The compensatory antioxidant response system with a focus on neuroprogressive disorders

https://doi.org/10.1016/j.pnpbp.2019.109708Get rights and content

Highlights

  • Oxidative and nitrosative stress (O&NS) are observed in neuroprogressive disorders.

  • Inflammation also contributes to the pathophysiology of neuroprogressive disorders.

  • Several counter-regulatory mechanisms that may protect brain cells in the context of O&NS and neuroinflammation.

  • The existence of a compensatory anti-oxidant response system is proposed.

Abstract

Major antioxidant responses to increased levels of inflammatory, oxidative and nitrosative stress (ONS) are detailed. In response to increasing levels of nitric oxide, S-nitrosylation of cysteine thiol groups leads to post-transcriptional modification of many cellular proteins and thereby regulates their activity and allows cellular adaptation to increased levels of ONS. S-nitrosylation inhibits the function of nuclear factor kappa-light-chain-enhancer of activated B cells, toll-like receptor-mediated signalling and the activity of several mitogen-activated protein kinases, while activating nuclear translocation of nuclear factor (erythroid-derived 2)-like 2 (Nrf2 or NFE2L2); in turn, the redox-regulated activation of Nrf2 leads to increased levels and/or activity of key enzymes and transporter systems involved in the glutathione system. The Nrf2/Kelch-like ECH-associated protein-1 axis is associated with upregulation of NAD(P)H:quinone oxidoreductase 1, which in turn has anti-inflammatory effects. Increased Nrf2 transcriptional activity also leads to activation of haem oxygenase-1, which is associated with upregulation of bilirubin, biliverdin and biliverdin reductase as well as increased carbon monoxide signalling, anti-inflammatory and antioxidant activity. Associated transcriptional responses, which may be mediated by retrograde signalling owing to elevated hydrogen peroxide, include the unfolded protein response (UPR), mitohormesis and the mitochondrial UPR; the UPR also results from increasing levels of mitochondrial and cytosolic reactive oxygen species and reactive nitrogen species leading to nitrosylation, glutathionylation, oxidation and nitration of crucial cysteine and tyrosine causing protein misfolding and the development of endoplasmic reticulum stress. It is shown how these mechanisms co-operate in forming a co-ordinated rapid and prolonged compensatory antioxidant response system.

Introduction

Prolonged cellular oxidative and nitrosative stress (ONS), as evidenced by excessive levels of reactive oxygen species (ROS) and of reactive nitrogen species (RNS), such as nitric oxide (NO or, more strictly, NO) and peroxynitrite (ONOO), can cause profound damage to proteins, lipids and DNA resulting in structural damage, the creation of self-antigens, dysfunction of signalling pathways, disruption of the epigenetic landscape and bioenergetic failure, ultimately resulting in apoptosis or necrosis (Morris et al., 2018e, Morris et al., 2018c). Importantly, a state of chronic ONS may arise as a result of overproduction of ROS and RNS and/or failure of cellular antioxidant defences (Morris and Maes, 2014; Paladino et al., 2018) and hence the optimal performance and timely activation of the latter is an indispensable element in cellular survival and unsurprisingly has been the subject of extensive research over several decades. Such research has revealed that the cellular antioxidant system consists of an array of enzymatic and non-enzymatic effector molecules operating under the influence of several transcription factors.

Antioxidant enzymes include superoxide dismutase (SOD) and catalase (CAT), which catalyse the dismutation of superoxide radicals into hydrogen peroxide, and hydrogen peroxide into water and oxygen, respectively (Sepasi Tehrani and Moosavi-Movahedi, 2018; Younus, 2018). Other glutathione (GSH) system enzymes such as glutathione peroxidase (GPx), glutathione transferase (GST) and glutathione reductase enable the operation of the GSH system and the clearance of electrophiles, lipid peroxides and xenobiotics (Allocati et al., 2018; Cardoso et al., 2017; Morris et al., 2014). Thioredoxin reductase is another pivotal enzyme as it enables the function of the thioredoxin system and in turn the actions of the peroxidase family which are the most abundant cytosolic antioxidant enzymes (Lee et al., 2013; Rhee, 2016). The non-enzymatic components of the cellular antioxidant defences include vitamin A, uric acid, vitamin E, lipoic acid, molecular hydrogen, carbon monoxide (CO), haem oxygenase-1 (HO-1), coenzyme Q10 (CoQ10 - the only natural membrane bound antioxidant), thioredoxin and GSH (Morris et al., 2014, Morris et al., 2013; Paladino et al., 2018; Parfenova et al., 2012).

