1,961
Views
9
CrossRef citations to date
0
Altmetric
Research Article

The role of modulation of antioxidant enzyme systems in the treatment of neurodegenerative diseases

, , , &
Pages 194-204 | Received 19 May 2016, Accepted 20 Jun 2016, Published online: 07 Jul 2016

Abstract

Oxidative stress is a much-appreciated phenomenon associated with the progression of neurodegenerative diseases (NDDs) due to imbalances in redox homeostasis. The poor correlations between the in vitro benefits and clinical trials of direct radical scavengers have prompted research into indirect antioxidant enzymes such as Nrf2. Activation of Nrf2 leads to the upregulation of a myriad of cytoprotective and antioxidant enzymes/proteins. Traditionally, early Nrf2-activators were studied as chemoprotective agents. There is a consequential lack of clinical trials testing Nrf2 activation in NDDs. However, there is abundant evidence of their utility in pre-clinical studies. Herein, we review the endogenous Nrf2 regulatory pathway and avenues for targeting this pathway. Furthermore, we provide updated information on pre-clinical studies for natural and synthetic Nrf2 activators. On the basis of our findings, we posit that successful therapeutics for NDDs rely on the design of potent synthetic Nrf2 activators with a careful combination of other neuroprotective activities.

Introduction

To date, there has been a myriad of theories mapping the molecular mechanisms that underlie the pathology of NDDs. Mitochondrial dysfunction, oxidative stress, chronic inflammation and accumulation of protein aggregates have been identified in unison as pathological factors associated with a broad spectrum of NDDs such as Alzheimer’s disease (AD), Parkinson’s disease (PD), Huntington’s disease (HD) and amyotrophic lateral sclerosis (ALS)Citation1,Citation2. Although mitochondrial dysfunction has been implicated as a common feature among these heterogeneous diseasesCitation3–6, the role of oxidative stress has been identified as a major hallmark and the key mitochondrial insult consistently linked to the development of NDDsCitation7. An imbalance between reactive oxygen species (ROS) production and the detoxification of reactive intermediates is associated with oxidative stressCitation2 and subsequent attack of free radical species on neuronal cells.

Having failed in clinical trials, free radical scavengers have been explored extensively as means to reduce the severity of ROS onslaught on neuronal cells. Transcription factor NF-E2-related factor 2 (Nrf2) is generally believed to be an adaptive cellular response to endogenous and environmental stressCitation8. Thus its emerging role in oxidative stress is that of a regulatory function against oxidantsCitation9. Activated by a variety of electrophiles and oxidants, Nrf2 is a basic leucine zipper/Cap ‘n’ collar protein, which regulates antioxidant response elements (ARE) containing genesCitation10,Citation11. The pharmacological activation and genetic up-regulation of Nrf2 has been shown to be fundamentally neuroprotective in a number of experimental modelsCitation12.

In relevant Nrf2 knock out mice, the deletion of the Nrf2 gene in mice resulted in neurodegeneration, widespread astrogliosisCitation13, increased sensitivity to neuroinflammation, severe dopaminergic dysfunctionCitation14, lack of ARE activationCitation15 and show the neuroprotective function of Nrf2. Nrf2 knockout mice displayed increased vulnerability to common neurotoxinsCitation16,Citation17 and the subsequent transplantation of Nrf2-overexpressing astrocytes protected against malonateCitation18 and 6-OHDACitation16 insults. Further to these reports, a number of natural and synthetic compounds have been shown to serve as Nrf2 activators and of benefit in several in vivo models. In this review, we posit that the induction of Nrf2 is an exciting therapeutic target and the design of synthetics targeting this pathway may provide effective tratments in NDDs.

ROS/RNS-mediated oxidative stress

Based on its high oxygen consumption and the relatively reduced regeneration of neuronal cells, the human brain is predisposed to oxidative stressCitation19. There is a paradoxical influence of oxygen due to its rather toxic nature via the accumulation of reactive oxygen species (ROS), reactive nitrogen species (RNS) and the disturbance of pro-oxidant equilibriumCitation20,Citation21. In neurodegeneration, reduced levels of antioxidant enzymes along with increased ROS/RNS in specific areas of the brain have been shown to damage several key macromolecules such as DNA, proteins and lipidsCitation22–26.

Reactive oxygen species (ROS) are oxygen-derived radicals such as superoxide (O•-2), hydroxyl (HO), alkoxyl (RO) and peroxyl (ROO), as well as non-radicals such as hydrogen peroxide (H2O2) and hypochlorous acid (HOCl)Citation27,Citation28. This list of reactive species can be expanded to include nitric oxide (NO), nitrogen dioxide (NO2) and the more recently recognized peroxynitrite (ONOO), which is classified as reactive nitrogen species (RNS)Citation29. The production of oxidants following specific physiological cues serves as important signalling molecules in immune function, during redox signalling, in autophagy and cell divisionCitation9,Citation30. In neurodegeneration, the uncontrolled production of ROS/RNS impairs neuronal function and contributes to neuronal deathCitation31 particularly in the brain, where ROS-mediated redox signalling is actively involved in neurogenesis, neuronal differentiation, plasticity and memory consolidationCitation32. Under physiological conditions, ROS production and elimination is tightly regulated under “redox homeostasis” with the maintenance of low ROS concentrationCitation33. The tight regulation of ROS production via redox homeostasis and the disturbance thereof causes overproduction of ROS and deficiency in enzymatic and non-enzymatic antioxidants. The generation of ROS/RNS in the brain is mainly due to premature “electron leaks” in the electron transport chain of the mitochondriaCitation34 and the production of superoxide during energy transduction in the mitochondrial has been linked with ageing and mitochondrial damageCitation35.

The mitochondrion is described as a dynamic organelle, playing a crucial role in promoting the intact regulation of cellular function by controlling energy metabolism, mitochondrial shape/number as well as its DNA contentCitation1. It has been shown to be of great importance in energy production and cell physiology and impairment thereof has been implicated in a number of NDDsCitation36–39. Dysfunctional mitochondria result in the reduction in the cellular supply of energy and failure in the maintenance of cellular homeostasis with the activation of cellular death pathwaysCitation40. The pathology of several other NDDs displays a vast spectrum of mitochondrial dysfunction including respiratory chain dysfunction and oxidative stress, reduced ATP production, calcium dysregulation, mitochondrial permeability transition pore opening, perturbation in mitochondrial dynamics and deregulated mitochondrial clearanceCitation41. PD has received the greatest attention in relation to mitochondrial dysfunction and mitochondrial ROS production with reductions in complex I activity as demonstrated in the substantia nigra, lymphocytes and plateletsCitation41. Typically, the overproduction of amyloid beta (Aβ) in the postmortem brains of AD patients was indicated to induce imbalances in mitochondrial fission and fusion processesCitation42,Citation43. Finally, imbalances in mitochondrial dynamics were also observed in ALS modelsCitation44,Citation45.

Several enzymes in the brain have been shown to contribute to the production of ROS/RNS. As part of physiological function, NADPH oxidase (NOX)Citation46, monoamine oxidase B (MAO-B)Citation47, nitric oxidase synthase (NOS)Citation48 and xanthine oxidase (XO)Citation49,Citation50 generate ROS/RNS. Typically, NOX-mediated production of ROS has been shown to play a critical role in neuronal apoptosisCitation51 with evidence of microglial and neuronal NOX involvement in NDDsCitation52,Citation53. Furthermore, the expression of monoamine oxidase (MAO) in patients with AD was reported to be significantly higher than in the healthy controls of the same ageCitation54. MAO-B elevation in murine brain astrocytes resulted in several hallmarks of PD pathology such as decreased mitochondrial complex I activity, increased mitochondrial oxidative stress, activation of microglia in substantia nigra and finally progressive loss of dopaminergic neuronsCitation55.

Amidst the failure of several antioxidants with free radical scavenging roles in human trials, the call for targeting innate cellular protective mechanisms have drawn much attention to the Nrf2 pathway and its activators. As the master regulator of cellular response to oxidative stress, Nrf2 appeared as a promising target as modulation of this target could provide an alternative to the drawbacks associated with exogenous antioxidant therapy. Therefore, Nrf2 activating compounds have been explored as a means of inducing these endogenous antioxidant enzymes. This review thus summarizes the current information available on these compounds highlighting the need to delve into synthetics with improved bioavailability as Nrf2 modulators for the treatment of NDDs.

Nrf2/ARE pathway

The transcriptional response to reactive chemical stress is via the induction of a number of cytoprotective enzymes, which is mediated by the cis-acting Antioxidant-Response-Element (ARE)Citation56. Nrf2 was identified as transcriptional factor acting on the ARE with in vivo studies implicating severely impaired ARE-dependent genes in knockout (Nrf−/−) miceCitation10,Citation57. Nrf2 is known to function independently and dependently of Keap-1 to control the inducible and constitutive expression of ARE.

