Next Article in Journal
RNA G-Quadruplex within the 5′-UTR of FEN1 Regulates mRNA Stability under Oxidative Stress
Next Article in Special Issue
Phenotypic Modulation of Cancer-Associated Antioxidant NQO1 Activity by Post-Translational Modifications and the Natural Diversity of the Human Genome
Previous Article in Journal
Characterization of Dextran Biosynthesized by Glucansucrase from Leuconostoc pseudomesenteroides and Their Potential Biotechnological Applications
Previous Article in Special Issue
Transcriptomic Profiling and Pathway Analysis of Mesenchymal Stem Cells Following Low Dose-Rate Radiation Exposure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Stress Activated MAP Kinases and Cyclin-Dependent Kinase 5 Mediate Nuclear Translocation of Nrf2 via Hsp90α-Pin1-Dynein Motor Transport Machinery

1
School of Medicine, University of Tsukuba, Tsukuba 305-8577, Japan
2
King’s British Heart Foundation Centre of Research Excellence, School of Cardiovascular and Metabolic Medicine & Sciences, Faculty of Life Sciences & Medicine, King’s College London, 150 Stamford Street, London SE1 9NH, UK
*
Author to whom correspondence should be addressed.
Antioxidants 2023, 12(2), 274; https://doi.org/10.3390/antiox12020274
Submission received: 25 December 2022 / Revised: 19 January 2023 / Accepted: 22 January 2023 / Published: 26 January 2023

Abstract

:
Non-lethal low levels of oxidative stress leads to rapid activation of the transcription factor nuclear factor-E2-related factor 2 (Nrf2), which upregulates the expression of genes important for detoxification, glutathione synthesis, and defense against oxidative damage. Stress-activated MAP kinases p38, ERK, and JNK cooperate in the efficient nuclear accumulation of Nrf2 in a cell-type-dependent manner. Activation of p38 induces membrane trafficking of a glutathione sensor neutral sphingomyelinase 2, which generates ceramide upon depletion of cellular glutathione. We previously proposed that caveolin-1 in lipid rafts provides a signaling hub for the phosphorylation of Nrf2 by ceramide-activated PKCζ and casein kinase 2 to stabilize Nrf2 and mask a nuclear export signal. We further propose a mechanism of facilitated Nrf2 nuclear translocation by ERK and JNK. ERK and JNK phosphorylation of Nrf2 induces the association of prolyl cis/trans isomerase Pin1, which specifically recognizes phosphorylated serine or threonine immediately preceding a proline residue. Pin1-induced structural changes allow importin-α5 to associate with Nrf2. Pin1 is a co-chaperone of Hsp90α and mediates the association of the Nrf2-Pin1-Hsp90α complex with the dynein motor complex, which is involved in transporting the signaling complex to the nucleus along microtubules. In addition to ERK and JNK, cyclin-dependent kinase 5 could phosphorylate Nrf2 and mediate the transport of Nrf2 to the nucleus via the Pin1-Hsp90α system. Some other ERK target proteins, such as pyruvate kinase M2 and hypoxia-inducible transcription factor-1, are also transported to the nucleus via the Pin1-Hsp90α system to modulate gene expression and energy metabolism. Notably, as malignant tumors often express enhanced Pin1-Hsp90α signaling pathways, this provides a potential therapeutic target for tumors.
Keywords:
Nrf2; Hsp90; ERK; JNK; Cdk5; Pin1; HO-1

Graphical Abstract

1. Introduction

Cells respond to mild oxidative stress, such as non-lethal levels of hydrogen peroxide (H2O2), and rapidly upregulate defense systems to afford protection against oxidative damage (reviewed in [1]). Among the redox-sensitive transcription factors, nuclear factor-E2-related factor 2 (Nrf2) plays an important role in the expression of genes involved in detoxification, glutathione (GSH) synthesis, and defenses against oxidative damage [2,3,4]. Upregulation and activation of Nrf2 are controlled through a complex transcriptional, translational, and post-translational network that ensures an increase in its activity during redox perturbation, inflammation, growth factor stimulation, and nutrient/energy fluxes, thereby enabling Nrf2 to orchestrate adaptive responses to diverse forms of stress (reviewed in [5,6,7,8,9]).
The rapid activation of Nrf2 by stress agents depends on both inhibition of degradation and/or enhanced Nrf2 translation in the cytoplasm and facilitated nuclear translocation of Nrf2 to the nucleus (Figure 1A). Nrf2 is unstable under unstressed conditions due to constant degradation via the 26S proteasome, which is mediated by its cytoplasmic binding partner Kelch-like ECH-associated protein 1 (Keap1). Keap1 has cysteine residues highly reactive with various types of electrophiles, which form Michael adducts with Keap1 to inhibit interaction with Nrf2 resulting in Nrf2 stabilization and accumulation (reviewed in [10,11,12]). Degradation of Nrf2 is also controlled by phosphorylation at the Neh6 domain by glycogen synthase kinase-3 (GSK-3) (reviewed in [6,7,8]). It enhances the degradation of Nrf2 mediated by β-transducin repeat-containing protein, present in the ubiquitin ligase complex. Therefore, inhibition of GSK-3 by PI3-K (phosphatidylinositol 3-kinase)/Akt (PKB) signaling is a prerequisite to induce Keap1-mediated Nrf2 stabilization by electrophiles (reviewed in [6,7,8]).
In addition to Nrf2 stabilization, oxidative stress induces rapid Nrf2 protein synthesis [13]. Treatment of rat cardiomyocytes, human HeLa, and other cells with H2O2 induces rapid upregulation of translational Nrf2 protein synthesis independent of Nrf2 protein stabilization in these cells [13,14]. Notably, Nrf2 mRNA has an internal ribosomal entry site within the 5′ untranslated region (5′UTR) of human Nrf2 mRNA [14], and H2O2 treatments have been shown to cause rapid translocation of La autoantigen (Sjögren Syndrome Antigen B) from the nucleus to perinuclear space to associate with ribosomes. Binding La autoantigen to 5′UTR of Nrf2 mRNA stabilizes it and induces an Internal Ribosome Entry Site mediated protein translation in HeLa cells [15]. However, increased Nrf2 levels in the cytoplasm do not automatically lead to nuclear accumulation of Nrf2, as the transport of Nrf2 into the nucleus is tightly controlled, requiring additional regulators such as importins [16].
Nuclear import of proteins is tightly controlled by the interaction with a cytosolic protein termed importin composed of α and β subunits [17,18,19]. Importin α recognizes and binds nuclear localization signals (NLSs) of karyophilic proteins, and importin β helps bind them to the nuclear envelope, with energy-dependent, small GTPase Ran-mediated translocation through the pore resulting in the accumulation of import substrate and importin-α in the nucleus [18,19,20]. Nrf2 has three functional Lys/Arg-rich NLS motifs in Neh1, Neh2, and Neh3 domains, respectively, and mutations of the three NLS motifs significantly impair nuclear translocation of Nrf2 in HepG2 cells, and ARE-reporter gene expression in human leukemia K562 cells treated with tertiary-butylhydroquinone (tBHQ) [16]. These authors further showed that anti-importin α5 and anti-importin β1, respectively, co-immunoprecipitated with Nrf2 shortly after treating K562 cells with tBHQ, and that the amount of Nrf2 co-immunoprecipitated by these antibodies in nuclear fractions increased over 30–60 min [16]. This study shows that the transcriptional activation of Nrf2 depends largely on importin-mediated nucleocytoplasmic shuttling. It is also noted that exportin mediates nuclear export through interaction with nuclear export signals (NESs) [21,22]. Nrf2 has at least two functional NESs in the leucine zipper domain [23] and the transactivation domain [24]. The former Nrf2 NES can be masked by the formation of heterodimers with functional partner small Maf proteins, and the latter is a redox-sensitive NES. Thus, maximal Nrf2 activation or nuclear accumulation is achieved by enhanced association with importin α5β1 in the cytoplasm to facilitate nuclear translocation and inhibition of exportin binding in the nucleus (Figure 1B).
Although the precise mechanism of importin-dependent Nrf2 nuclear translocation remains unclear, previous studies have established the importance of stress-activated mitogen-activated protein kinases (MAPKs), p38, extracellular signal-regulated kinase (ERK), and c-Jun N-terminal kinase (JNK) in the nuclear translocation and activation of Nrf2 (reviewed in [25,26]). MAPKs are proline-directed serine/threonine protein kinases activated by dual phosphorylation on threonine and tyrosine residues in response to a wide array of extracellular stimuli. They are essential components of signaling pathways that convert various extracellular signals into intracellular responses through serial phosphorylation cascades (reviewed in [27,28,29]). ERK, known alternatively as microtubule-associated protein-2/myelin basic protein kinase, is activated by numerous hormones, growth factors, and other extracellular stimuli (reviewed in [30,31]), whereas JNK and p38 are activated by distinct and overlapping sets of stress-related stimuli, including heat shock, inflammatory cytokines, ultraviolet, gamma irradiation, and hyperosmolarity (reviewed in [32]). Interestingly, tBHQ activates ERK2 and JNK1 in HepG2 and HeLa cells [33], and phenethyl isothiocyanate activates ERK2 and JNK1, which phosphorylate Nrf2, resulting in nuclear translocation in human prostate cancer PC-3 cells [34] (Figure 1C). Table 1 summarizes the role of MAPKs in Nrf2 activation and ARE-mediated reporter gene expression in different cell types stimulated with various stress agents reported between 2000 to 2012 [34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54]. Evidently, the dependence of Nrf2 activation on ERK, JNK, and p38 MAPKs differs significantly among cells and stress agents (see Table 1). The dependence of Nrf2 activation on MAP kinases is widely confirmed by later numerous studies using different types of cultured cells stimulated with other various natural and synthetic chemical agents that activate Nrf2. However, molecular mechanisms of MAP kinase-dependent Nrf2 activation remain unsolved.
In this review, we critically evaluate the differential role of p38, JNK, and ERK on the activation of Nrf2 by stress agents. We hypothesize that ERK/JNK signaling contributes to the nuclear translocation of Nrf2 and that simultaneous activation of p38 signaling activated under glutathione (GSH) deletion leads to maximal accumulation of Nrf2 in the nucleus (Figure 1D). We previously proposed that p38 signaling leads to the phosphorylation of Nrf2 by PKCζ and casein kinase 2 (CK2) to stabilize it and masks a nuclear exporting signal (NES) [1] as described in Section 2. Notably, previous studies show that stress-activated ERK/JNK signaling facilitates Nrf2 nuclear translocation [34,48,50]. We discuss the possible functional partners of ERK that regulate facilitated Nrf2 nuclear translocation, and propose a novel mechanism of Nrf2 nuclear translocation mediated by ERK, involving peptidylprolyl cis/trans isomerase (PPIase) NIMA-interacting 1 (Pin1) and the molecular chaperon heat shock protein 90α (Hsp90α) as discussed in Section 3, Section 4, Section 5 and Section 6. We further discuss cyclin-dependent kinase 5 (Cdk5)-mediated Nrf2 nuclear translocation in Section 7.

2. p38 Controls Glutathione Sensor Neutral Sphingomyelinase 2

Maintenance of the small antioxidant GSH at high levels is essential for protecting cells against oxidative damage. Cells respond to the downregulation of GSH under oxidative stress by activating Nrf2, which upregulates the expression of genes required for GSH synthesis to restore cellular GSH levels. Notably, neutral sphingomyelinase 2 (nSMase2) senses GSH levels, and its activity is inhibited by high levels of cellular GSH (>3mM), but depletion of cellular GSH induces nSMase2 activation to generate the lipid signaling molecule ceramide [55,56,57,58,59]. We previously proposed that depletion of GSH induces ceramide/PKCζ/CK2 signaling leading to phosphorylation of Nrf2 by these kinases [60,61,62]. Interestingly, oxidative stress induces activation of p38 MAP kinase, which causes trafficking of nSMase2 from perinuclear regions to the plasma membrane [63,64], thereby enhancing ceramide generation by nSMase2 under oxidative stress. We further speculated that ceramide/PKCζ/CK2 signaling phosphorylates Nrf2 tethered to caveolin 1 (Cav1) in membrane lipid rafts/caveolae and that phosphorylation by these kinases stabilizes Nrf2 and masks an NES which could favor Nrf2 nuclear localization [1] (Figure 2A).
However, this hypothesis requires an additional mechanism to facilitate the translocation of stabilized Nrf2 from cell membrane compartments to the nucleus. Facilitated nuclear translocation of Nrf2 requires importin binding to the NLSs and directional movement through the cytoplasm to nuclear membrane pores. Upon arrival at the nuclear pores, the complexes are transferred to the nuclear interior by importin-dependent facilitated diffusion [20]. Mechanisms and functional partners for the facilitated movement of Nrf2 from the cell periphery to nuclear pores remain unclear. We discuss the possible functional partners and mechanism of facilitated Nrf2 nuclear translocation in the following sections.

3. ERK/JNK and PPIase Pin1 Control Nrf2 Nuclear Translocation

Xu et al. [34] previously showed that ERK and JNK modulate the nuclear translocation of Nrf2 in human prostate cancer PC-3 cells. These authors pre-treated the cells in 0.5% serum-containing medium overnight, which induced sequestration of Nrf2 in the cytoplasm. Upon transfection of JNK1 and its activating kinase MKK into PC-3 cells, Nrf2 was localized in both the cytoplasm and nucleus. Similar results were obtained when PC-3 cells were transfected with ERK2 and its activating kinase MEK1. These authors suggested that JNK1 and ERK2 can phosphorylate Nrf2 and induce nuclear translocation.
Recent studies show that the PPIase Pin1 upregulates nuclear accumulation of Nrf2 [65,66,67]. PPIase catalyzes the conversion between cis and trans conformations of proline imidic peptide bonds, playing a role in protein folding, signal transduction, trafficking, assembly, and cell cycle regulation [68]. The three classes of PPIase are cyclophilins, FK506-binding proteins (FKBPs), and parvulins [68], and Pin1 belongs to the parvulin class of PPIase. Cyclophilins and FKBPs are called immunophilins as immunosuppressive drugs such as cyclosporin A, FK506, and rapamycin directly bind and inhibit these PPIases [68]. Previous studies show that immunophilins associate with steroid hormone receptors to modulate their functions (reviewed in [69]).
Pin1 is a unique PPIase that specifically recognizes phosphorylated serine or threonine immediately preceding a proline residue (pSer/Thr-Pro), isomerizes the peptide bond, and is known to play an important role in cell cycle progression (reviewed in [68,70]). Liang et al. [65] showed that Pin1 contributes to Nrf2 activation in pancreatic ductal adenocarcinoma cells with high K-ras activity [65]. Saeidi et al. [66] found that enhanced H-ras signaling in human breast cancer cells induces the association of Pin1 with Nrf2 to protect Nrf2 from Keap1-mediated degradation. Saeidi et al. [66] further showed that phosphorylation of human Nrf2 at Ser-215, -408, and -577 is essential for its interaction with Pin1 [66]. These authors showed that among MAP kinases, ERK and JNK, but not p38, phosphorylate these serine residues, suggesting Ras-ERK signaling promotes Pin1 association with ERK-phosphorylated Nrf2 to facilitate translocation to the nucleus. We speculated, based on these studies, that ERK2 and JNK initially phosphorylate Nrf2, resulting in the association of Pin1 to the p-Ser/Thr-Pro site(s). It seems plausible that a Pin1-mediated structural change in Nrf2 could expose NLS(s) for association with importin α5, which then recruits importin β1 to facilitate nuclear translocation (Figure 2B). We next discuss the possible additional partners for the ERK/JNK-Pin1-mediated Nrf2 nuclear translocation.

4. PPIase and Hsp90 Cooperate in the Nuclear Transport of Signaling Molecules

Another important function of some PPIases is their interaction with heat shock protein 90 (Hsp90) and the dynein/dynactin complex [71,72]. Hsp90 is the major molecular chaperone protecting many client proteins from denaturation and aggregation (reviewed in [73,74,75]). Notably, in addition to chaperone activity, Hsp90 controls nucleocytoplasmic trafficking of signaling molecules (reviewed in [76,77]). PPIases are regarded as co-chaperones of Hsp90 and are crucial for translocating hormone receptors, transcription factors, and signaling molecules [77,78,79,80,81].
Hsp90 has three functional domains, an N-terminal domain with ATPase, a middle domain for binding co-chaperones and clients, and a C-terminal domain for dimerization [82] (Figure 3A). The Hsp90 dimer forms two alternative structures, ATP-free open and ATP-bound closed structures [83,84] (Figure 3A). Notably, ATP hydrolysis is coupled with the chaperone activity and accompanies the release of associated clients. There are many co-chaperone proteins (cofactors) that regulate Hsp90 functions, including its ATPase activity [85]. Importantly, to keep a client associated with the benefits of long-distance translocation, a co-chaperone p23 inhibits ATPase activity [86] and is recruited to the Hsp90-FKBP52 complex [81,87].
Two isoforms of Hsp90, α and β, are abundantly expressed in the cytoplasm with similar chaperone activities. Proliferating cells express higher levels of Hsp90α than Hsp90β, as Hsp90α gene expression is controlled by c-Myc downstream of growth factor signaling [88]. As Hsp90α has higher dimer-forming potential than Hsp90β [89,90], and p23 co-chaperone preferentially associates with Hsp90α compared to Hsp90β [84], we propose that the Hsp90α dimer and p23 co-chaperone compose a backbone for the nuclear transport machinery for co-transport of PPIase and client proteins (Figure 3B).

5. Functional Interaction of Nrf2 with Hsp90

Ngo et al. [91] recently showed that Hsp90 (isoforms not identified) directly interacts with Nrf2 in a yeast model expression system and HeLa cells, and with Hsp90 preventing overexpressed Nrf2 from forming protein aggregates. This study shows that Nrf2 is a client of Hsp90 and suggests a possibility that under the normal/low levels of Nrf2, the Pin1-Nrf2 associates with Hsp90α dimer-p23, forming a multiprotein complex that can be transported efficiently to the nuclear pore complex via the dynein motor complex along microtubules (Figure 3C). To verify this hypothesis, further studies are required to detect stable interaction of the Nrf2-Pin1 complex with Hsp90α and to demonstrate the Pin1-mediated linkage of the signaling complex with the dynein motor system.
Interestingly, previous studies suggest functional interactions of Nrf2 with Hsp90. Jia et al. [92] observed that an Hsp90 ATPase inhibitor, 17-dimethylaminoethylamino-17-demethoxygeldanamycin (17-DMAG), upregulated nuclear Nrf2 and the expression of HO-1 in neuronal HT22 cells subjected to hypoxia/reoxygenation. Lazaro et al. [93] also observed that 17-DMAG upregulated Nrf2 activation in macrophages and vascular smooth muscle cells in atherosclerotic plaques in diabetic apolipoprotein E-deficient mice. In contrast to these reports, Hsp90 ATPase inhibitors, including 17-allylamino-demethoxygeldanamycin (17-AAG), caused the gradual death of human cancer cells expressing high levels of Nrf2 [94]. The Hsp90 inhibitor exhibits no toxicity to other cancer cells with normal levels of Nrf2 but induces toxicity in cells after treatment with diethyl maleate (100 μM) to upregulate Nrf2 levels [94]. This toxic effect of 17-AAG may reflect the importance of Hsp90 chaperone activity to prevent Nrf2 protein aggregation in Nrf2-overexpressing cells [91].

6. Pin1 Controls the Nuclear Translocation of Other ERK Substrates

In addition to Nrf2, the ERK-Pin1 system naturally controls the nuclear translocation of several other ERK substrates, which are important for energy metabolism and proliferation. The nuclear translocation of pyruvate kinase M2 (PKM2) is controlled by the ERK-Pin1 system [95]. PKM2 promotes glucose metabolism by aerobic glycolysis and contributes to anabolic metabolism [96]. PKM2 expression is upregulated in multiple cancer types and contributes to the Warburg effect (reviewed in [97,98,99,100,101]). Yang et al. [102] showed that activated ERK2 binds directly to PKM2 through the ERK2 docking groove and phosphorylates PKM2 at Ser-37, inducing a structural change of PKM2 from tetramer to dimer or monomer. PKM2 dimer recruits Pin1 for cis-trans isomerization of PKM2 and binding to importin α5 and translocation to the Hsp90 containing complex to the nucleus [102], supporting our hypothesis of ERK-Pin1-Hsp90 dependent Nrf2 nuclear translocation (Figure 3C).
The ERK-Pin1-Hsp90α machinery also controls the nuclear transport of hypoxia-inducible transcription factor-1 (HIF-1α). HIF-1α mediates the activation of networks of target genes involved in angiogenesis, erythropoiesis, and glycolysis (reviewed in [103,104,105]). Besides hypoxic conditions, phosphorylation of HIF-1α at Ser-451 by ERK is another central post-translational modification, which regulates its stability under both hypoxia and physiological normoxia (reviewed in [106]), and plays a crucial role in promoting tumor growth [107,108,109]. Jalouli et al. [108] showed that Pin1 associates with the p-Ser-Pro motif and regulates HIF-1α transcriptional activity. HIF-1α is a client of Hsp90α [110] and contains NLSs for importin α binding [111,112,113]. Mylonis et al. [114] further showed that phosphorylation of Ser-641/643 by ERK inhibits the association of exportin to the NES, resulting in the accumulation of HIF-1α in the nucleus. These studies indicate that the nuclear translocation of HIF-1α is largely dependent on the ERK-mediated Pin1-Hsp90α system.

7. Cdk5 Controls Nrf2 Nuclear Translocation through Pin1

In addition to ERK and JNK, another proline-directed serine/threonine kinase, cyclin-dependent kinase 5 (Cdk5), also controls Nrf2 activation. Jimenez-Blasco et al. [115] showed that treating astrocytes from fetal rat brains with 20 μM N-methyl-d-aspartate (NMDA) for 8 h induced Nrf2-dependent activation of antioxidant genes. These authors further showed that NMDA induces the phospholipase C-mediated endoplasmic reticulum release of Ca2+ and activation of PKCδ, which phosphorylates and stabilizes Cdk5 cofactor p35. Furthermore, the active p35/Cdk5 complex phosphorylated Nrf2 leading to Nrf2 nuclear translocation and ARE-mediated gene expression [115]. In another study, Lee et al. [116] showed that tBHQ induces Nrf2 activation and expression of NQO1 in IMR-32 human neuroblastoma cells independent of ERK. Interestingly, oxidative stress induces the upregulation of Cdk5 catalytic subunit p35 in IMR-32 cells [117], suggesting Cdk5/p35 instead of ERK plays a role in Pin1-mediated Nrf2 nuclear translocation in neuroblastoma cells. It has been reported that Cdk5 phosphorylates ubiquitin ligase TRIM59 leading to Pin1 and importin α5 association resulting in nuclear translocation [118]. These results suggest that Cdk5/Nrf2/Pin1 axis contributes to Nrf2 nuclear translocation, like ERK/Nrf2/Pin1 axis, irrespective of whether Cdk5 phosphorylates different Ser/Thr-Pro sites of Nrf2. Cdk5 regulates neuronal functions but is also associated with cancer development and has been considered a potential target for cancer treatment (reviewed in [119,120,121,122,123]).

8. Summary and Conclusions

MAP kinases contribute to the activation of the transcription factor Nrf2 (reviewed by Kong et al., 2001, 2002), noting that activation is dependent on the cell type and stress agent (see Table 1). We proposed that p38-mediated signaling could stabilize Nrf2 via ceramide-activated PKCζ phosphorylation and mask an NES by CK2 phosphorylation under oxidative stress accompanying GSH depletion (Figure 2A) [1]. In contrast, ERK2 and JNK1 induce Nrf2 translocation into the nucleus [34]. Concerning the mechanism underlying ERK/JNK-dependent Nrf2 nuclear translocation, we propose a novel concept that direct phosphorylation of Nrf2 by ERK/JNK induces assembly of the Hsp90α-Pin1-Nrf2 complex. Pin1 causes prolyl-isomerization of Nrf2, which allows importin α5 to associate the Hsp90α-Pin1-Nrf2 complex. Then, the Hsp90α-Pin1-Nrf2-importin containing signaling complex is carried by the dynein motor system toward the nuclear pore complex along microtubules (Figure 3C). In addition to ERK and JNK, Cdk5 also phosphorylates Nrf2 and helps Nrf2 nuclear translocation via Hsp90α-Pin1-dynein machinery. Importantly, Hsp90 contributes to Nrf2 activation in two ways, inhibition of denaturation or aggregation when Nrf2 is over-expressed [91] and facilitation of Nrf2 nuclear import. The Hsp90 chaperone activity depends on ATP hydrolysis, but it accompanies client release. However, ATP hydrolysis is suppressed during nuclear transport of Hsp90α-Pin1-Nrf2 containing multiprotein complex via association of co-chaperon p23 (Figure 3C). Thus, under oxidative stress, p38 and ERK/JNK and/or Cdk5 could work together to stabilize Nrf2 to facilitate the maximal accumulation of Nrf2 in the nucleus (Figure 1C,D).
Hsp90α is required for the proliferation, migration, and invasion of cancer cells in culture [124]. Hsp90 is considered a druggable target for cancer treatment [125,126]. Accumulating evidence indicates that Pin1 plays a key role in various cancers. Pin1-mediated β-catenin accumulation occurs in about 70% of hepatocellular carcinoma [127]. Intriguingly, cell proliferation, migration, and invasion are significantly inhibited in Pin1-silenced Hep-2 cells [128]. Silencing of Pin1 causes down-regulation of β-catenin and cyclin D1 expression [128] and significantly increases the sensitivity to cisplatin in HeLa cells [129]. Thus, ERK- and Cdk5-signaling coupled with Hsp90α-Pin1-dynein machinery may be a prime target of chemotherapy for tumors. Developing chemical agents that inhibit Pin1 rather than Hap90 ATPase activity could be a promising approach for cancer chemotherapy (reviewed in [130,131,132]). For instance, a Pin1 inhibitor all-trans retinoic acid has been used to treat acute promyelocytic leukemia in animal models and human patients [133], and all-trans retinoic acid reduced the growth of transplanted tamoxifen-resistant human breast cancer cells in mice [134]. Dubiella et al. [135] showed that a Pin1 inhibitor Sulfopin reduced tumor progression and conferred survival benefits in animal models, while Liu et al. [136] developed a delivery system of a Pin1 inhibitor AG177724 targeting cancer-associated fibroblasts and observed the inhibition of tumor growth in mice.
Kim et al. [137] raised a question concerning the role of Pin1 in the expression of HO-1 induced by nitric oxide (NO) in mouse vascular smooth muscle cells and embryonic fibroblasts. These authors observed that treatment of the cells with NO donor nitroprusside (3 mM) for 8 h upregulated HO-1 levels in control cells to a higher extent than in Pin1 deficient cells and argued that Nrf2/ARE-mediated transcriptional activity is negatively controlled by Pin1 in fibroblasts [137]. However, the expression of HO-1 is controlled by both Nrf2 and AP-1 via partially overlapping AREs and TPA-responsive elements in the gene promotor [138], and Mouawad et al. [139] showed that NO-dependent expression of HO-1 is controlled by transcription factors C/EBPβ and AP-1 in macrophages [139]. Therefore, we suggest the possibility that NO-activated AP-1 could induce HO-1 gene expression more efficiently in the absence of Pin1-mediated Nrf2 nuclear translocation.
Low molecular kinase inhibitors are widely used to examine the function of kinases, but some inhibitors exhibit unexpected side effects. For instance, the p38 inhibitor SB203580 is a potent agonist of aryl hydrocarbon receptor (AhR) [140,141], which is a ligand-activated transcription factor and key regulator of xenobiotic metabolism, and AhR activation induces Cyp1a1 gene expression via xenobiotic responsive element (XRE) in Hepa 1c1c7 and HepG2 cells [142]. A complex effect of SB203580 is observed in HepG2 cells, which express high AhR levels [143]. Yu et al. [143] observed that tBHQ (100 μM) activated p38 and upregulated NQO1 expression in HepG2 cells, suggesting the role of the p38/Nrf2/ARE axis for the NQO1 expression. However, these authors observed that the addition of the p38 inhibitor SB203580 (5 μM) with tBHQ upregulated expression of NQO1 levels higher than tBHQ alone in 24 h, and argued that the p38 kinase pathway functions as a negative regulator in the ARE-mediated induction of phase II detoxifying enzymes [143]. We believe this conclusion may be misleading due to neglecting the influence of AhR activation by SB203580. Notably, NQO1 gene expression can be controlled by both Nrf2/ARE and AhR/XRE [144], and Nrf2 gene expression can be upregulated by AhR/XRE signaling [145]. Treatment of hepatoma 1c1c7 cells with AhR agonist 2,3,7,8-tetrachlorodibenzo-p-dioxin (10 nM) leads to upregulation of Nrf2 mRNA in 2 h and Nrf2 protein levels in 6 h [145]. Notably, tBHQ is a ligand for AhR. It is known to induce expression of Cyp1a1 in hepatoma 1c1c7 and HepG2 cells [146], suggesting the possibility that SB203580 and tBHQ synergistically activate AhR/XRE signaling, leading to rapid NQO1 gene expression with delayed Nrf2/ARE-mediated NQO1 expression. Thus, Phase I and II xenobiotic metabolism is coordinately regulated by the cross-talk between AhR and Nrf2 via ARE and XRE elements, respectively (reviewed in [147]).
In summary, we propose that ERK, JNK, and Cdk5 control Nrf2 phosphorylation inducing a formation of a multiprotein complex containing Hsp90α-p23-Pin1-Nrf2-importins to associate with dynein motors to move from cell membrane signaling compartments toward nuclear pores along microtubules (Figure 3C). Therefore, the integrity of microtubules is important to ensure the nuclear translocation of Nrf2 and other signaling molecules. It is important to note that tau family proteins control the assembly and maintenance of the structural stability of microtubules and that proline-directed kinases, MAP kinases, GSK-3β, and Cdk5, can phosphorylate tau proteins and affect their functions (reviewed in [148,149,150]). GSK-3β phosphorylated tau reduces the affinity to microtubules, causing the dysfunction of microtubules [151,152]. Interestingly, Pin1 interacts with phosphorylated tau proteins, restores the ability of tau to bind microtubules and promote assembly in vitro [153], and downregulates the GSK-3β-mediated phosphorylation of tau [154,155]. These studies suggest that Pin1/tau axis is also important for microtubule-mediated facilitated translocation of Nrf2 to the nucleus. As tau controls axonal transport and neurite outgrowth in neurons, defects in tau function could lead to neurodegeneration [156]. Thus, Pin1 plays an important but opposite role in the pathogenesis of Alzheimer’s disease and many human cancers (reviewed in [157,158]).

Funding

The Great Britain SASAKAWA Foundation for a Butterfield Award (B131) supporting our UK-Japan collaboration between the University of Tsukuba (T.I., E.W.) and King’s College London (G.E.M.). The Japanese Society for Promotion of Science, KAKENHI (E.W. JP20K07421), British Heart Foundation (G.E.M.), and COST Action ‘BenBedPhar’ CA20121 (G.E.M.).

Conflicts of Interest

The authors declare they have no potential conflict of interest.

Abbreviations

AhR, aryl hydrocarbon receptor; ERK, extracellular signal-regulated kinase; FKBP, FK506-binding protein; GSH, glutathione; GSK-3, glycogen synthase kinase-3; Keap1, Kelch-like ECH-associated protein 1; HIF-1α, hypoxia-inducible transcription factor-1α; HO-1, heme oxygenase 1; Hsp90, heat shock protein 90; MAPK, mitogen-activated protein kinase; NES, nuclear export signal; NLS, nuclear localization signal; NMDA, N-methyl-d-aspartate; Nrf2, nuclear factor-E2-related factor 2; Pin1, peptidyl-prolyl cis/trans isomerase NIMA-interacting 1; PI3-K, phosphatidylinositol 3-kinase; PKM2, pyruvate kinase M2; PPIase, peptidyl-prolyl cis/trans isomerase; tBHQ, tert-butylhydroquinone; 5’UTR, 5’ untranslated region.

