ArticlePDF AvailableLiterature Review

Abstract and Figures

Activation of the KEAP1-NRF2 signaling pathway is an adaptive response to environmental and endogenous stresses and serves to render animals resistant to chemical carcinogenesis and other forms of toxicity, whereas disruption of the pathway exacerbates these outcomes. This pathway, which can be activated by sulfhydryl-reactive, small-molecule pharmacologic agents, regulates the inducible expression of an extended battery of cytoprotective genes, often by direct binding of the transcription factor to antioxidant response elements in the promoter regions of target genes. However, it is becoming evident that some of the protective effects may be mediated indirectly through cross talk with additional pathways affecting cell survival and other aspects of cell fate. These interactions provide a multi-tiered, integrated response to chemical stresses. This review highlights recent observations on the molecular interactions and their functional consequences between NRF2 and the arylhydrocarbon receptor (AhR), NF-κB, p53, and Notch1 signaling pathways.
Content may be subject to copyright.
FORUM REVIEW ARTICLE
When NRF2 Talks, Who’s Listening?
Nobunao Wakabayashi,
1,2
Stephen L. Slocum,
3
John J. Skoko,
4
Soona Shin,
4
and Thomas W. Kensler
1–4
Abstract
Activation of the KEAP1-NRF2 signaling pathway is an adaptive response to environmental and endogenous
stresses and serves to render animals resistant to chemical carcinogenesis and other forms of toxicity, whereas
disruption of the pathway exacerbates these outcomes. This pathway, which can be activated by sulfhydryl-
reactive, small-molecule pharmacologic agents, regulates the inducible expression of an extended battery of
cytoprotective genes, often by direct binding of the transcription factor to antioxidant response elements in the
promoter regions of target genes. However, it is becoming evident that some of the protective effects may be
mediated indirectly through cross talk with additional pathways affecting cell survival and other aspects of cell
fate. These interactions provide a multi-tiered, integrated response to chemical stresses. This review highlights
recent observations on the molecular interactions and their functional consequences between NRF2 and the
arylhydrocarbon receptor (AhR), NF-kB, p53, and Notch1 signaling pathways. Antioxid. Redox Signal. 13, 1649–
1663.
Introduction
NRF2 (NF-E2-related factor 2, NFE2L2) is a transcrip-
tion factor that mediates a broad-based set of adaptive
responses to intrinsic and extrinsic cellular stresses (73). As
such, NRF2 influences sensitivity to physiologic and patho-
logic processes affected by oxidative and electrophilic stres-
ses, such as those imposed by inflammation and exposures to
environmental toxicants. Carcinogenesis, chronic obstructive
pulmonary disease, obesogenesis, and neurodegeneration are
known to be affected by the Nrf2 genotype in murine models
(73).
Molecular details of the NRF2 signaling pathway have
emerged over the past decade, as reviewed elsewhere in this
issue. Notably, definition of the interacting partners of NRF2,
such as KEAP1 (Kelch-like ECH-associated protein 1), and the
means by which they sense and transduce chemical signals of
stress are under intense investigation in many laboratories
(48). Characterization of the immediate downstream target
genes of NRF2 has also been well enumerated, although
features that define the tissue and cell-type specificity in re-
sponses are poorly understood. The gene-expression re-
sponses to activation of NRF2 signaling, mediated through
interactions of NRF2 and the antioxidant response element(s)
(ARE: 50-NTGAG=CNNNGC-30) in the regulatory domains of
its target genes, result in a distinctive cytoprotective response.
Beyond the now-seminal response of catalyzing the detoxifi-
cation of carcinogens and other xenobiotics through conju-
gation and trapping processes, genomic analyses indicated
that gene families affected by Nrf2 (a) provide direct antiox-
idants; (b) encode enzymes that directly inactivate oxidants;
(c) increase levels of glutathione synthesis and regeneration;
(d) stimulate NADPH synthesis; (e) enhance toxin export
through the multidrug-response transporters; (f ) enhance the
recognition, repair, and removal of damaged proteins; (g)
elevate nucleotide excision repair; (h) regulate expression of
other transcription factors, growth factors and receptors, and
molecular chaperones; and (i) inhibit cytokine-mediated in-
flammation (48, 73).
It also is clear, though, that these direct genomic responses
do not account for all of the actions of pharmacologic agents
that activate NRF2 signaling, nor do they explain the range of
stress-response phenotypes observed when components of
the pathway, notably, Nrf2 or Keap1, have been genetically
disrupted in mice (162). Analyses of global transcriptional
responses have hinted toward interactions between NRF2
signaling and other prominent signaling pathways. These
interactions can occur in multiple forms. Post-translational
modifications such as phosphorylation play a major role in the
regulation of gene expression and function. These covalent
modifications control intracellular distribution, transcrip-
tional activity, and stability of transcription factors, including
NRF2 (81, 143). Some transcription factors can antagonize
NRF2 either by competing for binding to AREs or by
1
Department of Pharmacology & Chemical Biology, University of Pittsburgh, Pittsburgh, Pennsylvania.
2
Department of Environmental Health Sciences and
3
Department of Biochemistry and Molecular Biology, Bloomberg School of Public
Health,
4
Department of Pharmacology and Molecular Sciences, School of Medicine, Johns Hopkins University, Baltimore, Maryland.
ANTIOXIDANTS & REDOX SIGNALING
Volume 13, Number 11, 2010
ªMary Ann Liebert, Inc.
DOI: 10.1089=ars.2010.3216
1649
inhibiting NRF2 through a physical association. Small MAF
proteins, BACH1, and the immediate early proteins c-FOS
and FRA1 can compete with NRF2 for binding to AREs (103,
151). Furthermore, several transcription factors, including
activating transcription factor 3, proliferator-activated recep-
tor (PPAR)g, and retinoic acid receptor ahave been reported
to form inhibitory complexes with NRF2 (6, 13, 59, 155, 168).
In this review, we focus on recent evidence for transcriptional
cross-talk between NRF2 and the arylhydrocarbon receptor
(AhR), NF-kB, p53, and Notch pathways, as summarized
schematically in Fig. 1.
NRF2 Interactions with AhR Signaling
The aryl hydrocarbon receptor is a ligand-activated mem-
ber of the bHLH=PAS (basic helix-loop-helix=Per-Arnt-Sim)
family of transcription factors that mediates the biologic and
toxic effects of its xenobiotic ligands (44). When bound by
polycyclic aromatic hydrocarbons such as dioxins (TCDD),
AhR translocates from the cytoplasm to the nucleus, hetero-
dimerizes with AhR nuclear translocator (ARNT), and
activates transcription through the xenobiotic-responsive
element (XRE). The XRE consists of a canonic motif of
50-TNGCGTG-30. After nuclear export, AhR is degraded through
the 26S proteasome pathway. It is well recognized that both
AhR and NRF2 signaling regulate the expression of genes
affecting the metabolism of xenobiotics. In some instances, the
response elements recognizing these transcription factors can
be found in the regulatory domains of the same target genes,
such as Nqo1 (104). Ma et al. (95) reported that the inducible
expression of NQO1 by the potent AhR ligand TCDD de-
pends on both AhR and NRF2. Genetic experiments using
Ahr-, Arnt-, or Nrf2-deficient cells revealed that induction of
NQO1 by TCDD depended on the presence of AhR and
ARNT, whereas the basal and inducible expression required a
functional NRF2 pathway. Yaeger et al. (163) extended these
findings by using Nrf2-deficient mice to demonstrate the re-
quirement of NRF2 for the induction of Nqo1,Ugt1a6, and
Gsta1 transcripts, among other genes, by TCDD in mouse
liver. AhR–Nrf2 compound null mutant mice respond to nei-
ther AhR ligands nor prototypical NRF2 activators with target
genes, such as Nqo1 (105). A functional collaboration between
these pathways is also highlighted by the earlier findings of
Talalay and colleagues (115), in which they described the roles
of ‘‘monofunctional’’ and ‘‘bifunctional’’ activators of NRF2
signaling. ‘‘Monofunctional’’ inducers affect only NRF2 sig-
naling. ‘‘Bifunctional inducers,’’ such as b-napthoflavone,
may interact with the AhR directly to activate AhR-regulated
genes, such as Cyp1a1,Cyp1a2, and Cyp1b1, and then undergo
transformation by these enzymes to reactive intermediates
that then trigger NRF2 signaling (see Fig. 2). In this way, a
comprehensive set of responses to xenobiotic challenges can
be mounted (79).
The mechanisms underlying the linkages between AhR and
NRF2 signaling need further investigation. Miao et al. (98)
demonstrated that Nrf2 gene transcription is directly modu-
lated by AhR activation. DNA sequence analyses of the mu-
rine Nrf2 promoter revealed an XRE located at 712 bp and
two additional XRE-like motifs located further upstream.
Direct binding of AhR to the Nrf2 promoter was observed.
Moreover, silencing of AhR expression with siRNA obviated
induction of Nrf2 mRNA by TCDD. NRF2 also autoregulates
its own expression through an ARE-like element located in the
proximal region of its promoter, leading to persistent nuclear
accumulation of NRF2 and protracted induction of its target
genes (83).
NRF2 also directly modulates AhR signaling, highlighting
bidirectional interactions of these pathways (138). Constitutive
expression of AhR was affected by Nrf2 genotype. Moreover,
a pharmacologic activator of NRF2 signaling, CDDO-Im
{1-[2-cyano-3,12-dioxooleana-1,9(11)-dien-28-oyl]imidazole}, in-
duced Ahr,Cyp1a1,andCyp1b1 transcription in Nrf2
þ=þ
mouse
embryonic fibroblasts (MEFs) but not in Nrf2
-=-
MEFs. Reporter
analysis and chromatin immunoprecipitation assays revealed
that NRF2 directly binds to one ARE found in the 230 bp
region of the promoter of Ahr. Better understanding of this
bidirectional regulation of AhR and NRF2 pathways should
provide clues on the roles of NRF2 in complex diseases and in
uncharacterized phenotypes.
Cell-culture studies and knockout mouse models have
shown that, in addition to controlling the xenobiotic detoxi-
fication response, AhR activation leads to G
0
=G
1
arrest, di-
minished capacity for DNA replication, inhibition of cell
proliferation, and impaired cell differentiation. AhR may
function also as a tumor-suppressor gene that becomes si-
lenced during the process of tumor formation (32). Tran-
scriptional responses to ligand-dependent activation of AhR
signaling vary across cell types and by species, but consis-
tently influence cytochrome P450 and other monooxygenase
activities (17). Ligand-dependent and -independent AhR
signaling influences (a) positive regulation of transcription,
vascular development, organ morphogenesis, and metabolic
processes; (b) processes related to mRNA processing and
transport; (c) processes related to cell death, proliferation, and
regulation of apoptosis; (d) mitotic cell cycle; and (e) embry-
onic and other developmental processes (132).
FIG. 1. Possible means for regulation of cell survival and
other cell-fate responses through interactions of NRF2 with
additional cell-signaling pathways, including AhR, NF-jB,
p53, and Notch1. (For interpretation of the references to color
in this figure legend, the reader is referred to the web version
of this article at www.liebertonline.com=ars).
1650 WAKABAYASHI ET AL.
Ahr-null mice and Nrf2-null mice share some common
phenotypes. These transcription factors appear to function as
cell-survival factors; both types of knockout mice exhibit in-
creased sensitivity to infection (75, 147), carcinogenesis (32, 72,
117), altered responses to wounding (12, 18), and metabolic
syndrome (137, 144). Studies on responses to adipogenic
stimuli in vitro and high-fat stress diets in vivo have provided
some opportunities for evaluating the physiological interac-
tions between these pathways.
Shimba et al. (136) observed that TCDD treatment sup-
presses the conversion of 3T3-L1 cells into adipocytes. By
using MEF cell lines derived from Ahr-knockout mice,
Alexander et al. (5) reported that the AhR is a constitutive
inhibitor of triglyceride synthesis and an early regulator of
adipocyte differentiation. In addition, one of the phenotypes
of Ahr-null mice is transient fatty liver (133), implying an
in vivo regulatory role of AhR in the adipogenic process.
Shin et al. (138) postulated that NRF2 would inhibit adipo-
genesis through the interaction with the AhR pathway.
Nrf2
-=-
MEFs showed markedly accelerated adipogenesis on
stimulation, whereas Keap1
-=-
MEFs (which exhibit higher
NRF2 signaling) differentiated slowly compared with their
congenic wild-type MEFs. Ectopic expression of AhR and
dominant-positive Nrf2 in Nrf2
-=-
MEFs also substantially
delayed differentiation. To evaluate how NRF2 regulates
adipogenesis, differentiation markers involved at multiple
stages of adipogenesis were analyzed. CEBPs (CCAAT-
enhancer–binding proteins) and PPARs are the two families
of transcription factors that play critical roles in adipogen-
esis (159). CEBPband CEBPdfunction at an early phase of
the differentiation process by sensing adipogenic stimuli
and initiating expression of CEBPaand PPARg(122). CEBPa
and PPARgplay roles at a later stage by inducing and
maintaining expression of adipocyte-specific genes, such as
Fabp4. The elevation of CEBPbon adipogenic stimulation is
transient, whereas CEBPaand PPARgremain upregulated
for the duration of adipogenesis. Although the mechanism
by which AhR regulates adipogenesis has not been fully
characterized, recent work suggests that AhR affects dif-
ferentiation stages that follow CEBPbactivation (i.e., CEBPa
or PPARgupregulation. In 3T3-L1 preadipocytes, forced
expression of AhR resulted in lower induction of Fabp4 and
Cebpaupon differentiation than did that in control cells,
whereas the induction of Ppargwas not affected (136). Pparg
expression could be induced by differentiation in Ahr
-=-
MEFs, but levels were lower than those in Ahr
þ=þ
MEFs (5).
Similarly, mRNA levels of Cebpaand Fabp4 were higher in
Nrf2
-=-
MEFs than in Nrf2
þ=þ
MEFs, both before and after
differentiation, whereas induction of Pparg2was not af-
fected by the Nrf2 genotype. mRNA levels of Cebpbwere
lower in Nrf2
-=-
MEFs than in Nrf2
þ=þ
MEFs. In primary
Keap1
-=-
MEFs, disruption of Keap1 resulted in minimal in-
duction of Cebpa,Fabp4, and Pparg2upon differentiation.