Keap1-dependent regulation of Nrf2

When first discovered, Keap1 was identified as a cytoplasmic protein that binds Nrf2 at the N-terminal Neh2 domain ()Citation58. Afterwards, the Keap1 dimerisation via its BTB was found to be required for Nrf2 repression and its ETGE motif in the Neh2 domain identified as critical for binding Keap 1Citation59. At this point, Keap1 was viewed simply as a passive repressor of Nrf2 by virtue of sequestering it in cytoplasm and, therefore, preventing its nuclear translocation. This notion was supported by the fact that Nrf2 constitutively accumulates in the nuclei of Keap1-knockout miceCitation60. Therefore, the first plausible model of Nrf2 activation comprised the disruption of the Keap1–Nrf2 complex resulting in nuclear translocation of Nrf2.

Figure 1. The domains of Nrf2 and Keap1. Both Nrf2 and Keap1 are highly conserved between species. In Nrf2 six homology domains, Neh1–6, were identified. The domains of interest in Nrf2 are Neh1, containing DNA-binding basic leucine zipper domain; Neh2, containing Keap1-binding redox-sensitive degron; and Neh6, containing redox-insensitive phosphodegron. The domains of interest in Keap1 are the Broad complex, Tramtrack and Bric-a-brac (BTB), the intervening linker region (IVR or simply linker) and Kelch repeats domain (Kelch). Keap1 homodimerizes via its BTB domain and binds to Nrf2 via its Kelch domain. Electrophilic inducers react with Cys 151 in the BTB domain abolishing Keap1 repressor activity. The linker region contains a number of cysteine residues critical to constitutive Keap1 repressor activity.

Figure 1. The domains of Nrf2 and Keap1. Both Nrf2 and Keap1 are highly conserved between species. In Nrf2 six homology domains, Neh1–6, were identified. The domains of interest in Nrf2 are Neh1, containing DNA-binding basic leucine zipper domain; Neh2, containing Keap1-binding redox-sensitive degron; and Neh6, containing redox-insensitive phosphodegron. The domains of interest in Keap1 are the Broad complex, Tramtrack and Bric-a-brac (BTB), the intervening linker region (IVR or simply linker) and Kelch repeats domain (Kelch). Keap1 homodimerizes via its BTB domain and binds to Nrf2 via its Kelch domain. Electrophilic inducers react with Cys 151 in the BTB domain abolishing Keap1 repressor activity. The linker region contains a number of cysteine residues critical to constitutive Keap1 repressor activity.

Figure 2. Modes of Nrf2 activation. Under unstressed conditions Nrf2 is constantly recruited for ubiquitination by Keap1–Cul3–Rbx complex resulting in low constitutive level of activity. Electrophilic compounds or ROS covalently bind to reactive cysteines interfering with the normal function of the complex and resulting in accumulation of Nrf2. Nrf2 then translocates into nucleus and drives transcription of antioxidant/detoxifying enzymes. PPI inhibitors achieve the same goal by preventing Nrf2 binding to Keap1.

Figure 2. Modes of Nrf2 activation. Under unstressed conditions Nrf2 is constantly recruited for ubiquitination by Keap1–Cul3–Rbx complex resulting in low constitutive level of activity. Electrophilic compounds or ROS covalently bind to reactive cysteines interfering with the normal function of the complex and resulting in accumulation of Nrf2. Nrf2 then translocates into nucleus and drives transcription of antioxidant/detoxifying enzymes. PPI inhibitors achieve the same goal by preventing Nrf2 binding to Keap1.

Indeed, treatment with electrophilic inducers such as sulphoraphaneCitation61, diethyl maleateCitation62 and β-naphthoflavoneCitation63 caused nuclear accumulation of Nrf2 and transcription of its target genesCitation62. However, a number of electrophilic chemicals were shown to be unable to disrupt the Nrf2–Keap1 complex, suggesting that the Keap1–Nrf2 dissociation model is incorrectCitation64–66. Furthermore, the accumulation of Nrf2 requires its de novo synthesisCitation62, but the classic electrophilic inducers do not affect the rate of transcription of Nrf2Citation63,Citation67,Citation68. Indeed, Nrf2 was identified as a substrate of the ubiquitin–proteasome pathwayCitation69,Citation70. Furthermore, the binding of Keap1 to Nrf2 was found to actively target Nrf2 for proteasomal degradation at a higher rate than unbound Nrf2, suggesting a more active role of Keap1 in the repression of Nrf2 activityCitation61,Citation71. This was a breaking point, as soon after Keap1 was found to function as an adaptor bridging Nrf2 and Cul3–Rbx1 E3 ligase and targeting Nrf2 for ubiquitinationCitation64,72–74.

The Nrf2 activation by electrophilic inducers has been thought to result from direct alkylation or oxidation of reactive cysteine residues on Keap1 resulting in dissociation of the Keap1–Nrf2 complex. Indeed, a number of reactive cysteine residues have been identifiedCitation71,Citation75. Particularly, Keap1–C151S mutant was insensitive to tert-butylhydroquinone (tBHQ) and SFN treatments and continued to repress Nrf2 activity under both basal and induced conditions, while both Keap1–C273S and Keap1–C288S mutants completely lost the ability to repress Nrf2 under any conditionsCitation71. The critical role of C273 and C288 for constitutive repression of Nrf2 was later confirmed in vivoCitation76. Keap1 was reported to have a zinc atom coordinated to cysteine thiols, possibly in the linker region, making them more reactiveCitation77. This prompted another model, where electrophilic inducers replaced zinc and resulted in conformational change and release of Nrf2Citation77. Later, C151 was identified as the most reactive target towards Michael acceptorsCitation78. However, the activation of Nrf2 by arsenic was found to be independent of C151Citation79. In view of reports of multiple reactive cysteines, it was proposed that Keap1–Nrf2–Cul3 complex constitutes a multiple-sensing mechanismCitation80. The reactivity of Keap1 cysteine residues with various inducers has been reviewedCitation81.

Although the complete crystal structure of the Keap1–Nrf2–Cul3 complex has not been resolved yet, some individual domains and their interactions have been revealed. The crystal structure of the Kelch motif was identified as a six β-propeller structure resembling a “doughnut” shapeCitation82. In addition to the high affinity ETGE motif, a secondary low affinity DLG motif was identified in the Neh2 domainCitation83. Soon after, the crystal structures of Kelch motif bound to ETGE and DLG motifs revealed their binding at the centre groove of Kelch motifCitation84,Citation85. Keap1 was found to exist as a homodimer via its BTB motif in the Cul3 ligase complex, which led to the widely accepted “hinge and latch” modelCitation86,Citation87 According to this model, Nrf2 binds the Keap1 homodimer via its ETGE and DLG motifs positioning the middle lysine rich alpha helix of the Neh2 domain for ubiquitination by Cul3 ligase complex. This model explains the faster rates of Keap1-mediated ubiquitination and degradation of Nrf2 under unstressed conditions. In support of this model, a low-resolution Keap1 dimer structure revealed two large spheres attached by short linker arms to the sides of a small forked-stem structure, resembling a “cherry-bob”Citation88. The crystal structure of BTB domain of Keap1 has recently been elucidated, revealing a number of basic amino acids in close proximity of Cys-151, explaining its highly reactive natureCitation89. The authors further revealed the structure of bardoxolone, a potent Keap1 antagonist, covalently bound to Cys151. They further demonstrated that this interaction was sufficient to prevent Cul3 binding, and thus inhibit the proteasomal degradation of Nrf2. The latest structural findings on the interactions within the Nrf2–Keap1–Cul3 complex have been reviewed in detail by Canning and BullockCitation90.

Keap1-independent regulation of Nrf2

Various canonical kinases have long since been implicated in the regulation of Nrf2Citation63,91–93. Activation of PKC directly phosphorylates Nrf2 at Ser-40 and leads to its translocation into nucleusCitation93–95. Subsequently, this phosphorylation at Ser-40 was shown to be necessary for its dissociation from Keap1, but not required for Nrf2 accumulation in the nucleusCitation96.

GSK-3β was later identified as a common downstream effector of PI3K/AktCitation97,Citation98, PKCCitation99 and MAPK/ERKCitation100. Additionally, glycogen synthase kinase-3 beta (GSK-3β) was shown to negatively regulate Nrf2 by reducing its nuclear localisationCitation14,Citation97 and acting to phosphorylate a number of Src subfamily kinases, which translocate into the nucleus and phosphorylate Nrf2 leading to its nuclear exportCitation101,Citation102. All these results suggest that GSK-3β could be a potential target. The utility of GSK-3β/Nrf2 regulatory pathway was proven in vivo by the synergistic use of a GSK-3β inhibitor and an Nrf2 activator to protect the mice brain against a neurotoxinCitation103.