References

  1. Ishii, T.; Warabi, E.; Mann, G.E. Mechanisms underlying Nrf2 nuclear translocation by non-lethal levels of hydrogen peroxide: p38 MAPK-dependent neutral sphingomyelinase2 membrane trafficking and ceramide/PKCζ/CK2 signaling. Free Radic. Biol. Med. 2022, 191, 191–202. [Google Scholar] [CrossRef]
  2. Itoh, K.; Chiba, T.; Takahashi, S.; Ishii, T.; Igarashi, K.; Katoh, Y.; Oyake, T.; Hayashi, N.; Satoh, K.; Hatayama, I.; et al. An Nrf2/small Maf heterodimer mediates the induction of phase II detoxifying enzyme genes through antioxidant response elements. Biochem. Biophys. Res. Commun. 1997, 236, 313–322. [Google Scholar] [CrossRef]
  3. Ishii, T.; Itoh, K.; Takahashi, S.; Sato, H.; Yanagawa, T.; Katoh, Y.; Bannai, S.; Yamamoto, M. Transcription factor Nrf2 coordinately regulates a group of oxidative stress-inducible genes in macrophages. J. Biol. Chem. 2000, 275, 16023–16029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Chan, J.Y.; Kwong, M. Impaired expression of glutathione synthetic enzyme genes in mice with targeted deletion of the Nrf2 basic-leucine zipper protein. Biochim. Biophys. Acta 2000, 1517, 19–26. [Google Scholar] [CrossRef] [PubMed]
  5. Hayes, J.D.; Dinkova-Kostova, A.T. The Nrf2 regulatory network provides an interface between redox and intermediary metabolism. Trends Biochem. Sci. 2014, 39, 199–218. [Google Scholar] [CrossRef] [PubMed]
  6. Hayes, J.D.; Chowdhry, S.; Dinkova-Kostova, A.T.; Sutherland, C. Dual regulation of transcription factor Nrf2 by Keap1 and by the combined actions of β-TrCP and GSK-3. Biochem. Soc. Trans. 2015, 43, 611–620. [Google Scholar] [CrossRef] [Green Version]
  7. Cuadrado, A. Structural and functional characterization of Nrf2 degradation by glycogen synthase kinase 3/β-TrCP. Free Radic. Biol. Med. 2015, 88, 147–157. [Google Scholar] [CrossRef]
  8. Tebay, L.E.; Robertson, H.; Durant, S.T.; Vitale, S.R.; Penning, T.M.; Dinkova-Kostova, A.T.; Hayes, J.D. Mechanisms of activation of the transcription factor Nrf2 by redox stressors, nutrient cues, and energy status and the pathways through which it attenuates degenerative disease. Free Radic. Biol. Med. 2015, 88 Pt B, 108–146. [Google Scholar] [CrossRef] [Green Version]
  9. Yamamoto, M.; Kensler, T.W.; Motohashi, H. The KEAP1-NRF2 System: A Thiol-Based Sensor-Effector Apparatus for Maintaining Redox Homeostasis. Physiol. Rev. 2018, 98, 1169–1203. [Google Scholar] [CrossRef] [Green Version]
  10. Itoh, K.; Tong, K.I.; Yamamoto, M. Molecular mechanism activating Nrf2-Keap1 pathway in regulation of adaptive response to electrophiles. Free Radic. Biol. Med. 2004, 36, 1208–1213. [Google Scholar] [CrossRef] [PubMed]
  11. Kwak, M.K.; Wakabayashi, N.; Kensler, T.W. Chemoprevention through the Keap1-Nrf2 signaling pathway by phase 2 enzyme inducers. Mutat. Res. 2004, 555, 133–148. [Google Scholar] [CrossRef]
  12. Itoh, K.; Mimura, J.; Yamamoto, M. Discovery of the negative regulator of Nrf2, Keap1: A historical overview. Antioxid. Redox Signal. 2010, 13, 1665–1678. [Google Scholar] [CrossRef] [PubMed]
  13. Purdom-Dickinson, S.E.; Sheveleva, E.V.; Sun, H.; Chen, Q.M. Translational control of nrf2 protein in activation of antioxidant response by oxidants. Mol. Pharmacol. 2007, 72, 1074–1081. [Google Scholar] [CrossRef] [PubMed]
  14. Li, W.; Thakor, N.; Xu, E.Y.; Huang, Y.; Chen, C.; Yu, R.; Holcik, M.; Kong, A.N. An internal ribosomal entry site mediates redox-sensitive translation of Nrf2. Nucleic Acids Res. 2010, 38, 778–788. [Google Scholar] [CrossRef] [Green Version]
  15. Zhang, J.; Dinh, T.N.; Kappeler, K.; Tsaprailis, G.; Chen, Q.M. La autoantigen mediates oxidant induced de novo Nrf2 protein translation. Mol. Cell Proteom. 2012, 11, M111.015032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Theodore, M.; Kawai, Y.; Yang, J.; Kleshchenko, Y.; Reddy, S.P.; Villalta, F.; Arinze, I.J. Multiple nuclear localization signals function in the nuclear import of the transcription factor Nrf2. J. Biol. Chem. 2008, 283, 8984–8994, Erratum in J. Biol. Chem. 2008, 283, 14176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Görlich, D.; Prehn, S.; Laskey, R.A.; Hartmann, E. Isolation of a protein that is essential for the first step of nuclear protein import. Cell 1994, 79, 767–778. [Google Scholar] [CrossRef] [PubMed]
  18. Görlich, D.; Vogel, F.; Mills, A.D.; Hartmann, E.; Laskey, R.A. Distinct functions for the two importin subunits in nuclear protein import. Nature 1995, 377, 246–248. [Google Scholar] [CrossRef] [PubMed]
  19. Görlich, D.; Mattaj, I.W. Nucleocytoplasmic transport. Science 1996, 271, 1513–1518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Rexach, M.; Blobel, G. Protein import into nuclei: Association and dissociation reactions involving transport substrate, transport factors, and nucleoporins. Cell 1995, 83, 683–692. [Google Scholar] [CrossRef] [Green Version]
  21. Ullman, K.S.; Powers, M.A.; Forbes, D.J. Nuclear export receptors: From importin to exportin. Cell 1997, 90, 967–970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Weis, K. Importins and exportins: How to get in and out of the nucleus. Trends Biochem. Sci. 1998, 23, 185–189, Erratum in Trends Biochem. Sci. 1998, 23, 235. [Google Scholar] [CrossRef]
  23. Li, W.; Jain, M.R.; Chen, C.; Yue, X.; Hebbar, V.; Zhou, R.; Kong, A.N. Nrf2 Possesses a redox-insensitive nuclear export signal overlapping with the leucine zipper motif. J. Biol. Chem. 2005, 280, 28430–28438. [Google Scholar] [CrossRef] [Green Version]
  24. Li, W.; Yu, S.W.; Kong, A.N. Nrf2 possesses a redox-sensitive nuclear exporting signal in the Neh5 transactivation domain. J. Biol. Chem. 2006, 281, 27251–27263. [Google Scholar] [CrossRef] [Green Version]
  25. Kong, A.N.; Yu, R.; Chen, C.; Mandlekar, S.; Primiano, T. Signal transduction events elicited by natural products: Role of MAPK and caspase pathways in homeostatic response and induction of apoptosis. Arch. Pharm. Res. 2000, 23, 1–16. [Google Scholar] [CrossRef] [PubMed]
  26. Kong, A.N.; Owuor, E.; Yu, R.; Hebbar, V.; Chen, C.; Hu, R.; Mandlekar, S. Induction of xenobiotic enzymes by the MAP kinase pathway and the antioxidant or electrophile response element (ARE/EpRE). Drug Metab. Rev. 2001, 33, 255–271. [Google Scholar] [CrossRef]
  27. Wagner, E.F.; Nebreda, A.R. Signal integration by JNK and p38 MAPK pathways in cancer development. Nat. Rev. Cancer 2009, 9, 537–549. [Google Scholar] [CrossRef] [PubMed]
  28. Cuadrado, A.; Nebreda, A.R. Mechanisms and functions of p38 MAPK signalling. Biochem. J. 2010, 429, 403–417. [Google Scholar] [CrossRef] [Green Version]
  29. Guo, Y.J.; Pan, W.W.; Liu, S.B.; Shen, Z.F.; Xu, Y.; Hu, L.L. ERK/MAPK signalling pathway and tumorigenesis. Exp. Ther. Med. 2020, 19, 1997–2007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Cobb, M.H.; Robbins, D.J.; Boulton, T.G. ERKs, extracellular signal-regulated MAP-2 kinases. Curr. Opin. Cell Biol. 1991, 3, 1025–1032. [Google Scholar] [CrossRef]
  31. Ahn, N.G.; Seger, R.; Bratlien, R.L.; Krebs, E.G. Growth factor-stimulated phosphorylation cascades: Activation of growth factor-stimulated MAP kinase. Ciba Found. Symp. 1992, 164, 113–126; discussion 126–131. [Google Scholar] [CrossRef] [PubMed]
  32. Paul, A.; Wilson, S.; Belham, C.M.; Robinson, C.J.; Scott, P.H.; Gould, G.W.; Plevin, R. Stress-activated protein kinases: Activation, regulation and function. Cell Signal. 1997, 9, 403–410. [Google Scholar] [CrossRef]
  33. Yu, R.; Tan, T.H.; Kong, A.N. Butylated hydroxyanisole and its metabolite tert-butylhydroquinone differentially regulate mitogen-activated protein kinases. The role of oxidative stress in the activation of mitogen-activated protein kinases by phenolic antioxidants. J. Biol. Chem. 1997, 272, 28962–28970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Xu, C.; Yuan, X.; Pan, Z.; Shen, G.; Kim, J.H.; Yu, S.; Khor, T.O.; Li, W.; Ma, J.; Kong, A.N. Mechanism of action of isothiocyanates: The induction of ARE-regulated genes is associated with activation of ERK and JNK and the phosphorylation and nuclear translocation of Nrf2. Mol. Cancer Ther. 2006, 5, 1918–1926. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Yu, R.; Chen, C.; Mo, Y.Y.; Hebbar, V.; Owuor, E.D.; Tan, T.H.; Kong, A.N. Activation of mitogen-activated protein kinase pathways induces antioxidant response element-mediated gene expression via a Nrf2-dependent mechanism. J. Biol. Chem. 2000, 275, 39907–39913. [Google Scholar] [CrossRef] [Green Version]
  36. Zipper, L.M.; Mulcahy, R.T. Inhibition of ERK and p38 MAP kinases inhibits binding of Nrf2 and induction of GCS genes. Biochem. Biophys. Res. Commun. 2000, 278, 484–492. [Google Scholar] [CrossRef]
  37. Gong, P.; Hu, B.; Cederbaum, A.I. Diallyl sulfide induces heme oxygenase-1 through MAPK pathway. Arch. Biochem. Biophys. 2004, 432, 252–260. [Google Scholar] [CrossRef]
  38. Yeh, C.T.; Yen, G.C. Involvement of p38 MAPK and Nrf2 in phenolic acid-induced P-form phenol sulfotransferase expression in human hepatoma HepG2 cells. Carcinogenesis 2006, 27, 1008–1017. [Google Scholar] [CrossRef] [Green Version]
  39. Yao, P.; Nussler, A.; Liu, L.; Hao, L.; Song, F.; Schirmeier, A.; Nussler, N. Quercetin protects human hepatocytes from ethanol-derived oxidative stress by inducing heme oxygenase-1 via the MAPK/Nrf2 pathways. J. Hepatol. 2007, 47, 253–261. [Google Scholar] [CrossRef]
  40. Alam, J.; Wicks, C.; Stewart, D.; Gong, P.; Touchard, C.; Otterbein, S.; Choi, A.M.; Burow, M.E.; Tou, J. Mechanism of heme oxygenase-1 gene activation by cadmium in MCF-7 mammary epithelial cells. Role of p38 kinase and Nrf2 transcription factor. J. Biol. Chem. 2000, 275, 27694–27702. [Google Scholar] [CrossRef] [Green Version]
  41. Keum, Y.S.; Owuor, E.D.; Kim, B.R.; Hu, R.; Kong, A.N. Involvement of Nrf2 and JNK1 in the activation of antioxidant responsive element (ARE) by chemopreventive agent phenethyl isothiocyanate (PEITC). Pharm. Res. 2003, 20, 1351–1356. [Google Scholar] [CrossRef]
  42. Ogborne, R.M.; Rushworth, S.A.; O’Connell, M.A. Alpha-lipoic acid-induced heme oxygenase-1 expression is mediated by nuclear factor erythroid 2-related factor 2 and p38 mitogen-activated protein kinase in human monocytic cells. Arterioscler. Thromb. Vasc Biol. 2005, 25, 2100–2105. [Google Scholar] [CrossRef] [Green Version]
  43. Anwar, A.A.; Li, F.Y.; Leake, D.S.; Ishii, T.; Mann, G.E.; Siow, R.C. Induction of heme oxygenase 1 by moderately oxidized low-density lipoproteins in human vascular smooth muscle cells: Role of mitogen-activated protein kinases and Nrf2. Free Radic. Biol. Med. 2005, 39, 227–236. [Google Scholar] [CrossRef] [PubMed]
  44. Rushworth, S.A.; Ogborne, R.M.; Charalambos, C.A.; O’Connell, M.A. Role of protein kinase C delta in curcumin-induced antioxidant response element-mediated gene expression in human monocytes. Biochem. Biophys. Res. Commun. 2006, 341, 1007–1016. [Google Scholar] [CrossRef] [PubMed]
  45. Papaiahgari, S.; Kleeberger, S.R.; Cho, H.Y.; Kalvakolanu, D.V.; Reddy, S.P. NADPH oxidase and ERK signaling regulates hyperoxia-induced Nrf2-ARE transcriptional response in pulmonary epithelial cells. J. Biol. Chem. 2004, 279, 42302–42312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Pischke, S.E.; Zhou, Z.; Song, R.; Ning, W.; Alam, J.; Ryter, S.W.; Choi, A.M. Phosphatidylinositol 3-kinase/Akt pathway mediates heme oxygenase-1 regulation by lipopolysaccharide. Cell Mol. Biol. (Noisy-le-Grand) 2005, 51, 461–470. [Google Scholar]
  47. Choi, B.M.; Kim, S.M.; Park, T.K.; Li, G.; Hong, S.J.; Park, R.; Chung, H.T.; Kim, B.R. Piperine protects cisplatin-induced apoptosis via heme oxygenase-1 induction in auditory cells. J. Nutr. Biochem. 2007, 18, 615–622. [Google Scholar] [CrossRef]
  48. Manandhar, S.; Cho, J.M.; Kim, J.A.; Kensler, T.W.; Kwak, M.K. Induction of Nrf2-regulated genes by 3H-1, 2-dithiole-3-thione through the ERK signaling pathway in murine keratinocytes. Eur. J. Pharmacol. 2007, 577, 17–27. [Google Scholar] [CrossRef]
  49. Zhang, H.; Liu, H.; Iles, K.E.; Liu, R.M.; Postlethwait, E.M.; Laperche, Y.; Forman, H.J. 4-Hydroxynonenal induces rat gamma-glutamyl transpeptidase through mitogen-activated protein kinase-mediated electrophile response element/nuclear factor erythroid 2-related factor 2 signaling. Am. J. Respir. Cell Mol. Biol. 2006, 34, 174–181. [Google Scholar] [CrossRef] [Green Version]
  50. Lim, H.J.; Lee, K.S.; Lee, S.; Park, J.H.; Choi, H.E.; Go, S.H.; Kwak, H.J.; Park, H.Y. 15d-PGJ2 stimulates HO-1 expression through p38 MAP kinase and Nrf-2 pathway in rat vascular smooth muscle cells. Toxicol. Appl. Pharmacol. 2007, 223, 20–27. [Google Scholar] [CrossRef]
  51. Lin, A.H.; Chen, H.W.; Liu, C.T.; Tsai, C.W.; Lii, C.K. Activation of Nrf2 is required for up-regulation of the π class of glutathione S-transferase in rat primary hepatocytes with L-methionine starvation. J. Agric. Food Chem. 2012, 60, 6537–6545. [Google Scholar] [CrossRef]
  52. Balogun, E.; Hoque, M.; Gong, P.; Killeen, E.; Green, C.J.; Foresti, R.; Alam, J.; Motterlini, R. Curcumin activates the haem oxygenase-1 gene via regulation of Nrf2 and the antioxidant-responsive element. Biochem. J. 2003, 371, 887–895. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Buckley, B.J.; Marshall, Z.M.; Whorton, A.R. Nitric oxide stimulates Nrf2 nuclear translocation in vascular endothelium. Biochem. Biophys. Res. Commun. 2003, 307, 973–979. [Google Scholar] [CrossRef] [PubMed]
  54. Aburaya, M.; Tanaka, K.; Hoshino, T.; Tsutsumi, S.; Suzuki, K.; Makise, M.; Akagi, R.; Mizushima, T. Heme oxygenase-1 protects gastric mucosal cells against non-steroidal anti-inflammatory drugs. J. Biol. Chem. 2006, 281, 33422–33432. [Google Scholar] [CrossRef] [Green Version]
  55. Liu, B.; Hannun, Y.A. Inhibition of the neutral magnesium-dependent sphingomyelinase by glutathione. J. Biol. Chem. 1997, 272, 16281–16287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Yoshimura, S.; Banno, Y.; Nakashima, S.; Hayashi, K.; Yamakawa, H.; Sawada, M.; Sakai, N.; Nozawa, Y. Inhibition of neutral sphingomyelinase activation and ceramide formation by glutathione in hypoxic PC12 cell death. J. Neurochem. 1999, 73, 675–683. [Google Scholar] [CrossRef]
  57. Chatterjee, S.; Han, H.; Rollins, S.; Cleveland, T. Molecular cloning, characterization, and expression of a novel human neutral sphingomyelinase. J. Biol. Chem. 1999, 274, 37407–37412. [Google Scholar] [CrossRef] [Green Version]
  58. Bernardo, K.; Krut, O.; Wiegmann, K.; Kreder, D.; Micheli, M.; Schäfer, R.; Sickman, A.; Schmidt, W.E.; Schröder, J.M.; Meyer, H.E.; et al. Purification and characterization of a magnesium-dependent neutral sphingomyelinase from bovine brain. J. Biol. Chem. 2000, 275, 7641–7647. [Google Scholar] [CrossRef] [Green Version]
  59. Lavrentiadou, S.N.; Chan, C.; Kawcak, T.; Ravid, T.; Tsaba, A.; van der Vliet, A.; Rasooly, R.; Goldkorn, T. Ceramide-mediated apoptosis in lung epithelial cells is regulated by glutathione. Am. J. Respir. Cell Mol. Biol. 2001, 25, 676–684. [Google Scholar] [CrossRef] [Green Version]
  60. Ishii, T.; Warabi, E.; Mann, G.E. Circadian control of BDNF-mediated Nrf2 activation in astrocytes protects dopaminergic neurons from ferroptosis. Free Radic. Biol. Med. 2019, 133, 169–178. [Google Scholar] [CrossRef] [Green Version]
  61. Ishii, T.; Warabi, E. Mechanism of Rapid Nuclear Factor-E2-Related Factor 2 (Nrf2) Activation via Membrane-Associated Estrogen Receptors: Roles of NADPH Oxidase 1, Neutral Sphingomyelinase 2 and Epidermal Growth Factor Receptor (EGFR). Antioxidants 2019, 8, 69. [Google Scholar] [CrossRef] [Green Version]
  62. Ishii, T.; Warabi, E.; Mann, G.E. Mechanisms underlying unidirectional laminar shear stress-mediated Nrf2 activation in endothelial cells: Amplification of low shear stress signaling by primary cilia. Redox Biol. 2021, 46, 102103. [Google Scholar] [CrossRef] [PubMed]
  63. Levy, M.; Castillo, S.S.; Goldkorn, T. nSMase2 activation and trafficking are modulated by oxidative stress to induce apoptosis. Biochem. Biophys. Res. Commun. 2006, 344, 900–905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Clarke, C.J.; Truong, T.G.; Hannun, Y.A. Role for neutral sphingomyelinase-2 in tumor necrosis factor alpha-stimulated expression of vascular cell adhesion molecule-1 (VCAM) and intercellular adhesion molecule-1 (ICAM) in lung epithelial cells: p38 MAPK is an upstream regulator of nSMase2. J. Biol. Chem. 2007, 282, 1384–1396. [Google Scholar] [CrossRef] [Green Version]
  65. Liang, C.; Shi, S.; Liu, M.; Qin, Y.; Meng, Q.; Hua, J.; Ji, S.; Zhang, Y.; Yang, J.; Xu, J.; et al. PIN1 Maintains Redox Balance via the c-Myc/NRF2 Axis to Counteract Kras-Induced Mitochondrial Respiratory Injury in Pancreatic Cancer Cells. Cancer Res. 2019, 79, 133–145. [Google Scholar] [CrossRef] [Green Version]
  66. Saeidi, S.; Kim, S.J.; Han, H.J.; Kim, S.H.; Zheng, J.; Lee, H.B.; Han, W.; Noh, D.Y.; Na, H.K.; Surh, Y.J. H-Ras induces Nrf2-Pin1 interaction: Implications for breast cancer progression. Toxicol. Appl. Pharmacol. 2020, 402, 115121. [Google Scholar] [CrossRef] [PubMed]
  67. Saeidi, S.; Kim, S.J.; Guillen-Quispe, Y.N.; Jagadeesh, A.S.V.; Han, H.J.; Kim, S.H.; Zhong, X.; Piao, J.Y.; Kim, S.J.; Jeong, J.; et al. Peptidylprolyl cis-trans isomerase NIMA-interacting 1 directly binds and stabilizes Nrf2 in breast cancer. FASEB J. 2022, 36, e22068. [Google Scholar] [CrossRef]
  68. Göthel, S.F.; Marahiel, M.A. Peptidylprolyl cis-trans isomerases, a superfamily of ubiquitous folding catalysts. Cell Mol. Life Sci. 1999, 55, 423–436. [Google Scholar] [CrossRef]
  69. Ratajczak, T.; Cluning, C.; Ward, B.K. Steroid Receptor-Associated Immunophilins: A Gateway to Steroid Signalling. Clin. Biochem Rev. 2015, 36, 31–52. [Google Scholar]
  70. Rostam, M.A.; Piva, T.J.; Rezaei, H.B.; Kamato, D.; Little, P.J.; Zheng, W.; Osman, N. Peptidylprolyl isomerases: Functionality and potential therapeutic targets in cardiovascular disease. Clin. Exp. Pharmacol. Physiol. 2015, 42, 117–124. [Google Scholar] [CrossRef]
  71. Galigniana, M.D.; Radanyi, C.; Renoir, J.M.; Housley, P.R.; Pratt, W.B. Evidence that the peptidylprolyl isomerase domain of the hsp90-binding immunophilin FKBP52 is involved in both dynein interaction and glucocorticoid receptor movement to the nucleus. J. Biol. Chem. 2001, 276, 14884–14889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Galigniana, M.D.; Harrell, J.M.; Murphy, P.J.; Chinkers, M.; Radanyi, C.; Renoir, J.M.; Zhang, M.; Pratt, W.B. Binding of hsp90-associated immunophilins to cytoplasmic dynein: Direct binding and in vivo evidence that the peptidylprolyl isomerase domain is a dynein interaction domain. Biochemistry 2002, 41, 13602–13610. [Google Scholar] [CrossRef] [PubMed]
  73. Pearl, L.H.; Prodromou, C. Structure and mechanism of the Hsp90 molecular chaperone machinery. Annu Rev. Biochem. 2006, 75, 271–294. [Google Scholar] [CrossRef]
  74. Li, J.; Soroka, J.; Buchner, J. The Hsp90 chaperone machinery: Conformational dynamics and regulation by co-chaperones. Biochim. Biophys. Acta 2012, 1823, 624–635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Schopf, F.H.; Biebl, M.M.; Buchner, J. The HSP90 chaperone machinery. Nat. Rev. Mol. Cell Biol. 2017, 18, 345–360. [Google Scholar] [CrossRef] [PubMed]
  76. Pratt, W.B.; Toft, D.O. Regulation of signaling protein function and trafficking by the hsp90/hsp70-based chaperone machinery. Exp. Biol Med. (Maywood) 2003, 228, 111–133. [Google Scholar] [CrossRef]
  77. Pratt, W.B.; Galigniana, M.D.; Harrell, J.M.; DeFranco, D.B. Role of hsp90 and the hsp90-binding immunophilins in signalling protein movement. Cell Signal. 2004, 16, 857–872. [Google Scholar] [CrossRef]
  78. Galigniana, M.D.; Echeverría, P.C.; Erlejman, A.G.; Piwien-Pilipuk, G. Role of molecular chaperones and TPR-domain proteins in the cytoplasmic transport of steroid receptors and their passage through the nuclear pore. Nucleus 2010, 1, 299–308. [Google Scholar] [CrossRef] [Green Version]
  79. Chan, S.C.; Li, Y.; Dehm, S.M. Androgen receptor splice variants activate androgen receptor target genes and support aberrant prostate cancer cell growth independent of canonical androgen receptor nuclear localization signal. J. Biol. Chem. 2012, 287, 19736–19749. [Google Scholar] [CrossRef] [Green Version]
  80. Lee, J.; An, Y.S.; Kim, M.R.; Kim, Y.A.; Lee, J.K.; Hwang, C.S.; Chung, E.; Park, I.C.; Yi, J.Y. Heat Shock Protein 90 Regulates Subcellular Localization of Smads in Mv1Lu Cells. J. Cell Biochem. 2016, 117, 230–238. [Google Scholar] [CrossRef] [PubMed]
  81. Mazaira, G.I.; Piwien Pilipuk, G.; Galigniana, M.D. Corticosteroid receptors as a model for the Hsp90•immunophilin-based transport machinery. Trends Endocrinol Metab. 2021, 32, 827–838. [Google Scholar] [CrossRef] [PubMed]
  82. Li, J.; Buchner, J. Structure, function and regulation of the hsp90 machinery. Biomed. J. 2013, 36, 106–117. [Google Scholar] [CrossRef] [PubMed]
  83. Pearl, L.H.; Prodromou, C.; Workman, P. The Hsp90 molecular chaperone: An open and shut case for treatment. Biochem. J. 2008, 410, 439–453. [Google Scholar] [CrossRef] [Green Version]
  84. Synoradzki, K.; Miszta, P.; Kazlauskas, E.; Mickevičiūtė, A.; Michailovienė, V.; Matulis, D.; Filipek, S.; Bieganowski, P. Interaction of the middle domains stabilizes Hsp90α dimer in a closed conformation with high affinity for p23. Biol. Chem. 2018, 399, 337–345. [Google Scholar] [CrossRef]
  85. Taipale, M.; Tucker, G.; Peng, J.; Krykbaeva, I.; Lin, Z.Y.; Larsen, B.; Choi, H.; Berger, B.; Gingras, A.C.; Lindquist, S. A quantitative chaperone interaction network reveals the architecture of cellular protein homeostasis pathways. Cell 2014, 158, 434–448. [Google Scholar] [CrossRef] [Green Version]
  86. Rehn, A.B.; Buchner, J. p23 and Aha1. Subcell Biochem. 2015, 78, 113–131. [Google Scholar] [CrossRef] [PubMed]
  87. Hildenbrand, Z.L.; Molugu, S.K.; Herrera, N.; Ramirez, C.; Xiao, C.; Bernal, R.A. Hsp90 can accommodate the simultaneous binding of the FKBP52 and HOP proteins. Oncotarget 2011, 2, 43–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Li, W.; Miao, X.; Qi, Z.; Zeng, W.; Liang, J.; Liang, Z. Hepatitis B virus X protein upregulates HSP90alpha expression via activation of c-Myc in human hepatocarcinoma cell line, HepG2. Virol. J. 2010, 7, 45. [Google Scholar] [CrossRef] [Green Version]
  89. Nemoto, T.; Sato, N. Oligomeric forms of the 90-kDa heat shock protein. Biochem. J. 1998, 330 Pt 2, 989–995. [Google Scholar] [CrossRef] [Green Version]
  90. Kobayakawa, T.; Yamada, S.; Mizuno, A.; Nemoto, T.K. Substitution of only two residues of human Hsp90alpha causes impeded dimerization of Hsp90beta. Cell Stress Chaperones. 2008, 13, 97–104. [Google Scholar] [CrossRef] [Green Version]
  91. Ngo, V.; Brickenden, A.; Liu, H.; Yeung, C.; Choy, W.Y.; Duennwald, M.L. A novel yeast model detects Nrf2 and Keap1 interactions with Hsp90. Dis. Model Mech. 2022, 15, dmm.049258. [Google Scholar] [CrossRef]
  92. Jia, Z.; Dong, A.; Che, H.; Zhang, Y. 17-DMAG Protects Against Hypoxia-/Reoxygenation-Induced Cell Injury in HT22 Cells Through Akt/Nrf2/HO-1 Pathway. DNA Cell Biol. 2017, 36, 95–102. [Google Scholar] [CrossRef]
  93. Lazaro, I.; Oguiza, A.; Recio, C.; Lopez-Sanz, L.; Bernal, S.; Egido, J.; Gomez-Guerrero, C. Interplay between HSP90 and Nrf2 pathways in diabetes-associated atherosclerosis. Clin. Investig. Arterioscler. 2017, 29, 51–59. [Google Scholar] [CrossRef]
  94. Baird, L.; Suzuki, T.; Takahashi, Y.; Hishinuma, E.; Saigusa, D.; Yamamoto, M. Geldanamycin-Derived HSP90 Inhibitors Are Synthetic Lethal with NRF2. Mol. Cell Biol. 2020, 40, e00377-20. [Google Scholar] [CrossRef]
  95. Yang, W.; Lu, Z. Nuclear PKM2 regulates the Warburg effect. Cell Cycle 2013, 12, 3154–3158. [Google Scholar] [CrossRef] [Green Version]
  96. Vander Heiden, M.G.; Locasale, J.W.; Swanson, K.D.; Sharfi, H.; Heffron, G.J.; Amador-Noguez, D.; Christofk, H.R.; Wagner, G.; Rabinowitz, J.D.; Asara, J.M.; et al. Evidence for an alternative glycolytic pathway in rapidly proliferating cells. Science 2010, 329, 1492–1499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Chen, M.; Zhang, J.; Manley, J.L. Turning on a fuel switch of cancer: hnRNP proteins regulate alternative splicing of pyruvate kinase mRNA. Cancer Res. 2010, 70, 8977–8980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Bayley, J.P.; Devilee, P. The Warburg effect in 2012. Curr. Opin. Oncol. 2012, 24, 62–67. [Google Scholar] [CrossRef]
  99. Tamada, M.; Suematsu, M.; Saya, H. Pyruvate kinase M2: Multiple faces for conferring benefits on cancer cells. Clin. Cancer Res. 2012, 18, 5554–5561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Filipp, F.V. Cancer metabolism meets systems biology: Pyruvate kinase isoform PKM2 is a metabolic master regulator. J. Carcinog. 2013, 12, 14. [Google Scholar] [CrossRef] [PubMed]
  101. Wong, N.; Ojo, D.; Yan, J.; Tang, D. PKM2 contributes to cancer metabolism. Cancer Lett. 2015, 356, 184–191. [Google Scholar] [CrossRef] [PubMed]
  102. Yang, W.; Zheng, Y.; Xia, Y.; Ji, H.; Chen, X.; Guo, F.; Lyssiotis, C.A.; Aldape, K.; Cantley, L.C.; Lu, Z. ERK1/2-dependent phosphorylation and nuclear translocation of PKM2 promotes the Warburg effect. Nat. Cell Biol. 2012, 14, 1295–1304, Erratum in Nat. Cell Biol. 2013, 15, 124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Schofield, C.J.; Ratcliffe, P.J. Oxygen sensing by HIF hydroxylases. Nat. Rev. Mol. Cell Biol. 2004, 5, 343–354. [Google Scholar] [CrossRef]
  104. Kaelin, W.G. Proline hydroxylation and gene expression. Annu. Rev. Biochem. 2005, 74, 115–128. [Google Scholar] [CrossRef] [PubMed]
  105. Semenza, G.L. Hypoxia-inducible factor 1 (HIF-1) pathway. Sci STKE. 2007, 2007, cm8. [Google Scholar] [CrossRef]
  106. Keeley, T.P.; Mann, G.E. Defining Physiological Normoxia for Improved Translation of Cell Physiology to Animal Models and Humans. Physiol Rev. 2019, 99, 161–234. [Google Scholar] [CrossRef] [Green Version]
  107. Kallio, P.J.; Okamoto, K.; O’Brien, S.; Carrero, P.; Makino, Y.; Tanaka, H.; Poellinger, L. Signal transduction in hypoxic cells: Inducible nuclear translocation and recruitment of the CBP/p300 coactivator by the hypoxia-inducible factor-1alpha. EMBO J. 1998, 17, 6573–6586. [Google Scholar] [CrossRef]
  108. Jalouli, M.; Déry, M.A.; Lafleur, V.N.; Lamalice, L.; Zhou, X.Z.; Lu, K.P.; Richard, D.E. The prolyl isomerase Pin1 regulates hypoxia-inducible transcription factor (HIF) activity. Cell Signal. 2014, 26, 1649–1656. [Google Scholar] [CrossRef]
  109. Han, H.J.; Saeidi, S.; Kim, S.J.; Piao, J.Y.; Lim, S.; Guillen-Quispe, Y.N.; Choi, B.Y.; Surh, Y.J. Alternative regulation of HIF-1α stability through Phosphorylation on Ser451. Biochem. Biophys. Res. Commun. 2021, 545, 150–156. [Google Scholar] [CrossRef]
  110. Tang, X.; Chang, C.; Hao, M.; Chen, M.; Woodley, D.T.; Schönthal, A.H.; Li, W. Heat shock protein-90alpha (Hsp90α) stabilizes hypoxia-inducible factor-1α (HIF-1α) in support of spermatogenesis and tumorigenesis. Cancer Gene Ther. 2021, 28, 1058–1070, Erratum in Cancer Gene Ther. 2021, 28, 1071–1072. [Google Scholar] [CrossRef]
  111. Depping, R.; Steinhoff, A.; Schindler, S.G.; Friedrich, B.; Fagerlund, R.; Metzen, E.; Hartmann, E.; Köhler, M. Nuclear translocation of hypoxia-inducible factors (HIFs): Involvement of the classical importin alpha/beta pathway. Biochim. Biophys. Acta 2008, 1783, 394–404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Chachami, G.; Paraskeva, E.; Mingot, J.M.; Braliou, G.G.; Görlich, D.; Simos, G. Transport of hypoxia-inducible factor HIF-1alpha into the nucleus involves importins 4 and 7. Biochem. Biophys. Res. Commun. 2009, 390, 235–240. [Google Scholar] [CrossRef] [PubMed]
  113. Depping, R.; Jelkmann, W.; Kosyna, F.K. Nuclear-cytoplasmatic shuttling of proteins in control of cellular oxygen sensing. J. Mol. Med. 2015, 93, 599–608. [Google Scholar] [CrossRef]
  114. Mylonis, I.; Chachami, G.; Paraskeva, E.; Simos, G. Atypical CRM1-dependent nuclear export signal mediates regulation of hypoxia-inducible factor-1alpha by MAPK. J. Biol. Chem. 2008, 283, 27620–27627. [Google Scholar] [CrossRef] [Green Version]
  115. Jimenez-Blasco, D.; Santofimia-Castaño, P.; Gonzalez, A.; Almeida, A.; Bolaños, J.P. Astrocyte NMDA receptors’ activity sustains neuronal survival through a Cdk5-Nrf2 pathway. Cell Death Differ. 2015, 22, 1877–1889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Lee, J.M.; Hanson, J.M.; Chu, W.A.; Johnson, J.A. Phosphatidylinositol 3-kinase, not extracellular signal-regulated kinase, regulates activation of the antioxidant-responsive element in IMR-32 human neuroblastoma cells. J. Biol. Chem. 2001, 276, 20011–20016. [Google Scholar] [CrossRef] [Green Version]
  117. Strocchi, P.; Pession, A.; Dozza, B. Up-regulation of cDK5/p35 by oxidative stress in human neuroblastoma IMR-32 cells. J. Cell Biochem. 2003, 88, 758–765. [Google Scholar] [CrossRef]
