ArticlePDF Available

Heme oxygenase-1 induction by NRF2 requires inactivation of the transcriptional repressor BACH1

Authors:

Abstract and Figures

Oxidative stress activates the transcription factor NRF2, which in turn binds cis-acting antioxidant response element (ARE) enhancers and induces expression of protective antioxidant genes. In contrast, the transcriptional repressor BACH1 binds ARE-like enhancers in cells naïve to oxidative stress and antagonizes NRF2 binding until it becomes inactivated by pro-oxidants. Here, we describe the dynamic roles of BACH1 and NRF2 in the transcription of the heme oxygenase-1 (HMOX1) gene. HMOX1 induction, elicited by arsenite-mediated oxidative stress, follows inactivation of BACH1 and precedes activation of NRF2. BACH1 repression is dominant over NRF2-mediated HMOX1 transcription and inactivation of BACH1 is a prerequisite for HMOX1 induction. In contrast, thioredoxin reductase 1 (TXNRD1) is regulated by NRF2 but not by BACH1. By comparing the expression levels of HMOX1 with TXNRD1, we show that nuclear accumulation of NRF2 is not necessary for HMOX1 induction; rather, BACH1 inactivation permits NRF2 already present in the nucleus at low basal levels to bind the HMOX1 promoter and elicit HMOX1 induction. Thus, BACH1 confers an additional level of regulation to ARE-dependent genes that reveals a new dimension to the oxidative stress response.
Identification of NRF2 and BACH1 interactions with core ARE motifs of HMOX1. (A) The position of all 12 putative ARE motifs, relative to the HMOX1 transcription start site as annotated by the NCBI Homo sapiens Genome Map Viewer, Build 36.2. Open boxes = motif conforming to the consensus ARE sequence, shaded boxes = imperfect ARE motif. Boxed numerals indicate the number of motif repeats in that region. Arrows indicate the plus-strand orientation of each ARE motif relative to the consensus ARE sequence (RTGAYnnnGC). Relative DNA enrichment associated with NRF2 (B) and BACH1 (C) ChIP at each of the HMOX1 ARE-containing sites. Immunoprecipitated DNA was analyzed for enrichment by qRT–PCR using primers flanking ARE motifs at the positions indicated. Vertical bars and circle symbols graphically depict the relative magnitude of DNA enrichment at each position. Vertical bars indicate enrichment at positions −3928 bp (E1) and −9069 bp (E2) while circles correspond with regions where negligible DNA enrichment was observed. Values oriented vertically along the abscissa indicate the position of each DNA region amplified by qRT–PCR. Amplification was expressed in terms of ΔΔCt as calculated by normalizing the Ct for each primer in chromatin immunoprecipitated samples to the Ct obtained from the respective input DNA and expressed as a percentage relative to the PCR primer having the maximum enrichment (–9069 bp). Values represent the mean ± SEM of at least three independent experiments performed in triplicate.
… 
Content may be subject to copyright.
Nucleic Acids Research, 2007, 1–13
doi:10.1093/nar/gkm638
Heme oxygenase-1 induction by NRF2 requires
inactivation of the transcriptional repressor BACH1
John F. Reichard*, Gregory T. Motz and Alvaro Puga
Department of Environmental Health and Center for Environmental Genetics, University of Cincinnati, Cincinnati,
OH 45267-0056, USA
Received May 21, 2007; Revised July 12, 2007; Accepted August 1, 2007
ABSTRACT
Oxidative stress activates the transcription factor
NRF2, which in turn binds cis-acting antioxidant
response element (ARE) enhancers and induces
expression of protective antioxidant genes. In con-
trast, the transcriptional repressor BACH1 binds
ARE-like enhancers in cells naı¨ve to oxidative stress
and antagonizes NRF2 binding until it becomes
inactivated by pro-oxidants. Here, we describe the
dynamic roles of BACH1 and NRF2 in the transcrip-
tion of the heme oxygenase-1 (HMOX1) gene.