The antioxidant defences are also under transcriptional and post-transcriptional control with the most important players being the activation and nuclear translocation of nuclear factor (erythroid-derived 2)-like 2 (Nrf2 or NFE2L2), the activation of retrograde mitochondria to nuclear signalling pathways known as mitohormesis and S-nitrosylation of crucial structural and functional proteins (Miller et al., 2018; Morris et al., 2018a, Morris et al., 2018d).

The results of several published meta analyses confirm the presence of ONS in the brains and periphery of individuals suffering from major depressive disorder (MDD) (Black et al., 2015; Jiménez-Fernández et al., 2015; Liu et al., 2015). Individual authors have presented evidence of extensive lipid peroxidation as evidenced by high levels of malondialdehyde (MDA) and 4-hydroxy-2-trans-nonenal (HNE) whose levels correlate with the severity of symptoms indicating that this phenomenon has a causative role in the pathophysiology of the illness (Forlenza and Miller, 2006; Mazereeuw et al., 2015; Milaneschi et al., 2013; Yager et al., 2010). Other research teams have reported evidence of extensive DNA damage, elevated CAT and SOD levels, lowered levels of vitamin E and C and high levels of isoprostanes (Czarny et al., 2015; Forlenza and Miller, 2006; Gałecki et al., 2009; Maes et al., 2000). There is also evidence of nitrosative and oxidative mediated changes in the conformation of proteins rendering them immunogenic and hence acting as an additional source of elevated ROS and NO (Morris et al., 2015). Such evidence includes elevated levels of natural IgM autoantibodies directed at oxidatively modified epitopes and NO adducts (Maes et al., 2013, Maes et al., 2012, Maes et al., 2011). This evidence suggesting the presence of excessive levels of NO together with the presence of high levels of neuronal nitric oxide synthase (nNOS) and inducible nitric oxide, which appears to extend to all body compartments, together with evidence of widespread dysregulation in NO signalling is consistent with the existence of hypernitrosylation in these patients (Nasyrova et al., 2015; Zhou et al., 2018). It is also noteworthy that researchers have reported a positive association between higher levels of NO and NO metabolites in patients with MDD who have committed suicide compared to patients with MDD who have not (Kim et al., 2006; Lee et al., 2006; Vargas et al., 2013).

The results of meta analyses confirm the presence of ONS in the periphery and post mortem brains of individuals with bipolar disorder (BPD) (Andreazza et al., 2008; Brown et al., 2014; Gawryluk et al., 2011). The most common observations are oxidative and peroxidative damage to proteins lipids, DNA and RNA, elevation of thiobarbituric acid reactive substances (TBARS), CAT/SOD imbalance, disturbance in the GSH system and elevated levels of NO (Andreazza et al., 2008; Brown et al., 2014; Gawryluk et al., 2011; Selek et al., 2008). Importantly the severity of stress correlates with the severity and or duration of the illness and is reduced following successful therapeutic intervention via the administration of lithium or valproate suggesting that increased levels of ROS and RNS and or compromised antioxidant defence systems plays a major role in the pathogenesis and or pathophysiology of BPD (Jornada et al., 2011; Kauer-Sant'Anna et al., 2009; Selek et al., 2008; Yumru et al., 2009). Once again the high levels of nNOS in the brains and periphery of BPD patients is consistent with the presence of hypernitrosylation (Oliveira et al., 2008). There is also evidence of a positive association between levels of NO metabolites, and markers of oxidative stress and suicide risk in individuals suffering from this illness (Vargas et al., 2013).