CebpbmRNA levels remained higher in Keap1
-=-
MEFs than
in Keap1
þ=þ
MEFs, both before and after induction of adi-
pogenesis. Thus, NRF2 negatively modulates expression of
Cebpaand Pparg2but not Cebpbduring the course of adi-
pogenesis. The stages of differentiation that are affected by
NRF2 directly overlap with those affected by AhR, thereby
supporting the hypothesis that NRF2 inhibits adipogenesis
through cross talk with AhR signaling (see Fig. 3).
Treatment with CDDO-Im effectively prevented high-fat
diet–induced increases in body weight, adipose mass, and
hepatic lipid accumulation in Nrf2 wild-type mice but not in
Nrf2-disrupted mice (137). Wild-type mice on a high-fat diet
and treated with CDDO-Im exhibited higher oxygen con-
sumption and energy expenditure than did vehicle-treated
mice, whereas food intake was lower in CDDO-Im–treated
than in vehicle-treated mice. Levels of gene transcripts for
fatty acid synthesis enzymes were downregulated after
CDDO-Im treatment in the liver of wild-type mice. This in-
hibitory effect of CDDO-Im on lipogenic gene expression was
significantly reduced in Nrf2-disrupted mice. However, the
extent to which this pharmacologic activation of Nrf2 to
produce an anti-obesogenic effect is mediated through AhR, if
at all, is unclear.
FIG. 2. Regulation of xenobiotic me-
tabolism by AhR and NRF2. Xenobiotics
are metabolized by cytochrome P450s
(CYPs) into intermediates that are reactive
or nonreactive. Reactive intermediates can
be metabolized by cytoprotective enzymes
into less-toxic and often readily excretable
products. When activated by a ligand,
AhR in the cytoplasm complexes with
ARNT and translocates into the nucleus
and induces the transcription of CYPs by
binding to the xenobiotic response ele-
ments (XREs) in the promoters of target
genes. NRF2 bound to KEAP1 in the cy-
toplasm translocates into the nucleus
when the pathway is activated by exoge-
nous or endogenous electrophilic or free
radical stresses, and binds to the antioxi-
dant response elements (AREs) to induce
transcription of cytoprotective genes.
NRF2 CROSS TALK 1651
As with many transcription factors, the actions of the direct
target genes of AhR alone do not fully explain its toxicologic
and physiologic effects. In addition to interactions with NRF2
signaling, AhR extends its regulatory sweep by modulating
the function of other transcription factors, including estrogen
receptor (ERaand ERb) and androgen receptor (AR). These
cross-talk pathways are important mediators of the functions
of endogenous and exogenous AhR ligands. The liganded
AhR recently was shown to promote the ubiquitination and
proteasomal degradation of ERs and AR by assembling a
ubiquitin ligase complex, CUL4B
AHR
(108). Thus, complexes
of the AhR with ERs or AR appear to regulate transcription as
functional units by multiple mechanisms. As depicted in
Fig. 4, KEAP1 also serves as an adaptor protein for an E3
ubiquitin ligase complex (78). Thus, broad similarities exist in
the ways these pathways sense and respond to stresses. In-
deed, other stress response–signaling pathways such as
HIF-1aand NF-kB also interact with ubiquitin ligase systems
to facilitate proteolytic turnover of signaling components (109).
Direct and Indirect NRF2–NF-kB Interactions
Cross talk between NRF2 and nuclear factor k-light-chain-
enhancer of activated B cells (NF-kB) is an area of extensive
interest. NF-kB proteins are a family of transcription factors
involved in several processes such as inflammation, immune
response, apoptosis, development, and cell growth. Targets of
NF-kB include genes classified as chemokines, cytokines,
immunoreceptors, cell-adhesion molecules, stress-response
genes, regulators of apoptosis, growth factors, and tran-
scription factors, among many others. Tumor necrosis factor a
(TNF-a), inducible NOS (iNOS), interleukin-1 (IL-1), intra-
cellular adhesion molecule-1, and cyclooxygenase (COX-2)
have all been shown to be induced by NF-kB (112). Patholo-
gies related to alterations of the NF-kB pathway include al-
lergies (25), Alzheimer disease (97), autoimmunity (46),
obesity (40), atherosclerosis (125), arthritis (123), cancer (41),
Crohn’s disease (114), diabetes (164), and stroke (51).
All NF-kB proteins contain a conserved Rel homology do-
main responsible for dimerization and DNA binding to the
consensus sequence GGGRNNYYCC, (R ¼purine, N ¼any
base, Y ¼pyrimidine). The NF-kB family of proteins can be
divided into two distinct groups based on the presence of a
transactivation domain. RelA (p65), RelB, and c-Rel all con-
tain transactivation domains, whereas p50 and p52 do not,
and they require heterodimerization with the Rel proteins to
activate transcription. Inactive NF-kB is located in the cyto-
plasm associated with the negative regulator IkB, which can
conceal the nuclear-localization sequence or DNA-binding
domain of NF-kB to prevent transcription (148, 157, 165).
Activation of NF-kB is accomplished through phosphoryla-
tion of IkBbyIkB kinases (IKKs), which leads to the release
and nuclear translocation of NF-kB (22, 28, 150). A wide range
of agents have the ability to stimulate IKK to activate NF-kB,
including H
2
O
2
, TNF-a, IL-1, phorbol esters, hypoxia fol-
lowed by reoxygenation, ultraviolet radiation, or microbial
infection (112).
The NRF2 and NF-kB signaling pathways interface at
several points to control the transcription or function of
downstream target proteins. Many examples exist in which
activation and repression occur between members of the two
pathways through mechanisms of regulation ranging from
direct effects on the transcription factors themselves to pro-
tein–protein interactions and second-messenger effects on
target genes. Several cancer chemopreventive agents trigger
NRF2 signaling with a concomitant repression of NF-kB
FIG. 3. Schematic representation of stages of adipogen-
esis regulated by AhR and NRF2. Adipogenic stimuli such
as dexamethasone and isobutylmethylxanthine induce ex-
pression of the early markers CEBPband CEBPd, which
leads to induction of PPARgand CEBPa. NRF2 induces ex-
pression of Ahr mRNA by directly binding to the promoter
of the Ahr gene. AhR inhibits differentiation stages that fol-
low CEBPbactivation (i.e., CEBPaor PPARgupregulation).
FIG. 4. E3 ubiquitin ligase activities of KEAP1 and AHR.
KEAP1 binds with CUL3, RBX1, and E2 to facilitate the
ubiquitination of NRF2, leading its enhanced degradation by
the proteasome. Ligand-bound AhR assembles a CUL4B-
based E3 ubiquitin ligase complex to mediate a nongenomic
signaling pathway for influencing estrogen and androgen
actions. ER, estrogen receptor; AR, androgen receptor.
1652 WAKABAYASHI ET AL.
and its target genes. As examples, sulforaphane [(-)-1-
isothiocyanato-(4R)-methylsulfinyl)butane] reduces DNA
binding of NF-kB and decreases generation of NO, PGE
2
, and
TNF-ain Raw 264.7 macrophages without affecting IkBor
nuclear translocation of the transcription factor (50). 3H-1,2-
Dithiole-3-thione reduces the nuclear translocation and DNA
binding of NF-kB, along with changes in phosphorylation of
IkB in the hepatocytes of lipopolysaccharide (LPS)-treated
rats (69). Epigallocatechin-3-gallate induces NRF2 and re-
duces NF-kB, TNF-aand IL-1bin the lungs of bleomycin-
treated rats (140). Chalcone has been shown to induce NRF2
and inhibit the activation of NF-kB in endothelial cells (94).
The synthetic triterpenoid CDDO-Me can inhibit NF-kB ac-
tivity through repression of IKKb(1). Curcumin is able to
inhibit NF-kB by blocking phosphorylation and degradation
of IkB in macrophages challenged with LPS (113).
NRF2 and NF-kB can interact to have direct effects on gene
expression. NRF2 and NF-kB mediate the activation of
IkB and ARE sites, respectively, through antagonism of
transcription-factor binding to DNA. NF-kB was recently
shown to prevent the transcription of NRF2-dependent genes
by reducing available co-activator levels and promoting re-
cruitment of a co-repressor. p65 and Nrf2 both bind to the
CH1-KIX domain of CREB-binding protein (CBP), and after
phosphorylation of p65 at Ser276, NF-kB is able to suppress
transcription of ARE-dependent genes by preventing CBP
from binding to Nrf2 (93). A second mechanism of p65 tran-
scriptional repression of the ARE involving histone deacety-
lase 3 (HDAC3) has also been described. Overexpression of
p65 causes the recruitment of HDAC3 to the ARE by binding
to CBP or MafK. HDAC3 was shown to bind MafK in the
NRF2 dimerization region and to prevent the acetylation of
MafK by CBP (93). Further experiments showed that p65 is
able to affect NRF2-dependent transcription through the re-
cruitment of HDAC3 to the ARE by binding to CBP. HDAC3
was previously shown to bind and deacetylate CBP, resulting
in reduced CBP-mediated transcriptional coactivation (24). A
model for effects of reduced coactivator and recruitment of a
co-repressor is depicted in Fig. 5. Phosphorylation of p65 on
Ser276 reduces the available coactivator levels of CBP from
NRF2, which causes decreased transcription through the ARE
and also allows HDAC3 to bind CBP. The loss of CBP from the
MafK-NRF2 heterodimer may then cause destabilization and
allow HDAC3 to bind and deacetylate MafK. The HDAC
could be then be recruited to the ARE through the action of the
CBP-bound HDAC or a MafK homodimer–bound HDAC3
(93). Interestingly, this mechanism may be countered by in-
duction of NRF2, which leads to increased cytoplasmic-to-
nuclear translocation of the transcription factor. Experiments
have shown that repression of NRF2-dependent transcription
was eliminated by transfection of a dominant negative NRF2
that does not undergo cytoplasmic sequestration (93).
Several experiments have shown increased NF-kB activa-
tion in Nrf2
-=-
mice when compared with wild-type after
stimuli such as traumatic brain injury (65), LPS (147), TNF-a
(147), ovalbumin (119), and respiratory syncytial virus (23).
NRF2 has been implicated in NF-kB control through attenu-
ation of phosphorylated IkB, which causes NF-kB degrada-
tion. Nrf2
-=-
MEFs have higher levels of phosphorylated IkB
and greater IKK kinase activity when compared with Nrf2
þ=þ
MEFs after challenge with LPS or TNF-a(147). Interestingly,
overexpression of NRF2 seems to change only downstream
NF-kB targets, not NF-kBorIkB (21, 68, 90). Overexpression
of NRF2 in human aortic endothelial cells by using an ade-
noviral vector did not inhibit TNF-a–induced NF-kB activity
or IkB degradation, but induction of monocyte chemoat-
tractant protein-1 and VCAM-1 with TNF-awere inhibited
(20). These inconsistencies could be due to the loss of NRF2,
creating an environment with a diminished capacity to scav-
enge reactive oxygen species (ROS). The impaired expression
of antioxidative genes could potentially lead to increased ac-
tivation of the NF-kB pathway. Alternatively, enhancement of
NF-kB due to direct transcriptional action of NRF2 may occur.
In genetic studies, Nrf2
-=-
fibroblasts have shown reduced
steady-state protein levels of p50 and p65 when compared
with wild-type, whereas in vivo studies revealed that livers of
Nrf2
-=-
mice exhibited greatly reduced transcript levels of p65
(161). By contrast, TNF-aor LPS stimulation of Nrf2
-=-
fibro-
blasts caused increased activation of NF-kB (147). Enhance-
ment of NF-kB by NRF2 may be coordinated in a subset of the
NF-kB family members to drive expression of specific genes.
In Nrf2
-=-
fibroblasts p50 and p65 show reduced steady-state
protein levels to 30% of wild type, whereas c-Rel protein levels
were increased 400% (161). It is unknown whether NRF2 is
able specifically to control expression of NF-kB proteins that
lack a transactivation domain to repress transcription of NF-
kB target proteins (88, 91). The enhancement of NRF2 due to
direct transcriptional action of NF-kB may also be possible.
In silico transcription factor–binding site analysis has identi-
fied an NF-kB consensus binding site in the murine promoter
of NRF2 (102), which may increase transcription of NRF2 to
blunt prolonged NF-kB activity. As multiple agents and
stimuli such as ROS, LPS, flow shear stress, oxidized low-
density lipoprotein, and cigarette smoke have been shown to
induce both NRF2 and NF-kB activity (2, 7, 16, 42, 56, 77, 127),
a completely antagonistic effect of the two transcription fac-
tors against one another in all situations is unlikely.
Interactions between downstream targets of NRF2 and NF-
kB that ultimately lead to modulation of transcription factor
activity have been discovered, which adds complexity to the
two pathways. Three examples of NRF2 target genes that are
able to influence NF-kB activity are heme oxygenase-1 (HO-
1), NAD(P)H:quinone oxidoreductase (NQO1), and thior-
edoxin (TRX). HO-1, which breaks down heme to produce
biliverdin, CO, and free iron, is able to modulate activation of
NF-kB through the action of bilirubin and free iron (3, 66, 111,
135, 139, 145) (see Fig. 6). HO-1 has been shown specifically to
prevent phosphorylation of RelA through the action of free
iron at a site required for TNF-a–dependent NF-kB activation
(135). Induction of HO-1 can inhibit IkB degradation, whereas
inhibition of HO-1 increases p65 activity in HT-29 cells and
colonic mucosa after treatment with TNF-aor IL-1ß (66).
NQO1 has both positive and negative effects on NF-kB.
NQO1 overexpression reduces LPS-induced expression of the
downstream inflammatory genes TNF-aand IL-1 in THP-1
cells independent of NF-kB (128). Mice deficient of NQO1 in
bone marrow, spleen, and thymus showed reduced NF-kB
DNA binding under basal or LPS-treated conditions (61). TRX
also has both positive and negative effects on NF-kB. TRX can
modulate NF-kB activity through the reduction of cysteine
residues (26, 36, 47, 74, 96, 116). Nuclear TRX is required for
the reduction of cysteine residues on p50 to allow binding to
DNA, whereas cytoplasmic TRX is able to block the degra-
dation of IkB (54).
NRF2 CROSS TALK 1653
Target proteins of the NRF2 and NF-kB pathways and their
products can also interact to regulate enzyme activity. The
NF-kB target IL-10 (15) has been shown to induce HO-1 in
murine macrophages and in vivo (87). A feedback loop is
present between HO-1 and iNOS in macrophages to protect
the cell against injury due to excess NO. Expression of iNOS,
which leads to overproduction of NO, causes induction of
HO-1 (34, 60). HO-1 is then able to inhibit iNOS through the
action of free iron (156) and CO (131), as well as through the
overall reduction of heme (4) to prevent further production of
NO (141). Interestingly, NO can be scavenged by biliverdin
and bilirubin to further entangle the signaling pathways (71)
(Fig. 6). HO-1 has also been shown negatively to affect
VCAM-1 expression (9). Additional interactions between HO-
1 and the NF-kB pathway can be seen with the HO-1 product
CO. CO can repress the formation of TNF-a, IL-1b, and
macrophage inflammatory protein-1band increase IL-10 after
LPS stimulation (110).
