidney disease. (a) Unfolded proteins bind to BiP and allow the activation of IRE1 to occur through its dimerization and trans-autophosphorylation. IRE1 phosphorylation as indicated p in the yellow cycle. Activated IRE1 splices the mRNA of XBP1u to form spliced XBP1 (XBP1s) mRNA, which encodes XBP1s protein then translocates to the nucleus, inducing genes that are thought to be generally cytoprotective, such as chaperones, trafficking associators, ERAD-related proteins. (b) Evidence shows that XBP1 knockout makes cells sensitive to oxidative stress, an important insult to kidney injury. Combining the XBP1 knockout model with genetically induced ER stress in the kidney leads to severe albuminuria, glomerulosclerosis, and renal failure. (c) XBP1s could directly induce angiogenin (ANG) expression in renal epithelial cells, which has been proven to protect cells and attenuate adverse effects caused by ER stress. (d) Severe ER stress triggers cell death through the IRE1-TRAF2-JNK pathway in many kinds of kidney diseases. Quercetin, a natural flavonoid, has been demonstrated to inhibit the IRE1-TRAF2-JNK pathway in unilateral ureteral obstruction (UUO), diabetic nephropathy, and cadmium-induced kidney injury. It might be a strong therapeutic option for the targeting of UPRs.
Ijms 22 08674 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, J.-H.; Wu, C.-H.; Chiang, C.-K. Therapeutic Approaches Targeting Proteostasis in Kidney Disease and Fibrosis. Int. J. Mol. Sci. 2021, 22, 8674. https://doi.org/10.3390/ijms22168674

AMA Style

Chen J-H, Wu C-H, Chiang C-K. Therapeutic Approaches Targeting Proteostasis in Kidney Disease and Fibrosis. International Journal of Molecular Sciences. 2021; 22(16):8674. https://doi.org/10.3390/ijms22168674

Chicago/Turabian Style

Chen, Jia-Huang, Chia-Hsien Wu, and Chih-Kang Chiang. 2021. "Therapeutic Approaches Targeting Proteostasis in Kidney Disease and Fibrosis" International Journal of Molecular Sciences 22, no. 16: 8674. https://doi.org/10.3390/ijms22168674

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop