Next Article in Journal / Special Issue
Peroxiredoxin 6: The Protector of Male Fertility
Previous Article in Journal
Crystal Structure of Chloroplastic Thioredoxin f2 from Chlamydomonas reinhardtii Reveals Distinct Surface Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Peroxiredoxin 6 in Cell Signaling

by
José A. Arevalo
and
José Pablo Vázquez-Medina
*
Department of Integrative Biology, University of California, Berkeley, CA, 94705, USA
*
Author to whom correspondence should be addressed.
Antioxidants 2018, 7(12), 172; https://doi.org/10.3390/antiox7120172
Submission received: 7 November 2018 / Revised: 17 November 2018 / Accepted: 20 November 2018 / Published: 24 November 2018
(This article belongs to the Special Issue Peroxiredoxin 6 as a Unique Member of the Peroxiredoxin Family)

Abstract

:
Peroxiredoxin 6 (Prdx6, 1-cys peroxiredoxin) is a unique member of the peroxiredoxin family that, in contrast to other mammalian peroxiredoxins, lacks a resolving cysteine and uses glutathione and π glutathione S-transferase to complete its catalytic cycle. Prdx6 is also the only peroxiredoxin capable of reducing phospholipid hydroperoxides through its glutathione peroxidase (Gpx) activity. In addition to its peroxidase activity, Prdx6 expresses acidic calcium-independent phospholipase A2 (aiPLA2) and lysophosphatidylcholine acyl transferase (LPCAT) activities in separate catalytic sites. Prdx6 plays crucial roles in lung phospholipid metabolism, lipid peroxidation repair, and inflammatory signaling. Here, we review how the distinct activities of Prdx6 are regulated during physiological and pathological conditions, in addition to the role of Prdx6 in cellular signaling and disease.

1. Introduction

Peroxiredoxins are a ubiquitous family of highly conserved enzymes that share a catalytic mechanism in which a redox-active (peroxidatic) cysteine residue in the active site is oxidized by a peroxide [1]. In peroxiredoxins 1–5 (2-cys peroxiredoxins), the resulting sulfenic acid then reacts with another (resolving) cysteine residue, forming a disulfide that is subsequently reduced by an appropriate electron donor to complete a catalytic cycle [2,3]. In contrast to peroxiredoxins 1–5, peroxiredoxin 6 (Prdx6, 1-cys peroxiredoxin) lacks a resolving cysteine and uses glutathione (GSH) to complete its catalytic peroxidatic reaction [4,5]. Aside from the mammalian Prdx6, 1-cys peroxiredoxins that use GSH as the resolving thiol have been described in yeast [6].
Prdx6 is the only mammalian peroxiredoxin capable of reducing phospholipid hydroperoxides through its glutathione peroxidase (Gpx) activity, in addition to reducing short chain hydroperoxides [7]. Moreover, Prdx6 also expresses acidic calcium-independent phospholipase A2 (aiPLA2) and lysophosphatidylcholine acyl transferase (LPCAT) activities in separate catalytic sites [8,9]. Therefore, Prdx6 is a multi-tasking enzyme that participates in cellular signaling by modulating several pathways through its peroxidase, aiPLA2, and LPCAT activities. The three activities of Prdx6 are differentially regulated by the subcellular localization of the protein, substrate binding, and post-translational modifications [10]. Differential expression of the enzymatic activities of Prdx6 appears to have contrasting roles in several pathologies. Here, we review the current knowledge of the role of Prdx6 in cellular signaling. This publication is part of a forum on Prdx6 as a unique member of the peroxiredoxin family.

2. Regulation of the Enzymatic Activities of Prdx6

The subcellular localization of Prdx6 is crucial for the regulation of its enzymatic activities. Prdx6 exhibits maximal aiPLA2 activity at an acidic pH and maximal peroxidase activity at a neutral pH [10]. Hence, when Prdx6 is localized in lysosomal-type organelles such as lamellar bodies in type II pneumocytes, it functions as an aiPLA2, contributing to lung phospholipid metabolism [11]. In contrast, cytosolic Prdx6 functions mainly as a peroxidase under basal conditions [4]. Aside from the cytosol and lysosomal-type organelles, Prdx6 can move to the mitochondria, the plasma membrane, and possibly other organelles, after cellular stimulation with different agonists [12,13]. Prdx6 translocation to the mitochondria controls oxidative stress in the initial step of PINK1/Parkin-mediated mitophagy [14]. Prdx6 translocation to the plasma membrane has several biological effects involving its peroxidase and aiPLA2 activities, which will be described in detail in this manuscript. Prdx6 has also been detected in extracellular fluids, including plasma, bronchoalveolar lavage, and cerebrospinal fluid, but it is unknown whether extracellular Prdx6 functions as a peroxidase or as an aiPLA2 [11,15,16].
Several post-translational modifications regulate the enzymatic activities of Prdx6 by promoting changes in the intracellular localization of the protein or inactivating one or both catalytic sites. The phosphorylation of Prdx6 at threonine 177 by p38 mitogen-activated protein kinase (MAPK), after stimulation with phorbol ester, angiotensin II, or shear stress, induces a conformational change in the protein, promotes its translocation to the plasma membrane, and increases its aiPLA2 activity [12,17]. Prdx6 is also phosphorylated in a circadian manner, suggesting that a switch between peroxidase and aiPLA2 activities may occur naturally [18]. Similarly, MAPK activity is required for the targeting of Prdx6 to lysosomal-type organelles, which occurs through the binding of Prdx6 to the chaperone protein 14-3-3ε [19,20,21]. The hyperoxidation of Prdx6 at cysteine 47 inactivates the Gpx activity while increasing aiPLA2. In contrast to the hyperoxidation of 2-cys peroxiredoxins, the hyperoxidation of Prdx6 is irreversible [22]. The glutathionylation of Prdx6 by π GSH S-transferase (πGST) is required for the reduction of the oxidized cysteine and the completion of the peroxidatic catalytic cycle [23]. Prdx6 can also be aberrantly Sumoylated by Sumo1 at lysines 122 and 142 during oxidative stress. Aberrant sumoylation of Prdx6 reduces its gene expression and protein abundance [24,25,26].
The binding of Prdx6 to different substrates differentially regulates its GPx and aiPLA2 activities. At an acidic pH, when aiPLA2 is maximal, Prdx6 binds to reduced phospholipids. In contrast, Prdx6 binds only to oxidized phospholipids at cytosolic pH. The binding of Prdx6 to oxidized phospholipids requires the translocation of Prdx6 from the cytosol to the cell membrane, and has been observed after treating live cells with peroxides [11,27,28,29,30]. Cytosolic Prdx6 could also potentially translocate to oxidized intracellular organelle membranes, but this idea remains unexplored.
Aside from posttranslational modifications and subcellular localization, the enzymatic activities of Prdx6 can be regulated by interactions with other proteins. The direct interaction of Prdx6 with the surfactant protein A (SP-A) suppresses aiPLA2 in a process that is important for the regulation of lung phospholipid metabolism in lamellar bodies of type II pneumocytes [31,32,33]. Similarly, the interaction of Prdx6 with LIMP2 likely facilitates the binding of Prdx6 to phospholipids in acidic environments [34]. As described above, the binding of Prdx6 to πGST is necessary for the reduction of the oxidized cysteine and thus for the completion of the peroxidatic catalytic cycle, but not for aiPLA2 activity [23,35,36]. Prdx6 can also interact with the NADPH oxidase (Nox) subunits p67phox and Noxa1 [37,38]. These interactions seem to be necessary for optimal Nox activity. Interaction of phosphorylated Prdx6 with p67phox inhibits aiPLA2 and appears to be the mechanism that terminates Nox activation [39].