Recent meta-analyses have confirmed the presence of ONS in patients with schizophrenia even in patients during first episode and hence treatment naïve (Flatow et al., 2013; Fraguas et al., 2017; Sarandol et al., 2015). Importantly, levels of oxidative stress correlates with severity of positive symptoms in first episode patients (Martínez-Cengotitabengoa et al., 2012) and multiple lines of evidence suggest that increased levels of ROS and RNS and or failure of cellular antioxidant defences play a major causative role in the pathogenesis and pathophysiology of the illness (Bitanihirwe and Woo, 2011; O'Donnell et al., 2014; Wood, 2009). Individual studies report evidence of lipid peroxidation, protein carbonylation, DNA damage, disturbances in the GSH system, increased SOD, reduced serum antioxidant capacity (Bitanihirwe and Woo, 2011; O'Donnell et al., 2014; Wood, 2009). These markers of ONS have been observed in the in the brain and cerebrospinal fluid (CSF) and periphery in vivo and in the brain post mortem consistent with a state of hypernitrosylation (Akyol et al., 2002; Do et al., 2000; Prabakaran et al., 2007; Wang et al., 2009). It should be noted that while meta analyses consistently acknowledge the presence of ONS in schizophrenia there is considerable variance in the results of individual studies particularly those involving long term patients probably due at least in part to the confounding effects of anti-psychotics (Bošković et al., 2011). Increased ONS also appears to be a pathophysiological hallmark of schzophrenia but there is no evidence of a positive association between the extent of both parameters and increased suicide risk in this illness (Flatow et al., 2013).

Several authors have reported decreased Nrf-2 activity and nuclear translocation in brains of individuals with MDD and BPD (Hashimoto, 2018; Martín-Hernández et al., 2018; Zhang et al., 2018). There is also some suggestion that this might be true in the case of schizophrenia (Genc and Genc, 2009). Higher levels of kynurenine 3-monooxygenase (KMO) observed in patients with MDD and BPD with a resulting increase in quinolinic acid (QUIN) appear to restrict the activity of Nrf-2 (Parrott and O'Connor, 2015). Chronically upregulated KMO secondary to the presence of neuroinflammation is also increasingly recognised as a source of self-amplifying oxidative stress and mitochondrial dysfunction (Castellano-Gonzalez et al., 2019) and hence may inhibit the activity of Nrf-2 by the oxidation of functional cysteine residues residing within DJ-1 which acts as the “master activator” of the Nrf-2 system (Morris et al., 2018a). It should be noted that there is limited evidence of increased KMO activity in schizophrenia. However, the production of kynurenic acid (KYNA) appears to far outweigh the production of QUIN in the brains of individuals suffering from this disease (Morris et al., 2016). Furthermore, the high levels of ONS seen in the brains of patients with schizophrenia might well be sufficient to inactivate DJ-1 and thus downregulate Nrf-2 activity in these individuals. In any event decreased Nrf-2 activity may be a cause downregulated or dysregulated HO-1 expression in neuroprogressive illnesses as the activity of the OH-1/Bilirubin system is regulated the kelch-like ECH-associated protein-1 (KEAP1)/Nrf-2 system (Neis et al., 2018). The effect of reduced HO-1 activity has been emphasised in MDD where the extent of such reduction correlates with symptom severity and increases in markers of oxidative stress (Robaczewska et al., 2016). Compromised Nrf-2 activity can also exacerbate ROS mediated endoplasmic reticulum (ER) stress and compromise the ER stress mediated cellular survival systems as part of the unfolded protein response (Cullinan and Diehl, 2004; Digaleh et al., 2013) and the mitochondrial unfolded protein response (Morris et al., 2018b). This phenomenon is of clinical relevance as unopposed ER stress also appears to be an element driving the pathophysiology of neuroprogressive disorders (Muneer and Shamsher Khan, 2019). For example, ER stress leading to a compromised mitochondrial UPR (mtUPR) would appear to be one of the major drivers of neuronal apoptosis in schizophrenia (Drechsel and Patel, 2010). This phenomenon is also associated with the dysfunctional pattern of mitochondrial dynamics seen in individuals with this illness (Barksdale et al., 2014; Flippo and Strack, 2017). The presence of ER stress and activation of the UPR has also been reported in patients with schizophrenia and is thought to play a major role in the pathophysiology of this illness (Bengesser et al., 2016; Pfaffenseller et al., 2014). Perhaps predictably given the presence of ER stress there is also evidence of a dysfunctional mtUPR in BPD as revealed by compromised mitochondrial dynamics and an imbalance between mitochondrial fusion and fission protein levels (Kim et al., 2017b; Scaini et al., 2017). There is limited evidence of ER stress in patients with MDD although one team of authors have reported increased levels of proteins involved in the UPR post mortem (Ii Timberlake and Dwivedi, 2019).