An NF-kB target that has been shown to influence NRF2
activity is COX-2 and its downstream product 15-deoxy
D(12,14) prostaglandin J2 (15d-PGJ2). Induction of COX-2 by
shear stress has been shown to inhibit phosphatidylinositol 3-
kinase activity, causing reduction of the transcription of
NRF2, NQO1, HO-1, GCLR, and GSTM1 in human chon-
drocytes. Restoration of NRF2-dependent induction of the
genes was seen on treatment with the COX-2–specific inhib-
itors CAY10404 and NS398 (49). 15d-PGJ2 has been shown to
play a role in NRF2 modulation by binding to KEAP1 and
causing nuclear translocation of NRF2 in macrophages (63).
Further evidence of a direct effect of 15d-PGJ2 on NRF2 acti-
vation was found in vivo by using Nrf2
-=-
mice and the COX-2–
selective inhibitor NS-398 (99). 15d-PGJ2 has also been shown
to provide negative feedback through its own pathway by
binding IKK to prevent the activation of NF-kB in HeLa cells
and appears to do the same in human peripheral blood
monocytes (126).
NRF2 and NF-kB have a subset of target genes that are
regulated through the actions of both transcription factors on
ARE and kB sites. HO-1 has been shown to have a functional
ARE that can be activated by NRF2 (3), as well as a functional
NF-kB site, by using EMSA and DNase I footprinting analysis
(86). Glutamate-cysteine ligase catalytic subunit (GCLC) can
be activated by NRF2 through an ARE (101) or by NF-kB
through a kB site after ethanol exposure (76). The inhibitory G
protein, Gai2, is another example of a gene activated by both
NF-kB and NRF2 (8). Expression of the chemokine IL-8 can be
increased by both NRF2 and NF-kB, but not through the usual
mechanisms. NF-kB activates transcription of IL-8 by binding
to the kB site (82), whereas NRF2 instead uses a mechanism
that stabilizes IL-8 mRNA (166).
p53 and NRF2: Cooperation and Antagonism
in Protection Against Cancer
The examination of p53 after its discovery in 1979 (92) has
been extensive, with approximately 52,000 articles acces-
sioned when ‘‘p53’’ is searched on PubMed. This investigative
interest is not without reason, inasmuch as p53 is mutated in
upward of 50% of human cancers through single, compound,
or deletion mutations (55). This common frequency of muta-
tions makes p53 an attractive target for chemotherapeutics. If
p53 is to be targeted, however, the body of knowledge sur-
rounding the protein should be expanded to provide full
understanding of the interactions of p53 with other signaling
molecules and networks.
p53 regulates a host of processes within the cell, both di-
rectly, including the induction of apoptosis through the per-
meabilization of the outer mitochondrial membrane through
interaction with proapoptotic factors (43, 89, 100), and indi-
rectly, through the transcription of genes that effect changes
on the cell. Examples include both pro- and antiapoptotic
responses, the induction of cellular senescence, and the repair
of genotoxic damage (39). These various activities of p53 make
it an integral part of the cellular defense against transforma-
FIG. 5. Repression of NRF2
signaling by members of the
NF-jB pathway. p65 represses
NRF2-dependent transcription at
the ARE when phosphorylated
on S276 by either competition for
coactivator binding proteins or
by recruiting HDAC to the ARE,
which can deacetylate histone H4
and MafK.
FIG. 6. The interactions of the NRF2 and NF-jB target
proteins HO-1 and iNOS and their products can regulate
the activity of each other in macrophages to protect against
an overabundance of NO. Excess NO acts as a signal to
increase HO-1, which is able to scavenge NO and block the
activity of iNOS to prevent further production.
1654 WAKABAYASHI ET AL.
tion: p53 controls cell fates by regulating the activation of
apoptosis to remove a heavily damaged cell, induces terminal
differentiation to remove the threat of a moderately damaged
cell, or induces the transcription of cellular scavenging pro-
teins to reestablish cellular homeostasis.
Iida et al. (58) showed that Nrf2 and p53 collaborate to
protect against carcinogenesis (58), with the demonstrations
that, compared with wild-type mice, either Nrf2 knockout
mice or p53 heterozygotic mice are more susceptible to
nitrosamine-induced bladder carcinogenesis, and that the
Nrf2
-=-
::p53
þ=-
compound knockout mice are even more sus-
ceptible. This interaction is not unexpected, as when either the
cytoprotective response system of NRF2, or p53, a cell-fate
decision molecule is removed, the incidence of cancer in-
creases after carcinogen challenge (117, 146). Although broad
explanations based on general understandings of pathways
can be posited, the actual interactions between NRF2 and p53
in this setting and the collective effects on the expression and
function of their downstream target genes remain to be
established.
In a more-direct examination of the interaction between
p53 and NRF2, p53 has been shown to influence NRF2-based
transcription by Cimino and colleagues (33). Specifically, p53
suppresses the transcription of three NRF2 target genes dri-
ven through the ARE. These genes, x-ct,Nqo1, and Gst1a1, are
all involved in mounting an antioxidant response within the
cell, specifically neutralizing ROS. In the study, various cell
lines were transfected with expression vectors containing Nrf2
or p53. In all circumstances, the transformation with both Nrf2
and p53 results in a reduction of transcript levels of the target
genes when compared with cells transformed with Nrf2 alone.
This, in combination with experiments in which cells were
exposed to etoposide (a topoisomerase II–inhibiting chemo-
therapeutic drug) also shows that p53 has the ability to reg-
ulate negatively the NRF2 target gene transcripts. When p53 is
highly upregulated, through either overexpression or a strong
DNA-damaging agent, antioxidant response genes are
downregulated through p53-mediated disruption. Whether
or to what extent NRF2 regulates p53 signaling is not clear.
Transcriptome analyses, along with unpublished observa-
tions (M.K. Kwak and N. Wakabayashi), suggest that NRF2
contributes to the basal expression of MDM2, an inhibitor of
p53 (85). The potential additional level of p53 regulation
through NRF2 allows the dampening of MDM2-based pro-
teasomal degradation of p53 in circumstances in which p53
may be modulating the expression of NRF2 target genes.
Dampening of Mdm2 would allow an enhanced p53 signal.
This effect makes intuitive sense, as if p53 were mediating an
apoptotic response through ROS production, NRF2-driven
gene expression of anti-ROS enzymes would be counterpro-
ductive. This observation, however, is confounded by a recent
article examining the relation between p21, a direct down-
stream target of p53, and NRF2 (19).
p21 is involved with many different cellular processes, in-
cluding cell-cycle arrest, cell differentiation, senescence, apo-
ptosis and DNA replication and repair, and oxidative stress
(29, 31, 38, 86, 107). Chen et al. (19) recently showed a direct
interaction between p21 and NRF2, which may explain the
cytoprotective properties of p21 when faced with oxidative
stress. These investigators demonstrated in vitro that p21 is
able to interact with the DLG motif within NRF2, thereby
attenuating KEAP1-based ubiquitination and subsequent
proteasomal degradation. The antioxidative properties of p21
rely on the presence of NRF2, as an increase in p21 concen-
tration leads to increased transcript levels of ARE-regulated
genes, and the presence of p21 within cells yields an increase
in the half-life of NRF2. Additionally, it was observed in vivo
that both basal and induced levels of NRF2 protein are higher
in p21 wild-type mice than in p21-null mice, with induced
levels of NRF2 target gene proteins NQO1 and HO-1 being
higher in p21 wild-type mice. This discovery contributes sig-
nificantly to appreciation of the emerging network of signal-
ing between p53 and NRF2 by demonstrating that a direct
target of p53 upregulates NRF2 levels and activity.
The three puzzle pieces presented here come together in an
intricate fashion, demonstrating a tunable response by p53 to
induction by any number of stimuli, as shown in Fig. 7. When
p53 is strongly induced, the induction of apoptosis through
the expression of proapoptotic proteins through the genera-
tion of intracellular ROS is secured by the suppression of
antioxidant gene expression regulated by Nrf2, as well as the
suppression of basal levels of Mdm2, to ensure the strongest
possible cell-death signal both through direct levels of p53
and by its direct activities remaining uncompromised. The
upregulation of proapoptotic factors by p53 may rely on in-
creased levels of intracellular ROS production for signaling. In
this setting, ROS is not neutralized by NRF2 target genes such
as Nqo1 and Gst1a1because of transcriptional interference by
p53 (33), and the proteasomal degradation of p53 is hampered
because of the reduction in MDM2 levels.
On the other extreme, in the face of weak p53 induction, p21
is activated, stalling the cell cycle at the G
1
=S-phase check-
point (30, 45, 160) This stalling allows the induction of DNA
damage-repair functions by p53, the expression of proteins in
an attempt to lower intracellular ROS (10, 129), and the sta-
bilization of NRF2 by p21, thereby disrupting proteasomal
degradation of NRF2 by KEAP1. This disruption of NRF2
turnover allows the nuclear translocation of NRF2, and the
subsequent transcription of ARE-driven genes, leading to a
general cytoprotective response. The collective actions of p53
and NRF2 lead to the repair of the cell and evasion of apo-
ptosis. Links between NRF2 and the oxidative-stress–induced
inhibition of proliferation have been well described (120, 121);
in some instances, they may operate independent of p53.
The circumstance of a moderate induction of p53 allows
some speculation that a middle ground exists between the
promotion of programmed cell death and full cell repair and
restoration of homeostasis, namely, that of cellular senescence
through terminal differentiation. This balance is achieved
through p21 activation after p53 induction through signaling.
p21 blocks the cell cycle by stopping progression through the
G
1
=S checkpoint (30, 45, 160), and allows the stabilization of
NRF2 and an increased level of detoxification enzymes (19).
p53 almost certainly plays a role in activating additional
proteins that contribute to the process of rescuing a damaged
cell that may be useful in a state of terminal differentiation but
has experienced some insult that renders the cell a liability
when used for mitotic purposes.
This system of dynamic response to cellular stresses by p53,
p21, and NRF2 demonstrates a powerful mechanism for cel-
lular and organismal maintenance, through the repair, dif-
ferentiation, or destruction of damaged cells. In the case of
mild damage, NRF2-mediated detoxification of carcinogens
combined with a p21-mediated pause in the cell cycle and a
NRF2 CROSS TALK 1655
p53=p21-mediated induction of DNA damage-repair en-
zymes salvages cells from damage, whereas in the case of
extensive damage to a cell, p53 mediates the disruption of
NRF2-mediated antioxidant response proteins, allowing p53-
driven apoptosis, in toto reflecting a tunable response to en-
vironmental stresses.
NRF2 and NOTCH Signaling:
Cell-Survival Factor Meets Cell-Fate Factor
Several lines of differential microarray analyses using
RNAs isolated from MEFs prepared from Nrf2
-=-
and wild-
type mice have shown that Notch1 and its direct downstream
genes (Hes, Herp) (62) displayed decreased expression in
Nrf2
-=-
MEF. Studies from our laboratory demonstrate that
Notch1 signaling triggered through its activating ligands,
DLL1 and JAG1, is dependent on Nrf2 genotype. Moreover,
one or more functional ARE sequences exist in the promoter of
Notch1. Therefore, it appears that NRF2 directly regulates
Notch1 gene expression (153). It is now well established that
signals through the NOTCH receptor are involved in the de-
velopment of several cell types and that the modulation of
these signals can markedly affect differentiation, prolifera-
tion, and apoptotic events. Genetic ablation studies indicate
that Notch1 is crucial for early development and regrowth in a
variety of tissues (57, 154). Activation of the pathway has been
shown to be a potent inhibitor of differentiation in different
developmental contexts and has been associated with the
amplification of certain somatic stem cells (27, 35, 64, 80, 134,
142). Considering the significance of the NOTCH1 signal
cascade in developmental biology, this indicated the possi-
bility that NRF2 could be a key molecule affecting both em-
bryonic and adult tissue stem cell renewal, as well as cell fate.
Of course, this cross talk might not only be mediated by
transcriptional machinery, as NOTCH signaling is regulated
by various dynamic mechanisms, including posttranslational
modifications (149). Critical to the evaluation of cross talk
between transcription factors are assessments of their func-
tional consequences. Several physiological lines of evidence
support functional interactions between NRF2 and Notch1:
liver regeneration, keratinogenesis, osteoblastogenesis, and
adipogenesis (Fig. 8).
Nrf2
-=-
mice display an interesting phenotype in the process
of liver regeneration (11). Nrf2
-=-
mice have been challenged
with a two-thirds partial hepatectomy to view the impact of
genotype on the kinetics of liver regeneration. These knockout
mice showed a significant delay in the recovery of liver mass
after partial hepatectomy. We have observed that this regen-
eration-defective phenotype is rescued by hepatocyte-specific
expression of the NOTCH1 intracellular domain (NICD) in
Nrf2
-=-
mice. NICD functions as a transcription co-factor of
RBPJ. This finding provides clear evidence for the existence
of NRF2-Notch1 cross talk within the damaged liver, and
within the process of liver regeneration (153). In a rat model
in which siRNAs against NOTCH1 and JAG1 (a NOTCH1
ligand) were injected into the superior mesenteric vein before
the two-thirds partial hepatectomy, significantly suppressed
proliferation of hepatocytes was reported at days 2 to 4 of the
regenerative response (80). Apparently, the NRF2-NOTCH1
signaling pathway may be activated during liver regeneration
and is potentially contributing to signals affecting cell growth
and differentiation.
Keap1
-=-
mice, which exhibit hyperactivated NRF2 signal-
ing, have a thickened cornified layer on the suprabasal layer of
the esophagus (152), with a huge keratin mass derived from
the cornified layer around the limiting ridge and cardia. All
Keap1-null mice die within 3 weeks of malnutrition caused
from esophageal and forestomach obstruction. Through mi-
croscopic observation, a clear difference is noted between the
FIG. 7. p53 modulates a
gradient of responses to cel-
lular stress. When the cell is
exposed to a high level of
stress, p53 acts in a proa-
poptotic fashion, modulating
multiple pathways to secure a
full commitment to cell death,
whereas with a low level of
cellular stress, p53 modifies
gene expression to ensure cy-
toprotection after removal of
thestress.Theintermediatere-
sponse is also possible, with p53
retaining the ability to induce
cellular senescence through exit
of the cell cycle and induction
of terminal differentiation with
a concurrent cytoprotective
response.