3. Signaling by Prdx6 Peroxidase

Peroxiredoxins are well known for their role in cellular signaling as hydrogen peroxide (H2O2) sensors. As mentioned above, Prdx6 is the only peroxiredoxin capable of reducing phospholipid hydroperoxides in addition to reducing H2O2 and short-chain peroxides. Hydroperoxide reduction occurs in three steps: peroxidation (the reduction of the hydroperoxide, which oxidizes the cysteine to sulfenic acid), resolution (the reduction of sulfenic acid), and recycling (the regeneration of the cysteine active site) [40,41]. This process occurs at the conserved cysteine 47 catalytic site and requires glutathione and πGST [5,36].
Lipid peroxidation products such as malondialdehyde (MDA) and 4-hydroxynonenal (4HNE) have traditionally been viewed as toxic byproducts implicated in the etiology of several diseases [42,43,44,45]. However, increasing evidence shows that MDA and 4HNE also elicit biological activities that result in anti-inflammatory and pro-survival effects such as Nrf2 activation and UCP3 upregulation [46,47]. Therefore, lipid hydroperoxide reduction by Prdx6 is crucial to limit peroxidation-derived pathological effects and to regulate adaptive responses mediated by electrophilic lipid peroxidation products. The only other enzyme known to be able to reduce phospholipid hydroperoxides in mammalian cells is glutathione peroxidase 4 (Gpx4) [48,49]. The expression of Prdx6 and Gpx4 likely varies among tissues, cell types, and possibly developmental stages and disease progression, but it is currently unknown if these two proteins exhibit complementary functions or if one replaces the other in particular pathophysiological conditions.
As described above, Prdx6 is recruited to peroxidized cell membranes following oxidative stress, where it reduces and hydrolyzes the oxidized sn-2 fatty acyl or the sn-2 ester (alkyl) bond of oxidized phospholipids through its GPx and aiPLA2 activities [30,50]. Thus, Prdx6 actively participates in the prevention and repair of lipid peroxidation. The critical role of both the GPx and aiPLA2 activities of Prdx6 in the repair of lipid peroxidation has been extensively demonstrated in cells, lungs, and knock-in mouse models that express either Prdx6 GPx or aiPLA2 activity after treatment with peroxides and exposure to paraquat or hyperoxia [28,29,51].
A novel lysophosphatidylcholine acyl transferase (LPCAT) activity of Prdx6 coupled to aiPLA2 was recently described [9]. Prdx6 LPCAT acylates lysophosphatidylcholine (LPC) generated by aiPLA2 in a continuous process, without the release of the LPC intermediate generated by aiPLA2. Thus, LPCAT and aiPLA2 combine to replace an oxidized acyl chain with palmitoyl CoA, effectively repairing oxidized cell membranes. Hence, Prdx6 possesses the ability to reduce peroxidized cell membrane phospholipids through its GPx activity and to replace the oxidized sn-2 fatty acyl group through hydrolysis/reacylation by aiPLA2 and LPCAT, providing a complete system for the repair of peroxidized cell membranes [30].
The recent discovery of ferroptosis, a distinct cell death pathway characterized by the intracellular accumulation of lipid hydroperoxides [52,53,54,55], is generating increasing interest in the role of lipid peroxidation in the progression of several diseases. The role of GPx4, the only other enzyme capable of reducing lipid hydroperoxides, is evident in ferroptosis [54,55]. In contrast, the role of Prdx6 as a regulator of ferroptosis has not been studied, but undoubtedly warrants further investigation.

4. Signaling by aiPLA2

Studies using Prdx6 knock-in and knock-out mice, intact lung preparations and primary cells, and an aiPLA2 inhibitor (MJ33) show that, aside from its role in lung phospholipid metabolism, aiPLA2 is necessary for the activation of NADPH oxidase 2 (Nox2) during stimulation with phorbol ester, shear stress, angiotensin II, N-formylmethionine-leucyl-phenylalanine, or endotoxin [12,37,56,57,58,59,60]. The mechanism by which Prdx6 regulates Nox2 starts with the p38 MAPK-mediated phosphorylation of Prdx6, its translocation to the plasma membrane, and the generation of LPC by aiPLA2. LPC is then converted to lysophosphatidic acid (LPA) by lysophospholipase D. LPA signals through LPA receptors, activating the small GTPase Rac which moves from the cytosol to the membrane to activate Nox2 [12,60,61].
Initial studies discovered a direct interaction of Prdx6 with p67phox, a cytosolic component of Nox2, and suggested that aiPLA2 plays a critical role in Nox2 activation [56]. Later studies showed that binding occurs only between phosphorylated Prdx6 and non-phosphorylated p67phox and that such an interaction suppresses aiPLA2, likely terminating Nox2 activity [39]. Contrasting evidence exists about the role of Prdx6 in the regulation of other members of the NADPH oxidase family. On the one hand, MJ33 did not inhibit angiotensin II-stimulated oxidant generation in human pulmonary smooth muscle cells, which express high levels of NADPH oxidase 1 (Nox1) [58]. On the other hand, recent evidence showed that Prdx6 binds to Noxa1, a regulator of Nox1, and that knockdown of Prdx6 or inhibition of aiPLA2 with MJ33 blunts Nox1-derived oxidant generation and suppresses cell migration [38].
NADPH oxidases are important activators of redox-dependent inflammatory pathways [62]. Thus, aiPLA2 signaling plays a role in inflammation by regulating Nox-derived oxidant generation. The inhibition of aiPLA2 with MJ33 protects the lungs from hyperoxia, endotoxin, and ischemia/reperfusion injury by blunting oxidant generation and preventing inflammation [58,59,63]. Our unpublished observations suggest that knock-in mice deficient in aiPLA2 are resistant to sepsis-induced acute lung injury. Signaling effects of aiPLA2, independent from the regulation of NADPH oxidases such as the generation of inflammatory mediators derived from arachidonic acid, have been proposed [64,65]. aiPLA2, however, does not show a preference for arachidonic acid-containing phospholipids [66,67]. Therefore, it is unclear how aiPLA2 could regulate inflammatory pathways derived from arachidonic acid release.

5. Prdx6 Signaling in Disease

5.1. Cancer

Peroxiredoxins have a dual role in carcinogenesis, acting as both tumor suppressors and promoters [68,69,70]. Such a dual role appears to be consistent with the established role of oxidants in both cell proliferation and death [68,71,72]. Increased Prdx6 levels have been detected in various cancers, and the role of Prdx6 in lung cancer invasion is well known [73]. aiPLA2 promotes cell invasion and metastasis in various models, likely through the regulation of NADPH oxidase, which is needed for cell proliferation [74]. Mice that over-express Prdx6 show a greater increase in the growth of lung tumors compared to wild-type animals. Both aiPLA2 and Prdx6 peroxidase activities are implicated in lung tumor development through the regulation of redox-sensitive pathways such as MAPK, JNK, JAK/STAT, and AP-1 [75,76,77].

