Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2529 word(s) 2529 2021-08-16 05:41:23 |
2 format correct Meta information modification 2529 2021-09-26 03:21:53 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Wu, C.; Chen, J.H. Proteostasis in Kidney Disease&Fibrosis. Encyclopedia. Available online: https://encyclopedia.pub/entry/14534 (accessed on 19 April 2024).
Wu C, Chen JH. Proteostasis in Kidney Disease&Fibrosis. Encyclopedia. Available at: https://encyclopedia.pub/entry/14534. Accessed April 19, 2024.
Wu, Chia-Hsien, Jia Huang Chen. "Proteostasis in Kidney Disease&Fibrosis" Encyclopedia, https://encyclopedia.pub/entry/14534 (accessed April 19, 2024).
Wu, C., & Chen, J.H. (2021, September 24). Proteostasis in Kidney Disease&Fibrosis. In Encyclopedia. https://encyclopedia.pub/entry/14534
Wu, Chia-Hsien and Jia Huang Chen. "Proteostasis in Kidney Disease&Fibrosis." Encyclopedia. Web. 24 September, 2021.
Proteostasis in Kidney Disease&Fibrosis
Edit

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. 

unfolded protein responses endoplasmic reticulum stress fibrosis kidney proteostasis

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.

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 [82][83]. 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 [84] 

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 [85]. 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 [86]. 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 [87]. 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 [88]. 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 [89]. 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 [84][90].

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.
  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.
  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.
  4. Wang, S.; Kaufman, R.J. The impact of the unfolded protein response on human disease. J. Cell Biol. 2012, 197, 857–867.
  5. Inagi, R.; Ishimoto, Y.; Nangaku, M. Proteostasis in endoplasmic reticulum—New mechanisms in kidney disease. Nat. Rev. Nephrol. 2014, 10, 369–378.
  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.
  7. Hetz, C.; Chevet, E.; Harding, H. Targeting the unfolded protein response in disease. Nat. Rev. Drug Discov. 2013, 12, 703–719.
  8. Ron, D.; Walter, P. Signal integration in the endoplasmic reticulum unfolded protein response. Nat. Rev. Mol. Cell Biol. 2007, 8, 519–529.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  17. Hollien, J.; Weissman, J.S. Decay of Endoplasmic Reticulum-Localized mRNAs During the Unfolded Protein Response. Science 2006, 313, 104–107.
  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.
  19. Wek, R.C.; Cavener, D.R. Translational control and the unfolded protein response. Antioxid. Redox Signal. 2007, 9, 2357–2372.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  33. Meusser, B.; Hirsch, C.R.; Jarosch, E.; Sommer, T. ERAD: The long road to destruction. Nat. Cell Biol. 2005, 7, 766–772.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  44. Cybulsky, A.V. Endoplasmic reticulum stress, the unfolded protein response and autophagy in kidney diseases. Nat. Rev. Nephrol. 2017, 13, 681–696.
  45. Senft, D.; Ronai, Z.A. UPR, autophagy, and mitochondria crosstalk underlies the ER stress response. Trends Biochem. Sci. 2015, 40, 141–148.
  46. Mizushima, N. Autophagy: Process and function. Genes Dev. 2007, 21, 2861–2873.
  47. Hoyerhansen, M.; Jaattela, M. Connecting endoplasmic reticulum stress to autophagy by unfolded protein response and calcium. Cell Death Differ. 2007, 14, 1576–1582.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  59. Nangaku, M. Mechanisms of Tubulointerstitial Injury in the Kidney: Final Common Pathways to End-stage Renal Failure. Intern. Med. 2004, 43, 9–17.
  60. Doyle, J.; Forni, L.G. Acute kidney injury: Short-term and long-term effects. Crit. Care 2016, 20, 1–7.
  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.
  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.
  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.
  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.
  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.
  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.
  67. Inoue, T.; Maekawa, H.; Inagi, R. Organelle crosstalk in the kidney. Kidney Int. 2019, 95, 1318–1325.
  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.
  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.
  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.
  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.
  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.
  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.
  74. Jun, M. Antioxidants for chronic kidney disease. Nephrology 2013, 18, 576–578.
  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.
  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.
  77. Chen, J.; Zhang, Z.; Cai, L. Diabetic Cardiomyopathy and Its Prevention by Nrf2: Current Status. Diabetes Metab. J. 2014, 38, 337–345.
  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.
  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.
  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.
  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.
  82. 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.
  83. 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.
  84. 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.
  85. Molitch, M.E.; DeFronzo, R.A.; Franz, M.J.; Keane, W.F. American Diabetes Association Nephropathy in Diabetes. Diabetes Care 2004, 27, S79–S83.
  86. 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.
  87. Mezzano, S.A.; Ruiz-Ortega, M.; Egido, J. Angiotensin II and Renal Fibrosis. Hypertension 2001, 38, 635–638.
  88. 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.
  89. 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.
  90. 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.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : ,
View Times: 373
Revisions: 2 times (View History)
Update Date: 26 Sep 2021
1000/1000