  118. Sang, Y.; Li, Y.; Zhang, Y.; Alvarez, A.A.; Yu, B.; Zhang, W.; Hu, B.; Cheng, S.Y.; Feng, H. CDK5-dependent phosphorylation and nuclear translocation of TRIM59 promotes macroH2A1 ubiquitination and tumorigenicity. Nat. Commun. 2019, 10, 4013. [Google Scholar] [CrossRef] [Green Version]
  119. Pozo, K.; Bibb, J.A. The Emerging Role of Cdk5 in Cancer. Trends Cancer. 2016, 2, 606–618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Liu, W.; Li, J.; Song, Y.S.; Li, Y.; Jia, Y.H.; Zhao, H.D. Cdk5 links with DNA damage response and cancer. Mol. Cancer. 2017, 16, 60. [Google Scholar] [CrossRef] [Green Version]
  121. Oner, M.; Lin, E.; Chen, M.C.; Hsu, F.N.; Shazzad Hossain Prince, G.M.; Chiu, K.Y.; Teng, C.J.; Yang, T.Y.; Wang, H.Y.; Yue, C.H.; et al. Future Aspects of CDK5 in Prostate Cancer: From Pathogenesis to Therapeutic Implications. Int. J. Mol. Sci. 2019, 20, 3881. [Google Scholar] [CrossRef] [Green Version]
  122. Do, P.A.; Lee, C.H. The Role of CDK5 in Tumours and Tumour Microenvironments. Cancers 2020, 13, 101. [Google Scholar] [CrossRef] [PubMed]
  123. Sharma, S.; Sicinski, P. A kinase of many talents: Non-neuronal functions of CDK5 in development and disease. Open Biol. 2020, 10, 190287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Vartholomaiou, E.; Madon-Simon, M.; Hagmann, S.; Mühlebach, G.; Wurst, W.; Floss, T.; Picard, D. Cytosolic Hsp90α and its mitochondrial isoform Trap1 are differentially required in a breast cancer model. Oncotarget 2017, 8, 17428–17442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Barginear, M.F.; Van Poznak, C.; Rosen, N.; Modi, S.; Hudis, C.A.; Budman, D.R. The heat shock protein 90 chaperone complex: An evolving therapeutic target. Curr. Cancer Drug Targets 2008, 8, 522–532. [Google Scholar] [CrossRef]
  126. Barrott, J.J.; Haystead, T.A. Hsp90, an unlikely ally in the war on cancer. FEBS J. 2013, 280, 1381–1396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Pang, R.; Yuen, J.; Yuen, M.F.; Lai, C.L.; Lee, T.K.; Man, K.; Poon, R.T.; Fan, S.T.; Wong, C.M.; Ng, I.O.; et al. PIN1 overexpression and beta-catenin gene mutations are distinct oncogenic events in human hepatocellular carcinoma. Oncogene 2004, 23, 4182–4186. [Google Scholar] [CrossRef] [Green Version]
  128. Fan, G.; Wang, L.; Xu, J.; Jiang, P.; Wang, W.; Huang, Y.; Lv, M.; Liu, S. Knockdown of the prolyl isomerase Pin1 inhibits Hep-2 cell growth, migration, and invasion by targeting the β-catenin signaling pathway. Biochem. Cell Biol. 2018, 96, 734–741. [Google Scholar] [CrossRef]
  129. Wang, T.; Liu, Z.; Shi, F.; Wang, J. Pin1 modulates chemo-resistance by up-regulating FoxM1 and the involvements of Wnt/β-catenin signaling pathway in cervical cancer. Mol. Cell Biochem. 2016, 413, 179–187. [Google Scholar] [CrossRef]
  130. Xu, G.G.; Etzkorn, F.A. Pin1 as an anticancer drug target. Drug News Perspect. 2009, 22, 399–407. [Google Scholar] [CrossRef]
  131. Zhou, X.Z.; Lu, K.P. The isomerase PIN1 controls numerous cancer-driving pathways and is a unique drug target. Nat. Rev. Cancer 2016, 16, 463–478. [Google Scholar] [CrossRef] [PubMed]
  132. Wu, W.; Xue, X.; Chen, Y.; Zheng, N.; Wang, J. Targeting prolyl isomerase Pin1 as a promising strategy to overcome resistance to cancer therapies. Pharmacol. Res. 2022, 184, 106456. [Google Scholar] [CrossRef] [PubMed]
  133. Wei, S.; Kozono, S.; Kats, L.; Nechama, M.; Li, W.; Guarnerio, J.; Luo, M.; You, M.H.; Yao, Y.; Kondo, A.; et al. Active Pin1 is a key target of all-trans retinoic acid in acute promyelocytic leukemia and breast cancer. Nat. Med. 2015, 21, 457–466. [Google Scholar] [CrossRef] [PubMed]
  134. Huang, S.; Chen, Y.; Liang, Z.M.; Li, N.N.; Liu, Y.; Zhu, Y.; Liao, D.; Zhou, X.Z.; Lu, K.P.; Yao, Y.; et al. Targeting Pin1 by All-Trans Retinoic Acid (ATRA) Overcomes Tamoxifen Resistance in Breast Cancer via Multifactorial Mechanisms. Front. Cell Dev. Biol. 2019, 7, 322. [Google Scholar] [CrossRef] [PubMed]
  135. Dubiella, C.; Pinch, B.J.; Koikawa, K.; Zaidman, D.; Poon, E.; Manz, T.D.; Nabet, B.; He, S.; Resnick, E.; Rogel, A.; et al. Sulfopin is a covalent inhibitor of Pin1 that blocks Myc-driven tumors in vivo. Nat. Chem Biol. 2021, 17, 954–963. [Google Scholar] [CrossRef]
  136. Liu, J.; Wang, Y.; Mu, C.; Li, M.; Li, K.; Li, S.; Wu, W.; Du, L.; Zhang, X.; Li, C.; et al. Pancreatic tumor eradication via selective Pin1 inhibition in cancer-associated fibroblasts and T lymphocytes engagement. Nat. Commun. 2022, 13, 4308. [Google Scholar] [CrossRef]
  137. Kim, S.E.; Lee, M.Y.; Lim, S.C.; Hien, T.T.; Kim, J.W.; Ahn, S.G.; Yoon, J.H.; Kim, S.K.; Choi, H.S.; Kang, K.W. Role of Pin1 in neointima formation: Down-regulation of Nrf2-dependent heme oxygenase-1 expression by Pin1. Free Radic. Biol. Med. 2010, 48, 1644–1653. [Google Scholar] [CrossRef]
  138. Alam, J.; Cook, J.L. How many transcription factors does it take to turn on the heme oxygenase-1 gene? Am. J. Respir. Cell Mol. Biol. 2007, 36, 166–174. [Google Scholar] [CrossRef] [Green Version]
  139. Mouawad, C.A.; Mrad, M.F.; Al-Hariri, M.; Soussi, H.; Hamade, E.; Alam, J.; Habib, A. Role of nitric oxide and CCAAT/enhancer-binding protein transcription factor in statin-dependent induction of heme oxygenase-1 in mouse macrophages. PLoS ONE. 2013, 8, e64092. [Google Scholar] [CrossRef]
  140. Murray, I.A.; Patterson, A.D.; Perdew, G.H. Aryl hydrocarbon receptor ligands in cancer: Friend and foe. Nat. Rev. Cancer 2014, 14, 801–814. [Google Scholar] [CrossRef]
  141. Pollet, M.; Krutmann, J.; Haarmann-Stemmann, T. Commentary: Usage of Mitogen-Activated Protein Kinase Small Molecule Inhibitors: More Than Just Inhibition! Front. Pharmacol. 2018, 9, 935. [Google Scholar] [CrossRef]
  142. Korashy, H.M.; Anwar-Mohamed, A.; Soshilov, A.A.; Denison, M.S.; El-Kadi, A.O. The p38 MAPK inhibitor SB203580 induces cytochrome P450 1A1 gene expression in murine and human hepatoma cell lines through ligand-dependent aryl hydrocarbon receptor activation. Chem Res. Toxicol. 2011, 24, 1540–1548. [Google Scholar] [CrossRef] [Green Version]
  143. Yu, R.; Mandlekar, S.; Lei, W.; Fahl, W.E.; Tan, T.H.; Kong, A.N. p38 mitogen-activated protein kinase negatively regulates the induction of phase II drug-metabolizing enzymes that detoxify carcinogens. J. Biol. Chem. 2000, 275, 2322–2327. [Google Scholar] [CrossRef] [Green Version]
  144. Nioi, P.; Hayes, J.D. Contribution of NAD(P)H:quinone oxidoreductase 1 to protection against carcinogenesis, and regulation of its gene by the Nrf2 basic-region leucine zipper and the arylhydrocarbon receptor basic helix-loop-helix transcription factors. Mutat Res. 2004, 555, 149–171. [Google Scholar] [CrossRef]
  145. Miao, W.; Hu, L.; Scrivens, P.J.; Batist, G. Transcriptional regulation of NF-E2 p45-related factor (NRF2) expression by the aryl hydrocarbon receptor-xenobiotic response element signaling pathway: Direct cross-talk between phase I and II drug-metabolizing enzymes. J. Biol. Chem. 2005, 280, 20340–20348. [Google Scholar] [CrossRef] [Green Version]
  146. Gharavi, N.; El-Kadi, A.O. tert-Butylhydroquinone is a novel aryl hydrocarbon receptor ligand. Drug Metab. Dispos. 2005, 33, 365–372. [Google Scholar] [CrossRef] [Green Version]
  147. Köhle, C.; Bock, K.W. Coordinate regulation of Phase I and II xenobiotic metabolisms by the Ah receptor and Nrf2. Biochem Pharmacol. 2007, 73, 1853–1862. [Google Scholar] [CrossRef] [PubMed]
  148. Mandelkow, E.M.; Biernat, J.; Drewes, G.; Gustke, N.; Trinczek, B.; Mandelkow, E. Tau domains, phosphorylation, and interactions with microtubules. Neurobiol. Aging. 1995, 16, 355–362; discussion 362–363. [Google Scholar] [CrossRef]
  149. Johnson, G.V.; Stoothoff, W.H. Tau phosphorylation in neuronal cell function and dysfunction. J. Cell Sci. 2004, 117, 5721–5729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Dehmelt, L.; Halpain, S. The MAP2/Tau family of microtubule-associated proteins. Genome Biol. 2005, 6, 204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Wagner, U.; Utton, M.; Gallo, J.M.; Miller, C.C. Cellular phosphorylation of tau by GSK-3 beta influences tau binding to microtubules and microtubule organisation. J. Cell Sci. 1996, 109 Pt 6, 1537–1543. [Google Scholar] [CrossRef] [PubMed]
  152. Lovestone, S.; Hartley, C.L.; Pearce, J.; Anderton, B.H. Phosphorylation of tau by glycogen synthase kinase-3 beta in intact mammalian cells: The effects on the organization and stability of microtubules. Neuroscience 1996, 73, 1145–1157. [Google Scholar] [CrossRef] [PubMed]
  153. Lu, P.J.; Wulf, G.; Zhou, X.Z.; Davies, P.; Lu, K.P. The prolyl isomerase Pin1 restores the function of Alzheimer-associated phosphorylated tau protein. Nature 1999, 399, 784–788. [Google Scholar] [CrossRef]
  154. Min, S.H.; Cho, J.S.; Oh, J.H.; Shim, S.B.; Hwang, D.Y.; Lee, S.H.; Jee, S.W.; Lim, H.J.; Kim, M.Y.; Sheen, Y.Y.; et al. Tau and GSK3beta dephosphorylations are required for regulating Pin1 phosphorylation. Neurochem. Res. 2005, 30, 955–961. [Google Scholar] [CrossRef]
  155. Ma, S.L.; Pastorino, L.; Zhou, X.Z.; Lu, K.P. Prolyl isomerase Pin1 promotes amyloid precursor protein (APP) turnover by inhibiting glycogen synthase kinase-3β (GSK3β) activity: Novel mechanism for Pin1 to protect against Alzheimer disease. J. Biol. Chem. 2012, 287, 6969–6973. [Google Scholar] [CrossRef] [Green Version]
  156. Santacruz, K.; Lewis, J.; Spires, T.; Paulson, J.; Kotilinek, L.; Ingelsson, M.; Guimaraes, A.; DeTure, M.; Ramsden, M.; McGowan, E.; et al. Tau suppression in a neurodegenerative mouse model improves memory function. Science 2005, 309, 476–481. [Google Scholar] [CrossRef] [Green Version]