HMOX1 induction, elicited by arsenite-mediated
oxidative stress, follows inactivation of BACH1 and
precedes activation of NRF2. BACH1 repression is
dominant over NRF2-mediated HMOX1 transcription
and inactivation of BACH1 is a prerequisite for
HMOX1 induction. In contrast, thioredoxin reduc-
tase 1 (TXNRD1) is regulated by NRF2 but not by
BACH1. By comparing the expression levels of
HMOX1 with TXNRD1, we show that nuclear accu-
mulation of NRF2 is not necessary for HMOX1
induction; rather, BACH1 inactivation permits NRF2
already present in the nucleus at low basal levels
to bind the HMOX1 promoter and elicit HMOX1
induction. Thus, BACH1 confers an additional level
of regulation to ARE-dependent genes that reveals
a new dimension to the oxidative stress response.
INTRODUCTION
Reactive oxygen species (ROS) pose a serious threat to all
aerobic organisms that maintain redox homeostasis in the
face of constant exposure to environmental oxidants. ROS
are harmful due to their reactivity with many cellular
macromolecules. To maintain cells in a state of redox
balance, biochemical antioxidants and a host of anti-
oxidant enzymatic reactions supply the needed reduction
potential. Antioxidant defenses exist in a balance with
endogenous oxidants, and it is the disruption of this
balance that characterizes the pathogenesis of many
human diseases and aging.
Arsenite, the trivalent form of inorganic arsenic, is an
environmental contaminant of major concern. Arsenic is
a potent electrophilic inducer of oxidative stress with
many of its effects attributable to its affinity for soft
nucleophiles, such as cysteine residues in glutathione
(GSH) and proteins (1). Arsenite exposure results in
rapid oxidation of glutathione (2) thereby disrupting
intracellular redox status (3). In response to the oxidative
stress mediated by arsenite, cells induce the antioxidant
battery of protective enzymes, of which heme oxygenase-1
(HMOX1) and thioredoxin reductase-1 (THXRD1) are
two well-recognized members.
Transactivation of HMOX1 and of other antioxidant
genes is regulated by binding of the transcription factor
NRF2 (Nuclear factor erythoid-derived 2 related factor 2)
to a cis-acting enhancer element known as the antioxidant
response element (ARE). Activation of NRF2 requires its
translocation to the nucleus, formation of a transcription-
ally active complex through dimerization with small MAF
(sMAF) proteins and binding to ARE enhancer motifs (4).
The ARE is one form of MAF response element (MARE)
having the core sequence RTGAYNNNGC (reverse
complement: GCNNNRTGAY) (5) and additional flank-
ing nucleotides that increase the specificity of NRF2
recognition (6,7). Some variations of the ARE motif can
be recognized by other regulatory factors, in addition to
NRF2, including NRF1 (8), NRF3 (9) and BACH1 (
BTB
and CNC homolog 1) (10). Thus, variation of ARE motif
sequences contribute to overlapping DNA binding by
factors that compete with NRF2 for ARE binding (11).
This arrangement appears to exist between NRF2 and
BACH1.
BACH1 is a transcriptional repressor (12) that is
conserved and ubiquitously expressed in tissues (13,14)
though its global activity in gene regulation is poorly
characterized. Like several other MAF-related trans-
cription factors, particularly NRF2, BACH1 hetero-
dimerizes with sMAF proteins in order to bind DNA.
In the basal state, BACH1/sMAF heterodimers interact
with the MARE-like enhancer sites recognized by NRF2
*To whom correspondence should be addressed. Tel: +1 513 558 0712; Fax: +1 513 558 0925; Email: john.reichard@childrens.harvard.edu
ß 2007 The Author(s)
This is an Open Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License (http://creativecommons.org/licenses/
by-nc/2.0/uk/) which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.
Nucleic Acids Research Advance Access published October 16, 2007
or NF-E2 to inhibit expression of the corresponding genes
(13,15). As a repressive transcription factor, BACH1
allows gene induction upon its release from enhancer
elements. Originally, BACH1 was described as a heme-
regulated repressor of b-globin genes (12) and HMOX1
(14–18). More recently, BACH1 has been suggested to
play a role as a sensor of oxidative stress. Human BACH1
is a thiol-rich protein possessing 34 interspersed cysteine
amino acids, of which two are responsible for BACH1
inactivation by oxidants (17). Therefore, it is possible that
heme and oxidants trigger gene induction simply by
relieving BACH1 repression (19). In this regard, it has
been demonstrated that BACH1 plays a role in redox
induction of HMOX1 (17) and NQO1 (10), though the
exact mechanism of this repression is not clear.