FIG. 8. Possible effects of NRF2–Notch interactions on
cell fate.
1656 WAKABAYASHI ET AL.
processes of cornified layer thickening in the esophagus ver-
sus the foregut. In the case of the esophagus, cell numbers in
the basal layer seem to be unchanged, with the cornified layer
exhibiting thickening. This appears to be a prodifferentiation
mode of hyperkeratosis. Within the limiting ridge of the
greater curvature and the cardiac portion of the forestomach,
however, a proliferative form of hyperkeratosis is observed,
with proliferation of undifferentiated cells. This phenotype
might be explained by the excess Notch signaling produced by
excessive NRF2 activation. NOTCH signaling is expressed in
the gastrointestinal tract. Specifically, in both the esophagus
and forestomach, NOTCH1 and JAG2 display the highest
levels of expression within their families in the basal layer (70),
where stem and progenitor cells are located (130). Progenitor
cells are differentiated to keratinocytes in the suprabasal layer
and denucleated in the cornified layer, with Notch signaling
being weakly expressed in the suprabasal layer. Interestingly,
NOTCH signaling induces terminal differentiation to activate
p21 gene expression directly in keratinocytes (118, 158). This
function of the NOTCH signal might be stronger in the
esophagus and forestomach of Keap1-null mice.
Hinoi et al. (52) demonstrated that NRF2 negatively regu-
lates osteoblast differentiation in the MC3T3-E1 cell system.
They also showed that NRF2 might inhibit cellular differen-
tiation and maturation in chondrocytes by using the pre-
chondrogenic cell line ATDC5, derived from the mouse
teratocarcinoma AT805 (53). As a proposed mechanism of
these phenomena, they conjectured that the inhibition of
RUNX2 (one of the essential factors for bone formation and
chondral maturation) was caused by either direct or indirect
interaction with NRF2. Conversely, Notch-signaling directly
influences osteogenesis and chondrogenesis (14). Interestingly,
these phenomena are all related to bHLH-transcription factors,
especially HES and HERP, which are primary target genes of
NOTCH signaling. These proteins function as negative regu-
lators of target gene expression through direct cis-element
(E-box, N-box)–binding machinery. They also facilitate pro-
teasomal degradation through heterodimerization with other
bHLH transcription factors. The heterodimer partners of HES
or HERP primarily promote differentiation. Thus, they can lead
to maintenance of current cell status. HES1 directly binds to the
osteocalcin promoter and represses osteocalcin gene expression
(167). Additionally, HES1=HERP2 physically interacts with
RUNX2 in osteoblasts (37). It is also possible that NOTCH1
expression in osteoblasts is influenced by the NRF2 transcrip-
tion factor, much as NOTCH1 is in MEF.
As discussed earlier, AhR and its downstream gene ex-
pression is reduced in Nrf2
-=-
MEF (138). This dampened gene
expression impairs adipocyte differentiation from MEF. Nrf
-=-
MEF show markedly accelerated adipogenesis on stimula-
tion, whereas Keap1
-=-
MEF, which demonstrate higher levels
of NRF2 signaling, differentiate slowly compared with their
congenic wild-type MEF. Within this system, ectopic expres-
sion of AhR rescues adipogenesis seen in Nrf2
-=-
MEF.
However, recently Ross et al. (124) reported that HES1, a
primary downstream gene product of NOTCH1 signaling,
inhibits the differentiation of adipocytes from 3T3-L1 cells.
They also demonstrated that both the bHLH and the WRPW
(67) [which is a co-repressor (GROUCHO=TLE) binding
motif ] domains in HES1 are required for its inhibitory effect.
This study raised the possibility that the enhanced adipo-
genesis shown in Nrf2
-=-
MEF is caused by a weaker output
of both NOTCH1 and AhR signaling than within wild-type
MEF.
Conclusions
The NRF2 pathway has been studied in diverse contexts
because of potential roles in stroke recovery, neurodegen-
eration, chronic obstructive pulmonary disease, and cancer
prevention in multiple organs. The profiling of gene-expression
changes mediated by the NRF2 pathway has been well
documented. The interpretation of such studies, however, has
been hampered by the inability to distinguish genes modu-
lated directly by NRF2 from genes induced secondarily by
NRF2-sensitive transcription factors and the mitigating im-
pacts of tissue-specific responses. Nonetheless, it is clear that
the protective effects of upregulation of NRF2 signaling can
take several forms. Protection can be immediate, reflecting in-
duction of genes directly regulated through NRF2 binding to
AREs in target genes (e.g., the innate immune response and
elevated cytoprotective responses to blunt cytokine surges or
detoxify reactive intermediates, respectively) (73). The pro-
tective effects can be secondary through induction of macro-
molecular damage repair=removal systems (proteasome,
DNA repair) (84, 106). Last, the protective effects can be ter-
tiary through activation of tissue repair=regeneration path-
ways. In these latter cases, involvement in cross talk with
additional pathways affecting cell survival and other aspects
of cell fate most certainly play important collaborating roles.
This review has highlighted several such collaborators (and
sometime competitors): AhR, NF-kB, p53, and Notch1. Future
studies evaluating the concordance between global mRNA
expression profiles at the transcriptomic levels (i.e., oligonu-
cleotide microarray analyses) and genome-wide NRF2-DNA
binding analysis using high-throughput methods (i.e., chro-
matin-immunoprecipitation with parallel sequencing) should
better establish a systems-level view of NRF2-dependent
processes occurring within cells. Functional importance will
require well-crafted molecular genetic studies in cell culture,
and more important, in vivo, to define the critical cross-talk
interactions between NRF2 and other transcription factors.
Acknowledgments
This work was supported by NIH grants CA39416 and
CA94076 (TWK) and the Maryland Cigarette Restitution
Fund. SLS was supported by T32 ES07141 and T32 CA009110;
JJS by T32 GM08763; and SS was the recipient of a Samsung
Scholarship (Samsung Foundation of Culture).
References
1. Ahmad R, Raina D, Meyer C, Kharbanda S, and Kufe D.
Triterpenoid CDDO-Me blocks the NF-kappaB pathway by
direct inhibition of IKKbeta on Cys-179. J Biol Chem 281:
35764–35769, 2006.
2. Ahn KS and Aggarwal BB. Transcription factor NF-
kappaB: a sensor for smoke and stress signals. Ann N Y
Acad Sci 1056: 218–2133, 2005.
3. Alam J, Stewart D, Touchard C, Boinapally S, Choi AM,
and Cook JL. Nrf2, a Cap’n’Collar transcription factor,
regulates induction of the heme oxygenase-1 gene. J Biol
Chem 274: 26071–26078, 1999.
4. Albakri QA and Stuehr DJ. Intracellular assembly of in-
ducible NO synthase is limited by nitric oxide-mediated
NRF2 CROSS TALK 1657
changes in heme insertion and availability. J Biol Chem 271:
5414–5421, 1996.
5. Alexander DL, Ganem LG, Fernandez-Salguero P, Gonza-
lez F, and Jefcoate CR. Aryl-hydrocarbon receptor is an
inhibitory regulator of lipid synthesis and of commitment
to adipogenesis. J Cell Sci 111: 3311–3322, 1998.
6. Ansell PJ, Lo SC, Newton LG, Espinosa-Nicholas C, Zhang
DD, Liu JH, Hannink M, and Lubahn DB. Repression of
cancer protective genes by 17beta-estradiol: ligand-dependent
interaction between human Nrf2 and estrogen receptor
alpha. Mol Cell Endocrinol 243: 27–34, 2005.
7. Anwar AA, Li FY, Leake DS, Ishii T, Mann GE, and Siow
RC. Induction of heme oxygenase 1 by moderately oxi-
dized low-density lipoproteins in human vascular smooth
muscle cells: role of mitogen-activated protein kinases and
Nrf2. Free Radic Biol Med 39: 227–236, 2005.
8. Arinze IJ and Kawai Y. Transcriptional activation of the
human Galphai2 gene promoter through nuclear factor-
kappaB and antioxidant response elements. J Biol Chem 280:
9786–9795, 2005.
9. Banning A, Brigelius-Flohe R. NF-kappaB, Nrf2, and HO-1
interplay in redox-regulated VCAM-1 expression. Antioxid
Redox Signal 7: 889–899, 2005.
10. Bensaad K, Tsuruta A, Selak MA, Vidal MN, Nakano K,
Bartrons R, Gottlieb E, Vousden KH. TIGAR, a p53-
inducible regulator of glycolysis and apoptosis. Cell 126:
107–120, 2006.
11. Beyer TA, Xu W, Teupser D, auf dem Keller U, Bugnon P,
Hildt E, Thiery J, Kan YW, and Werner S. Impaired liver
regeneration in Nrf2 knockout mice: role of ROS-mediated
insulin=IGF-1 resistance. EMBO J 27: 212–223, 2008.
12. Braun S, Hanselmann C, Gassmann MG, auf dem Keller U,
Born-Berclaz C, Chan K, Kan YW, and Werner S. Nrf2
transcription factor, a novel target of keratinocyte growth
factor action which regulates gene expression and inflam-
mation in the healing skin wound. Mol Cell Biol 22: 5492–
5505, 2002.
13. Brown SL, Sekhar KR, Rachakonda G, Sasi S, and Freeman
ML. Activating transcription factor 3 is a novel repressor of
the nuclear factor erythroid-derived 2-related factor 2 (Nrf2)-
regulated stress pathway. Cancer Res 68: 364–368, 2008.
14. Canalis E. Notch signaling in osteoblasts. Sci Signal 1: pe17,
2008.
15. Cao S, Zhang X, Edwards JP, and Mosser DM. NF-kappaB1
(p50) homodimers differentially regulate pro- and anti-
inflammatory cytokines in macrophages. J Biol Chem 281:
26041–26050, 2006.
16. Carayol N, Chen J, Yang F, Jin T, Jin L, States D, and Wang
CY. A dominant function of IKK=NF-kappaB signaling in
global lipopolysaccharide-induced gene expression. J Biol
Chem 281: 31142–31151, 2006.
17. Carlson EA, McCulloch C, Koganti A, Goodwin SB, Sutter
TR, and Silkworth JB. Divergent transcriptomic responses
to aryl hydrocarbon receptor agonists between rat and
human primary hepatocytes. Toxicol Sci 112: 257–272, 2009.
18. Carvajal-Gonzalez JM, Roman AC, Cerezo-Guisado MI,
Rico-Leo EM, Martin-Partido G, and Fernandez-Salguero
PM. Loss of dioxin-receptor expression accelerates wound
healing in vivo by a mechanism involving TGFbeta. J Cell
Sci 122: 1823–1833, 2009.
19. Chen W, Sun Z, Wang XJ, Jiang T, Huang Z, Fang D, and
Zhang DD. Direct interaction between Nrf2 and
p21(Cip1=WAF1) upregulates the Nrf2-mediated antioxi-
dant response. Mol Cell 34: 663–673, 2009.
20. Chen XL, Dodd G, Thomas S, Zhang X, Wasserman MA,
Rovin BH, and Kunsch C. Activation of Nrf2=ARE path-
way protects endothelial cells from oxidant injury and in-
hibits inflammatory gene expression. Am J Physiol Heart
Circ Physiol 290: H1862–H1870, 2006.
21. Chen XL, Varner SE, Rao AS, Grey JY, Thomas S, Cook CK,
Wasserman MA, Medford RM, Jaiswal AK, and Kunsch C.
Laminar flow induction of antioxidant response element-
mediated genes in endothelial cells: a novel anti-inflammatory
mechanism. J Biol Chem 278: 703–711, 2003.
22. Chen ZJ, Parent L, and Maniatis T. Site-specific phosphor-
ylation of IkappaBalpha by a novel ubiquitination-dependent
protein kinase activity. Cell 84: 853–862, 1996.
23. Cho HY, Imani F, Miller-DeGraff L, Walters D, Melendi
GA, Yamamoto M, Polack FP, and Kleeberger SR. Antiviral
activity of Nrf2 in a murine model of respiratory syncytial
virus disease. Am J Respir Crit Care Med 179: 138–150, 2009.
24. Chuang HC, Chang CW, Chang GD, Yao TP, and Chen H.
Histone deacetylase 3 binds to and regulates the GCMa
transcription factor. Nucleic Acids Res 34: 1459–1469, 2006.
25. Cousins DJ, McDonald J, and Lee TH. Therapeutic ap-
proaches for control of transcription factors in allergic dis-
ease. J Allergy Clin Immunol 121: 803–809; quiz 810–811, 2008.
26. Das KC. c-Jun NH2-terminal kinase-mediated redox-
dependent degradation of IkappaB: role of thioredoxin in
NF-kappaB activation. J Biol Chem 276: 4662–4670, 2001.
27. de la Pompa JL, Wakeham A, Correia KM, Samper E,
Brown S, Aguilera RJ, Nakano T, Honjo T, Mak TW, Ros-
sant J, and Conlon RA. Conservation of the Notch signal-
ling pathway in mammalian neurogenesis. Development
124: 1139–1148, 1997.
28. DiDonato JA, Hayakawa M, Rothwarf DM, Zandi E, and
Karin M. A cytokine-responsive IkappaB kinase that acti-
vates the transcription factor NF-kappaB. Nature 388: 548–
554, 1997.
29. Dotto GP. p21(WAF1=Cip1): more than a break to the cell
cycle? Biochim Biophys Acta 1471: M43–M56, 2000.
30. el-Deiry WS, Tokino T, Velculescu VE, Levy DB, Parsons R,
Trent JM, Lin D, Mercer WE, Kinzler KW, and Vogelstein B.
WAF1, a potential mediator of p53 tumor suppression. Cell
75: 817–825, 1993.
31. Esposito F, Cuccovillo F, Russo L, Casella F, Russo T, and
Cimino F. A new p21waf1=cip1 isoform is an early event of
cell response to oxidative stress. Cell Death Differ 5: 940–945,
1998.
32. Fan Y, Boivin GP, Knudsen ES, Nebert DW, Xia Y, and
Puga A. The aryl hydrocarbon receptor functions as a
tumor suppressor of liver carcinogenesis. Cancer Res 70:
212–222, 2010.
33. Faraonio R, Vergara P, Di Marzo D, Pierantoni MG, Na-
politano M, Russo T, and Cimino F. p53 suppresses the
Nrf2-dependent transcription of antioxidant response
genes. J Biol Chem 281: 39776–39784, 2006.