More recently another regulatory pathway was reported. The Neh6 domain of Nrf2 contains a cluster of serine residues that are phosphorylated by GSK-3Citation104,Citation105. The β-transducin repeat containing protein (β-TrCP) adapter protein recognizes phosphorylated Nrf2 and targets it for ubiquitination by SKP1–CUL1–F-box (SCF), a Cul1 E3 ubiquitin ligaseCitation104. The inhibition of GSK-3 in vivo and in vitro was shown to lead to activation of Nrf2, possibly via disruption of β-TrCP–Nrf2Citation105,Citation106. Within the Neh6 domain, two short β-TrCP recognition motifs have been identified, of which one requires phosphorylation by GSK-3Citation107. These findings prompted an updated “dual degradation” model of Nrf2 regulation, which would explain the varying levels of Nrf2 activity under stressed and unstressed conditionsCitation104. The increased Nrf2 activity caused by downregulation of the two repressors Keap1 and β-TrCP has been correlated to longevity in miceCitation108.

Nrf2 activators

The ability of Nrf2 activators to mediate a global antioxidant response presents a highly attractive strategy to counteract chronic oxidative stress at the cellular level. With the myriad of in vitro and in vivo models showcasing its regulatory role and impact in neuroprotection, natural and synthetic Nrf2 activators have been reported with varying modes of Nrf2 activation (). These activators function as indirect inhibitors of the Keap1–Nrf2 complex by interacting with the cysteine sulphydryl groups. In theory, these compounds can promote an innate control of genes encoding a vast array of antioxidants and proteins overcoming drawbacks associated with exogenous antioxidants.

Natural Nrf2 activators

A number of natural Nrf2 activators () have been reported yet, despite strong rationale and evidence backing their use, their development has been met with a number of hurdles amidst their poor bioavailability, solubility issues, high cost of human trials and stability problems.

Figure 3. Structures of natural of Nrf2-activators.

Figure 3. Structures of natural of Nrf2-activators.

The natural polyphenolic drug curcumin is a compound extracted from the turmeric plant with a distinct structure, in that it possesses two methoxyphenol groups linked to a β-diketone bridgeCitation109. This confers both Michael acceptor and metal chelation propertiesCitation110. Curcumin has been known as an antioxidant for over a decade and has thus been evaluated in NDDs for its role as a free radical scavenger of the superoxide and hydroxyl ions.

Recently, its selective role on the Nrf2–Keap1 ARE pathway has been the focus of study. In rotenone-treated mice, an Nrf2-specific short pin RNA or phosphoinositide 3-kinase inhibitor demonstrated a marked attenuation in tyrosine hydroxylase and glutathioneCitation111. As a result, the neuroprotective effects of curcumin against oxidative damage were eliminated. In a likewise manner, resveratrol was used as a polyphenol and examined for its neuroprotective behaviourCitation112. Resveratrol was shown in this study to function as an Nrf2 activator. Furthermore, the authors indicated that due to the depletion of Nrf2 in astrocytes, there was a reduction in neuroprotection. In Nrf2−/− astrocytes, a decreased mitochondrial antioxidant expression and an inability to up-regulate cellular antioxidants following preconditioned treatment were also recorded. There is much interest in utilizing curcumin and resveratrol for AD therapy due to their influence on amyloid beta, antioxidant and anti-inflammatory propertiesCitation113,Citation114. On the molecular level, as a Michael acceptor, the electrophilic α,β-unsaturated carbonyl groups of curcumin is believed to selectively interact with cysteines on Keap1 enabling the release of Nrf2 for actionCitation115.

Several clinical trials proceeded to evaluate the efficacy of curcumin administration in various pathological forms of NDDs but these results were not encouraging. Typical examples are the trials coded NCT00164749 and ACT00099710 for ADCitation116,Citation117. In the first study, on one hand, patients with AD were evaluated using a combination of curcumin and ginko in the treatment of AD (NCT00164749). Results indicated no differences between the treatment groups in Aβ levels and mini-mental stage examination (MMSE) scores. On the other hand, patients with mild to moderate AD were treated with curcumin C3 complex (ACT00099710). This study also indicated no difference between treatment groups and although the doses were tolerated (2 g and 4 g daily), the authors reported low bioavailability. A number of trials have been conducted evaluating reservatrol in the treatment of AD (NCT01504854 and NCT00678431). The study, NCT01504854, reported no difference between groups when AD duration was measured in years with no alteration in results following a post-hoc re-analysisCitation118. The diverse off-target effects of reservatrol were discussed which only engaged the central molecular target in the nM range. This raises concerns generally about the lack of specificity of the natural activators, thus calling for the design of synthetics with potent activity and reduced off-target effects.

Another natural drug of interest is puerarin, which is an antioxidant of kudzu roots, which was previously evaluated and reported to possess neuroprotective role. Li et al.Citation119 recently showed that puerarin treatment of rat substantia nigra neurons attenuated 6-OHDA caused dopaminergic neuronal degeneration, precipitated as a result of oxidative stress. In this study, glutathione and catalase levels in the substantia nigra were gradually elevated and Keap1 Nrf-2 was shown to be progressively elevated following treatment. The authors suggested that puerarin activated the nuclear translocation within neurons via the Nrf2/ARE pathway which then triggered the synthesis of antioxidant enzymes to mediate ROS attack/lesions.

Gastrodin, on the contrary, was evaluated for its neuroprotective effects in PD models. In an MPTP mice model of PD, gastrodin pretreatment was shown to forestall neurotoxicityCitation120. Induced oxidative stress was prevented and gastrodin treatment in MPTP intoxicated mice resulted in a robust increase in heme oxidase 1, superoxide dismutase, glutathione levels as well as the nuclear translocation of Nrf2 with the implication of the ERK1/2 phosphorylation pathway.

The flavonoid, baicalein, was also tested in a 6-OHDA neurotoxic model and shown to display neuroprotection in vitroCitation121. The flavonoid was useful only as a prophylactic, as it failed to exhibit restoration after 6-OHDA induced cellular damage. Treatment with baicalein showed elevated Nrf2/hemo oxygenase and reduced Keap1 expression in a time and dose–response manner with the activation of the PKCα and PI3K/AKT pathway.

Sulphoraphane (SFN) is an isothiocyanate and sulphoxythiocarbamate that has been recognized for its antioxidant properties. It was evaluated in several in vivo models of AD, PD and cell models of oxidative and nitrosative stress. Previously, SFN and allyl disulphide were shown to enhance glutathione and suppress the loss of neurons in both parkin and α-synuclein Drosophila modelsCitation12. This raised the possibility of their use as therapeutics in PD. The potency of SFN in stimulating the Nrf2/ARE pathway has been shown using behavioural tests that rely on the Y-mazeCitation122. Regrettably, they were found to be of no use on Aβ aggregation, a principal pathophysiological characteristic of neuronal dysfunction in ADCitation122. Interestingly, SFN protected neurotoxicity following 6-hydroxydopamine treatments in PD with its neuroprotective effect linked to its modulatory role on neuronal pathways and enhancement on glutathione-S-transferase and glutathione reductaseCitation123. Conversely, using allicin (diallyl thiosulphide) Aβ-memory impairment was ameliorated and activated Nrf2 hippocampal level in aging mice was observedCitation124,Citation125. Allicin has been shown to rescue tunicamycin-induced neuronal stress following pretreatmentCitation126. In this study, allicin acted by moderately elevating RNA-dependent protein kinase (PRK)-like ER-resident kinase (PERK) and its downstream substrate, Nrf2.

Although earlier epidemiological studies revealed reduced risk between tobacco user and coffee drinkers, Trinh et al.Citation127 demonstrated that caffeine and nicotine were not neuroprotective in Drosophila strains as measured by the number of dopamine neurons in 20-day old transgenic flies. Rather, the results revealed that decaffeinated coffee and nicotine-free tobacco provided neuroprotection in the α-synuclein and parkin Drosophila model. In this model, 0.15% decaffeinated coffee provided more neuroprotection than 0.03% nicotine tobacco; the converse was observed in the parkin PD model. The Nrf2 activating component was identified as cafestol in coffee and results implicated its neuroprotective effects with the up-regulation of glutathione. The serum cholesterol elevation effects of cafestol in this work highlight the need to design structural synthetics that lacked this side effect. The Nrf-2 activating compounds in coffee and tobacco were presumed capable of crossing the blood–brain barrier and thus highly attractive therapeutic agents.

Synthetic Nrf2 activators

Synthetic oleanane triterpenoids are the most studied synthetic activators of Nrf2 (). They have been shown to possess anti-inflammatory, cytoprotective, antiproliferative and pro-apoptotic properties making them potential candidates for the treatment of a wide range of diseasesCitation128.

Figure 4. Structures of the synthetic oleanane triterpenoids.

Figure 4. Structures of the synthetic oleanane triterpenoids.

Initially, 2-cyano-3,12-dioxooleana-1,9(11)-dien-28-oic acid (CDDO) was developed as a highly potent anti-inflammatory agent based on the structure of oleanolic acidCitation129,Citation130. However, the poor oral bioavailability of CDDO led to the development of many other derivatives. The details of their development and structure-activity relationship studies have been recently reviewed by Sporn et al.Citation131. Of note are the two Michael acceptor enone motifs in the rings that are critical for their anti-inflammatory activityCitation132. Mechanistically, CDDO and its derivatives are highly potent Nrf2 activatorsCitation77,Citation133 and NFκB pathway inhibitorsCitation134,Citation135.