The relationship between BACH1 inactivation and
NRF2 activation during the induction of antioxidant
genes is unknown. NRF2 coordinates induction of genes
through its interaction with ARE enhancer motifs,
frequently located 5
0
to the transcriptional start site
(TSS) of several well-characterized antioxidant genes (5).
In the absence of oxidative stress, the cytosolic protein
KEAP1 (Kelch-like ECH Associated Protein 1) directs E3
ligase-dependent proteasomal degradation of newly
synthesized NRF2. As a result of this continuous
degradation, NRF2 is effectively maintained at very low
cellular levels. KEAP1, like BACH1, is a thiol-rich
protein, and oxidation of a few of KEAP1’s cysteines
blocks NRF2 degradation (20–22). Consequently, activa-
tion of NRF2 is dependent on nuclear accumulation of
de novo synthesized protein for subsequent binding of
ARE motifs (23). At a few known genes, including
HMOX1, BACH1 binds ARE motifs to the exclusion
of NRF2.
Here, we have tested whether oxidative stress induced
by arsenite exposure can trigger gene induction by simply
releasing repressive BACH1 from ARE motifs. We find
that NRF2 and BACH1 bind to two distal ARE enhancer
sites far upstream of the HMOX1 TSS and that BACH1
removal is necessary for NRF2-mediated HMOX1 gene
induction. In contrast to HMOX1, TXNRD1, which has
a single ARE motif located 9 bp upstream of the TSS,
is regulated by NRF2 but not by BACH1. Comparison of
TXNRD1 expression with that of HMOX1 demonstrates
that BACH1 repression is dominant over NRF2-mediated
transcription. The dynamic interplay between BACH1
and NRF-2 can produce distinct regulatory expression
patterns at different oxidative stress-induced genes.
MATERIALS AND METHODS
Cells and chemical treatments
HaCaT (24) human keratinocytes were grown in DMEM
(Mediatech, Herndon, VA, USA) supplemented with 5%
fetal bovine serum (Sigma Aldrich, St Louis, MO, USA)
and 1% penicillin–streptomycin solution (Invitrogen,
Carlsbad, CA, USA). Cells were incubated at 378C with
5% CO
2
and grown to 90% confluence before treat-
ment. Aqueous solutions of NaAsO
2
(hereafter referred
to as arsenite) (Sigma Aldrich) were prepared from a
1000 stock in ddH
2
O.
Whole cell extracts, nuclear extracts and immunoblotting
For preparation of whole cell extracts, cells were washed
and harvested in PBS containing 1 Complete Protease
Inhibitor
Õ
(Roche Diagnostics Corporation, Indianapolis,
IN) and lyzed by sonication (Sonic Dismembrator 60,
Fisher Scientific) on ice 3 10 s in 300 ml NETN buffer
(100 mM NaCl, 20 mM Tris pH 8.0, 1 mM EDTA, 0.5%
NP-40, 1 Complete Protease Inhibitor). Nuclear and
cytosolic extracts were prepared using a nuclear extraction
kit from Panomics (Fremont, CA, USA) according to
the manufacturer’s protocol. Protein concentrations were
measured using the Bradford assay and concentrations
adjusted to 1 mg/ml. Proteins were separated by SDS–
PAGE, transferred to PVDF and probed for NRF2
(H300; Santa Cruz Biotechnology), b-actin (Sigma),
HMOX1 (C20; Santa Cruz Biotechnology), in blocking
buffer containing 3% NFDM in PBST (0.1 M PBS
with 0.2% Tween 20). Blots were probed for BACH1
(C20; Santa Cruz Biotechnology) in blocking buffer
containing 1% BSA in PBST. After washing, the blots
were incubated with species-appropriate HRP-conjugated
secondary antibody (Santa Cruz), incubated with chemi-
luminescent reagent (PicoWest Super Signal, Pierce
Rickford, IL, USA) and visualized by exposing to film
(GE Healthcare, Piscataway, NJ, USA).