5.2. Inflammation

Prdx6 appears to have a dual role in inflammatory disease. One the one hand, as described above, aiPLA2 is necessary for the activation of the pro-inflammatory NADPH oxidase. On the other hand, Prdx6 is crucial to counteract increased reactive oxygen species (ROS) generation and to repair oxidized cell membranes following oxidative stress. Hyperoxidized Prdx6 levels are high in the peripheral blood mononuclear cells of moderate-to-severe asthma patients, where ROS generation is also elevated [78]. Moreover, LPS-treated Prdx6-over-expressing mice show less renal apoptosis and leukocyte infiltration than wild-type controls [79]. Similarly, Prdx6 null mice show increased lung inflammation after intratracheal instillation of lipopolysaccharide (LPS) [80]. Lung inflammation and mortality induced by paraquat or hyperoxia are also increased in Prdx6 null mice [81,82]. In contrast, Prdx6 null mice show reduced inflammation, but increased hepatic oxidative damage and mitochondrial dysfunction after experimental ischemia/reperfusion [83]. Moreover, the inhibition of aiPLA2 with MJ33 reduces inflammation after hyperoxia or intratracheal LPS exposure [59,63]. Therefore, it is likely that Prdx6 plays a dual role in inflammatory conditions, both as an activator of inflammatory pathways related to NADPH oxidase through aiPLA2 and as a protective/regulatory mechanism through Prdx6 peroxidase activity.
The link between Prdx6 and NfκB, which is one of the most prominent redox-regulated pro-inflammatory regulators, has been known for some time [84]. Prdx6 protects against ischemia/reperfusion injury during liver transplantation by downregulating NfκB [85]. Similarly, Prdx6 expression is inversely correlated with NfκB during Clonorchis sinensis infection [86]. In contrast, extracellular Prdx6 signals through a toll-like receptor (TLR) after focal cerebral ischemia/reperfusion [87]. Other extracellular oxidized peroxiredoxins released via the exosomal pathway can bind to TLR, contributing to sustained inflammatory responses that involve NfκB signaling. The specific role of Prdx6 in this process, however, remains unknown [88]. Our unpublished observations show that the genetic inactivation (knock-in) of aiPLA2 prevents the nuclear translocation of NfκB in lung endothelial cells stimulated with LPS. Thus, it appears that Prdx6 peroxidase and aiPLA2 may play contrasting roles in NfκB signaling.

5.3. Metabolic Disease

Prdx6 is linked to the development of type 2 diabetes. Prdx6 null mice show low insulin and elevated blood glucose levels, compared to wild-type mice, after a glucose tolerance test. Similarly, insulin signaling and pancreatic β cell morphology are impaired in Prdx6 knockout mice. Moreover, Prdx6 null mice develop dyslipidemia [89]. Prdx6 appears to also be linked to the generation of hydroxy fatty acids, which are antidiabetic/anti-inflammatory lipid mediators synthesized in white adipose tissue via de novo lipogenesis [90]. The specific role of Prdx6 peroxidase and aiPLA2 activities in the development and progression of metabolic disease remains unexplored.

5.4. Ocular Damage

The role of Prdx6 in ocular oxidative damage has been extensively investigated [91,92]. Prdx6 expression in the eye declines with age, increasing the risk of cataract formation [26,93,94]. Moreover, Prdx6 prevents oxidative stress in human retinal pigment epithelial cells by activating the Phosphoinositide 3-kinase/Protein kinase B (PI3K/AKT) pathway [95] and blunting transforming growth factor beta (TGF-β) signaling [96]. Prdx6 is therefore necessary to prevent cataract formation and to limit aging-induced oxidative stress in the eye. The exposure of human lens epithelial cells to ultraviolet-B light results in Prdx6 hyperoxidation, increased ROS generation, and cellular toxicity [92,97]. The function of Prdx6 in the eye is the topic of a separate review article, which is part of this forum.

5.5. Brain Injury and Neurodegeneration

The role of Prdx6 in neurodegenerative diseases is controversial [98]. Following a traumatic brain injury, oxidized Prdx6 is detected in the cerebrospinal fluid [16,99]. Moreover, Prdx6 signaling through TLR4 increases after an ischemic stroke, leading to NfκB-mediated inflammation and cell death [87,100,101]. Similarly, the pharmacological inhibition of aiPLA2 with MJ33 reduces inflammation in an experimental stroke model, supporting the idea that aiPLA2 has a pro-inflammatory role in the brain under such conditions [102].
The impairment of brain neurogenesis is linked to Alzheimer’s (AD) and Parkinson’s disease (PD) [103,104]. Recent work shows that Prdx6 has an inhibitory effect on neurogenesis [105]. In contrast, Prdx6 reduces oxidative stress and counteracts ROS generation, suggesting that Prdx6 plays a neuroprotective role in AD models [103]. Similarly, Prdx6 reduces inflammation and protects against disruption to the blood–brain barrier in a model of multiple sclerosis [106]. Prdx6 hyperoxidation is observed in a PD mouse model [104]. As discussed above, the hyperoxidation of Prdx6 is irreversible, and inactivates the Gpx function. Therefore, the dual role of Prdx6 in neurodegeneration seems to be linked to the differential expression of Prdx6 peroxidase and aiPLA2 activities.

5.6. Male Infertility

Prdx6 plays a crucial role in spermatozoon fertilizing ability [107,108]. Lower levels of Prdx6 are associated with impaired sperm function and poor DNA integrity in infertile men [109]. Sperm motility, viability, fertilization, and blastocyst rates are lower in Prdx6 null mice than they are in wild-type controls [110]. The role of Prdx6 in male fertility is extensively discussed in a separate review article, which is part of this forum.

6. Conclusions and Future Directions

Prdx6 is an intriguing enzyme that contributes to the regulation of a plethora of signaling pathways related to its diverse catalytic activities. The contrasting biological roles of Prdx6 are likely explained by the delicate differential expression of its peroxidase and aiPLA2 activities. Further investigations on the role of Prdx6 in cell signaling should focus on understanding the mechanisms that control the shift between Prdx6 peroxidase and aiPLA2 activities under physiological and pathological conditions. Similarly, the role of extracellular Prdx6 in inflammatory signaling remains underexplored. Moreover, the role of Prdx6 as a complete lipid peroxidation repair enzyme warrants further studies to elucidate how Prdx6 regulates cellular signaling by lipid peroxidation products, as well as the possible role of Prdx6 in novel pathways of cell death.

Author Contributions

J.A.A. and J.P.V.-M. contributed equally to all aspects of this manuscript.

Funding

J.P.V.-M. is supported by University of California startup funds.