Finally compromised Nrf-2 activity also exerts a detrimental effect on another branch of the cellular antioxidant defence network referred to as mitohormensis (Coleman et al., 2018).

The objective of this paper is to discuss the antioxidant responses instigated by activation of Nrf2, mitohormesis and S-nitrosylation and how they co-operate in forming a co-ordinated rapid and prolonged compensatory antioxidant response system in the face of increasing levels of RNS and ROS and how they may be bolstered by selective therapeutic interventions.

Section snippets

Methods

An extensive literature search was carried out of the following databases: PubMed/MEDLINE, Scopus and Google Scholar Web of Science. The following Boolean search strategy was employed: oxidative stress OR nitrosative stress OR inflammation AND {antioxidant OR Nrf2 OR Nrf-2 OR mitohormesis OR nitrosylation OR Keap OR Keap1 OR NFkappaB OR MAPK OR TLR OR NADH:quinone oxidoreductase OR NADPH:quinone oxidoreductase OR NQO1 haem oxygenase-1 OR heme oxygenase-1 OR haem oxygenase OR heme oxygenase OR

S-nitrosylation

Increasing levels of NO which typically occur in a cellular environment with the development of nitrosative stress leads to a corresponding increase in the density of proteins with key functional thiol groups modified by the covalent addition of NO originating from dinitrogen trioxide (N2O3), via a mechanism described as S-nitrosylation (Banerjee, 2012; Hill and Bhatnagar, 2012; Paulsen and Carroll, 2010; Winterbourn and Hampton, 2008). This form of post transcriptional modification, hereafter

Nrf2 activation

Briefly, under normal unstressed conditions, Nrf2 is continually polyubiquinated by a complex of cullin-3 (Cul3), ubiquitin E3 ligase and KEAP1, with the latter molecule binding with Neh2 (Nrf2-ECH homology 2) sites within certain domains in the tertiary structure of the transcription factor (Kansanen et al., 2013; Taguchi et al., 2011). This level of polyubiquitination and subsequent proteasomal degradation effectively sequester effectively sequester Nrf2 in the cytosol (Kaspar and Jaiswal,

Upregulation of NAD(P)H:quinone oxidoreductase 1

Another enzyme involved in phase II detoxification upregulated by the Nrf2/KEAP1 axis is NAD(P)H:quinone oxidoreductase 1 (NQO1) (Roubalová et al., 2017; Yamaguchi et al., 2010). However, the anti-inflammatory effects of NQO1 are not confined to the detoxification of xenobiotics and electrophiles and this enzyme plays a number of direct roles in decreasing oxidative stress (Zhu and Li, 2012). Direct effects include the reduction of quinone residues in a plethora of different proteins and direct

Upregulation of haem oxygenase-1

Increased Nrf2 transcriptional activity also leads to the activation of haem oxygenase-1 (HO-1), another pivotal molecular entity which plays an indispensable role in cell survival in an environment of increasing inflammation and ONS (Loboda et al., 2016; Na and Surh, 2014). Activation of microsomal HO-1 occurs as an adaptive strategy in most, if not all, tissues in the face of increasing cellular stress driven by a wide range of inflammatory and oxidative stimuli. Once activated, this enzyme