34. Foresti R, Clark JE, Green CJ, and Motterlini R. Thiol
compounds interact with nitric oxide in regulating heme
oxygenase-1 induction in endothelial cells. Involvement of
superoxide and peroxynitrite anions. J Biol Chem 272:
18411–18417, 1997.
35. Fre S, Huyghe M, Mourikis P, Robine S, Louvard D, and
Artavanis-Tsakonas S. Notch signals control the fate of
immature progenitor cells in the intestine. Nature 435: 964–
968, 2005.
36. Freemerman AJ, Gallegos A, and Powis G. Nuclear factor
kappaB transactivation is increased but is not involved in
1658 WAKABAYASHI ET AL.
the proliferative effects of thioredoxin overexpression in
MCF-7 breast cancer cells. Cancer Res 59: 4090–4094, 1999.
37. Garg V, Muth AN, Ransom JF, Schluterman MK, Barnes R,
King IN, Grossfeld PD, and Srivastava D. Mutations in
NOTCH1 cause aortic valve disease. Nature 437: 270–274,
2005.
38. Gartel AL and Tyner AL. The role of the cyclin-dependent
kinase inhibitor p21 in apoptosis. Mol Cancer Ther 1: 639–
649, 2002.
39. Gatz SA and Wiesmuller L. p53 in recombination and re-
pair. Cell Death Differ 13: 1003–1016, 2006.
40. Gil A, Maria Aguilera C, Gil-Campos M, and Canete R.
Altered signalling and gene expression associated with the
immune system and the inflammatory response in obesity.
Br J Nutr 98(suppl 1): S121–S126, 2007.
41. Gilmore T, Gapuzan ME, Kalaitzidis D, and Starczynowski
D. Rel=NF-kappa B=I kappa B signal transduction in the
generation and treatment of human cancer. Cancer Lett 181:
1–9, 2002.
42. Go YM, Gipp JJ, Mulcahy RT, and Jones DP. H2O2-
dependent activation of GCLC-ARE4 reporter occurs by
mitogen-activated protein kinase pathways without oxi-
dation of cellular glutathione or thioredoxin-1. J Biol Chem
279: 5837–5845, 2004.
43. Green DR and Kroemer G. Cytoplasmic functions of the
tumour suppressor p53. Nature 458: 1127–1130, 2009.
44. Gu YZ, Hogenesch JB, and Bradfield CA. The PAS super-
family: sensors of environmental and developmental sig-
nals. Annu Rev Pharmacol Toxicol 40: 519–561, 2000.
45. Harper JW, Adami GR, Wei N, Keyomarsi K, and Elledge
SJ. The p21 Cdk-interacting protein Cip1 is a potent inhib-
itor of G1 cyclin-dependent kinases. Cell 75: 805–816, 1993.
46. Hayashi T and Faustman D. Defective function of the
proteasome in autoimmunity: involvement of impaired
NF-kappaB activation. Diabetes Technol Ther 2: 415–428,
2000.
47. Hayashi T, Ueno Y, and Okamoto T. Oxidoreductive reg-
ulation of nuclear factor kappa B. Involvement of a cellular
reducing catalyst thioredoxin. J Biol Chem 268: 11380–11388,
1993.
48. Hayes JD and McMahon M. NRF2 and KEAP1 mutations:
permanent activation of an adaptive response in cancer.
Trends Biochem Sci 34: 176–188, 2009.
49. Healy ZR, Lee NH, Gao X, Goldring MB, Talalay P, Kensler
TW, and Konstantopoulos K. Divergent responses of
chondrocytes and endothelial cells to shear stress: cross-
talk among COX-2, the phase 2 response, and apoptosis.
Proc Natl Acad Sci U S A 102: 14010–14015, 2005.
50. Heiss E, Herhaus C, Klimo K, Bartsch H, and Gerhauser C.
Nuclear factor kappa B is a molecular target for sulfor-
aphane-mediated anti-inflammatory mechanisms. J Biol
Chem 276: 32008–32015, 2001.
51. Herrmann O, Baumann B, de Lorenzi R, Muhammad S,
Zhang W, Kleesiek J, Malfertheiner M, Kohrmann M, Po-
trovita I, Maegele I, Beyer C, Burke JR, Hasan MT, Bujard
H, Wirth T, Pasparakis M, and Schwaninger M. IKK me-
diates ischemia-induced neuronal death. Nat Med 11: 1322–
1329, 2005.
52. Hinoi E, Fujimori S, Wang L, Hojo H, Uno K, and Yoneda
Y. Nrf2 negatively regulates osteoblast differentiation via
interfering with Runx2-dependent transcriptional activa-
tion. J Biol Chem 281: 18015–18024, 2006.
53. Hinoi E, Takarada T, Fujimori S, Wang L, Iemata M, Uno
K, and Yoneda Y. Nuclear factor E2 p45-related factor 2
negatively regulates chondrogenesis. Bone 40: 337–344,
2007.
54. Hirota K, Murata M, Sachi Y, Nakamura H, Takeuchi J,
Mori K, and Yodoi J. Distinct roles of thioredoxin in the
cytoplasm and in the nucleus: a two-step mechanism of
redox regulation of transcription factor NF-kappaB. J Biol
Chem 274: 27891–27897, 1999.
55. Hollstein M, Sidransky D, Vogelstein B, and Harris CC. p53
mutations in human cancers. Science 253: 49–53, 1991.
56. Hosoya T, Maruyama A, Kang MI, Kawatani Y, Shibata T,
Uchida K, Warabi E, Noguchi N, Itoh K, and Yamamoto M.
Differential responses of the Nrf2-Keap1 system to laminar
and oscillatory shear stresses in endothelial cells. J Biol
Chem 280: 27244–27250, 2005.
57. Huppert SS, Le A, Schroeter EH, Mumm JS, Saxena MT,
Milner LA, and Kopan R. Embryonic lethality in mice
homozygous for a processing-deficient allele of Notch1.
Nature 405: 966–970, 2000.
58. Iida K, Itoh K, Maher JM, Kumagai Y, Oyasu R, Mori Y,
Shimazui T, Akaza H, and Yamamoto M. Nrf2 and
p53 cooperatively protect against BBN-induced urinary
bladder carcinogenesis. Carcinogenesis 28: 2398–2403,
2007.
59. Ikeda Y, Sugawara A, Taniyama Y, Uruno A, Igarashi K,
Arima S, and Ito S, Takeuchi K. Suppression of rat throm-
boxane synthase gene transcription by peroxisome pro-
liferator-activated receptor gamma in macrophages via an
interaction with NRF2. J Biol Chem 275: 33142–33150, 2000.
60. Immenschuh S, Tan M, and Ramadori G. Nitric oxide
mediates the lipopolysaccharide dependent regulation of
the heme oxygenase-1 gene expression in cultured rat
Kupffer cells. J Hepatol 30: 61–69, 1999.
61. Iskander K, Li J, Han S, Zheng B, and Jaiswal AK. NQO1
and NQO2 regulation of humoral immunity and autoim-
munity. J Biol Chem 281: 30917–30924, 2006.
62. Iso T, Kedes L, and Hamamori Y. HES and HERP families:
multiple effectors of the Notch signaling pathway. J Cell
Physiol 194: 237–255, 2003.
63. Itoh K, Mochizuki M, Ishii Y, Ishii T, Shibata T, Kawamoto
Y, Kelly V, Sekizawa K, Uchida K, and Yamamoto M.
Transcription factor Nrf2 regulates inflammation by me-
diating the effect of 15-deoxy-Delta(12,14)-prostaglandin
j(2). Mol Cell Biol 24: 36–45, 2004.
64. Jensen CH, Jauho EI, Santoni-Rugiu E, Holmskov U, Teis-
ner B, Tygstrup N, and Bisgaard HC. Transit-amplifying
ductular (oval) cells and their hepatocytic progeny are
characterized by a novel and distinctive expression of
delta-like protein=preadipocyte factor 1=fetal antigen 1. Am
J Pathol 164: 1347–1359, 2004.
65. Jin W, Zhu L, Guan Q, Chen G, Wang QF, Yin HX, Hang
CH, Shi JX, and Wang HD. Influence of Nrf2 genotype on
pulmonary NF-kappaB activity and inflammatory response
after traumatic brain injury. Ann Clin Lab Sci 38: 221–227,
2008.
66. Jun CD, Kim Y, Choi EY, Kim M, Park B, Youn B, Yu K,
Choi KS, Yoon KH, Choi SC, Lee MS, Park KI, Choi M,
Chung Y, and Oh J. Gliotoxin reduces the severity of tri-
nitrobenzene sulfonic acid-induced colitis in mice: evidence
of the connection between heme oxygenase-1 and the
nuclear factor-kappaB pathway in vitro and in vivo. In-
flamm Bowel Dis 12: 619–629, 2006.
67. Kageyama R, Ohtsuka T, and Tomita K. The bHLH gene
Hes1 regulates differentiation of multiple cell types. Mol
Cells 10: 1–7, 2000.
NRF2 CROSS TALK 1659
68. Kanninen K, Heikkinen R, Malm T, Rolova T, Kuhmonen S,
Leinonen H, Yla-Herttuala S, Tanila H, Levonen AL,
Koistinaho M, and Koistinaho J. Intrahippocampal injection
of a lentiviral vector expressing Nrf2 improves spatial
learning in a mouse model of Alzheimer’s disease. Proc Natl
Acad Sci U S A 106: 16505–10510, 2009.
69. Karuri AR, Huang Y, Bodreddigari S, Sutter CH, Roebuck
BD, Kensler TW, and Sutter TR. 3H-1,2-dithiole-3-thione
targets nuclear factor kappaB to block expression of in-
ducible nitric-oxide synthase, prevents hypotension, and
improves survival in endotoxemic rats. J Pharmacol Exp
Ther 317: 61–67, 2006.
70. Katoh M and Katoh M. Notch signaling in gastrointestinal
tract (review). Int J Oncol 30: 247–251, 2007.
71. Kaur H, Hughes MN, Green CJ, Naughton P, Foresti R, and
Motterlini R. Interaction of bilirubin and biliverdin with
reactive nitrogen species. FEBS Lett 543: 113–119, 2003.
72. Kawajiri K, Kobayashi Y, Ohtake F, Ikuta T, Matsush-
imaY,MimuraJ,PetterssonS,PollenzRS,SakakiT,
Hirokawa T, Akiyama T, Kurosumi M, Poellinger L,
Kato S, and Fujii-Kuriyama Y. Aryl hydrocarbon re-
ceptor suppresses intestinal carcinogenesis in ApcMin=þ
mice with natural ligands. Proc Natl Acad Sci U S A 106:
13481–13486, 2009.
73. Kensler TW, Wakabayashi N, and Biswal S. Cell survival
responses to environmental stresses via the Keap1-Nrf2-
ARE pathway. Annu Rev Pharmacol Toxicol 47: 89–116,
2007.
74. Kim YC, Masutani H, Yamaguchi Y, Itoh K, Yamamoto M,
and Yodoi J. Hemin-induced activation of the thioredoxin
gene by Nrf 2: a differential regulation of the antioxidant
responsive element by a switch of its binding factors. J Biol
Chem 276: 18399–18406, 2001.
75. Kimura A, Naka T, Nakahama T, Chinen I, Masuda K,
Nohara K, Fujii-Kuriyama Y, and Kishimoto T. Aryl hy-
drocarbon receptor in combination with Stat1 regulates
LPS-induced inflammatory responses. J Exp Med 206: 2027–
2035, 2009.
76. Kimura T, Kawasaki Y, Okumura F, Sone T, Natsuki R, and
Isobe M. Ethanol-induced expression of glutamate-cysteine
ligase catalytic subunit gene is mediated by NF-kappaB.
Toxicol Lett 185: 110–115, 2009.
77. Knorr-Wittmann C, Hengstermann A, Gebel S, Alam J, and
Muller T. Characterization of Nrf2 activation and heme
oxygenase-1 expression in NIH3T3 cells exposed to aque-
ous extracts of cigarette smoke. Free Radic Biol Med 39:
1438–1448, 2005.
78. Kobayashi A, Kang MI, Okawa H, Ohtsuji M, Zenke Y,
Chiba T, Igarashi K, and Yamamoto M. Oxidative stress
sensor Keap1 functions as an adaptor for Cul3-based E3
ligase to regulate proteasomal degradation of Nrf2. Mol Cell
Biol 24: 7130–7139, 2004.
79. Kohle C and Bock KW. Coordinate regulation of Phase I
and II xenobiotic metabolisms by the Ah receptor and Nrf2.
Biochem Pharmacol 73: 1853–1862, 2007.
80. Kohler C, Bell AW, Bowen WC, Monga SP, Fleig W, and
Michalopoulos GK. Expression of Notch-1 and its ligand
Jagged-1 in rat liver during liver regeneration. Hepatology
39: 1056–1065, 2004.
81. Kong AN, Owuor E, Yu R, Hebbar V, Chen C, Hu R, and
Mandlekar S. Induction of xenobiotic enzymes by the MAP
kinase pathway and the antioxidant or electrophile re-
sponse element (ARE=EpRE). Drug Metab Rev 33: 255–271,
2001.
82. Kunsch C and Rosen CA. NF-kappa B subunit-specific
regulation of the interleukin-8 promoter. Mol Cell Biol 13:
6137–6146, 1993.
83. Kwak MK, Itoh K, Yamamoto M, and Kensler TW. En-
hanced expression of the transcription factor Nrf2 by cancer
chemopreventive agents: role of antioxidant response ele-
ment-like sequences in the nrf2 promoter. Mol Cell Biol 22:
2883–2892, 2002.
84. Kwak MK, Wakabayashi N, Greenlaw JL, Yamamoto M,
and Kensler TW. Antioxidants enhance mammalian pro-
teasome expression through the Keap1-Nrf2 signaling
pathway. Mol Cell Biol 23: 8786–8794, 2003.
85. Kwak MK, Wakabayashi N, Itoh K, Motohashi H, Yama-
moto M, Kensler TW. Modulation of gene expression by
cancer chemopreventive dithiolethiones through the
Keap1-Nrf2 pathway: identification of novel gene clusters
for cell survival. J Biol Chem 278: 8135–8145, 2003.
86. Lavrovsky Y, Schwartzman ML, Levere RD, Kappas A, and
Abraham NG. Identification of binding sites for transcrip-
tion factors NF-kappa B and AP-2 in the promoter region of
the human heme oxygenase 1 gene. Proc Natl Acad Sci U S A
91: 5987–5991, 1994.
87. Lee TS and Chau LY. Heme oxygenase-1 mediates the anti-
inflammatory effect of interleukin-10 in mice. Nat Med 8:
240–246, 2002.