CDDO imidazolide (CDDO-Im) induced Nrf2-dependent antioxidant genes, attenuated LPS-induced proinflammatory cytokine expression and decreased mortality in miceCitation136. CDDO-methylamide (CDDO-MA) was able to cross the blood–brain barrier and significantly improved spatial memory retention and reduced plaque burden, Aβ42 levels, microgliosis, and oxidative stress in Aβ-mice model of ADCitation137. In another study, CDDO-MA demonstrated significant neuroprotection against MPTP- and 3-nitropropionic acid (3-NP)-induced nigrostriatal dopaminergic neurodegeneration in mice and ratsCitation138. Oral administration of either CDDO ethyl amide (CDDO-EA) or CDDO trifluoroethyl amide (CDDO-TFEA) blocked MPTP-induced dopaminergic neurotoxicity in miceCitation139. Both CDDO-EA and CDDO-TFEA penetrated blood–brain barrier, enhanced motor performance and extended the survival of G93A SOD1 transgenic ALS miceCitation140. When orally administered to N171-82Q Huntington transgenic HD mice, both CDDO-EA and CDDO-TFEA upregulated Nrf2/ARE-induced genes in the brain and peripheral tissues, reduced oxidative stress, improved motor impairment and survivalCitation141. CDDO methyl (CDDO-Me or bardoxolone methyl) administration to mice following ischemia–reperfusion injury upregulated Nrf2 and HO-1, decreased infarct volume and improved neurological symptomsCitation142. As these results indicate, without a doubt, CDDO derivatives are highly potent activators of Nrf2 and inducers of its target genes both in vitro and in vivo.

To date, there are no clinical trials testing synthetic oleanane triterpenoids for the treatment of NDDs. However, there are a number of clinical trials that tested bardoxolone methyl (CDDO-Me) for the treatment of cancer and chronic kidney disease (CKD). Although bardoxolone methyl was found to be generally safe in earlier trials, a phase 3 trial (NCT01351675) evaluating its use for the treatment of chronic kidney disease was terminated in 2012 due to increased rates of heart-related adverse eventsCitation143. There is a concern, however, of whether late stage CKD was the right clinical indication to pursueCitation144. It is feasible that drugs with improved specificity and better toxicity profile could prove beneficial in exploiting the protective and preventive functions of Nrf2 pathway in treating patients with early stages of degenerative diseases instead.

Dimethyl fumarate (DMF) is an orally bioavailable chemically simple compound with a Michael acceptor functionality. In two randomized, double blind, placebo-controlled phase 3 studies (NCT00420212 and NCT00451451), dimethyl fumarate (BG-12) significantly reduced the proportion of patients who had a relapseCitation145,Citation146. The United States Food and Drug Administration (FDA) and the European Medicines Agency (EMA) have subsequently approved Dimethyl fumarate in 2013 and 2014, respectively, for adult patients with relapsing-remitting MS. Although its mechanism of action in humans is not fully clear, there is a plenty of animal and cell-based reports demonstrating the involvement of Nrf2 activationCitation147–149.

DMF protected rat neural progenitor cells and motor neurons against H2O2-induced oxidative stressCitation149. Both DMF and its primary metabolite, monomethyl fumarate, significantly protected neuronal and astrocyte cultures against H2O2-induced toxicity in an Nrf2-dependent mannerCitation147,Citation150. In two HD mice models, the treatment with DMF preserved the motor function and the morphology of neurons in the striatum and the motor cortexCitation151.

Carbon monoxide (CO), the product of HO-1 metabolism, mediates potent anti-inflammatory and anti-apoptotic effectsCitation152,Citation153. For this reason, a number of synthetic hybrids () combining Nrf2-activating and CO-releasing functionalities, named “HYCOs”, have been developedCitation154,Citation155.

Figure 5. Structures of Nrf2-activating carbon monoxide synthetic hybrids.

Figure 5. Structures of Nrf2-activating carbon monoxide synthetic hybrids.

The Nrf2-activating component is the well-known Michael acceptor enone motif, while the CO liberating component is a cobalt carbonyl complex. Interestingly, CO is only liberated when cobalt is oxidized under conditions of oxidative stress. HYCOs () markedly increased Nrf2/HO-1 expression, liberated CO and exerted anti-inflammatory activity in vitro and in vivo. At least in vitro, HYCO-1 reduces inflammation and NO production more effectively than dimethyl fumarate, suggesting the synergistic/additive effect of the two functions.

Chalcones are natural compounds abundant in edible plants and possess anti-inflammatory and anti-cancer activities among other biological activitiesCitation156,Citation157. The chalcone scaffold is a well-known Michael acceptor having enone functionality sandwiched between two aromatic rings. Notably, the α,β-double bond and the carbonyl group are critical for the induction of HO-1Citation158,Citation159.

Structure–activity relationship studies around the chalcone scaffold yielded 1 () as a potent activator of Nrf2 in vitro and in miceCitation160. Recently, its heterocyclic analogue 2 demonstrated improved aqueous solubility and oral bioavailability, which translated into higher activity in vivoCitation161. Synthetic chalcones 3a and 3b at 10 μM were non-toxic and upregulated g-Glutamylcysteine Ligase Catalytic Subunit (GCLC) and haeme oxygenase-1 (HO-1) up to 3- and 25-folds, respectively, and attenuated H2O2-induced apoptosis of in vitroCitation150. E-α-(4-methoxyphenyl)-2′,3,4,4′-tetramethoxychalconeCitation4 at 30 μM caused translocation of Nrf2 into nucleus, potently induced HO-1 and downregulated NFκB activity with no observed toxicity at higher concentrationsCitation151.

Figure 6. Structures of chalcone scaffolds with Nrf2 activity.

Figure 6. Structures of chalcone scaffolds with Nrf2 activity.

The high electrophilicity of traditional Nrf2-activators poses the risk of “off-target” effects due to their ability to react with the cysteines of other proteins. Therefore, there has been an increased interest in the direct targeting of Nrf2/Keap1 protein–protein interaction (PPI) as an alternative strategy (recently reviewed by WellsCitation162 and Abed et al.Citation163). High-throughput screening using fluorescence polarisation (FP) assay yielded several micromolar hitsCitation164,Citation165. Remarkably, the structure-based design by Jiang et al.Citation166 led to the discovery of the first nanomolar inhibitor of Keap1–Nrf2 PPI in FP assay. However, all these compounds have only micromolar activities in cell-based assays. Very recently a fragment-based approach yielded KI-696 (), a highly active Nrf2 activator in vitro and in vivoCitation167. Targeting Keap1–Nrf2 PPI is certainly feasible and we have yet to see the full potential of this approach.

Figure 7. Chemical structure of KI-696.

Figure 7. Chemical structure of KI-696.

Conclusion and perspective

Following the failure of radical scavengers in NDDs, recent efforts have been aimed at the identification of alternative targets of which the master regulator Nrf2 took the centre stage. The activation of Nrf2 by any of the discussed approaches in this review can up-regulate key antioxidant/detoxifying enzymes and restore the redox homeostasis within a diseased neuron.

In this review, we highlight a number of reported Nrf2-activators with the emphasis on pre-clinical evidence. The most active and well-studied activators possess an electrophilic Michael acceptor motif that covalently binds to reactive cysteines in Keap1 disrupting its Nrf2-repressor activity. This property alone is not sufficient, as such molecules could possess off-target effects as discussed. A logical approach would be to design a molecule around the Michael acceptor motif to couple high reactivity and selectivity obtained from its crystal structure information. However, this approach is hindered by the lack of complete crystal structures of Nrf2 and its repressor proteins. Nonetheless, we are starting to see first rational designs come to light based on partial crystal structures. In the last couple of years, PPI inhibitors are gaining momentum and the latest reports show very promising lead compounds. We are seeing an increased number of rational drug design approaches that could lead to novel synthetic molecules.

The utility of Nrf2 as a therapeutic target has been clearly established in animal models of neurodegeneration. Traditionally, clinical trials evaluating natural Nrf2 activators, with the exception of curcumin, have focused on other diseases such as cancer. There is a striking lack of clinical trials evaluating these compounds for treatment of NDDs. However, clinical trials utilizing curcumin in both PD and AD demonstrated limited efficacy, likely due to its known poor bioavailability. DMF on the contrary, has been recently approved for the treatment of MS and its beneficial action is at least in part due to the activation of Nrf2Citation168. Conversely bardoxolone was pulled out of a human trial due to its cardiovascular risk further stressing the need to design highly selective Nrf2 activators.

We posit that the utility of Nrf2-activators may lack the ability to fully restore neuronal function but may be useful for preserving surviving neurons. Therefore, testing Nrf2 activation in early stage patients could prove more relevant to this pathway. It is logical to conclude that even a marginal improvement in late-stage patients could imply a substantial benefit in the form of delayed disease progression of neurodegeneration.