Upregulation of BR, BV and BVR activity

The weight of evidence suggests that BR is a significant player in the cellular antioxidant system and exerts profound cytoprotective effects, particularly in neural tissue, even at concentrations which are an order of magnitude lower than those of GSH (Rigato et al., 2005; Sedlak et al., 2009). The effectiveness of BR as an antioxidant is thrown into stark relief by data demonstrating that an experimentally induced decrease of this substance in the brains of rodents dramatically increases

CO signalling

CO binds to cytochrome c oxidase (COX) in vivo, most notably in the haem a3 and bc1 complex (Queiroga et al., 2012). The result of CO binding to COX is a decreased rate of mitochondrial respiration with a concomitant increase in superoxide leakage from complex III as well as a denuded turnover of adenosine diphosphate (ADP) (Zuckerbraun et al., 2007). CO-mediated hydrogen peroxide production in mitochondria and subsequent egress into the cytosol oxidises functional cysteine thiol groups on a

ER stress and the UPR

Increasing levels of mitochondrial and cytosolic ROS and RNS can lead to the nitrosylation, glutathionylation, oxidation and nitration of crucial cysteine and tyrosine residues leading to protein misfolding and the development of ER stress (Li et al., 2010; Pedruzzi et al., 2004) – reviewed (Chen et al., 2018). ER stress in turn results in increasing ROS production leading to ever increasing levels of ROS and ER stress, ultimately resulting in cellular death by apoptosis or necrosis (Cao and

Underpinning mechanisms

Elevated but sub-lethal levels of mtROS induce a range of adaptive cytoprotective responses aimed at combatting the damaging effects of ROS and RNS and restoring cellular homeostasis in a process described as mitohormesis (Barbieri et al., 2013; Ristow, 2014; Yun and Finkel, 2014). Unsurprisingly, this phenomenon also involves retrograde mitochondria to nucleus signalling, probably in the form of elevated mitochondrial hydrogen peroxide levels subsequently exported into the cytosol. Elevated

Coenzyme Q10

The results of several meta-analyses and clinical trials have confirmed a significant decrease in peripheral inflammation and oxidative stress following CoQ10 administration in animals and humans (Fan et al., 2017; Sanoobar et al., 2015, Sanoobar et al., 2013; Zhai et al., 2017).

There would appear to be several mechanisms underpinning these observations such as the downregulation of NF-κB and PICs (Huo et al., 2018; Tsai et al., 2012). Several authors investigating the effects of CoQ10

Conclusion

This paper has given details of how the body mounts a number of powerful and effective countermeasures to inflammatory and ONS, including S-nitrosylation, Nrf2 activation, NQO1 upregulation, HO-1 upregulation, upregulation of BR, BV and BVR activity, CO signalling, the UPR, mitohormesis and the mtUPR. These countermeasures have been shown to act in a co-ordinated, rapid and prolonged manner. Taken together, these pathways appear to comprise a putative compensatory antioxidant defence network,

Acknowledgements

MB is supported by a NHMRC Senior Principal Research Fellowship 1059660 and 1156072.

References (362)

  • S. Bindu et al.

    Translocation of heme oxygenase-1 to mitochondria is a novel cytoprotective mechanism against non-steroidal anti-inflammatory drug-induced mitochondrial oxidative stress, apoptosis, and gastric mucosal injury

    J. Biol. Chem.

    (2015)
  • C. Biswas et al.

    Nuclear heme oxygenase-1 (HO-1) modulates subcellular distribution and activation of Nrf2, impacting metabolic and anti-oxidant defenses

    J. Biol. Chem.

    (2014)
  • B.K.Y. Bitanihirwe et al.

    Oxidative stress in schizophrenia: an integrated approach

    Neurosci. Biobehav. Rev.

    (2011)
  • C.N. Black et al.

    Is depression associated with increased oxidative stress? A systematic review and meta-analysis

    Psychoneuroendocrinology

    (2015)
  • N.C. Brown et al.

    An updated meta-analysis of oxidative stress markers in bipolar disorder

    Psychiatry Res.

    (2014)
  • C. Cantó et al.

    The NAD+ precursor nicotinamide riboside enhances oxidative metabolism and protects against high-fat diet-induced obesity

    Cell Metab.