88. Lernbecher T, Muller U, and Wirth T. Distinct NF-kappa
B=Rel transcription factors are responsible for tissue-specific
and inducible gene activation. Nature 365: 767–770, 1993.
89. Letai AG. Diagnosing and exploiting cancer’s addiction to
blocks in apoptosis. Nat Rev Cancer 8: 121–132, 2008.
90. Levonen AL, Inkala M, Heikura T, Jauhiainen S, Jyrkkanen
HK, Kansanen E, Maatta K, Romppanen E, Turunen P,
Rutanen J, and Yla-Herttuala S. Nrf2 gene transfer induces
antioxidant enzymes and suppresses smooth muscle cell
growth in vitro and reduces oxidative stress in rabbit aorta
in vivo. Arterioscler Thromb Vasc Biol 27: 741–747, 2007.
91. Lin R, Gewert D, and Hiscott J. Differential transcriptional
activation in vitro by NF-kappa B=Rel proteins. J Biol Chem
270: 3123–3131, 1995.
92. Linzer DI and Levine AJ. Characterization of a 54K dalton
cellular SV40 tumor antigen present in SV40-transformed cells
and uninfected embryonal carcinomacells. Cell 17:43–52, 1979.
93. Liu GH, Qu J, and Shen X. NF-kappaB=p65 antagonizes
Nrf2-ARE pathway by depriving CBP from Nrf2 and fa-
cilitating recruitment of HDAC3 to MafK. Biochim Biophys
Acta 1783: 713–727, 2008.
94. Liu YC, Hsieh CW, Wu CC, and Wung BS. Chalcone
inhibits the activation of NF-kappaB and STAT3 in endo-
thelial cells via endogenous electrophile. Life Sci 80: 1420–
1430, 2007.
95. Ma Q, Kinneer K, Bi Y, Chan JY, and Kan YW. Induction of
murine NAD(P)H:quinone oxidoreductase by 2,3,7,8-tetra-
chlorodibenzo-p-dioxin requires the CNC (cap ’n’ collar)
basic leucine zipper transcription factor Nrf2 (nuclear factor
erythroid 2-related factor 2): cross-interaction between AhR
(aryl hydrocarbon receptor) and Nrf2 signal transduction.
Biochem J 377: 205–213, 2004.
96. Matthews JR, Wakasugi N, Virelizier JL, Yodoi J, and Hay
RT. Thioredoxin regulates the DNA binding activity of NF-
kappa B by reduction of a disulphide bond involving cys-
teine 62. Nucleic Acids Res 20: 3821–3830, 1992.
97. Mattson MP and Camandola S. NF-kappaB in neuronal
plasticity and neurodegenerative disorders. J Clin Invest
107: 247–254, 2001.
1660 WAKABAYASHI ET AL.
98. Miao W, Hu L, Scrivens PJ, and Batist G. Transcriptional
regulation of NF-E2 p45-related factor (NRF2) expression
by the aryl hydrocarbon receptor-xenobiotic response ele-
ment signaling pathway: direct cross-talk between phase I
and II drug-metabolizing enzymes. J Biol Chem 280: 20340–
20348, 2005.
99. Mochizuki M, Ishii Y, Itoh K, Iizuka T, Morishima Y, Ki-
mura T, Kiwamoto T, Matsuno Y, Hegab AE, Nomura A,
Sakamoto T, Uchida K, Yamamoto M, and Sekizawa K.
Role of 15-deoxy delta(12,14) prostaglandin J2 and Nrf2
pathways in protection against acute lung injury. Am J
Respir Crit Care Med 171: 1260–1266, 2005.
100. Moll UM, Marchenko N, and Zhang XK. p53 and Nur77=TR3:
transcription factors that directly target mitochondria for cell
death induction. Oncogene 25: 4725–4743, 2006.
101. Mulcahy RT, Wartman MA, Bailey HH, and Gipp JJ.
Constitutive and beta-naphthoflavone-induced expression
of the human gamma-glutamylcysteine synthetase heavy
subunit gene is regulated by a distal antioxidant response
element=TRE sequence. J Biol Chem 272: 7445–7454, 1997.
102. Nair S, Doh ST, Chan JY, Kong AN, and Cai L. Regulatory
potential for concerted modulation of Nrf2- and Nfkb1-
mediated gene expression in inflammation and carcino-
genesis. Br J Cancer 99: 2070–2082, 2008.
103. Nguyen T, Huang HC, and Pickett CB. Transcriptional regu-
lation of the antioxidant response element: activation by Nrf2
and repression by MafK. JBiolChem275: 15466–15473, 2000.
104. Nioi P and Hayes JD. Contribution of NAD(P)H:quinone
oxidoreductase 1 to protection against carcinogenesis, and
regulation of its gene by the Nrf2 basic-region leucine
zipper and the arylhydrocarbon receptor basic helix-
loop-helix transcription factors. Mutat Res 555: 149–171, 2004.
105. Noda S, Harada N, Hida A, Fujii-Kuriyama Y, Motohashi
H, and Yamamoto M. Gene expression of detoxifying en-
zymes in AhR and Nrf2 compound null mutant mouse.
Biochem Biophys Res Commun 303: 105–111, 2003.
106. O’Dwyer PJ, Johnson SW, Khater C, Krueger A, Matsu-
moto Y, Hamilton TC, and Yao KS. The chemopreventive
agent oltipraz stimulates repair of damaged DNA. Cancer
Res 57: 1050–1053, 1997.
107. O’Reilly MA. Redox activation of p21Cip1=WAF1=Sdi1: a
multifunctional regulator of cell survival and death. Anti-
oxid Redox Signal 7: 108–118, 2005.
108. Ohtake F, Baba A, Takada I, Okada M, Iwasaki K, Miki H,
Takahashi S, Kouzmenko A, and Nohara K, Chiba T, Fujii-
Kuriyama Y, Kato S. Dioxin receptor is a ligand-dependent
E3 ubiquitin ligase. Nature 446: 562–566, 2007.
109. Ohtake F, Fujii-Kuriyama Y, and Kato S. AhR acts as an E3
ubiquitin ligase to modulate steroid receptor functions.
Biochem Pharmacol 77: 474–484, 2009.
110. Otterbein LE, Bach FH, Alam J, Soares M, Tao Lu H, Wysk
M, Davis RJ, Flavell RA, and Choi AM. Carbon monoxide
has anti-inflammatory effects involving the mitogen-
activated protein kinase pathway. Nat Med 6: 422–428, 2000.
111. Pae HO, Oh GS, Lee BS, Rim JS, Kim YM, and Chung HT.
3-Hydroxyanthranilic acid, one of L-tryptophan metabo-
lites, inhibits monocyte chemoattractant protein-1 secretion
and vascular cell adhesion molecule-1 expression via heme
oxygenase-1 induction in human umbilical vein endothelial
cells. Atherosclerosis 187: 274–284, 2006.
112. Pahl HL. Activators and target genes of Rel=NF-kappaB
transcription factors. Oncogene 18: 6853–6866, 1999.
113. Pan MH, Lin-Shiau SY, Lin JK. Comparative studies on the
suppression of nitric oxide synthase by curcumin and its
hydrogenated metabolites through down-regulation of
IkappaB kinase and NFkappaB activation in macrophages.
Biochem Pharmacol 60: 1665–1676, 2000.
114. Pena AS and Penate M. Genetic susceptibility and regulation
of inflammation in Crohn’s disease relationship with the in-
nate immune system. Rev Esp Enferm Dig 94: 351–360, 2002.
115. Prochaska HJ and Talalay P. Regulatory mechanisms of
monofunctional and bifunctional anticarcinogenic enzyme
inducers in murine liver. Cancer Res 48: 4776–4782, 1988.
116. Qin J, Clore GM, Kennedy WM, Huth JR, and Gronenborn
AM. Solution structure of human thioredoxin in a mixed
disulfide intermediate complex with its target peptide from
the transcription factorNF kappa B. Structure 3: 289–297, 1995.
117. Ramos-Gomez M, Kwak MK, Dolan PM, Itoh K, Yama-
moto M, Talalay P, and Kensler TW. Sensitivity to carci-
nogenesis is increased and chemoprotective efficacy of
enzyme inducers is lost in nrf2 transcription factor-deficient
mice. Proc Natl Acad Sci U S A 98: 3410–3415, 2001.
118. Rangarajan A, Talora C, Okuyama R, Nicolas M, Mam-
mucari C, Oh H, Aster JC, Krishna S, Metzger D, Chambon
P, Miele L, Aguet M, Radtke F, and Dotto GP. Notch sig-
naling is a direct determinant of keratinocyte growth arrest
and entry into differentiation. EMBO J 20: 3427–3436, 2001.
119. Rangasamy T, Guo J, Mitzner WA, Roman J, Singh A, Fryer
AD, Yamamoto M, Kensler TW, Tuder RM, Georas SN, and
Biswal S. Disruption of Nrf2 enhances susceptibility to se-
vere airway inflammation and asthma in mice. J Exp Med
202: 47–59, 2005.
120. Reddy NM, Kleeberger SR, Bream JH, Fallon PG, Kensler
TW, Yamamoto M, and Reddy SP. Genetic disruption of the
Nrf2 compromises cell-cycle progression by impairing GSH-
induced redox signaling. Oncogene 27: 5821–5832, 2008.
121. Reddy NM, Kleeberger SR, Yamamoto M, Kensler TW,
Scollick C, Biswal S, and Reddy SP. Genetic dissection of
the Nrf2-dependent redox signaling-regulated transcrip-
tional programs of cell proliferation and cytoprotection.
Physiol Genomics 32: 74–81, 2007.
122. Rosen ED. The transcriptional basis of adipocyte develop-
ment. Prostaglandins Leukot Essent Fatty Acids 73: 31–34, 2005.
123. Roshak AK, Callahan JF, and Blake SM. Small-molecule
inhibitors of NF-kappaB for the treatment of inflammatory
joint disease. Curr Opin Pharmacol 2: 316–321, 2002.
124. Ross DA, Hannenhalli S, Tobias JW, Cooch N, Shiekhattar
R, and Kadesch T. Functional analysis of Hes-1 in pre-
adipocytes. Mol Endocrinol 20: 698–705, 2006.
125. Ross JS, Stagliano NE, Donovan MJ, Breitbart RE, and
Ginsburg GS. Atherosclerosis: a cancer of the blood ves-
sels? Am J Clin Pathol 1160(suppl): S97–S107, 2001.
126. Rossi A, Kapahi P, Natoli G, Takahashi T, Chen Y, Karin M,
and Santoro MG. Anti-inflammatory cyclopentenone
prostaglandins are direct inhibitors of IkappaB kinase.
Nature 403: 103–108, 2000.
127. Rushworth SA, Chen XL, Mackman N, Ogborne RM, and
O’Connell MA. Lipopolysaccharide-induced heme oxyge-
nase-1 expression in human monocytic cells is mediated via
Nrf2 and protein kinase C. J Immunol 175: 4408–4415, 2005.
128. Rushworth SA, MacEwan DJ, and O’Connell MA. Lipo-
polysaccharide-induced expression of NAD(P)H:quinone
oxidoreductase 1 and heme oxygenase-1 protects against
excessive inflammatory responses in human monocytes.
J Immunol 181: 6730–6737, 2008.
129. Sablina AA, Budanov AV, Ilyinskaya GV, Agapova LS,
Kravchenko JE, and Chumakov PM. The antioxidant function
of the p53 tumor suppressor. Nat Med 11: 1306–1313, 2005.
NRF2 CROSS TALK 1661
130. Sander GR and Powell BC. Expression of notch receptors
and ligands in the adult gut. J Histochem Cytochem 52: 509–
516, 2004.
131. Sarady JK, Zuckerbraun BS, Bilban M, Wagner O, Usheva
A, Liu F, Ifedigbo E, Zamora R, Choi AM, and Otterbein
LE. Carbon monoxide protection against endotoxic shock
involves reciprocal effects on iNOS in the lung and liver.
FASEB J 18: 854–856, 2004.
132. Sartor MA, Schnekenburger M, Marlowe JL, Reichard JF,
Wang Y, Fan Y, Ma C, Karyala S, Halbleib D, Liu X,
Medvedovic M, and Puga A. Genomewide analysis of aryl
hydrocarbon receptor binding targets reveals an extensive
array of gene clusters that control morphogenetic and de-
velopmental programs. Environ Health Perspect 117: 1139–
1146, 2009.
133. Schmidt JV, Su GH, Reddy JK, Simon MC, and Bradfield
CA. Characterization of a murine Ahr null allele: involve-
ment of the Ah receptor in hepatic growth and develop-
ment. Proc Natl Acad Sci U S A 93: 6731–6736, 1996.
134. Schroder N and Gossler A. Expression of Notch pathway
components in fetal and adult mouse small intestine. Gene
Expr Patterns 2: 247–250, 2002.
135. Seldon MP, Silva G, Pejanovic N, Larsen R, Gregoire IP,
Filipe J, Anrather J, and Soares MP. Heme oxygenase-1
inhibits the expression of adhesion molecules associated
with endothelial cell activation via inhibition of NF-kappaB
RelA phosphorylation at serine 276. J Immunol 179: 7840–
7851, 2007.
136. Shimba S, Wada T, and Tezuka M. Arylhydrocarbon re-
ceptor (AhR) is involved in negative regulation of adipose
differentiation in 3T3-L1 cells: AhR inhibits adipose dif-
ferentiation independently of dioxin. J Cell Sci 114: 2809–
2817, 2001.
137. Shin S, Wakabayashi J, Yates MS, Wakabayashi N, Dolan
PM, Aja S, Liby KT, Sporn MB, Yamamoto M, and Kensler
TW. Role of Nrf2 in prevention of high-fat diet-induced
obesity by synthetic triterpenoid CDDO-imidazolide. Eur J
Pharmacol 620: 138–144, 2009.
138. Shin S, Wakabayashi N, Misra V, Biswal S, Lee GH,
Agoston ES, Yamamoto M, and Kensler TW. NRF2 mod-
ulates aryl hydrocarbon receptor signaling: influence on
adipogenesis. Mol Cell Biol 27: 7188–7197, 2007.
139. Soares MP, Seldon MP, Gregoire IP, Vassilevskaia T, Berberat
PO, Yu J, Tsui TY, and Bach FH. Heme oxygenase-1 modu-
lates the expression of adhesion molecules associated with
endothelial cell activation. J Immunol 172: 3553–23563, 2004.