  157. Driver, J.A.; Zhou, X.Z.; Lu, K.P. Pin1 dysregulation helps to explain the inverse association between cancer and Alzheimer’s disease. Biochim. Biophys. Acta 2015, 1850, 2069–2076. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Lanni, C.; Masi, M.; Racchi, M.; Govoni, S. Cancer and Alzheimer’s disease inverse relationship: An age-associated diverging derailment of shared pathways. Mol. Psychiatry 2021, 26, 280–295. [Google Scholar] [CrossRef]
Figure 1. MAP kinases p38, ERK, JNK, and GSH depletion control activation of the transcription factor Nrf2. (A) Rapid activation of Nrf2 under oxidative stress depends on two separate steps, stabilization and upregulation of Nrf2 protein and nuclear translocation. (B) Nuclear import of Nrf2 is controlled by the association of importin α5β1 to its nuclear localization signals (NLSs), while nuclear export of Nrf2 is controlled by exportin binding to its nuclear export signals (NESs). (C) Non-lethal levels of oxidative stress, such as H2O2, induce GSH depletion and activation of MAP kinases (reviewed in [1]). ERK and JNK control Nrf2 nuclear translocation, and p38 signaling induces stabilization of Nrf2 and masking of an NES. (D) In unstressed conditions, Nrf2 (green dots) levels are low and mainly localized in the cytoplasm but not in the nucleus due to functional nuclear export signals. ERK/JNK activation could cause nuclear transport of Nrf2 across the nuclear membrane but does not enable Nrf2 to accumulate in the nucleus due to the functional nuclear export signals. As p38 signaling masks a nuclear export signal, simultaneous activation of ERK and p38 signaling pathways under GSH-depleting conditions induces effective nuclear accumulation/activation of Nrf2.
Figure 1. MAP kinases p38, ERK, JNK, and GSH depletion control activation of the transcription factor Nrf2. (A) Rapid activation of Nrf2 under oxidative stress depends on two separate steps, stabilization and upregulation of Nrf2 protein and nuclear translocation. (B) Nuclear import of Nrf2 is controlled by the association of importin α5β1 to its nuclear localization signals (NLSs), while nuclear export of Nrf2 is controlled by exportin binding to its nuclear export signals (NESs). (C) Non-lethal levels of oxidative stress, such as H2O2, induce GSH depletion and activation of MAP kinases (reviewed in [1]). ERK and JNK control Nrf2 nuclear translocation, and p38 signaling induces stabilization of Nrf2 and masking of an NES. (D) In unstressed conditions, Nrf2 (green dots) levels are low and mainly localized in the cytoplasm but not in the nucleus due to functional nuclear export signals. ERK/JNK activation could cause nuclear transport of Nrf2 across the nuclear membrane but does not enable Nrf2 to accumulate in the nucleus due to the functional nuclear export signals. As p38 signaling masks a nuclear export signal, simultaneous activation of ERK and p38 signaling pathways under GSH-depleting conditions induces effective nuclear accumulation/activation of Nrf2.
Antioxidants 12 00274 g001
Figure 2. Stress-activated p38 controls nSMase2 activation, and ERK and JNK mediate Pin1-dependent Nrf2 nuclear translocation. (A) MAP kinase p38 induces the transfer of neutral sphingomyelinase 2 (nSMase2) from the perinuclear region to the plasma membrane. Glutathione (GSH) depletion causes ceramide/PKCζ/CK2 signaling, which induces Nrf2 phosphorylation and stabilization and masks a nuclear export signal [1]. (B) Phosphorylation of Nrf2 by ERK/JNK leads to an association with Pin1, which causes a cis/trans structural change at the preceding proline residue (pSer/Thr-Pro), allowing an association with importin α5 at the exposed nuclear localization signal. Then, importin β1 associates with importin α5, allowing Nrf2 to translocate to the nucleus.
Figure 2. Stress-activated p38 controls nSMase2 activation, and ERK and JNK mediate Pin1-dependent Nrf2 nuclear translocation. (A) MAP kinase p38 induces the transfer of neutral sphingomyelinase 2 (nSMase2) from the perinuclear region to the plasma membrane. Glutathione (GSH) depletion causes ceramide/PKCζ/CK2 signaling, which induces Nrf2 phosphorylation and stabilization and masks a nuclear export signal [1]. (B) Phosphorylation of Nrf2 by ERK/JNK leads to an association with Pin1, which causes a cis/trans structural change at the preceding proline residue (pSer/Thr-Pro), allowing an association with importin α5 at the exposed nuclear localization signal. Then, importin β1 associates with importin α5, allowing Nrf2 to translocate to the nucleus.
Antioxidants 12 00274 g002
Figure 3. Proposed mechanism of Hsp90-mediated nuclear translocation of Nrf2. (A) Hsp90 is composed of three domains, an N-terminal domain with ATPase activity, a middle domain for binding co-chaperones and clients, and a C-terminal domain for dimerization. Hsp90 dimer takes two forms, e.g., open and ATP-bound closed ring forms. Co-chaperone p23 inhibits Hsp90 ATPase activity. (B) Hsp90α dimer-p23 complex is the basal structure for transporting PPIase-associated signal molecules such as steroid receptors to the nucleus. (C) Hsp90α-p23-Pin1-Nrf2-importin-α5β1 containing complex associates with a dynein motor complex, which carries the cargo toward the nuclear pore complex (NPC) along microtubules.
Figure 3. Proposed mechanism of Hsp90-mediated nuclear translocation of Nrf2. (A) Hsp90 is composed of three domains, an N-terminal domain with ATPase activity, a middle domain for binding co-chaperones and clients, and a C-terminal domain for dimerization. Hsp90 dimer takes two forms, e.g., open and ATP-bound closed ring forms. Co-chaperone p23 inhibits Hsp90 ATPase activity. (B) Hsp90α dimer-p23 complex is the basal structure for transporting PPIase-associated signal molecules such as steroid receptors to the nucleus. (C) Hsp90α-p23-Pin1-Nrf2-importin-α5β1 containing complex associates with a dynein motor complex, which carries the cargo toward the nuclear pore complex (NPC) along microtubules.
Antioxidants 12 00274 g003
Table 1. MAP kinase-dependent Nrf2 activation.
Table 1. MAP kinase-dependent Nrf2 activation.
CellsActivatorsMAPK DependenceReferences
Human hepatoma HepG2Sodium arsenite and mercury chlorideJNK-dependent ARE reporter gene and HO-1 expression[35]
Pyrrolidine dithiocarbamateERK and p38 inhibitors PD98059 and SB202190 reduced about 50% in γ-glutamylcystein synthetase expression[36]
Diallyl sulfideERK- and p38- dependent Nrf2 nuclear translocation and HO-1 expression[37]
Gallic acidp38 inhibitor reduced ARE-dependent P-form of phenol sulfotransferase expression[38]
Human hepatocyteQuercetinp38- and ERK-dependent Nrf2 activation and HO-1 expression[39]
Human mammary epithelia MCF-7Cadmium chloridep38-dependent but ERK-independent HO-1 expression[40]
Human HeLaPhenethyl isothiocyanateJNK-dependent ARE-reporter gene expression[41]
Human monocytic THP-1α-Lipoic acidp38 inhibitor significantly reduced Nrf2 dependent HO-1 expression[42]
Human aortic smooth muscleOxidized low-density lipoproteinERK, p38, and JNK inhibitors respectively reduced HO-1 expression and Nrf2 nuclear translocation[43]
Human prostate carcinoma PC-3Phenethyl isothiocyanateERK and JNK phosphorylate Nrf2 and induce nuclear translocation of Nrf2[34]
Human monocyteCurcuminp38 inhibitor but not ERK inhibitor reduced ARE-dependent GCLM and HO-1 mRNA expression[44]
Mouse alveolar epithelial C10HyperoxiaHyperoxia activates NADPH oxidase, which results in ERK-dependent Nrf2 activation[45]
Mouse macrophage RAW 264.7Lipopolysaccharidep38 inhibitor significantly reduced Nrf2 dependent HO-1 expression[46]
Mouse cochlearPiperineJNK inhibitor significantly reduced ARE-reporter gene expression and HO-1 expression[47]
Mouse keratinocyte3H-1,2-dithiole-3-thioneERK inhibitor but not p38 inhibitor suppressed Nrf2 nuclear translocation and ARE-reporter gene expression[48]
Rat epithelial L24-hydroxynonenalERK- and p38-dependent EPRE-mediated γ-glutamyl transpeptidase expression[49]
Rat vascular smooth muscle15d-PGJ2p38 inhibitor abolished Nrf2 dependent HO-1 expression[50]
Rat primary hepatocytesMethionine restrictionERK-dependent Nrf2 nuclear translocation and GSH-S-transferase π expression[51]
Rat kidney epithelial NRK-52ECurcuminp38 inhibitor reduced about 50% of HO-1 expression, but ERK and JNK inhibitors did not suppress HO-1 expression[52]
Bovine aortic endothelialSpermine NONOate (NO donor)p38 and ERK inhibitors SB203580 and PD98059 respectively reduce HO-1 expression[53]
Guinea pig gastric mucosalIndomethacinp38 inhibitor significantly reduced Nrf2 nuclear accumulation and HO-1 expression[54]
Abbreviations: ARE, antioxidant response element; HO-1, heme oxygenase-1; GCLM, glutamate-cysteine ligase modifier subunit; EPRE, electrophile response element.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ishii, T.; Warabi, E.; Mann, G.E. Stress Activated MAP Kinases and Cyclin-Dependent Kinase 5 Mediate Nuclear Translocation of Nrf2 via Hsp90α-Pin1-Dynein Motor Transport Machinery. Antioxidants 2023, 12, 274. https://doi.org/10.3390/antiox12020274

AMA Style

Ishii T, Warabi E, Mann GE. Stress Activated MAP Kinases and Cyclin-Dependent Kinase 5 Mediate Nuclear Translocation of Nrf2 via Hsp90α-Pin1-Dynein Motor Transport Machinery. Antioxidants. 2023; 12(2):274. https://doi.org/10.3390/antiox12020274

Chicago/Turabian Style

Ishii, Tetsuro, Eiji Warabi, and Giovanni E. Mann. 2023. "Stress Activated MAP Kinases and Cyclin-Dependent Kinase 5 Mediate Nuclear Translocation of Nrf2 via Hsp90α-Pin1-Dynein Motor Transport Machinery" Antioxidants 12, no. 2: 274. https://doi.org/10.3390/antiox12020274

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