    (2012)
  • J.F. Carr et al.

    The nuclear form of heme oxygenase (HO)-1 preferentially promotes glycolysis in hyperoxia but fails to protect against heme toxicity

    Free Radic. Biol. Med.

    (2016)
  • H.-K. Choi et al.

    Inhibition of liver fibrosis by solubilized coenzyme Q10: role of Nrf2 activation in inhibiting transforming growth factor-beta1 expression

    Toxicol. Appl. Pharmacol.

    (2009)
  • V.-A. Chouinard et al.

    Brain bioenergetics and redox state measured by 31P magnetic resonance spectroscopy in unaffected siblings of patients with psychotic disorders

    Schizophr. Res.

    (2017)
  • F.J. Cronje et al.

    Carbon monoxide actuates O2-limited heme degradation in the rat brain

    Free Radic. Biol. Med.

    (2004)
  • S.B. Cullinan et al.

    PERK-dependent activation of Nrf2 contributes to redox homeostasis and cell survival following endoplasmic reticulum stress

    J. Biol. Chem.

    (2004)
  • J.G. Dickhout et al.

    Integrated stress response modulates cellular redox state via induction of cystathionine γ-lyase

    J. Biol. Chem.

    (2012)
  • A.T. Dinkova-Kostova et al.

    The emerging role of Nrf2 in mitochondrial function

    Free Radic. Biol. Med.

    (2015)
  • D.A. Drechsel et al.

    Respiration-dependent H2O2 removal in brain mitochondria via the thioredoxin/peroxiredoxin system

    J. Biol. Chem.

    (2010)
  • B. Erol et al.

    Coenzyme Q10 treatment reduces lipid peroxidation, inducible and endothelial nitric oxide synthases, and germ cell-specific apoptosis in a rat model of testicular ischemia/reperfusion injury

    Fertil. Steril.

    (2010)
  • L. Fan et al.

    Effects of coenzyme Q10 supplementation on inflammatory markers: a systematic review and meta-analysis of randomized controlled trials

    Pharmacol. Res.

    (2017)
  • G. Filomeni et al.

    p38MAPK and ERK1/2 dictate cell death/survival response to different pro-oxidant stimuli via p53 and Nrf2 in neuroblastoma cells SH-SY5Y

    Biochem. Pharmacol.

    (2012)
  • C.J. Fiorese et al.

    The transcription factor ATF5 mediates a mammalian mitochondrial UPR

    Curr. Biol.

    (2016)
  • J. Flatow et al.

    Meta-analysis of oxidative stress in schizophrenia

    Biol. Psychiatry

    (2013)
  • K.H. Flippo et al.

    An emerging role for mitochondrial dynamics in schizophrenia

    Schizophr. Res.

    (2017)
  • M.W. Foster et al.

    Protein S-nitrosylation in health and disease: a current perspective

    Trends Mol. Med.

    (2009)
  • A.A. Fouad et al.

    Hepatoprotective effect of coenzyme Q10 in rats with acetaminophen toxicity

    Environ. Toxicol. Pharmacol.

    (2012)
  • S. Fourquet et al.

    Activation of NRF2 by nitrosative agents and H2O2 involves KEAP1 disulfide formation

    J. Biol. Chem.

    (2010)
  • M. Fulco et al.

    Glucose restriction inhibits skeletal myoblast differentiation by activating SIRT1 through AMPK-mediated regulation of Nampt

    Dev. Cell

    (2008)
  • P. Gałecki et al.

    Lipid peroxidation and antioxidant protection in patients during acute depressive episodes and in remission after fluoxetine treatment

    Pharmacol. Rep.

    (2009)
  • K. Genc et al.

    Oxidative stress and dysregulated Nrf2 activation in the pathogenesis of schizophrenia

    Biosci. Hypotheses

    (2009)
  • S.A. Abd El-Aal et al.

    CoQ10 augments rosuvastatin neuroprotective effect in a model of global ischemia via inhibition of NF-κB/JNK3/Bax and activation of Akt/FOXO3A/Bim cues

    Front. Pharmacol.