RNA isolation and real-time RT–PCR
Total RNA was extracted using NucleoSpin RNA II
columns (Macherey-Nage, Bethlehem, PA) according to the
manufacturer’s protocol. c DNA was synthesized by reverse
transcription of 1 mg total RNA in a total volume of 20 ml
containing 1 reverse transcriptase buffer (Invitrogen,
Carlsbad, CA, USA), 25 mg/ml oligo (dT)
12–18
(Invitrogen),
0.5 mM dNTP mix (GeneChoice, Frederick, MD), 10 mM
dithiothreitol (Invitrogen), 20 U of RNase inhibitor
(RNasin
Õ
; Promega, Madison WI) and 100 U of
SuperScript
II reverse transcriptase (Invitrogen). Resulting
cDNA products were diluted in a final volume of 200 mlanda
2 ml aliquot was used as template for subsequent quantifica-
tion by real-time PCR amplification. Quantitative real-time
PCR was performed in a 25 ml reaction mixture containing
1 Clonetech QTaq polymerase reaction mix (Mountain
View,CA),1 SybrGreen (Invitrogen) as a marker of
amplification, 0.1 mMofeachprimerand2mloftemplate
cDNA. Products were amplified with human HMOX1
primers (forward: 5
0
-CTCAAACCTCCAAAAGCC-3
0
and
reverse: 5
0
-TCAAAAACCACCCCAACCC-3
0
), TXNRD1
primers (forward: 5
0
-CTTTTTCATTCCTGCTACTC-
TACC-3
0
and reverse: 5
0
-CTCTCTCCTTTTCCC
TTTTCC-3
0
), BACH1 primers (forward: 5
0
-TGCGATGT-
CACCATCTTTGT-3
0
and reverse: 5
0
-CCTGGCCTAC-
GATTCTTGAG-3
0
), NRF2 primers (forward: 5
0
-GAGA
GCCCAGTCTTCATTGC-3
0
and reverse: 5
0
-TGCT
CAATGTCCTGTTGCAT-3
0
). Amplifications were per-
formed using an Opticon 2 Real-Time PCR Detection
System (MJ research). Cycle threshold (C
t
)ofeachsample
was automatically determined to be the first cycle at which a
2
Nucleic Acids Research, 2007
significant increase in optical signal above an arbitrary
baseline was detected. Amplification of b-actin cDNA in the
same samples was used as an internal control for all PCR
amplification reactions. Relative mRNA expression was
quantified using the comparative C
t
(C
t
) method and
expressed as 2
C
t
. Each assay was done in triplicate.
Chromatin immunoprecipitation (ChIP) analysis
Nuclei were isolated from formaldehyde (1% final) fixed
cells by lysing cells in buffer containing 5 mM PIPES
(pH 8.0), 85 mM KCl, 0.5% NP-40 and 1 Complete
Protease Inhibitor
Õ
. DNA was fragmented by sonication
(Diagenode Bioruptor) in buffer containing 50 mM Tris–
HCl pH 8.1, 10 mM EDTA, 1% SDS, 1 Complete
Protease Inhibitor
Õ
. DNA-cross-linked proteins were
immunoprecipitated from precleared samples with 2–5 mg
of each specific antibodies, including anti-NRF2, anti-
BACH1 or rabbit IgG control (Millipore), or for use as
total input chromatin. Antibodies were pulled down with
protein A beads at 48C overnight. Recovered beads were
re-suspended in 1 dialysis buffer [2 mM EDTA, 50 mM
Tris–HCl (pH 8.0), 0.2% Sarkosyl and 1 Complete
Protease Inhibitor
Õ
] and washed twice with dialysis buffer,
following which pellets were washed four times with IP
wash buffer [100 mM Tris–HCl (pH 9.0), 500 mM LiCl,
1% NP-40, 1% deoxycholic acid and 1 Complete
Protease Inhibitor
Õ
]. Antibody/protein/DNA complexes
were eluted in IP elution buffer (50 mM NaHCO
3
,1%
SDS) with vigorous shaking. Immunoprecipitated DNA
and total input chromatin were diluted to 120 ml with
water, brought up to 0.3 M NaCl. Cross-linking was
reversed at 658C overnight in the presence of RNase A
(10 mg), followed by the addition of RNase A followed by
protease K digestion at 458C for 2 h. DNA was purified
using a Qiaquick PCR purification kit per manufacturer
protocols. Samples were evaluated for enrichment by
quantitative real-time PCR (qRT–PCR) in a 25 ml reaction
mixture containing 1 BD QTaq polymerase reaction mix
(BD Biosciences), 1 SybrGreen (Invitrogen) as a marker
of DNA amplification, and 0.1 mM of each primer
(Supplementary Table 1). ChIP enrichment was evaluated
using primers to regions upstream of the human HMOX1
or TXNRD1 genes (Table S1). Relative efficiency of each
PCR primer was determined using input DNA and
adjusted accordingly. The DNA in each ChIPed sample
was normalized to the corresponding input chromatin
(C
t
) and enrichment was defined as the change in C
t
in
treated samples relative to untreated controls (C
t
),
relative to the IgG negative control. Exponential C
t
values were converted to linear values (2
C
t
) for
graphical presentation as either as fold change or percent
change, where indicated.