Acknowledgments

We thank Aron Fisher, Sheldon Feinstein, and Shampa Chatterjee (University of Pennsylvania) for discussions and input about Prdx6 biology.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Perkins, A.; Nelson, K.J.; Parsonage, D.; Poole, L.B.; Karplus, P.A. Peroxiredoxins: Guardians against oxidative stress and modulators of peroxide signaling. Trends Biochem. Sci. 2015, 40, 435–445. [Google Scholar] [CrossRef] [PubMed]
  2. Rhee, S.G.; Kil, I.S. Multiple functions and regulation of mammalian peroxiredoxins. Annu. Rev. Biochem. 2017, 86, 749–775. [Google Scholar] [CrossRef] [PubMed]
  3. Rhee, S.G.; Kang, S.W.; Chang, T.S.; Jeong, W.; Kim, K. Peroxiredoxin, a novel family of peroxidases. IUBMB Life 2001, 52, 35–41. [Google Scholar] [CrossRef] [PubMed]
  4. Kang, S.W.; Baines, I.C.; Rhee, S.G. Characterization of a mammalian peroxiredoxin that contains one conserved cysteine. J. Biol. Chem. 1998, 273, 6303–6311. [Google Scholar] [CrossRef] [PubMed]
  5. Manevich, Y.; Feinstein, S.I.; Fisher, A.B. Activation of the antioxidant enzyme 1-CYS peroxiredoxin requires glutathionylation mediated by heterodimerization with pi GST. Proc. Natl. Acad. Sci. USA 2004, 101, 3780–3785. [Google Scholar] [CrossRef] [PubMed]
  6. Pedrajas, J.R.; McDonagh, B.; Hernandez-Torres, F.; Miranda-Vizuete, A.; Gonzalez-Ojeda, R.; Martinez-Galisteo, E.; Padilla, C.A.; Barcena, J.A. Glutathione is the resolving thiol for thioredoxin peroxidase activity of 1-Cys peroxiredoxin without being consumed during the catalytic cycle. Antioxid. Redox Signal. 2016, 24, 115–128. [Google Scholar] [CrossRef] [PubMed]
  7. Fisher, A.B.; Dodia, C.; Manevich, Y.; Chen, J.W.; Feinstein, S.I. Phospholipid hydroperoxides are substrates for non-selenium glutathione peroxidase. J. Biol. Chem. 1999, 274, 21326–21334. [Google Scholar] [CrossRef] [PubMed]
  8. Huang, C.F.; Sun, Z.J.; Zhao, Y.F.; Chen, X.M.; Jia, J.; Zhang, W.F. Increased expression of peroxiredoxin 6 and cyclophilin A in squamous cell carcinoma of the tongue. Oral Dis. 2011, 17, 328–334. [Google Scholar] [CrossRef] [PubMed]
  9. Fisher, A.B.; Dodia, C.; Sorokina, E.M.; Li, H.; Zhou, S.; Raabe, T.; Feinstein, S.I. A novel lysophosphatidylcholine acyl transferase activity is expressed by peroxiredoxin 6. J. Lipid Res. 2016, 57, 587–596. [Google Scholar] [CrossRef] [PubMed]
  10. Fisher, A.B. Peroxiredoxin 6: A bifunctional enzyme with glutathione peroxidase and phospholipase A(2) activities. Antioxid. Redox Signal. 2011, 15, 831–844. [Google Scholar] [CrossRef] [PubMed]
  11. Fisher, A.B.; Dodia, C.; Chander, A.; Jain, M. A competitive inhibitor of phospholipase A2 decreases surfactant phosphatidylcholine degradation by the rat lung. Biochem. J. 1992, 288 Pt 2, 407–411. [Google Scholar] [CrossRef]
  12. Chatterjee, S.; Feinstein, S.I.; Dodia, C.; Sorokina, E.; Lien, Y.C.; Nguyen, S.; Debolt, K.; Speicher, D.; Fisher, A.B. Peroxiredoxin 6 phosphorylation and subsequent phospholipase A2 activity are required for agonist-mediated activation of NADPH oxidase in mouse pulmonary microvascular endothelium and alveolar macrophages. J. Biol. Chem. 2011, 286, 11696–11706. [Google Scholar] [CrossRef] [PubMed]
  13. Min, Y.; Wi, S.M.; Shin, D.; Chun, E.; Lee, K.Y. Peroxiredoxin-6 negatively regulates bactericidal activity and NF-kappaB activity by interrupting TRAF6-ECSIT complex. Front. Cell. Infect. Microbiol. 2017, 7, 94. [Google Scholar] [CrossRef] [PubMed]
  14. Ma, S.; Zhang, X.; Zheng, L.; Li, Z.; Zhao, X.; Lai, W.; Shen, H.; Lv, J.; Yang, G.; Wang, Q.; et al. Peroxiredoxin 6 is a crucial factor in the initial step of mitochondrial clearance and is upstream of the PINK1-Parkin pathway. Antioxid. Redox Signal. 2016, 24, 486–501. [Google Scholar] [CrossRef] [PubMed]
  15. Wattiez, R.; Noel-Georis, I.; Cruyt, C.; Broeckaert, F.; Bernard, A.; Falmagne, P. Susceptibility to oxidative stress: Proteomic analysis of bronchoalveolar lavage from ozone-sensitive and ozone-resistant strains of mice. Proteomics 2003, 3, 658–665. [Google Scholar] [CrossRef] [PubMed]
  16. Buonora, J.E.; Mousseau, M.; Jacobowitz, D.M.; Lazarus, R.C.; Yarnell, A.M.; Olsen, C.H.; Pollard, H.B.; Diaz-Arrastia, R.; Latour, L.; Mueller, G.P. Autoimmune Profiling Reveals Peroxiredoxin 6 as a Candidate Traumatic Brain Injury Biomarker. J. Neurotrauma 2015, 32, 1805–1814. [Google Scholar] [CrossRef] [PubMed]
  17. Wu, Y.; Feinstein, S.I.; Manevich, Y.; Chowdhury, I.; Pak, J.H.; Kazi, A.; Dodia, C.; Speicher, D.W.; Fisher, A.B. Mitogen-activated protein kinase-mediated phosphorylation of peroxiredoxin 6 regulates its phospholipase A(2) activity. Biochem. J. 2009, 419, 669–679. [Google Scholar] [CrossRef] [PubMed]
  18. Reddy, A.B.; Karp, N.A.; Maywood, E.S.; Sage, E.A.; Deery, M.; O’Neill, J.S.; Wong, G.K.; Chesham, J.; Odell, M.; Lilley, K.S.; et al. Circadian orchestration of the hepatic proteome. Curr. Biol. 2006, 16, 1107–1115. [Google Scholar] [CrossRef] [PubMed]
  19. Sorokina, E.M.; Feinstein, S.I.; Milovanova, T.N.; Fisher, A.B. Identification of the amino acid sequence that targets peroxiredoxin 6 to lysosome-like structures of lung epithelial cells. Am. J. Physiol. Lung Cell. Mol. Physiol. 2009, 297, 871–880. [Google Scholar] [CrossRef] [PubMed]
  20. Sorokina, E.M.; Dodia, C.; Zhou, S.; Tao, J.Q.; Gao, L.; Raabe, T.; Feinstein, S.I.; Fisher, A.B. Mutation of Serine 32 to Threonine in Peroxiredoxin 6 Preserves Its Structure and Enzymatic Function but Abolishes Its Trafficking to Lamellar Bodies. J. Biol. Chem. 2016, 291, 9268–9280. [Google Scholar] [CrossRef] [PubMed]
  21. Sorokina, E.M.; Feinstein, S.I.; Zhou, S.; Fisher, A.B. Intracellular targeting of peroxiredoxin 6 to lysosomal organelles requires MAPK activity and binding to 14-3-3epsilon. Am. J. Physiol. Cell Physiol. 2011, 300, C1430–C1441. [Google Scholar] [CrossRef] [PubMed]