In view of the poor bioavailability and lack of specificity of the natural activators, the development of highly potent rationally designed synthetic Nrf2 activators could serve as promising leads in the treatment of NDDs. A highly potent and selective Nrf2 activator could be of high value to address some of the hallmarks of neurodegeneration, such as oxidative stress. However, a more complete and effective treatment may require a careful combination of Nrf2 activating component with other neuroprotective action(s) to better address the complexity of neurodegeneration.

Declaration of interest

The authors report that they have no conflict of interest.

References

  • Denzer I, Munch G, Friedland K. Modulation of mitochondrial dysfunction in neurodegenerative diseases via activation of nuclear factor erythroid-2-related factor 2 by food-derived compounds. Pharmacol Res 2016;103:80–94
  • Chen X, Guo C, Kong J. Oxidative stress in neurodegenerative diseases. Neural Regen Res 2012;7:376–85
  • Moreira PI, Cardoso SM, Santos MS, Oliveira CR. The key role of mitochondria in Alzheimer's disease. J Alzheimers Dis 2006;9:101–10
  • Nunomura A, Perry G, Aliev G, et al. Oxidative damage is the earliest event in Alzheimer disease. J Neuropathol Exp Neurol 2001;60:759–67
  • Henchcliffe C, Beal MF. Mitochondrial biology and oxidative stress in Parkinson disease pathogenesis. Nat Clin Pract Neurol 2008;4:600–9
  • Bueler H. Impaired mitochondrial dynamics and function in the pathogenesis of Parkinson's disease. Exp Neurol 2009;218:235–46
  • Lin MT, Beal MF. Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature 2006;443:787–95
  • Joshi G, Johnson JA. The Nrf2-ARE pathway: a valuable therapeutic target for the treatment of neurodegenerative diseases. Recent Pat CNS Drug Discov 2012;7:218–29
  • Ma Q. Role of nrf2 in oxidative stress and toxicity. Annu Rev Pharmacol Toxicol 2013;53:401–26
  • McMahon M, Itoh K, Yamamoto M, et al. The Cap‘n‘Collar basic leucine zipper transcription factor Nrf2 (NF-E2 p45-related factor 2) controls both constitutive and inducible expression of intestinal detoxification and glutathione biosynthetic enzymes. Cancer Res 2001;61:3299–307
  • Alam J, Stewart D, Touchard C, et al. Nrf2, a Cap‘n’Collar transcription factor, regulates induction of the heme oxygenase-1 gene. J Biol Chem 1999;274:26071–8
  • Barone MC, Sykiotis GP, Bohmann D. Genetic activation of Nrf2 signaling is sufficient to ameliorate neurodegenerative phenotypes in a Drosophila model of Parkinson's disease. Dis Model Mech 2011;4:701–7
  • Hubbs AF, Benkovic SA, Miller DB, et al. Vacuolar leukoencephalopathy with widespread astrogliosis in mice lacking transcription factor Nrf2. Am J Pathol 2007;170:2068–76
  • Rojo AI, Innamorato NG, Martin-Moreno AM, et al. Nrf2 regulates microglial dynamics and neuroinflammation in experimental Parkinson's disease. GLIA 2010;58:588–98
  • Kraft AD, Johnson DA, Johnson JA. Nuclear factor E2-related factor 2-dependent antioxidant response element activation by tert-butylhydroquinone and sulforaphane occurring preferentially in astrocytes conditions neurons against oxidative insult. J Neurosci 2004;24:1101–12
  • Jakel RJ, Townsend JA, Kraft AD, Johnson JA. Nrf2-mediated protection against 6-hydroxydopamine. Brain Res 2007;1144:192–201
  • Chen PC, Vargas MR, Pani AK, et al. Nrf2-mediated neuroprotection in the MPTP mouse model of Parkinson's disease: critical role for the astrocyte. Proc Natl Acad Sci USA 2009;106:2933–8
  • Calkins MJ, Jakel RJ, Johnson DA, et al. Protection from mitochondrial complex II inhibition in vitro and in vivo by Nrf2-mediated transcription. Proc Natl Acad Sci USA 2005;102:244–9
  • Facecchia K, Fochesato LA, Ray SD, et al. Oxidative toxicity in neurodegenerative diseases: role of mitochondrial dysfunction and therapeutic strategies. J Toxicol 2011;2011:683728
  • Sehitoglu MH, Han H, Kalin P, et al. Pistachio (Pistacia vera L.) gum: a potent inhibitor of reactive oxygen species. J Enzyme Inhib Med Chem 2015;30:264–9
  • Topal F, Nar M, Gocer H, et al. Antioxidant activity of taxifolin: an activity–structure relationship. J Enzyme Inhib Med Chem 2016;31:674–83
  • Good PF, Hsu A, Werner P, et al. Protein nitration in Parkinson's disease. J Neuropathol Exp Neurol 1998;57:338–42
  • Hensley K, Maidt ML, Yu Z, et al. Electrochemical analysis of protein nitrotyrosine and dityrosine in the Alzheimer brain indicates region-specific accumulation. J Neurosci 1998;18:8126–32
  • Dexter DT, Carter CJ, Wells FR, et al. Basal lipid peroxidation in substantia nigra is increased in Parkinson's disease. J Neurochem 1989;52:381–9
  • Giasson BI, Duda JE, Murray IV, et al. Oxidative damage linked to neurodegeneration by selective alpha-synuclein nitration in synucleinopathy lesions. Science 2000;290:985–9
  • Horiguchi T, Uryu K, Giasson BI, et al. Nitration of tau protein is linked to neurodegeneration in tauopathies. Am J Pathol 2003;163:1021–31
  • Burdon RH, Rice-Evans C. Free radicals and the regulation of mammalian cell proliferation. Free Radic Res Commun 1989;6:345–58
  • Burdon RH. Superoxide and hydrogen peroxide in relation to mammalian cell proliferation. Free Radic Biol Med 1995;18:775–94
  • Gulcin I. Antioxidant activity of food constituents: an overview. Arch Toxicol 2012;86:345–91
  • Zuo L, Zhou T, Pannell BK, et al. Biological and physiological role of reactive oxygen species-the good, the bad and the ugly. Acta Physiol (Oxf) 2015;214:329–48
  • Anderson G, Maes M. Neurodegeneration in Parkinson's disease: interactions of oxidative stress, tryptophan catabolites and depression with mitochondria and sirtuins. Mol Neurobiol 2014;49:771–83
  • Borquez DA, Urrutia PJ, Wilson C, et al. Dissecting the role of redox signaling in neuronal development. J Neurochem 2016;137:506–17
  • Wang K, Zhang T, Dong Q, et al. Redox homeostasis: the linchpin in stem cell self-renewal and differentiation. Cell Death Dis 2013;4:e537
  • MP M. How mitochondria produce reactive oxygen species. Biochemisrty 2009;417:1–13
  • Cui H, Kong Y, Zhang H. Oxidative stress, mitochondrial dysfunction, and aging. J Signal Transduct 2012;2012:646354
  • Muller WE, Eckert A, Kurz C, et al. Mitochondrial dysfunction: common final pathway in brain aging and Alzheimer's disease-therapeutic aspects. Mol Neurobiol 2010;41:159–71
  • Ryan BJ, Hoek S, Fon EA, Wade-Martins R. Mitochondrial dysfunction and mitophagy in Parkinson's: from familial to sporadic disease. Trends Biochem Sci 2015;40:200–10
  • Reddy TT, Kano A, Maruyama A, et al. Synthesis and characterization of semi-interpenetrating polymer networks based on polyurethane and N-isopropylacrylamide for wound dressing. J Biomed Mater Res Part B: Appl Biomater 2009;88:32–40
  • Grabacka MM, Gawin M, Pierzchalska M. Phytochemical modulators of mitochondria: the search for chemopreventive agents and supportive therapeutics. Pharmaceuticals (Basel) 2014;7:913–42
  • Banerjee R, Starkov AA, Beal MF, Thomas B. Mitochondrial dysfunction in the limelight of Parkinson's disease pathogenesis. Biochim Biophys Acta 2009;1792:651–63
  • Gandhi S, Abramov AY. Mechanism of oxidative stress in neurodegeneration. Oxid Med Cell Longev 2012;2012:428010
  • Wang X, Su B, Siedlak SL, et al. Amyloid-beta overproduction causes abnormal mitochondrial dynamics via differential modulation of mitochondrial fission/fusion proteins. Proc Natl Acad Sci USA 2008;105:19318–23
  • Wang X, Su B, Lee HG, et al. Impaired balance of mitochondrial fission and fusion in Alzheimer's disease. J Neurosci 2009;29:9090–103