    (2017)
  • A. Agarwal et al.

    Adaptive responses to tissue injury: role of heme oxygenase-1

    Trans. Am. Clin. Climatol. Assoc.

    (2013)
  • A.M. Aldbass et al.

    Protective and therapeutic potency of N-acetyl-cysteine on propionic acid-induced biochemical autistic features in rats

    J. Neuroinflammation

    (2013)
  • N. Allocati et al.

    Glutathione transferases: substrates, inihibitors and pro-drugs in cancer and neurodegenerative diseases

    Oncogenesis

    (2018)
  • E.J. Anderson et al.

    Mitochondria as a source and target of lipid peroxidation products in healthy and diseased heart

    Clin. Exp. Pharmacol. Physiol.

    (2012)
  • K. Aquilano et al.

    p53 orchestrates the PGC-1α-mediated antioxidant response upon mild redox and metabolic imbalance

    Antioxid. Redox Signal.

    (2013)
  • T. Arnould et al.

    Mitochondria retrograde signaling and the UPRmt: where are we in mammals?

    Int. J. Mol. Sci.

    (2015)
  • R. Banerjee

    Redox outside the box: linking extracellular redox remodeling with intracellular redox metabolism

    J. Biol. Chem.

    (2012)
  • A. Banning et al.

    The GI-GPx gene is a target for Nrf2

    Mol. Cell. Biol.

    (2005)
  • S. Bansal et al.

    Mitochondria-targeted heme oxygenase-1 induces oxidative stress and mitochondrial dysfunction in macrophages, kidney fibroblasts and in chronic alcohol hepatotoxicity

    Redox Biol.

    (2014)
  • D.E. Baranano et al.

    Biliverdin reductase: a major physiologic cytoprotectant

    Proc. Natl. Acad. Sci. U. S. A.

    (2002)
  • E. Barbieri et al.

    Mitohormesis in muscle cells: a morphological, molecular, and proteomic approach

    Muscles Ligaments Tendons J.

    (2013)
  • K.A. Barksdale et al.

    Synaptic proteins in the postmortem anterior cingulate cortex in schizophrenia: relationship to treatment and treatment response

    Neuropsychopharmacology

    (2014)
  • E. Barone et al.

    Oxidative and Nitrosative modifications of biliverdin reductase-a in the brain of subjects with Alzheimer's disease and amnestic mild cognitive impairment

    J. Alzheimers Dis.

    (2011)
  • Cited by (20)

    • Intertwined associations between oxidative and nitrosative stress and endocannabinoid system pathways: Relevance for neuropsychiatric disorders

      2022, Progress in Neuro-Psychopharmacology and Biological Psychiatry
      Citation Excerpt :

      It is also noteworthy that several other research teams have reported direct NFκB inhibition following increased activity of CB2 (Correa et al., 2010; Niu et al., 2017; Zhu et al., 2019). There have also been reports of Nrf2 activation following activation of CB2 (Jayarajan et al., 2019; Omar et al., 2009), which is interesting in the light of established cross talk between the Nrf2 and NFκB systems (Morris et al., 2019b). The importance of NFκB in the activation of NOX2 and NOX4 is well documented and has already been noted above.

    • Polyphenols as adjunctive treatments in psychiatric and neurodegenerative disorders: Efficacy, mechanisms of action, and factors influencing inter-individual response

      2021, Free Radical Biology and Medicine
      Citation Excerpt :

      This mechanism may also explain, at least in part, dose-dependent decreases in NF-κB activity following quercetin supplementation which is associated with reduced IKKβ phosphorylation [494,495]. Readers interested in the role of NF-κB in cellular survival are invited to consult the work of [496] and those interested in the role of Nrf-2 as the master regulator of the cellular antioxidant responses are invited to consult the work of [497]. Lastly, the o-quinone derivative also activates mitochondrial-mediated apoptosis by stimulating release of cytochrome c into the cytoplasm and the activation of caspase 3, caspase 9 and the activation of PARP-1 [498].

    View all citing articles on Scopus
    View full text