  22. Kim, S.Y.; Jo, H.Y.; Kim, M.H.; Cha, Y.Y.; Choi, S.W.; Shim, J.H.; Kim, T.J.; Lee, K.Y. H2O2-dependent hyperoxidation of peroxiredoxin 6 (Prdx6) plays a role in cellular toxicity via up-regulation of iPLA2 activity. J. Biol. Chem. 2008, 283, 33563–33568. [Google Scholar] [CrossRef] [PubMed]
  23. Zhou, S.; Lien, Y.C.; Shuvaeva, T.; DeBolt, K.; Feinstein, S.I.; Fisher, A.B. Functional interaction of glutathione S-transferase pi and peroxiredoxin 6 in intact cells. Int. J. Biochem. Cell Biol. 2013, 45, 401–407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Chhunchha, B.; Kubo, E.; Fatma, N.; Singh, D.P. Sumoylation-deficient Prdx6 gains protective function by amplifying enzymatic activity and stability and escapes oxidative stress-induced aberrant Sumoylation. Cell Death Dis. 2017, 8, e2525. [Google Scholar] [CrossRef] [PubMed]
  25. Chhunchha, B.; Fatma, N.; Kubo, E.; Singh, D.P. Aberrant sumoylation signaling evoked by reactive oxygen species impairs protective function of Prdx6 by destabilization and repression of its transcription. FEBS J. 2014, 281, 3357–3381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Chhunchha, B.; Kubo, E.; Singh, P.; Singh, D.P. Sumoylation-deficient Prdx6 repairs aberrant Sumoylation-mediated Sp1 dysregulation-dependent Prdx6 repression and cell injury in aging and oxidative stress. Aging (Albany NY) 2018, 10, 2284–2315. [Google Scholar] [CrossRef] [PubMed]
  27. Manevich, Y.; Shuvaeva, T.; Dodia, C.; Kazi, A.; Feinstein, S.I.; Fisher, A.B. Binding of peroxiredoxin 6 to substrate determines differential phospholipid hydroperoxide peroxidase and phospholipase A(2) activities. Arch. Biochem. Biophys. 2009, 485, 139–149. [Google Scholar] [CrossRef] [PubMed]
  28. Lien, Y.C.; Feinstein, S.I.; Dodia, C.; Fisher, A.B. The roles of peroxidase and phospholipase A2 activities of peroxiredoxin 6 in protecting pulmonary microvascular endothelial cells against peroxidative stress. Antioxid. Redox Signal. 2012, 16, 440–451. [Google Scholar] [CrossRef] [PubMed]
  29. Li, H.; Benipal, B.; Zhou, S.; Dodia, C.; Chatterjee, S.; Tao, J.Q.; Sorokina, E.M.; Raabe, T.; Feinstein, S.I.; Fisher, A.B. Critical role of peroxiredoxin 6 in the repair of peroxidized cell membranes following oxidative stress. Free Radic. Biol. Med. 2015, 87, 356–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Fisher, A.B. Peroxiredoxin 6 in the repair of peroxidized cell membranes and cell signaling. Arch. Biochem. Biophys. 2017, 617, 68–83. [Google Scholar] [CrossRef] [PubMed]
  31. Fisher, A.B.; Dodia, C.; Chander, A. Inhibition of lung calcium-independent phospholipase A2 by surfactant protein A. Am. J. Physiol. 1994, 267, 335. [Google Scholar] [CrossRef] [PubMed]
  32. Wu, Y.Z.; Manevich, Y.; Baldwin, J.L.; Dodia, C.; Yu, K.; Feinstein, S.I.; Fisher, A.B. Interaction of surfactant protein A with peroxiredoxin 6 regulates phospholipase A2 activity. J. Biol. Chem. 2006, 281, 7515–7525. [Google Scholar] [CrossRef] [PubMed]
  33. Krishnaiah, S.Y.; Dodia, C.; Sorokina, E.M.; Li, H.; Feinstein, S.I.; Fisher, A.B. Binding sites for interaction of peroxiredoxin 6 with surfactant protein A. Biochim. Biophys. Acta 2016, 1864, 419–425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Kook, S.; Wang, P.; Young, L.R.; Schwake, M.; Saftig, P.; Weng, X.; Meng, Y.; Neculai, D.; Marks, M.S.; Gonzales, L.; et al. Impaired lysosomal integral membrane protein 2-dependent Peroxiredoxin 6 delivery to lamellar bodies accounts for altered alveolar phospholipid content in adaptor protein-3-deficient pearl mice. J. Biol. Chem. 2016, 291, 8414–8427. [Google Scholar] [CrossRef] [PubMed]
  35. Zhou, S.; Sorokina, E.M.; Harper, S.; Li, H.; Ralat, L.; Dodia, C.; Speicher, D.W.; Feinstein, S.I.; Fisher, A.B. Peroxiredoxin 6 homodimerization and heterodimerization with glutathione S-transferase pi are required for its peroxidase but not phospholipase A2 activity. Free Radic. Biol. Med. 2016, 94, 145–156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Ralat, L.A.; Misquitta, S.A.; Manevich, Y.; Fisher, A.B.; Colman, R.F. Characterization of the complex of glutathione S-transferase pi and 1-cysteine peroxiredoxin. Arch. Biochem. Biophys. 2008, 474, 109–118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Ambruso, D.R.; Ellison, M.A.; Thurman, G.W.; Leto, T.L. Peroxiredoxin 6 translocates to the plasma membrane during neutrophil activation and is required for optimal NADPH oxidase activity. Biochim. Biophys. Acta 2012, 1823, 306–315. [Google Scholar] [CrossRef] [PubMed]
  38. Kwon, J.; Wang, A.; Burke, D.J.; Boudreau, H.E.; Lekstrom, K.J.; Korzeniowska, A.; Sugamata, R.; Kim, Y.S.; Yi, L.; Ersoy, I.; et al. Peroxiredoxin 6 (Prdx6) supports NADPH oxidase1 (Nox1)-based superoxide generation and cell migration. Free Radic. Biol. Med. 2016, 96, 99–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Krishnaiah, S.Y.; Dodia, C.; Feinstein, S.I.; Fisher, A.B. p67(phox) terminates the phospholipase A(2)-derived signal for activation of NADPH oxidase (NOX2). FASEB J. 2013, 27, 2066–2073. [Google Scholar] [CrossRef] [PubMed]
  40. Wood, Z.A.; Schroder, E.; Robin Harris, J.; Poole, L.B. Structure, mechanism and regulation of peroxiredoxins. Trends Biochem. Sci. 2003, 28, 32–40. [Google Scholar] [CrossRef]
  41. Ralat, L.A.; Manevich, Y.; Fisher, A.B.; Colman, R.F. Direct evidence for the formation of a complex between 1-cysteine peroxiredoxin and glutathione S-transferase pi with activity changes in both enzymes. Biochemistry 2006, 45, 360–372. [Google Scholar] [CrossRef] [PubMed]
  42. Yagi, K. Lipid peroxides and human diseases. Chem. Phys. Lipids 1987, 45, 337–351. [Google Scholar] [CrossRef]
  43. Dexter, D.T.; Carter, C.J.; Wells, F.R.; Javoy-Agid, F.; Agid, Y.; Lees, A.; Jenner, P.; Marsden, C.D. Basal lipid peroxidation in substantia nigra is increased in Parkinson’s disease. J. Neurochem. 1989, 52, 381–389. [Google Scholar] [CrossRef] [PubMed]
  44. Halliwell, B.; Chirico, S. Lipid peroxidation: Its mechanism, measurement, and significance. Am. J. Clin. Nutr. 1993, 57, 725S. [Google Scholar] [CrossRef] [PubMed]