  • Jiang Z, Wang W, Perry G, et al. Mitochondrial dynamic abnormalities in amyotrophic lateral sclerosis. Transl Neurodegener 2015;4:14
  • Santos D, Esteves AR, Silva DF, et al. The impact of mitochondrial fusion and fission modulation in sporadic Parkinson's disease. Mol Neurobiol 2015;52:573–86
  • Bedard K, Krause KH. The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev 2007;87:245–313
  • Edmondson DE. Hydrogen peroxide produced by mitochondrial monoamine oxidase catalysis: biological implications. Curr Pharm Des 2014;20:155–60
  • Alderton WK, Cooper CE, Knowles RG. Nitric oxide synthases: structure, function and inhibition. Biochem J 2001;357:593–615
  • Shahraki A, Stone TW. Blockade of presynaptic adenosine A1 receptor responses by nitric oxide and superoxide in rat hippocampus. Eur J Neurosci 2004;20:719–28
  • Ukeda DK H, Maeda S, Sawamura M. Spectrophotometric assay for superoxide dismutase based on the reduction of highly water-soluble tetrazolium salts by xanthine–xanthine oxidase. Biosci Biotechnol Biochem 1999;63:485–8
  • Tammariello SP, Quinn MT, Estus S. NADPH oxidase contributes directly to oxidative stress and apoptosis in nerve growth factor-deprived sympathetic neurons. J Neurosci 2000;20:RC53
  • Hernandes MS, Britto LR. NADPH oxidase and neurodegeneration. Curr Neuropharmacol 2012;10:321–7
  • Sorce S, Krause KH. NOX enzymes in the central nervous system: from signaling to disease. Antioxid Redox Signal 2009;11:2481–504
  • Emilsson L, Saetre P, Balciuniene J, et al. Increased monoamine oxidase messenger RNA expression levels in frontal cortex of Alzheimer's disease patients. Neurosci Lett 2002;326:56–60
  • Mallajosyula JK, Kaur D, Chinta SJ, et al. MAO-B elevation in mouse brain astrocytes results in Parkinson's pathology. PLoS One 2008;3:e1616
  • Nguyen T, Nioi P, Pickett CB. The Nrf2-antioxidant response element signaling pathway and its activation by oxidative stress. J Biol Chem 2009;284:13291–5
  • Itoh K, Chiba T, Takahashi S, 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;236:313–22
  • Itoh K, Wakabayashi N, Katoh Y, et al. Keap1 represses nuclear activation of antioxidant responsive elements by Nrf2 through binding to the amino-terminal Neh2 domain. Genes Dev 1999;13:76–86
  • Zipper LM, Mulcahy RT. The Keap1 BTB/POZ dimerization function is required to sequester Nrf2 in cytoplasm. J Biol Chem 2002;277:36544–52
  • Wakabayashi N, Itoh K, Wakabayashi J, et al. Keap1-null mutation leads to postnatal lethality due to constitutive Nrf2 activation. Nat Genet 2003;35:238–45
  • McMahon M, Itoh K, Yamamoto M, Hayes JD. Keap1-dependent proteasomal degradation of transcription factor Nrf2 contributes to the negative regulation of antioxidant response element-driven gene expression. J Biol Chem 2003;278:21592–600
  • Itoh K, Wakabayashi N, Katoh Y, et al. Keap1 regulates both cytoplasmic-nuclear shuttling and degradation of Nrf2 in response to electrophiles. Genes Cells 2003;8:379–91
  • Nguyen T, Sherratt PJ, Huang HC, et al. Increased protein stability as a mechanism that enhances Nrf2-mediated transcriptional activation of the antioxidant response element. Degradation of Nrf2 by the 26 S proteasome. J Biol Chem 2003;278:4536–41
  • Zhang DD, Lo SC, Cross JV, et al. Keap1 is a redox-regulated substrate adaptor protein for a Cul3-dependent ubiquitin ligase complex. Mol Cell Biol 2004;24:10941–53
  • Eggler AL, Liu G, Pezzuto JM, et al. Modifying specific cysteines of the electrophile-sensing human Keap1 protein is insufficient to disrupt binding to the Nrf2 domain Neh2. Proc Natl Acad Sci USA 2005;102:10070–5
  • Zhang DD. Mechanistic studies of the Nrf2-Keap1 signaling pathway. Drug Metab Rev 2006;38:769–89
  • Ishii T, Itoh K, Takahashi S, et al. Transcription factor Nrf2 coordinately regulates a group of oxidative stress-inducible genes in macrophages. J Biol Chem 2000;275:16023–9
  • Wild AC, Moinova HR, Mulcahy RT. Regulation of gamma-glutamylcysteine synthetase subunit gene expression by the transcription factor Nrf2. J Biol Chem 1999;274:33627–36
  • Sekhar KR, Yan XX, Freeman ML. Nrf2 degradation by the ubiquitin proteasome pathway is inhibited by KIAA0132, the human homolog to INrf2. Oncogene 2002;21:6829–34
  • Stewart D, Killeen E, Naquin R, et al. Degradation of transcription factor Nrf2 via the ubiquitin-proteasome pathway and stabilization by cadmium. J Biol Chem 2003;278:2396–402
  • Zhang DD, Hannink M. Distinct cysteine residues in Keap1 are required for Keap1-dependent ubiquitination of Nrf2 and for stabilization of Nrf2 by chemopreventive agents and oxidative stress. Mol Cell Biol 2003;23:8137–51
  • Kobayashi A, Kang MI, Okawa H, et al. Oxidative stress sensor Keap1 functions as an adaptor for Cul3-based E3 ligase to regulate proteasomal degradation of Nrf2. Mol Cell Biol 2004;24:7130–9
  • Cullinan SB, Gordan JD, Jin J, et al. The Keap1–BTB protein is an adaptor that bridges Nrf2 to a Cul3-based E3 ligase: oxidative stress sensing by a Cul3–Keap1 ligase. Mol Cell Biol 2004;24:8477–86
  • Furukawa M, Xiong Y. BTB protein Keap1 targets antioxidant transcription factor Nrf2 for ubiquitination by the Cullin 3-Roc1 ligase. Mol Cell Biol 2005;25:162–71
  • Dinkova-Kostova AT, Holtzclaw WD, Cole RN, et al. Direct evidence that sulfhydryl groups of Keap1 are the sensors regulating induction of phase 2 enzymes that protect against carcinogens and oxidants. Proc Natl Acad Sci USA 2002;99:11908–13
  • Yamamoto T, Suzuki T, Kobayashi A, et al. Physiological significance of reactive cysteine residues of Keap1 in determining Nrf2 activity. Mol Cell Biol 2008;28:2758–70
  • Dinkova-Kostova AT, Holtzclaw WD, Wakabayashi N. Keap1, the sensor for electrophiles and oxidants that regulates the phase 2 response, is a zinc metalloprotein. Biochemistry 2005;44:6889–99
  • Luo Y, Eggler AL, Liu D, et al. Sites of alkylation of human Keap1 by natural chemoprevention agents. J Am Soc Mass Spectrom 2007;18:2226–32
  • Wang XJ, Sun Z, Chen W, et al. Activation of Nrf2 by arsenite and monomethylarsonous acid is independent of Keap1–C151: enhanced Keap1–Cul3 interaction. Toxicol Appl Pharmacol 2008;230:383–9
  • Kobayashi M, Li L, Iwamoto N, et al. The antioxidant defense system Keap1–Nrf2 comprises a multiple sensing mechanism for responding to a wide range of chemical compounds. Mol Cell Biol 2009;29:493–502
  • Sekhar KR, Rachakonda G, Freeman ML. Cysteine-based regulation of the CUL3 adaptor protein Keap1. Toxicol Appl Pharmacol 2010;244:21–6
  • Li X, Zhang D, Hannink M, Beamer LJ. Crystal structure of the Kelch domain of human Keap1. J Biol Chem 2004;279:54750–8
  • Katoh Y, Iida K, Kang MI, et al. Evolutionary conserved N-terminal domain of Nrf2 is essential for the Keap1-mediated degradation of the protein by proteasome. Arch Biochem Biophys 2005;433:342–50
  • Lo SC, Li X, Henzl MT, et al. Structure of the Keap1: Nrf2 interface provides mechanistic insight into Nrf2 signaling. EMBO J 2006;25:3605–17
  • Padmanabhan B, Tong KI, Ohta T, et al. Structural basis for defects of Keap1 activity provoked by its point mutations in lung cancer. Mol Cell 2006;21:689–700
  • McMahon M, Thomas N, Itoh K, et al. Dimerization of substrate adaptors can facilitate cullin-mediated ubiquitylation of proteins by a “el for the Nrf2-Keap1 complex” mechanism: a two-site interaction model for the Nrf2-Keap1. J Biol Chem 2006;281:24756–68
  • Tong KI, Kobayashi A, Katsuoka F, Yamamoto M. Two-site substrate recognition model for the Keap1–Nrf2 system: a hinge and latch mechanism. Biol Chem 2006;387:1311–20