140. Sriram N, Kalayarasan S, and Sudhandiran G. Epigalloca-
techin-3-gallate augments antioxidant activities and inhib-
its inflammation during bleomycin-induced experimental
pulmonary fibrosis through Nrf2-Keap1 signaling. Pulm
Pharmacol Ther 22: 221–236, 2009.
141. Srisook K and Cha YN. Super-induction of HO-1 in mac-
rophages stimulated with lipopolysaccharide by prior de-
pletion of glutathione decreases iNOS expression and NO
production. Nitric Oxide 12: 70–79, 2005.
142. Stier S, Cheng T, Dombkowski D, Carlesso N, and Scadden
DT. Notch1 activation increases hematopoietic stem cell
self-renewal in vivo and favors lymphoid over myeloid
lineage outcome. Blood 99: 2369–2378, 2002.
143. Surh YJ, Kundu JK, and Na HK. Nrf2 as a master redox
switch in turning on the cellular signaling involved in
the induction of cytoprotective genes by some chemo-
preventive phytochemicals. Planta Med 74: 1526–1539,
2008.
144. Tanaka Y, Aleksunes LM, Yeager RL, Gyamfi MA, Esterly
N, Guo GL, and Klaassen CD. NF-E2-related factor 2 in-
hibits lipid accumulation and oxidative stress in mice fed a
high-fat diet. J Pharmacol Exp Ther 325: 655–664, 2008.
145. Tenhunen R, Marver HS, and Schmid R. The enzymatic
conversion of heme to bilirubin by microsomal heme oxy-
genase. Proc Natl Acad Sci USA 61: 748–755, 1968.
146. Tennant RW, French JE, and Spalding JW. Identifying
chemical carcinogens and assessing potential risk in short-
term bioassays using transgenic mouse models. Environ
Health Perspect 103: 942–950, 1995.
147. Thimmulappa RK, Lee H, Rangasamy T, Reddy SP, Ya-
mamoto M, Kensler TW, and Biswal S. Nrf2 is a critical
regulator of the innate immune response and survival
during experimental sepsis. J Clin Invest 116: 984–995, 2006.
148. Thompson JE, Phillips RJ, Erdjument-Bromage H, Tempst
P, and Ghosh S. I kappa B-beta regulates the persistent
response in a biphasic activation of NF-kappa B. Cell 80:
573–582, 1995.
149. Tien AC, Rajan A, and Bellen HJ. A Notch updated. J Cell
Biol 184: 621–629, 2009.
150. Traenckner EB, Pahl HL, Henkel T, Schmidt KN, Wilk S,
and Baeuerle PA. Phosphorylation of human I kappa B-
alpha on serines 32 and 36 controls I kappa B-alpha pro-
teolysis and NF-kappa B activation in response to diverse
stimuli. EMBO J 14: 2876–2883, 1995.
151. Venugopal R and Jaiswal AK. Nrf1 and Nrf2 positively and
c-Fos and Fra1 negatively regulate the human antioxidant
response element-mediated expression of NAD(P)H:qui-
none oxidoreductase1 gene. Proc Natl Acad Sci U S A 93:
14960–14965, 1996.
152. Wakabayashi N, Itoh K, Wakabayashi J, Motohashi H,
Noda S, Takahashi S, Imakado S, Kotsuji T, Otsuka F, Roop
DR, Harada T, Engel JD, and Yamamoto M. Keap1-null
mutation leads to postnatal lethality due to constitutive
Nrf2 activation. Nat Genet 35: 238–245, 2003.
153. Wakabayashi N, Shin S, Slocum S, Agoston ES, Waka-
bayashi J, Kwak MK, Misra V, Biswal S, Yamamoto M,
and Kensler TW. Regulation of Notch1 signaling by Nrf2:
implications for tissue regeneration. Sci Signal (in press,
2010).
154. Wang XD, Shou J, Wong P, French DM, and Gao WQ.
Notch1-expressing cells are indispensable for prostatic
branching morphogenesis during development and re-
growth following castration and androgen replacement.
J Biol Chem 279: 24733–24744, 2004.
155. Wang XJ, Hayes JD, Henderson CJ, and Wolf CR. Iden-
tification of retinoic acid as an inhibitor of transcrip-
tion factor Nrf2 through activation of retinoic acid
receptor alpha. Proc Natl Acad Sci USA 104: 19589–19594,
2007.
156. Weiss G, Werner-Felmayer G, Werner ER, Grunewald K,
Wachter H, and Hentze MW. Iron regulates nitric oxide
synthase activity by controlling nuclear transcription. J Exp
Med 180: 969–976, 1994.
157. Whiteside ST, Epinat JC, Rice NR, and Israel A. I kappa B
epsilon, a novel member of the I kappa B family, controls RelA
and cRel NF-kappa B activity. EMBO J 16: 1413–1426, 1997.
158. Wilson A and Radtke F. Multiple functions of Notch sig-
naling in self-renewing organs and cancer. FEBS Lett 580:
2860–2868, 2006.
159. Wu Z, Bucher NL, and Farmer SR. Induction of peroxisome
proliferator-activated receptor gamma during the conver-
sion of 3T3 fibroblasts into adipocytes is mediated by
1662 WAKABAYASHI ET AL.
C=EBPbeta, C=EBPdelta, and glucocorticoids. Mol Cell Biol
16: 4128–4136, 1996.
160. Xiong Y, Hannon GJ, Zhang H, Casso D, Kobayashi R, and
Beach D. p21 is a universal inhibitor of cyclin kinases.
Nature 366: 701–704, 1993.
161. Yang H, Magilnick N, Lee C, Kalmaz D, Ou X, Chan JY,
and Lu SC. Nrf1 and Nrf2 regulate rat glutamate-cysteine
ligase catalytic subunit transcription indirectly via NF-
kappaB and AP-1. Mol Cell Biol 25: 5933–5946, 2005.
162. Yates MS, Tran QT, Dolan PM, Osburn WO, Shin S,
McCulloch CC, Silkworth JB, Taguchi K, Yamamoto M,
Williams CR, Liby KT, Sporn MB, Sutter TR, and Kensler
TW. Genetic versus chemoprotective activation of Nrf2
signaling: overlapping yet distinct gene expression profiles
between Keap1 knockout and triterpenoid-treated mice.
Carcinogenesis 30: 1024–1031, 2009.
163. Yeager RL, Reisman SA, Aleksunes LM, and Klaassen CD.
Introducing the ‘‘TCDD-inducible AhR-Nrf2 gene battery.’
Toxicol Sci 111: 2382–2346, 2009.
164. Yuan M, Konstantopoulos N, Lee J, Hansen L, Li ZW,
Karin M, and Shoelson SE. Reversal of obesity- and diet-
induced insulin resistance with salicylates or targeted dis-
ruption of IKKbeta. Science 293: 1673–1677, 2001.
165. Zabel U and Baeuerle PA. Purified human I kappa B can
rapidly dissociate the complex of the NF-kappa B tran-
scription factor with its cognate DNA. Cell 61: 255–265,
1990.
166. Zhang X, Chen X, Song H, Chen HZ, and Rovin BH.
Activation of the Nrf2=antioxidant response pathway
increases IL-8 expression. Eur J Immunol 35: 3258–3267,
2005.
167. Zhang Y, Lian JB, Stein JL, van Wijnen AJ, and Stein GS.
The Notch-responsive transcription factor Hes-1 attenuates
osteocalcin promoter activity in osteoblastic cells. J Cell
Biochem 108: 651–659, 2009.
168. Zhou W, Lo SC, Liu JH, Hannink M, and Lubahn DB.
ERRbeta: a potent inhibitor of Nrf2 transcriptional activity.
Mol Cell Endocrinol 278: 52–62, 2007.
Address correspondence to:
Thomas W. Kensler
Department of Pharmacology & Molecular Sciences
E1352 BST
University of Pittsburgh
Pittsburgh, PA 15261
E-mail: tkensler@pitt.edu
Date of first submission to ARS Central, March 26, 2010; date
of acceptance, April 1, 2010.
Abbreviations Used
15d-PGJ2 ¼15-deoxy D(12,14) prostaglandin J2
AhR ¼arylhydrocarbon receptor
ARE ¼antioxidant response element
ARNT ¼AhR nuclear translocator
CBP ¼CREB-binding protein
CDDO-Im ¼1-(2-cyano-3,12-dioxooleana-1,9[11]-
dien-28-oyl)imidazole
CEBPs ¼CCAAT-enhancer-binding proteins
COX-2 ¼cyclooxygenase-2
GST ¼glutathione S-transferase
HO-1 ¼heme oxygenase-1
IKK ¼IkB kinase
IL-1 ¼interleukin-1
iNOS ¼inducible NOS
KEAP1 ¼Kelch-like ECH-associated protein 1
LPS ¼lipopolysaccharide
MEF ¼mouse embryonic fibroblast
NF-kB¼nuclear factor kappa-light-chain-enhancer
of activated B cells
NICD ¼notch1 intracellular domain
NQO1 ¼NAD(P)H: quinone-acceptor 1
NRF2 ¼NF-E2–related factor 2
PPAR ¼peroxisome proliferator-activated
receptor
ROS ¼reactive oxygen species
Sulforaphane ¼(-)-1-isothiocyanato-(4R)-
methylsulfinyl)butane
TNF-a¼tumor necrosis factor a
TRX ¼thioredoxin
XRE ¼xenobiotic-responsive element
NRF2 CROSS TALK 1663
... In these contexts, the growth and the progression of SCLC was demonstrated to be related to the well-known molecular KEAP1/NRF2 axis, a master regulator of antioxidants and cellular stress responses, also implicated in the resistance of tumor cells against chemo-and radiotherapies [18,19]. The NRF2 is a key regulator of the cell's adaptative response to radical oxidant species and xenobiotics through the interaction with its negative regulator, Keap1, and contributes to cancer development, progression, and chemoresistance [18,20,21]. Molecular dysfunction of the KEAP1/NRF2 axis was mainly investigated in NSCLC, whereas few papers have elucidated the impact of this mechanism deregulation in SCLCs [18]. ...
Article
Full-text available
The KEAP1/NRF2 pathway is a master regulator of several redox-sensitive genes implicated in the resistance of tumor cells against therapeutic drugs. The dysfunction of the KEAP1/NRF2 system has been correlated with neoplastic patients’ outcomes and responses to conventional therapies. In lung tumors, the growth and the progression of cancer cells may also involve the intersection between the molecular NRF2/KEAP1 axis and other pathways, including NOTCH, with implications for antioxidant protection, survival of cancer cells, and drug resistance to therapies. At present, the data concerning the mechanism of aberrant NRF2/NOTCH crosstalk as well as its genetic and epigenetic basis in SCLC are incomplete. To better clarify this point and elucidate the contribution of NRF2/NOTCH crosstalk deregulation in tumorigenesis of SCLC, we investigated genetic and epigenetic dysfunctions of the KEAP1 gene in a subset of SCLC cell lines. Moreover, we assessed its impact on SCLC cells’ response to conventional chemotherapies (etoposide, cisplatin, and their combination) and NOTCH inhibitor treatments using DAPT, a γ-secretase inhibitor (GSI). We demonstrated that the KEAP1/NRF2 axis is epigenetically controlled in SCLC cell lines and that silencing of KEAP1 by siRNA induced the upregulation of NRF2 with a consequent increase in SCLC cells’ chemoresistance under cisplatin and etoposide treatment. Moreover, KEAP1 modulation also interfered with NOTCH1, HES1, and DLL3 transcription. Our preliminary data provide new insights about the downstream effects of KEAP1 dysfunction on NRF2 and NOTCH deregulation in this type of tumor and corroborate the hypothesis of a cooperation of these two pathways in the tumorigenesis of SCLC.
... Promising work has shown that bovine cells respond to putative NRF2 agonists and that NRF2 activation can limit oxidative dysfunction and improve cell viability [12][13][14]. In monogastric models, the known targets of NRF2 are well-established [15]; however, very few studies have incorporated a direct measure of NRF2 activity or expanded their transcriptomic analysis of NRF2 modulation beyond these known target genes. ...
Article
Full-text available
Changes during the production cycle of dairy cattle can leave these animals susceptible to oxidative stress and reduced antioxidant health. In particular, the periparturient period, when dairy cows must rapidly adapt to the sudden metabolic demands of lactation, is a period when the production of damaging free radicals can overwhelm the natural antioxidant systems, potentially leading to tissue damage and reduced milk production. Central to the protection against free radical damage and antioxidant defense is the transcription factor NRF2, which activates an array of genes associated with antioxidant functions and cell survival. The objective of this study was to evaluate the effect that two natural NRF2 modulators, the NRF2 agonist sulforaphane (SFN) and the antagonist brusatol (BRU), have on the transcriptome of immortalized bovine mammary alveolar cells (MACT) using both the RT-qPCR of putative NRF2 target genes, as well as RNA sequencing approaches. The treatment of cells with SFN resulted in the activation of many putative NRF2 target genes and the upregulation of genes associated with pathways involved in cell survival, metabolism, and antioxidant function while suppressing the expression of genes related to cellular senescence and DNA repair. In contrast, the treatment of cells with BRU resulted in the upregulation of genes associated with inflammation, cellular stress, and apoptosis while suppressing the transcription of genes involved in various metabolic processes. The analysis also revealed several novel putative NRF2 target genes in bovine. In conclusion, these data indicate that the treatment of cells with SFN and BRU may be effective at modulating the NRF2 transcriptional network, but additional effects associated with cellular stress and metabolism may complicate the effectiveness of these compounds to improve antioxidant health in dairy cattle via nutrigenomic approaches.
... Kelch-like ECH-associated protein 1 (Keap1) regulates the Nrf2 signaling pathway promoting Nrf2 degradation through ubiquitination. 17 According to the molecular docking studies performed between these SL analogues and the Kelch domain of Keap1 (PDB ID: 2flu), (S)-EGO10, (R)-EGO10, and (R)-IND showed higher binding affinities at the Nrf2 binding site of the Keap1, −8.22, −7.66, and −7.13 kcal/mol, respectively. EGCG is also used as a positive control due to its structural similarity to the SL analogues. ...
Article
Full-text available
Phytohormones have significant roles in redox metabolism, inflammatory responses, and cellular survival mechanisms within the microenvironment of the mammalian brain. Herein, we identified the mammalian molecular targets of three representative strigolactone (SL) analogues structurally derived from apocarotenoids and the functional equivalent of plant hormones. All tested SL analogues have an inhibitory effect on NLRP3 inflammasome-mediated IL-1β release in murine microglial cells. However, IND and EGO10 became prominent among them due to their high potency at low micromolar doses. All SL analogues dose-dependently suppressed the release and expression of proinflammatory factors. For EGO10 and IND, IC50 values for iNOS-associated NO secretion were as low as 1.72 and 1.02 μM, respectively. In silico analyses revealed that (S)-EGO10 interacted with iNOS, NLRP3, and Keap1 ligands with the highest binding affinities among all stereoisomeric SL analogues. Although all compounds were effective in microglial Mox phenotype polarization, 4-Br-debranone exhibited a differential pattern for upregulating Nrf2-driven downstream enzymes.