  45. Gallelli, C.A.; Calcagnini, S.; Romano, A.; Koczwara, J.B.; de Ceglia, M.; Dante, D.; Villani, R.; Giudetti, A.M.; Cassano, T.; Gaetani, S. Modulation of the Oxidative Stress and Lipid Peroxidation by Endocannabinoids and Their Lipid Analogues. Antioxidants 2018, 7, 93. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, Y.; Sano, M.; Shinmura, K.; Tamaki, K.; Katsumata, Y.; Matsuhashi, T.; Morizane, S.; Ito, H.; Hishiki, T.; Endo, J.; et al. 4-hydroxy-2-nonenal protects against cardiac ischemia-reperfusion injury via the Nrf2-dependent pathway. J. Mol. Cell. Cardiol. 2010, 49, 576–586. [Google Scholar] [CrossRef] [PubMed]
  47. Wang, Y.N.; Gao, L.; Wu, S.Y.; Qin, S. Low-dose 4-Hydroxy-2-Nonenal (HNE) reperfusion therapy displays cardioprotective effects in mice after myocardial infarction that are abrogated by Genipin. Med. Sci. Monit. 2018, 24, 3702–3709. [Google Scholar] [CrossRef] [PubMed]
  48. Imai, H.; Nakagawa, Y. Biological significance of phospholipid hydroperoxide glutathione peroxidase (PHGPx, GPx4) in mammalian cells. Free Radic. Biol. Med. 2003, 34, 145–169. [Google Scholar] [CrossRef]
  49. Scheerer, P.; Borchert, A.; Krauss, N.; Wessner, H.; Gerth, C.; Hohne, W.; Kuhn, H. Structural basis for catalytic activity and enzyme polymerization of phospholipid hydroperoxide glutathione peroxidase-4 (GPx4). Biochemistry 2007, 46, 9041–9049. [Google Scholar] [CrossRef] [PubMed]
  50. Ayala, A.; Munoz, M.F.; Arguelles, S. Lipid peroxidation: Production, metabolism, and signaling mechanisms of malondialdehyde and 4-hydroxy-2-nonenal. Oxid. Med. Cell. Longev. 2014, 2014, 360438. [Google Scholar] [CrossRef] [PubMed]
  51. Fisher, A.B.; Vasquez-Medina, J.P.; Dodia, C.; Sorokina, E.M.; Tao, J.Q.; Feinstein, S.I. Peroxiredoxin 6 phospholipid hydroperoxidase activity in the repair of peroxidized cell membranes. Redox Biol. 2018, 14, 41–46. [Google Scholar] [CrossRef] [PubMed]
  52. Dixon, S.J.; Lemberg, K.M.; Lamprecht, M.R.; Skouta, R.; Zaitsev, E.M.; Gleason, C.E.; Patel, D.N.; Bauer, A.J.; Cantley, A.M.; Yang, W.S.; et al. Ferroptosis: An iron-dependent form of nonapoptotic cell death. Cell 2012, 149, 1060–1072. [Google Scholar] [CrossRef] [PubMed]
  53. Yu, Y.; Xie, Y.; Cao, L.; Yang, L.; Yang, M.; Lotze, M.T.; Zeh, H.J.; Kang, R.; Tang, D. The ferroptosis inducer erastin enhances sensitivity of acute myeloid leukemia cells to chemotherapeutic agents. Mol. Cell. Oncol. 2015, 2, e1054549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Yang, W.S.; Stockwell, B.R. Ferroptosis: Death by lipid peroxidation. Trends Cell Biol. 2016, 26, 165–176. [Google Scholar] [CrossRef] [PubMed]
  55. Tang, Q.; Bai, L.; Zou, Z.; Meng, P.; Xia, Y.; Cheng, S.; Mu, S.; Zhou, J.; Wang, X.; Qin, X.; et al. Ferroptosis is newly characterized form of neuronal cell death in response to arsenite exposure. Neurotoxicology 2018, 67, 27–36. [Google Scholar] [CrossRef] [PubMed]
  56. Leavey, P.J.; Gonzalez-Aller, C.; Thurman, G.; Kleinberg, M.; Rinckel, L.; Ambruso, D.W.; Freeman, S.; Kuypers, F.A.; Ambruso, D.R. A 29-kDa protein associated with p67phox expresses both peroxiredoxin and phospholipase A2 activity and enhances superoxide anion production by a cell-free system of NADPH oxidase activity. J. Biol. Chem. 2002, 277, 45181–45187. [Google Scholar] [CrossRef] [PubMed]
  57. Ellison, M.A.; Thurman, G.W.; Ambruso, D.R. Phox activity of differentiated PLB-985 cells is enhanced, in an agonist specific manner, by the PLA2 activity of Prdx6-PLA2. Eur. J. Immunol. 2012, 42, 1609–1617. [Google Scholar] [CrossRef] [PubMed]
  58. Lee, I.; Dodia, C.; Chatterjee, S.; Zagorski, J.; Mesaros, C.; Blair, I.A.; Feinstein, S.I.; Jain, M.; Fisher, A.B. A novel nontoxic inhibitor of the activation of NADPH oxidase reduces reactive oxygen species production in mouse lung. J. Pharmacol. Exp. Ther. 2013, 345, 284–296. [Google Scholar] [CrossRef] [PubMed]
  59. Lee, I.; Dodia, C.; Chatterjee, S.; Feinstein, S.I.; Fisher, A.B. Protection against LPS-induced acute lung injury by a mechanism-based inhibitor of NADPH oxidase (type 2). Am. J. Physiol. Lung Cell. Mol. Physiol. 2014, 306, 635–644. [Google Scholar] [CrossRef] [PubMed]
  60. Vazquez-Medina, J.P.; Dodia, C.; Weng, L.; Mesaros, C.; Blair, I.A.; Feinstein, S.I.; Chatterjee, S.; Fisher, A.B. The phospholipase A2 activity of peroxiredoxin 6 modulates NADPH oxidase 2 activation via lysophosphatidic acid receptor signaling in the pulmonary endothelium and alveolar macrophages. FASEB J. 2016, 30, 2885–2898. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Chatterjee, S.; Browning, E.A.; Hong, N.; DeBolt, K.; Sorokina, E.M.; Liu, W.; Birnbaum, M.J.; Fisher, A.B. Membrane depolarization is the trigger for PI3K/Akt activation and leads to the generation of ROS. Am. J. Physiol. Heart Circ. Physiol. 2012, 302, 105–114. [Google Scholar] [CrossRef] [PubMed]
  62. Lee, I.T.; Yang, C.M. Role of NADPH oxidase/ROS in pro-inflammatory mediators-induced airway and pulmonary diseases. Biochem. Pharmacol. 2012, 84, 581–590. [Google Scholar] [CrossRef] [PubMed]
  63. Benipal, B.; Feinstein, S.I.; Chatterjee, S.; Dodia, C.; Fisher, A.B. Inhibition of the phospholipase A2 activity of peroxiredoxin 6 prevents lung damage with exposure to hyperoxia. Redox Biol. 2015, 4, 321–327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Kim, S.Y.; Chun, E.; Lee, K.Y. Phospholipase A(2) of peroxiredoxin 6 has a critical role in tumor necrosis factor-induced apoptosis. Cell Death Differ. 2011, 18, 1573–1583. [Google Scholar] [CrossRef] [PubMed]
  65. Schmitt, A.; Schmitz, W.; Hufnagel, A.; Schartl, M.; Meierjohann, S. Peroxiredoxin 6 triggers melanoma cell growth by increasing arachidonic acid-dependent lipid signalling. Biochem. J. 2015, 471, 267–279. [Google Scholar] [CrossRef] [PubMed]