RESULTS
BACH1 inactivation precedes NRF2 activation
To characterize activation of the antioxidant response
following arsenite exposure, we first characterized
cellular and subcellular changes in NRF2 and
BACH1 disposition. In control cells, NRF2 is present at
very low levels in whole cell extracts attributable to
ongoing proteasomal degradation mediated by KEAP1
(25). After treatment with 25 mM arsenite, NRF2 accu-
mulates to high levels in whole cell extracts consistent with
oxidative inactivation of KEAP1 (23) (Figure 1A and B).
Importantly, NRF2 accumulation is negligible during the
first hour after arsenite treatment, and does not reach
maximum levels until 3 h after treatment. This initial lag
period represents the time necessary for translational
synthesis of new NRF2 protein (26), which precedes
nuclear translocation and subsequent ARE-mediated gene
induction. In comparison to NRF2, BACH1 gradually
accumulates in whole cell extracts over 5 h after treatment.
The extent to which BACH1 accumulation is attributable
to protein stabilization is unknown; however, no change
in BACH1 mRNA expression was observed (data not
shown).
Nuclear localization of both NRF2 and BACH1 is
essential for their transcriptional activity and can be used
as an indicator of their activation. To investigate their
temporal activation following arsenite treatment, changes
in nuclear and cytosolic distribution were examined
(Figure 1C and D). Nuclear NRF2 accumulation parallels
that observed in whole cell extracts. Prior to treatment,
very low levels of NRF2 are present in nuclear fractions.
Consistent with NRF2 levels in whole cell extracts,
nuclear NRF2 accumulation is negligible during the first
hour after arsenite treatment but becomes highly elevated
by 3 h after treatment. Throughout the time course, and
despite the accumulation of NRF2 in whole cell and
nuclear fractions, there is little or no NRF2 accumulation
in the cytoplasmic fraction. Rather, whenever detected,
NRF2 is always associated with nuclear fractions,
consistent with the rapid transport of NRF2 into the
nucleus after its synthesis. Detection of NRF2 in nuclear
fractions of control cells shows that even in the absence
of oxidant exposure a low level of NRF2 resides in the
nucleus.
In contrast to NRF2, BACH1 is rapidly exported from
the nucleus following arsenite treatment, reaching minimal
nuclear levels by 30 min after treatment, corresponding
with an increase in cytosolic levels. BACH1 levels in the
cytosol increase by 2-fold during the first hour after
arsenite treatment and reach maximum observed levels
by 3 h after treatment (Figure 1C and D).
Oxidative stress and arsenite are known to strongly
induce HMOX1. Time-course analyses following 25 mM
arsenite treatment shows a rapid induction of HMOX1
mRNA following an initial 1-h lag period (Figure 1F),
coinciding with the time required for nuclear localiza-
tion of de novo synthesized NRF2 protein. This induction
continues in a linear fashion through 8 h after treat-
ment without achieving a steady-state plateau. Together,
these results show that HMOX1 transcription is
preceded by BACH1 inactivation and occurs in parallel
with, rather than following, NRF2 activation, suggest-
ing that BACH1 inactivation is the antecedent
event corresponding with transcriptional initiation of
HMOX1.