Article
Silk fibroin from Bombyx mori has traditionally been utilized for garment production; however, recently, it has become feasible to manufacture edible and cosmetic products using this material. Gamma‐ray irradiation can improve physiological activity through the modification of the structure of proteins in biological materials. In this study, we examined whether gamma ray‐irradiated fibroin protein (gFP) has a protective effect on tumor necrosis factor alpha (TNF‐ α )‐induced cellular stress through a change in protein structure. First, we confirmed that the structure of FP was changed by gamma‐ray irradiation using electrophoresis, UV spectra and Fourier‐transform infrared spectroscopy (FT‐IR). We further investigated the cytoprotective potential of 20 kGy‐irradiated fibroin protein (gFP20) in human skin keratinocytes (HaCaT cells) exposed to extracellular stress. gFP20 effectively decreased TNF‐ α ‐induced matrix metalloproteinase‐1 (MMP‐1) overexpression and restored reduced type‐1 procollagen in HaCaT cells. This reduction occurred concomitantly with suppressed nuclear factor‐ κ B (NF‐ κ B) nuclear translocation, resulting in a decreased level of inflammatory mediator and pro‐inflammatory cytokines. Furthermore, gFP20 protected cells from TNF‐ α ‐induced oxidative stress by attenuating reactive oxygen species (ROS) overexpression and increasing the production of the antioxidant enzyme heme oxygenase‐1 (HO‐1) through the stimulation of the translocation of nuclear factor erythroid 2‐related factor 2 (Nrf2) into the nucleus. Taken together, our findings indicate that low‐dose irradiated fibroin protein (gFP20) could be considered as a functional material for skincare products.
Article
Full-text available
NAD(P)H Quinone Dehydrogenase 1 (NQO1) plays a pivotal role in the regulation of neuronal function and synaptic plasticity, cellular adaptation to oxidative stress, neuroinflammatory and degenerative processes, and tumorigenesis in the central nervous system (CNS). Impairment of the NQO1 activity in the CNS can result in abnormal neurotransmitter release and clearance, increased oxidative stress, and aggravated cellular injury/death. Furthermore, it can cause disturbances in neural circuit function and synaptic neurotransmission. The abnormalities of NQO1 enzyme activity have been linked to the pathophysiological mechanisms of multiple neurological disorders, including Parkinson's disease, Alzheimer's disease, epilepsy, multiple sclerosis, cerebrovascular disease, traumatic brain injury, and brain malignancy. NQO1 contributes to various dimensions of tumorigenesis and treatment response in various brain tumors. The precise mechanisms through which abnormalities in NQO1 function contribute to these neurological disorders continue to be a subject of ongoing research. Building upon the existing knowledge, the present study reviews current investigations describing the role of NQO1 dysregulations in various neurological disorders. This study emphasizes the potential of NQO1 as a biomarker in diagnostic and prognostic approaches, as well as its suitability as a target for drug development strategies in neurological disorders.
Article
BACKGROUND Many chemotherapeutic drugs, including paclitaxel, produce neuropathic pain in patients with cancer, which is a dose-dependent adverse effect. Such chemotherapy-induced neuropathic pain (CINP) is difficult to treat with existing drugs. Nuclear factor erythroid 2-related factor 2 (Nrf2) is a major regulator of antioxidative responses and activates phosphorylated Nrf2 (pNrf2). We determined the analgesic effects of bardoxolone methyl (BM), an Nrf2 activator, and the role of pNrf2 on CINP. METHODS CINP was induced in rats by intraperitoneally injecting paclitaxel on 4 alternate days in rats. BM was injected systemically as single or repeated injections after pain fully developed. RNA transcriptome, mechanical hyperalgesia, levels of inflammatory mediators and pNrf2, and location of pNrf2 in the dorsal root ganglia (DRG) were measured by RNA sequencing, von Frey filaments, Western blotting, and immunohistochemistry in rats and human DRG samples. In addition, the mitochondrial functions in 50B11 DRG neuronal cells were measured by fluorescence assay. RESULTS Our RNA transcriptome of CINP rats showed a downregulated Nrf2 pathway in the pain condition. Importantly, single and repeated systemic injections of BM ameliorated CINP. Paclitaxel increased inflammatory mediators, but BM decreased them and increased pNrf2 in the DRG. In addition, paclitaxel decreased mitochondrial membrane potential and increased mitochondrial volume in 50B11 cells, but BM restored them. Furthermore, pNrf2 was expressed in neurons and satellite cells in rat and human DRG. CONCLUSIONS Our results demonstrate the analgesic effects of BM by Nrf2 activation and the fundamental role of pNrf2 on CINP, suggesting a target for CINP and a therapeutic strategy for patients.
Article
Full-text available
Induction of phase 2 enzymes, which neutralize reactive electrophiles and act as indirect antioxidants, appears to be an effective means for achieving protection against a variety of carcinogens in animals and humans. Transcriptional control of the expression of these enzymes is mediated, at least in part, through the antioxidant response element (ARE) found in the regulatory regions of their genes. The transcription factor Nrf2, which binds to the ARE, appears to be essential for the induction of prototypical phase 2 enzymes such as glutathione S-transferases (GSTs) and NAD(P)H: quinone oxidoreductase (NQO1), Constitutive hepatic and gastric activities of CST and NQO1 were reduced by 50-80% in nrf2-deficient mice compared with wild-type mice. Moreover, the 2- to 5-fold induction of these enzymes in wild-type mice by the chemoprotective agent oltipraz, which is currently in clinical trials, was almost completely abrogated in the nrf2-deficient mice. In parallel with the enzymatic changes, nrf2-deficient mice had a significantly higher burden of gastric neoplasia after treatment with benzo[a]pyrene than did wild-type mice. Oltipraz significantly reduced multiplicity of gastric neoplasia in wild-type mice by 55%, but had no effect on tumor burden in nrf2-deficient mice. Thus, Nrf2 plays a central role in the regulation of constitutive and inducible expression of phase 2 enzymes in vivo and dramatically influences susceptibility to carcinogenesis. Moreover, the total loss of anticarcinogenic efficacy of oltipraz in the nrf2-disrupted mice highlights the prime importance of elevated phase 2 gene expression in chemoprotection by this and similar enzyme inducers.
Article
Full-text available
Activated macrophages express high levels of Nrf2, a transcription factor that positively regulates the gene expression of antioxidant and detoxication enzymes. In this study, we examined how Nrf2 contributes to the anti-inflammatory process. As a model system of acute inflammation, we administered carrageenan to induce pleurisy and found that in Nrf2-deficient mice, tissue invasion by neutrophils persisted during inflammation and the recruitment of macrophages was delayed. Using an antibody against 15-deoxy-Δ12,14-prostaglandin J2 (15d-PGJ2), it was observed that macrophages from pleural lavage accumulate 15d-PGJ2. We show that in mouse peritoneal macrophages 15d-PGJ2 can activate Nrf2 by forming adducts with Keap1, resulting in an Nrf2-dependent induction of heme oxygenase 1 and peroxiredoxin I (PrxI) gene expression. Administration of the cyclooxygenase 2 inhibitor NS-398 to mice with carrageenan-induced pleurisy caused persistence of neutrophil recruitment and, in macrophages, attenuated the 15d-PGJ2 accumulation and PrxI expression. Administration of 15d-PGJ2 into the pleural space of NS-398-treated wild-type mice largely counteracted both the decrease in PrxI and persistence of neutrophil recruitment. In contrast, these changes did not occur in the Nrf2-deficient mice. These results demonstrate that Nrf2 regulates the inflammation process downstream of 15d-PGJ2 by orchestrating the recruitment of inflammatory cells and regulating the gene expression within those cells.
Article
The role of Notch signaling in growth/differentiation control of mammalian epithelial cells is still poorly defined. We show that keratinocyte-specific deletion of the Notch1 gene results in marked epidermal hyperplasia and deregulated expression of multiple differentiation markers. In differentiating primary keratinocytes in vitro endogenous Notch1 is required for induction of p21WAF1/Cip1 expression, and activated Notch1 causes growth suppression by inducing p21WAF1/Cip1 expression. Activated Notch1 also induces expression of 'early' differentiation markers, while suppressing the late markers. Induction of p21WAF1/Cip1 expression and early differentiation markers occur through two different mechanisms. The RBP-Jkappa protein binds directly to the endogenous p21 promoter and p21 expression is induced specifically by activated Notch1 through RBP-Jkappa-dependent transcription. Expression of early differentiation markers is RBP-Jkappa-independent and can be induced by both activated Notch1 and Notch2, as well as the highly conserved ankyrin repeat domain of the Notch1 cytoplasmic region. Thus, Notch signaling triggers two distinct pathways leading to keratinocyte growth arrest and differentiation.
Article
Glutathione (GSH) is an abundant cellular non-protein sulfhydryl that functions as an important protectant against reactive oxygen species and electrophiles, is involved in the detoxification of xenobiotics, and contributes to the maintenance of cellular redox balance, The rate-limiting enzyme in the de novo synthesis of glutathione is gamma-glutamylcysteine synthetase (GCS), a heterodimer consisting of heavy and light subunits expressing catalytic and regulatory functions, respectively, Exposure of HepG2 cells to beta-naphthoflavone (beta-NF) resulted in a time- and dose-dependent increase in the steady-state mRNA levels for both subunits, In order to identify sequences mediating the constitutive and induced expression of the heavy subunit gene, a series of deletion mutants created from the 5'-flanking region (-3802 to +465) were cloned into a luciferase reporter vector (pGL3-Basic) and transfected into HepG2 cells, Constitutive expression was maximally directed by sequences between -202 and +22 as well as by elements between -3802 to -2752, The former sequence contains a consensus TATA box, Increased luciferase expression following exposure to 10 mu M beta-NF was only detected in cells transfected with a reporter vector containing the full-length -3802:+465 fragment, Hence, elements directing constitutive and induced expression of the GCS heavy subunit are present in the distal portion of the 5'-flanking region, between positions -3802 and -2752, Sequence analysis revealed the presence of several putative consensus response elements in this region, including two potential antioxidant response elements (ARE3 and ARE4), separated by 34 base pairs. When cloned into the thymidine kinase-luciferase vector, pT81-luciferase, and transfected into HepG2 cells, both ARE3 and ARE4 increased basal luciferase expression approximately 20-fold. When cloned in tandem in their native arrangement the increase in luciferase activity was in excess of 100-fold, suggesting a strong interaction between the two sequences, Luciferase expression was elevated in beta-NF-treated cells transfected with the AREA-th-luciferase vector and all DNA fragments containing ARE4. In contrast, ARE3 did not direct increased luciferase expression in response to beta-NF nor did it significantly modify the magnitude of induction directed by ARE4. The influence of the ARE4 oligonucleotide on constitutive and induced expression was eliminated by introduction of a single base mutation, converting the core ARE sequence in ARE4 fi om 5'-GTGACTCAGCG-3' to 5'-GGGACTCAGCG-3'. When introduced into the full-length -3802:+465 segment, the same single base mutation also eliminated both functions, Collectively the data indicate that the constitutive and P-NF-induced expression of the human GCS heavy subunit gene is mediated by a distal ARE sequence containing an embedded tetradecanoylphorbol-13-acetate-responsive element.
Article
1. 1. Kimura, 2. et al . 2009. J. Exp. Med. doi: 10.1084/jem.20090560 [OpenUrl][1][Abstract/FREE Full Text][2] [1]: {openurl}?query=rft_id%253Dinfo%253Adoi%252F10.1084%252Fjem.20090560%26rft_id%253Dinfo%253Apmid%252F19703987%26rft.genre%253Darticle%26rft_val_fmt%253Dinfo
Chapter
Sixteen years have passed since the description of the nuclear factor-кB (NF-кB) as a regulator of к light-chain gene expression in murine B lymphocytes (Sen & Baltimore, 1986a). During that time, over 4,000 publications have appeared, characterizing the family of Rel/NF-кB transcription factors involved in the control of a large number of normal and pathological cellular processes. The physiological functions of NF-кB proteins include immunological and inflammatory responses, developmental processes, cellular growth and modulating effects on apoptosis. In addition, these factors are activated in a number of diseases, including cancer, arthritis, acute and chronic inflammatory states, asthma, as well as neurodegenerative and heart diseases.
Article
Recently, it was reported that nitric oxide (NO) directly controls intracellular iron metabolism by activating iron regulatory protein (IRP), a cytoplasmic protein that regulates ferritin translation. To determine whether intracellular iron levels themselves affect NO synthase (NOS), we studied the effect of iron on cytokine-inducible NOS activity and mRNA expression in the murine macrophage cell line J774A.1. We show here that NOS activity is decreased by about 50% in homogenates obtained from cells treated with interferon gamma plus lipopolysaccharide (IFN-gamma/LPS) in the presence of 50 microM ferric iron [Fe(3+)] as compared with extracts from cells treated with IFN-gamma/LPS alone. Conversely, addition of the iron chelator desferrioxamine (100 microM) at the time of stimulation with IFN-gamma/LPS increases NOS activity up to 2.5-fold in J774 cells. These effects of changing the cellular iron state cannot be attributed to a general alteration of the IFN-gamma/LPS signal, since IFN-gamma/LPS-mediated major histocompatibility complex class II antigen expression is unaffected. Furthermore, neither was the intracellular availability of the NOS cofactor tetrahydrobiopterin altered by treatment with Fe(3+) or desferrioxamine, nor do these compounds interfere with the activity of the hemoprotein NOS in vitro. We demonstrate that the mRNA levels for NOS are profoundly increased by treatment with desferrioxamine and reduced by Fe(3+). The half-life of NOS mRNA appeared not to be significantly altered by administration of ferric ion, and NOS mRNA stability was only slightly prolonged by desferrioxamine treatment. Nuclear run-off experiments demonstrate that nuclear transcription of cytokine-inducible NOS mRNA is strongly increased by desferrioxamine whereas it is decreased by Fe(3+). Thus, this transcriptional response appears to account quantitatively for the changes in enzyme activity. Our results suggest the existence of a regulatory loop between iron metabolism and the NO/NOS pathway.