  66. Kim, T.S.; Sundaresh, C.S.; Feinstein, S.I.; Dodia, C.; Skach, W.R.; Jain, M.K.; Nagase, T.; Seki, N.; Ishikawa, K.; Nomura, N.; et al. Identification of a human cDNA clone for lysosomal type Ca2+-independent phospholipase A2 and properties of the expressed protein. J. Biol. Chem. 1997, 272, 2542–2550. [Google Scholar] [CrossRef] [PubMed]
  67. Akiba, S.; Dodia, C.; Chen, X.; Fisher, A.B. Characterization of acidic Ca(2+)-independent phospholipase A2 of bovine lung. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 1998, 120, 393–404. [Google Scholar] [CrossRef]
  68. Waris, G.; Ahsan, H. Reactive oxygen species: Role in the development of cancer and various chronic conditions. J. Carcinog. 2006, 5, 14. [Google Scholar] [CrossRef] [PubMed]
  69. Rolfs, F.; Huber, M.; Gruber, F.; Bohm, F.; Pfister, H.J.; Bochkov, V.N.; Tschachler, E.; Dummer, R.; Hohl, D.; Schafer, M.; et al. Dual role of the antioxidant enzyme peroxiredoxin 6 in skin carcinogenesis. Cancer Res. 2013, 73, 3460–3469. [Google Scholar] [CrossRef] [PubMed]
  70. Park, M.H.; Jo, M.; Kim, Y.R.; Lee, C.K.; Hong, J.T. Roles of peroxiredoxins in cancer, neurodegenerative diseases and inflammatory diseases. Pharmacol. Ther. 2016, 163, 1–23. [Google Scholar] [CrossRef] [PubMed]
  71. Fleury, C.; Mignotte, B.; Vayssiere, J.L. Mitochondrial reactive oxygen species in cell death signaling. Biochimie 2002, 84, 131–141. [Google Scholar] [CrossRef]
  72. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg effect: The metabolic requirements of cell proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef] [PubMed]
  73. Walsh, B.; Pearl, A.; Suchy, S.; Tartaglio, J.; Visco, K.; Phelan, S.A. Overexpression of Prdx6 and resistance to peroxide-induced death in Hepa1-6 cells: Prdx suppression increases apoptosis. Redox Rep. 2009, 14, 275–284. [Google Scholar] [CrossRef] [PubMed]
  74. Ho, J.N.; Lee, S.B.; Lee, S.S.; Yoon, S.H.; Kang, G.Y.; Hwang, S.G.; Um, H.D. Phospholipase A2 activity of peroxiredoxin 6 promotes invasion and metastasis of lung cancer cells. Mol. Cancer Ther. 2010, 9, 825–832. [Google Scholar] [CrossRef] [PubMed]
  75. Benhar, M.; Engelberg, D.; Levitzki, A. ROS, stress-activated kinases and stress signaling in cancer. EMBO Rep. 2002, 3, 420–425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Yun, H.M.; Park, K.R.; Lee, H.P.; Lee, D.H.; Jo, M.; Shin, D.H.; Yoon, D.Y.; Han, S.B.; Hong, J.T. PRDX6 promotes lung tumor progression via its GPx and iPLA2 activities. Free Radic. Biol. Med. 2014, 69, 367–376. [Google Scholar] [CrossRef] [PubMed]
  77. Yun, H.M.; Park, K.R.; Park, M.H.; Kim, D.H.; Jo, M.R.; Kim, J.Y.; Kim, E.C.; Yoon, D.Y.; Han, S.B.; Hong, J.T. PRDX6 promotes tumor development via the JAK2/STAT3 pathway in a urethane-induced lung tumor model. Free Radic. Biol. Med. 2015, 80, 136–144. [Google Scholar] [CrossRef] [PubMed]
  78. Kwon, H.S.; Bae, Y.J.; Moon, K.A.; Lee, Y.S.; Lee, T.; Lee, K.Y.; Kim, T.B.; Park, C.S.; Moon, H.B.; Cho, Y.S. Hyperoxidized peroxiredoxins in peripheral blood mononuclear cells of asthma patients is associated with asthma severity. Life Sci. 2012, 90, 502–508. [Google Scholar] [CrossRef] [PubMed]
  79. Lee, D.H.; Park, J.H.; Han, S.B.; Yoon, D.Y.; Jung, Y.Y.; Hong, J.T. Peroxiredoxin 6 overexpression attenuates lipopolysaccharide-induced acute kidney injury. Oncotarget 2017, 8, 51096–51107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Yang, D.; Song, Y.; Wang, X.; Sun, J.; Ben, Y.; An, X.; Tong, L.; Bi, J.; Wang, X.; Bai, C. Deletion of peroxiredoxin 6 potentiates lipopolysaccharide-induced acute lung injury in mice. Crit. Care Med. 2011, 39, 756–764. [Google Scholar] [CrossRef] [PubMed]
  81. Wang, Y.; Feinstein, S.I.; Manevich, Y.; Ho, Y.S.; Fisher, A.B. Lung injury and mortality with hyperoxia are increased in peroxiredoxin 6 gene-targeted mice. Free Radic. Biol. Med. 2004, 37, 1736–1743. [Google Scholar] [CrossRef] [PubMed]
  82. Wang, Y.; Feinstein, S.I.; Manevich, Y.; Ho, Y.S.; Fisher, A.B. Peroxiredoxin 6 gene-targeted mice show increased lung injury with paraquat-induced oxidative stress. Antioxid. Redox. Signal. 2006, 8, 229–237. [Google Scholar] [CrossRef] [PubMed]
  83. Eismann, T.; Huber, N.; Shin, T.; Kuboki, S.; Galloway, E.; Wyder, M.; Edwards, M.J.; Greis, K.D.; Shertzer, H.G.; Fisher, A.B.; et al. Peroxiredoxin-6 protects against mitochondrial dysfunction and liver injury during ischemia-reperfusion in mice. Am. J. Physiol. Gastrointest. Liver Physiol. 2009, 296, 266–274. [Google Scholar] [CrossRef] [PubMed]
  84. Chhunchha, B.; Fatma, N.; Kubo, E.; Rai, P.; Singh, S.P.; Singh, D.P. Curcumin abates hypoxia-induced oxidative stress based-ER stress-mediated cell death in mouse hippocampal cells (HT22) by controlling Prdx6 and NF-kappaB regulation. Am. J. Physiol. Cell Physiol. 2013, 304, 636–655. [Google Scholar] [CrossRef] [PubMed]
  85. Tu, Q.; Xiong, Y.; Fan, L.; Qiao, B.; Xia, Z.; Hu, L.; Wang, Y.; Peng, G.; Ye, Q. Peroxiredoxin 6 attenuates ischemia and hypoxia induced liver damage of braindead donors. Mol. Med. Rep. 2016, 13, 753–761. [Google Scholar] [CrossRef] [PubMed]
  86. Pak, J.H.; Son, W.C.; Seo, S.B.; Hong, S.J.; Sohn, W.M.; Na, B.K.; Kim, T.S. Peroxiredoxin 6 expression is inversely correlated with nuclear factor-kappaB activation during Clonorchis sinensis infestation. Free Radic. Biol. Med. 2016, 99, 273–285. [Google Scholar] [CrossRef] [PubMed]
  87. Kuang, X.; Wang, L.F.; Yu, L.; Li, Y.J.; Wang, Y.N.; He, Q.; Chen, C.; Du, J.R. Ligustilide ameliorates neuroinflammation and brain injury in focal cerebral ischemia/reperfusion rats: Involvement of inhibition of TLR4/peroxiredoxin 6 signaling. Free Radic. Biol. Med. 2014, 71, 165–175. [Google Scholar] [CrossRef] [PubMed]
  88. Vazquez-Medina, J.P. Chapter 3—Redox Signaling and the Onset of the Inflammatory Cascade. Immun. Inflamm. Health Dis. 2018, 37–42. [Google Scholar] [CrossRef]
  89. Pacifici, F.; Arriga, R.; Sorice, G.P.; Capuani, B.; Scioli, M.G.; Pastore, D.; Donadel, G.; Bellia, A.; Caratelli, S.; Coppola, A.; et al. Peroxiredoxin 6, a novel player in the pathogenesis of diabetes. Diabetes 2014, 63, 3210–3220. [Google Scholar] [CrossRef] [PubMed]