Nucleic Acids Research, 2007 3
Bach1
Nuclear
Actin
Nrf2
Cytosol
Hours
Whole cell lysate
Hours 0.5 1 2 3 4 50
0 0.5 1 3 0 0.5 1 3 N0
0.5 1 2 3 4 50
Bach1
Nrf2
Actin
A
% Maximum
0
20
40
60
80
100
120
NRF2
BACH1
Whole cell lysate
Time (Hours)
B
C
D
BACH1
Time (Hours)
0 0.5 1 3 0 0.5 1 3
0
20
40
60
80
100
120
Nuclear
Cytosol
NRF2
% Maximum
Nuclear
Hours
02 46 8
Fold Control
0
100
200
300
400
500
F
Arsenite (µM)
Fold Control
1 5 25 50
0
100
200
300
400
E
Figure 1. Activation of antioxidant response and induction of HMOX1. (A) Representative immunoblots of total cellular proteins (20 mg) illustrating
the effect of 25 mM arsenite on NRF2 and BACH1 protein levels. (B) Graphical representation of three separate experiments showing changes in
expression of NRF2 and BACH1 proteins normalized to b-actin levels. (C) Representative immunoblots of nuclear (10 mg) and cytosolic proteins
(20 mg) illustrating changes in subcellular localization of NRF2 and BACH1 following arsenite treatment (25 mM). (D) Graphical representation of
three separate experiments showing changes in nuclear and cytosolic NRF2 and BACH1 proteins levels normalized to b-actin levels. The presence
of NRF2 in nuclear extract at 0 h (N0) is provided as a positive immunoblot control for cytosolic NRF2. Quantification represents the mean SEM
of three independent experiments. (E) Dose response of HMOX1 mRNA expression in HaCaT cells following a 6-h continuous treatment with
the indicated concentration of arsenite. (F) Time course of HMOX1 mRNA expression in HaCaT cells following treatment with 25 mM arsenite.
HMOX1 expression was determined using quantitative real-time PCR (qRT–PCR). HMOX1 mRNA concentrations were normalized to
b-actin mRNA and expressed as fold change relative to untreated controls. Values represent at least three independent experiments quantified
in triplicate.
4 Nucleic Acids Research, 2007
BACH1 and NRF2 bind to the same ARE motifs
in the HMOX1 promoter
To identify the maximum possible BACH1- and NRF2-
binding sites, we searched 10 kb upstream of the HMOX1
TSS for core ARE motifs conforming to the sequence
RTGAYNNNGC or its reverse complement (5). Twelve
consensus elements were identified (Table 1) and each
of these sites were investigated for NRF2 and BACH1
interactions by ChIP analysis (Figure 2A). Of the 12 ARE
motifs, NRF2 and BACH1 interact with the same two
sites containing multiple ARE motifs; one, a more
proximal site located at 3928 bp upstream of the TSS
(E1) and the other a more distal site at 8979 bp upstream
(E2). NRF2 binds both of these sites after arsenite
treatment (Figure 2B) while BACH1 binds both of them
in arsenic naı
¨
ve cells (Figure 2C). The proximal E1
element is composed of two slightly different ARE core
motifs having the sequences GCtgcGTCAT and GCtga
GTCAC and separated by 54 nt. The more distal E2 site
consists of four identical repeats of the core ARE motif
having the sequence GCtraGTCAC, each separated by
19 nt. An additional single motif is located at 9491 bp,
415 bp further upstream from the end of this ARE tetrad
for a total of five elements in this DNA region. In control
cells, BACH1 binding at the E2 site is 5-fold greater than
at E1. Following arsenite treatment, NRF2 binding at the
distal E2 tetrad is greater than at the E1 site by 2.5-fold.
None of the other five remaining ARE motifs were bound
by either NRF2 or BACH1. These data show that BACH1
and NRF2 undergo reciprocal binding at the E1 and E2
enhancers following oxidative stress initiated by arsenite
treatment.
Differential regulation of BACH1 and NRF2 activity
To investigate the relative contributions of BACH1 and
NRF2 to HMOX1 expression, we differentially regulated
their activities by treating cells with the proteasome
inhibitor MG132 or with hemin. MG132 indirectly
causes NRF2 activation by inhibiting KEAP1-dependent
proteasomal degradation (26). Hemin inactivates BACH1
through interaction with its multiple heme-binding motifs
leading to a conformational change and nuclear export
(27,28). Treatment of HaCaT cells with MG132 for 3 h
induces extensive NRF2 accumulation in whole cell lysates
relative to untreated control cells (Figure 3A). This
increase in NRF2 is almost entirely localized to the
nuclear fraction (Figure 3B) with negligible levels detected
in the cytosolic fraction (Figure 3C). In contrast to NRF2,
BACH1 levels are only slightly affected by MG132
treatment in whole cell lysates or nuclear distribution
(Figure 3). Hemin treatment contrasts with MG132
treatment in that it prominently triggers BACH1 inactiva-
tion, which manifests as a significant decrease in whole cell
BACH1 protein levels (Figure 3A) and as its disappear-
ance from the nuclear fraction (Figure 3B). Interestingly,
this effect is not associated so much with nuclear efflux to
the cytosolic compartment as occurs following arsenite
treatment, but rather is associated with a net loss of total
BACH1 (Figure 3A). Co-treatment of MG132 and hemin
blocks this net loss of BACH1 (Figure 3A) and results in
its cytoplasmic accumulation due to efflux from the
nucleus (Figure 3C). Thus, inactivation of BACH1 by
hemin triggers both its nuclear export and subsequent
proteasomal degradation. In contrast to hemin, arsenite
only mediates subcellular redistribution of BACH1,
as indicated by nuclear efflux, without notable protein
loss in whole cell lysates (Figure 3). Importantly, hemin
treatment has only minimal effects on NRF2 activation,
as indicated by the relative absence of NRF2 accumula-
tion in either whole cell extracts or nuclear fractions.