  • Ogura T, Tong KI, Mio K, et al. Keap1 is a forked-stem dimer structure with two large spheres enclosing the intervening, double glycine repeat, and C-terminal domains. Proc Natl Acad Sci USA 2010;107:2842–7
  • Cleasby A, Yon J, Day PJ, et al. Structure of the BTB domain of Keap1 and its interaction with the triterpenoid antagonist CDDO. PLoS One 2014;9:e98896
  • Canning P, Sorrell FJ, Bullock AN. Structural basis of Keap1 interactions with Nrf2. Free Radic Biol Med 2015;88:101–7
  • Yu R, Lei W, Mandlekar S, et al. Role of a mitogen-activated protein kinase pathway in the induction of phase II detoxifying enzymes by chemicals. J Biol Chem 1999;274:27545–52
  • Lee JM, Hanson JM, Chu WA, Johnson JA. 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;276:20011–16
  • Huang HC, Nguyen T, Pickett CB. Regulation of the antioxidant response element by protein kinase C-mediated phosphorylation of NF-E2-related factor 2. Proc Natl Acad Sci USA 2000;97:12475–80
  • Huang HC, Nguyen T, Pickett CB. Phosphorylation of Nrf2 at Ser-40 by protein kinase C regulates antioxidant response element-mediated transcription. J Biol Chem 2002;277:42769–74
  • Numazawa S, Ishikawa M, Yoshida A, et al. Atypical protein kinase C mediates activation of NF-E2-related factor 2 in response to oxidative stress. Am J Physiol Cell Physiol 2003;285:C334–42
  • Bloom DA, Jaiswal AK. Phosphorylation of Nrf2 at Ser40 by protein kinase C in response to antioxidants leads to the release of Nrf2 from INrf2, but is not required for Nrf2 stabilization/accumulation in the nucleus and transcriptional activation of antioxidant response element-mediated NAD(P)H: quinone oxidoreductase-1 gene expression. J Biol Chem 2003;278:44675–82
  • Salazar M, Rojo AI, Velasco D, et al. Glycogen synthase kinase-3beta inhibits the xenobiotic and antioxidant cell response by direct phosphorylation and nuclear exclusion of the transcription factor Nrf2. J Biol Chem 2006;281:14841–51
  • Li L, Dong H, Song E, et al. Nrf2/ARE pathway activation, HO-1 and NQO1 induction by polychlorinated biphenyl quinone is associated with reactive oxygen species and PI3K/AKT signaling. Chem Biol Interact 2014;209:56–67
  • Espada S, Rojo AI, Salinas M, Cuadrado A. The muscarinic M1 receptor activates Nrf2 through a signaling cascade that involves protein kinase C and inhibition of GSK-3beta: connecting neurotransmission with neuroprotection. J Neurochem 2009;110:1107–19
  • Tan Y, Ichikawa T, Li J, et al. Diabetic downregulation of Nrf2 activity via ERK contributes to oxidative stress-induced insulin resistance in cardiac cells in vitro and in vivo. Diabetes 2011;60:625–33
  • Jain AK, Jaiswal AK. GSK-3beta acts upstream of Fyn kinase in regulation of nuclear export and degradation of NF-E2 related factor 2. J Biol Chem 2007;282:16502–10
  • Niture SK, Jain AK, Shelton PM, Jaiswal AK. Src subfamily kinases regulate nuclear export and degradation of transcription factor Nrf2 to switch off Nrf2-mediated antioxidant activation of cytoprotective gene expression. J Biol Chem 2011;286:28821–32
  • Rojo AI, Sagarra MR, Cuadrado A. GSK-3beta down-regulates the transcription factor Nrf2 after oxidant damage: relevance to exposure of neuronal cells to oxidative stress. J Neurochem 2008;105:192–202
  • Rada P, Rojo AI, Chowdhry S, et al. SCF/{beta}-TrCP promotes glycogen synthase kinase 3-dependent degradation of the Nrf2 transcription factor in a Keap1-independent manner. Mol Cell Biol 2011;31:1121–33
  • Rada P, Rojo AI, Evrard-Todeschi N, et al. Structural and functional characterization of Nrf2 degradation by the glycogen synthase kinase 3/beta-TrCP axis. Mol Cell Biol 2012;32:3486
  • Rojo AI, Medina-Campos ON, Rada P, et al. Signaling pathways activated by the phytochemical nordihydroguaiaretic acid contribute to a Keap1-independent regulation of Nrf2 stability: role of glycogen synthase kinase-3. Free Radic Biol Med 2012;52:473–87
  • Chowdhry S, Zhang Y, McMahon M, et al. Nrf2 is controlled by two distinct beta-TrCP recognition motifs in its Neh6 domain, one of which can be modulated by GSK-3 activity. Oncogene 2013;32:3765–81
  • Lewis KN, Wason E, Edrey YH, et al. Regulation of Nrf2 signaling and longevity in naturally long-lived rodents. Proc Natl Acad Sci USA 2015;112:3722–7
  • Lee WH, Loo CY, Bebawy M, et al. Curcumin and its derivatives: their application in neuropharmacology and neuroscience in the 21st century. Curr Neuropharmacol 2013;11:338–78
  • Ak T, Gulcin I. Antioxidant and radical scavenging properties of curcumin. Chem Biol Interact 2008;174:27–37
  • Cui Q, Li X, Zhu H. Curcumin ameliorates dopaminergic neuronal oxidative damage via activation of the Akt/Nrf2 pathway. Mol Med Rep 2016;13:1381–8
  • Narayanan SV, Dave KR, Saul I, Perez-Pinzon MA. Resveratrol preconditioning protects against cerebral ischemic injury via nuclear erythroid 2-related factor 2. Stroke 2015;46:1626–32
  • Braidy N, Jugder BE, Poljak A, et al. Resveratrol as a potential therapeutic candidate for the treatment and management of Alzheimer's disease. Curr Top Med Chem 2016;16:1951–60
  • Ringman JM, Frautschy SA, Cole GM, et al. A potential role of the curry spice curcumin in Alzheimer's disease. Curr Alzheimer Res 2005;2:131–6
  • Dinkova-Kostova AT, Talalay P. Relation of structure of curcumin analogs to their potencies as inducers of Phase 2 detoxification enzymes. Carcinogenesis 1999;20:911–14
  • Baum L, Lam CW, Cheung SK, et al. Six-month randomized, placebo-controlled, double-blind, pilot clinical trial of curcumin in patients with Alzheimer disease. J Clin Psychopharmacol 2008;28:110–13
  • Ringman JM, Frautschy SA, Teng E, et al. Oral curcumin for Alzheimer's disease: tolerability and efficacy in a 24-week randomized, double blind, placebo-controlled study. Alzheimers Res Ther 2012;4:43
  • Turner RS, Thomas RG, Craft S, et al. A randomized, double-blind, placebo-controlled trial of resveratrol for Alzheimer disease. Neurology 2015;85:1383–91
  • Li R, Zheng N, Liang T, et al. Puerarin attenuates neuronal degeneration and blocks oxidative stress to elicit a neuroprotective effect on substantia nigra injury in 6-OHDA-lesioned rats. Brain Res 2013;1517:28–35
  • Wang XL, Xing GH, Hong B, et al. Gastrodin prevents motor deficits and oxidative stress in the MPTP mouse model of Parkinson's disease: involvement of ERK1/2-Nrf2 signaling pathway. Life Sci 2014;114:77–85
  • Kim S, Lee HG, Park SA, et al. Keap1 cysteine 288 as a potential target for diallyl trisulfide-induced Nrf2 activation. PLoS One 2014;9:e85984
  • Kim HV, Kim HY, Ehrlich HY, et al. Amelioration of Alzheimer's disease by neuroprotective effect of sulforaphane in animal model. Amyloid 2013;20:7–12
  • Morroni F, Tarozzi A, Sita G, et al. Neuroprotective effect of sulforaphane in 6-hydroxydopamine-lesioned mouse model of Parkinson's disease. Neurotoxicology 2013;36:63–71
  • Li XH, Li CY, Xiang ZG, et al. Allicin can reduce neuronal death and ameliorate the spatial memory impairment in Alzheimer's disease models. Neurosciences (Riyadh) 2010;15:237–43
  • Li XH, Li CY, Lu JM, et al. Allicin ameliorates cognitive deficits ageing-induced learning and memory deficits through enhancing of Nrf2 antioxidant signaling pathways. Neurosci Lett 2012;514:46–50
  • Zhu YF, Li XH, Yuan ZP, et al. Allicin improves endoplasmic reticulum stress-related cognitive deficits via PERK/Nrf2 antioxidative signaling pathway. Eur J Pharmacol 2015;762:239–46
  • Trinh K, Andrews L, Krause J, et al. Decaffeinated coffee and nicotine-free tobacco provide neuroprotection in Drosophila models of Parkinson's disease through an NRF2-dependent mechanism. J Neurosci 2010;30:5525–32