  90. Kuda, O.; Brezinova, M.; Silhavy, J.; Landa, V.; Zidek, V.; Dodia, C.; Kreuchwig, F.; Vrbacky, M.; Balas, L.; Durand, T.; et al. Nrf2-Mediated Antioxidant Defense and Peroxiredoxin 6 Are Linked to Biosynthesis of Palmitic Acid Ester of 9-Hydroxystearic Acid. Diabetes 2018, 67, 1190–1199. [Google Scholar] [CrossRef] [PubMed]
  91. Wahlig, S.; Lovatt, M.; Mehta, J.S. Functional role of peroxiredoxin 6 in the eye. Free Radic. Biol. Med. 2018, 126, 210–220. [Google Scholar] [CrossRef] [PubMed]
  92. Shibata, S.; Shibata, N.; Shibata, T.; Sasaki, H.; Singh, D.P.; Kubo, E. The role of Prdx6 in the protection of cells of the crystalline lens from oxidative stress induced by UV exposure. Jpn. J. Ophthalmol. 2016, 60, 408–418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Kubo, E.; Miyazawa, T.; Fatma, N.; Akagi, Y.; Singh, D.P. Development- and age-associated expression pattern of peroxiredoxin 6, and its regulation in murine ocular lens. Mech. Ageing Dev. 2006, 127, 249–256. [Google Scholar] [CrossRef] [PubMed]
  94. Hasanova, N.; Kubo, E.; Kumamoto, Y.; Takamura, Y.; Akagi, Y. Age-related cataracts and Prdx6: Correlation between severity of lens opacity, age and the level of Prdx 6 expression. Br. J. Ophthalmol. 2009, 93, 1081–1084. [Google Scholar] [CrossRef] [PubMed]
  95. Zha, X.; Wu, G.; Zhao, X.; Zhou, L.; Zhang, H.; Li, J.; Ma, L.; Zhang, Y. PRDX6 Protects ARPE-19 Cells from oxidative damage via PI3K/AKT signaling. Cell. Physiol. Biochem. 2015, 36, 2217–2228. [Google Scholar] [CrossRef] [PubMed]
  96. Lovicu, F.J.; Schulz, M.W.; Hales, A.M.; Vincent, L.N.; Overbeek, P.A.; Chamberlain, C.G.; McAvoy, J.W. TGFβ induces morphological and molecular changes similar to human anterior subcapsular cataract. Br. J. Ophthalmol. 2002, 86, 220–226. [Google Scholar] [CrossRef] [PubMed]
  97. Manevich, Y.; Sweitzer, T.; Pak, J.H.; Feinstein, S.I.; Muzykantov, V.; Fisher, A.B. 1-Cys peroxiredoxin overexpression protects cells against phospholipid peroxidation-mediated membrane damage. Proc. Natl. Acad. Sci. USA 2002, 99, 11599–11604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Hattori, F.; Oikawa, S. Peroxiredoxins in the central nervous system. Subcell. Biochem. 2007, 44, 357–374. [Google Scholar] [PubMed]
  99. Manevich, Y.; Hutchens, S.; Halushka, P.V.; Tew, K.D.; Townsend, D.M.; Jauch, E.C.; Borg, K. Peroxiredoxin VI oxidation in cerebrospinal fluid correlates with traumatic brain injury outcome. Free Radic. Biol. Med. 2014, 72, 210–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Shichita, T.; Hasegawa, E.; Kimura, A.; Morita, R.; Sakaguchi, R.; Takada, I.; Sekiya, T.; Ooboshi, H.; Kitazono, T.; Yanagawa, T.; et al. Peroxiredoxin family proteins are key initiators of post-ischemic inflammation in the brain. Nat. Med. 2012, 18, 911–917. [Google Scholar] [CrossRef] [PubMed]
  101. Shichita, T.; Sakaguchi, R.; Suzuki, M.; Yoshimura, A. Post-ischemic inflammation in the brain. Front. Immunol. 2012, 3, 132. [Google Scholar] [CrossRef] [PubMed]
  102. Shanshan, Y.; Beibei, J.; Li, T.; Minna, G.; Shipeng, L.; Li, P.; Yong, Z. Phospholipase A2 of Peroxiredoxin 6 Plays a Critical Role in Cerebral Ischemia/Reperfusion Inflammatory Injury. Front. Cell. Neurosci. 2017, 11, 99. [Google Scholar] [CrossRef] [PubMed]
  103. Power, J.H.; Asad, S.; Chataway, T.K.; Chegini, F.; Manavis, J.; Temlett, J.A.; Jensen, P.H.; Blumbergs, P.C.; Gai, W.P. Peroxiredoxin 6 in human brain: Molecular forms, cellular distribution and association with Alzheimer’s disease pathology. Acta Neuropathol. 2008, 115, 611–622. [Google Scholar] [CrossRef] [PubMed]
  104. Yun, H.M.; Choi, D.Y.; Oh, K.W.; Hong, J.T. PRDX6 Exacerbates Dopaminergic Neurodegeneration in a MPTP Mouse Model of Parkinson’s Disease. Mol. Neurobiol. 2015, 52, 422–431. [Google Scholar] [CrossRef] [PubMed]
  105. Yeo, I.J.; Park, M.H.; Son, D.J.; Kim, J.Y.; Nam, K.T.; Hyun, B.K.; Kim, S.Y.; Jung, M.H.; Song, M.J.; Chun, H.O.; et al. PRDX6 Inhibits Neurogenesis through Downregulation of WDFY1-Mediated TLR4 Signal. Mol. Neurobiol. 2018. [Google Scholar] [CrossRef] [PubMed]
  106. Yun, H.M.; Park, K.R.; Kim, E.C.; Hong, J.T. PRDX6 controls multiple sclerosis by suppressing inflammation and blood brain barrier disruption. Oncotarget 2015, 6, 20875–20884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Lee, D.; Moawad, A.R.; Morielli, T.; Fernandez, M.C.; O’Flaherty, C. Peroxiredoxins prevent oxidative stress during human sperm capacitation. Mol. Hum. Reprod. 2017, 23, 106–115. [Google Scholar] [CrossRef] [PubMed]
  108. Fernandez, M.C.; O’Flaherty, C. Peroxiredoxin 6 is the primary antioxidant enzyme for the maintenance of viability and DNA integrity in human spermatozoa. Hum. Reprod. 2018. [Google Scholar] [CrossRef] [PubMed]
  109. Gong, S.; San Gabriel, M.C.; Zini, A.; Chan, P.; O’Flaherty, C. Low amounts and high thiol oxidation of peroxiredoxins in spermatozoa from infertile men. J. Androl. 2012, 33, 1342–1351. [Google Scholar] [CrossRef] [PubMed]
  110. Moawad, A.R.; Fernandez, M.C.; Scarlata, E.; Dodia, C.; Feinstein, S.I.; Fisher, A.B.; O’Flaherty, C. Deficiency of peroxiredoxin 6 or inhibition of its phospholipase A2 activity impair the in vitro sperm fertilizing competence in mice. Sci. Rep. 2017, 7, 12994. [Google Scholar] [CrossRef] [PubMed] [Green Version]

Share and Cite

MDPI and ACS Style

Arevalo, J.A.; Vázquez-Medina, J.P. The Role of Peroxiredoxin 6 in Cell Signaling. Antioxidants 2018, 7, 172. https://doi.org/10.3390/antiox7120172

AMA Style

Arevalo JA, Vázquez-Medina JP. The Role of Peroxiredoxin 6 in Cell Signaling. Antioxidants. 2018; 7(12):172. https://doi.org/10.3390/antiox7120172

Chicago/Turabian Style

Arevalo, José A., and José Pablo Vázquez-Medina. 2018. "The Role of Peroxiredoxin 6 in Cell Signaling" Antioxidants 7, no. 12: 172. https://doi.org/10.3390/antiox7120172

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