  • Liby KT, Sporn MB. Synthetic oleanane triterpenoids: multifunctional drugs with a broad range of applications for prevention and treatment of chronic disease. Pharmacol Rev 2012;64:972–1003
  • Honda T, Rounds BV, Gribble GW, et al. Design and synthesis of 2-cyano-3,12-dioxoolean-1,9-dien-28-oic acid, a novel and highly active inhibitor of nitric oxide production in mouse macrophages. Bioorg Med Chem Lett 1998;8:2711–14
  • Suh N, Wang Y, Honda T, et al. A novel synthetic oleanane triterpenoid, 2-cyano-3,12-dioxoolean-1,9-dien-28-oic acid, with potent differentiating, antiproliferative, and anti-inflammatory activity. Cancer Res 1999;59:336–41
  • Sporn MB, Liby KT, Yore MM, et al. New synthetic triterpenoids: potent agents for prevention and treatment of tissue injury caused by inflammatory and oxidative stress. J Nat Prod 2011;74:537–45
  • Honda T, Rounds BV, Bore L, et al. Synthetic oleanane and ursane triterpenoids with modified rings A and C: a series of highly active inhibitors of nitric oxide production in mouse macrophages. J Med Chem 2000;43:4233–46
  • Liby K, Hock T, Yore MM, et al. The synthetic triterpenoids, CDDO and CDDO-imidazolide, are potent inducers of heme oxygenase-1 and Nrf2/ARE signaling. Cancer Res 2005;65:4789–98
  • Ahmad R, Raina D, Meyer C, et al. Triterpenoid CDDO-Me blocks the NF-kappaB pathway by direct inhibition of IKKbeta on Cys-179. J Biol Chem 2006;281:35764–9
  • Yore MM, Liby KT, Honda T, et al. The synthetic triterpenoid 1-[2-cyano-3,12-dioxooleana-1,9(11)-dien-28-oyl]imidazole blocks nuclear factor-kappaB activation through direct inhibition of IkappaB kinase beta. Mol Cancer Ther 2006;5:3232
  • Thimmulappa RK, Scollick C, Traore K, et al. Nrf2-dependent protection from LPS induced inflammatory response and mortality by CDDO-Imidazolide. Biochem Biophys Res Commun 2006;351:883–9
  • Dumont M, Wille E, Calingasan NY, et al. Triterpenoid CDDO-methylamide improves memory and decreases amyloid plaques in a transgenic mouse model of Alzheimer's disease. J Neurochem 2009;109:502–12
  • Yang L, Calingasan NY, Thomas B, et al. Neuroprotective effects of the triterpenoid, CDDO methyl amide, a potent inducer of Nrf2-mediated transcription. PLoS One 2009;4:e5757
  • Kaidery NA, Banerjee R, Yang L, et al. Targeting Nrf2-mediated gene transcription by extremely potent synthetic triterpenoids attenuate dopaminergic neurotoxicity in the MPTP mouse model of Parkinson's disease. Antioxid Redox Signal 2013;18:139–57
  • Neymotin A, Calingasan NY, Wille E, et al. Neuroprotective effect of Nrf2/ARE activators, CDDO ethylamide and CDDO trifluoroethylamide, in a mouse model of amyotrophic lateral sclerosis. Free Radic Biol Med 2011;51:88–96
  • Stack C, Ho D, Wille E, et al. Triterpenoids CDDO-ethyl amide and CDDO-trifluoroethyl amide improve the behavioral phenotype and brain pathology in a transgenic mouse model of Huntington's disease. Free Radic Biol Med 2010;49:147–58
  • Takagi T, Kitashoji A, Iwawaki T, et al. Temporal activation of Nrf2 in the penumbra and Nrf2 activator-mediated neuroprotection in ischemia-reperfusion injury. Free Radic Biol Med 2014;72:124–33
  • de Zeeuw D, Akizawa T, Audhya P, et al. Bardoxolone methyl in type 2 diabetes and stage 4 chronic kidney disease. N Engl J Med 2013;369:2492–503
  • Zhang DD. Bardoxolone brings Nrf2-based therapies to light. Antioxid Redox Signal 2013;19:517–18
  • Gold R, Kappos L, Arnold DL, et al. Placebo-controlled phase 3 study of oral BG-12 for relapsing multiple sclerosis. N Engl J Med 2012;367:1098–107
  • Fox RJ, Miller DH, Phillips JT, et al. Placebo-controlled phase 3 study of oral BG-12 or glatiramer in multiple sclerosis. N Engl J Med 2012;367:1087–97
  • Scannevin RH, Chollate S, Jung MY, et al. Fumarates promote cytoprotection of central nervous system cells against oxidative stress via the nuclear factor (erythroid-derived 2)-like 2 pathway. J Pharmacol Exp Ther 2012;341:274–84
  • Ellrichmann G, Petrasch-Parwez E, Lee DH, et al. Efficacy of fumaric acid esters in the R6/2 and YAC128 models of Huntington's disease. PLoS One 2011;6:e16172
  • Wang Q, Chuikov S, Taitano S, et al. Dimethyl fumarate protects neural stem/progenitor cells and neurons from oxidative damage through Nrf2-ERK1/2 MAPK pathway. Int J Mol Sci 2015;16:13885–907
  • Wu JZ, Cheng CC, Shen LL, et al. Synthetic chalcones with potent antioxidant ability on H(2)O(2)-induced apoptosis in PC12 cells. Int J Mol Sci 2014;15:18525–39
  • Kaufmann KB, Al-Rifai N, Ulbrich F, et al. The cytoprotective effects of E-alpha-(4-methoxyphenyl)-2′,3,4,4′-tetramethoxychalcone (E-alpha-p-OMe-C6H4-TMC) – a novel and non-cytotoxic HO-1 inducer. PLoS One 2015;10:e0142932
  • Otterbein LE, Bach FH, Alam J, et al. Carbon monoxide has anti-inflammatory effects involving the mitogen-activated protein kinase pathway. Nat Med 2000;6:422–8
  • Gullotta F, di Masi A, Ascenzi P. Carbon monoxide: an unusual drug. IUBMB Life 2012;64:378–86
  • Wilson JL, Fayad Kobeissi S, Oudir S, et al. Design and synthesis of new hybrid molecules that activate the transcription factor Nrf2 and simultaneously release carbon monoxide. Chemistry 2014;20:14698–704
  • Nikam A, Ollivier A, Rivard M, et al. Diverse Nrf2 activators coordinated to cobalt carbonyls induce heme oxygenase-1 and release carbon monoxide in vitro and in vivo. J Med Chem 2016;59:756–62
  • Go ML, Wu X, Liu XL. Chalcones: an update on cytotoxic and chemoprotective properties. Curr Med Chem 2005;12:481–99
  • Nowakowska Z. A review of anti-infective and anti-inflammatory chalcones. Eur J Med Chem 2007;42:125–37
  • Sawle P, Moulton BE, Jarzykowska M, et al. Structure–activity relationships of methoxychalcones as inducers of heme oxygenase-1. Chem Res Toxicol 2008;21:1484–94
  • Jin F, Jin XY, Jin YL, et al. Structural requirements of 2′,4′,6′-tris(methoxymethoxy) chalcone derivatives for anti-inflammatory activity: the importance of a 2′-hydroxy moiety. Arch Pharm Res 2007;30:1359–67
  • Kumar V, Kumar S, Hassan M, et al. Novel chalcone derivatives as potent Nrf2 activators in mice and human lung epithelial cells. J Med Chem 2011;54:4147–59
  • Lounsbury N, Mateo G, Jones B, et al. Heterocyclic chalcone activators of nuclear factor (erythroid-derived 2)-like 2 (Nrf2) with improved in vivo efficacy. Bioorg Med Chem 2015;23:5352–9
  • Wells G. Peptide and small molecule inhibitors of the Keap1–Nrf2 protein–protein interaction. Biochem Soc Trans 2015;43:674–9
  • Abed DA, Goldstein M, Albanyan H, et al. Discovery of direct inhibitors of Keap1–Nrf2 protein–protein interaction as potential therapeutic and preventive agents. Acta Pharm Sin B 2015;5:285–99
  • Hu L, Magesh S, Chen L, et al. Discovery of a small-molecule inhibitor and cellular probe of Keap1–Nrf2 protein–protein interaction. Bioorg Med Chem Lett 2013;23:3039–43
  • Marcotte D, Zeng W, Hus JC, et al. Small molecules inhibit the interaction of Nrf2 and the Keap1 Kelch domain through a non-covalent mechanism. Bioorg Med Chem 2013;21:4011–19
  • Jiang ZY, Lu MC, Xu LL, et al. Discovery of potent Keap1–Nrf2 protein–protein interaction inhibitor based on molecular binding determinants analysis. J Med Chem 2014;57:2736–45
  • Davies TG, Wixted WE, Coyle JE, et al. Monoacidic inhibitors of the Kelch-like ECH-associated protein 1: nuclear factor erythroid 2-related factor 2 (KEAP1:NRF2) protein–protein interaction with high cell potency identified by fragment-based discovery. J Med Chem 2016;59:3991–4006
  • Fox RJ, Kita M, Cohan SL, et al. BG-12 (dimethyl fumarate): a review of mechanism of action, efficacy, and safety. Curr Med Res Opin 2014;30:251–62

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.