Co-treatment with hemin plus MG132 produces a
combined pattern of NRF2 nuclear translocation and
BACH1 efflux similar to that elicited by arsenite. Consis-
tent with these findings, the expression of HMOX1 protein
(Figure 3A) is associated only with treatments that result
in BACH1 efflux from the nucleus, but not with NRF2
activation alone.
The elimination rate of NRF2 exceeds its rate of
synthesis (26); hence, NRF2 activation requires both
inhibition of proteasomal degradation and continuous
NRF2 synthesis. To characterize the roles of NRF2 in
basal and inducible expression of HMOX1 and TXHRD1,
NRF2 activation was prevented by pretreating cells
for 30 min with cycloheximide (CHX) to block protein
synthesis prior to treatment with MG132. Blocking
ongoing protein synthesis prior to inhibition of NRF2
degradation by MG132 prevents accumulation of de novo
synthesized NRF2 while eliciting depletion of basal
nuclear NRF2. The right half of Figure 3A shows that
pretreatment of HaCaT cells with 5 mM CHX blocks
synthesis of new NRF2, thereby preventing its activation
when followed by MG132 or arsenite treatment. The fact
that CHX pretreatment results in the elimination of NRF2
Table 1. DNA sequences consistent with the consensus ARE motifs
located within 10 Kb upstream of the HMOX1 TSS and 1 Kb upstream
of TXNRD1 TSS
Postulated antioxidant response elements of the HMOX1
and TXNRD1 genes
PCR primer location
for ChIP (bp)
Motif sequence Motif location
(bp)
Consensus motif:
GCnnnRTCAY or
CGnnnYAGTR
HMOX1
9069 GTGACagaGC 9491
GCtgaGTCAC 9066
GCtaaGTCAC 9037
GCtgaGTCAC 9008
GCtgaGTCAC 8979
7232 GCcttGTCAC 7104
6060 GCagaATCAT 6008
GCtgaATCAT 5967
3928 GCtgcGTCAT 3992
GCtgaGTCAC 3928
1319 GCgtgGTCAC 107
148 GCaaaATCAC 200
TXNRD1
91 & 36 GCtttGTCAT 19
PCR primer and motif positions correspond to the most distant 5
0
nucleotide from the TSS on the forward strand. Sequences were
considered candidate ARE motifs regardless of orientation or strand.
Nucleic Acids Research, 2007 5
(Figure 3) establishes that under these conditions NRF2
cannot be the mediator of gene transcription. In parallel
treatments, CHX has no effect on hemin- or arsenite-
induced export of BACH1 from the nucleus, though total
BACH1 levels increase in the presence of MG132 relative
to untreated and CHX-treated controls. These data
support a role for proteasomal degradation in the turn-
over of BACH1 and indicate that BACH1 has a much
longer half-life than NRF2 in cells naı
¨
ve to oxidative
stress or hemin.
The effect of differentially modulating BACH1 or
NRF2 activities was examined in relation to HMOX1
expression. HMOX1 is expressed in control cells at nearly
undetectable levels and its induction coincides with
the disappearance of nuclear BACH1 but not nuclear
translocation of NRF2. Treatment with MG132 triggers
prominent nuclear accumulation of NRF2 but fails to
induce HMOX1 expression (Figure 3A). In contrast,
treatment with either arsenite or hemin leads to nuclear
export of BACH1 and to a prominent increase in HMOX1
expression. Combined treatment with MG132 and hemin
augments HMOX1 protein levels, suggesting that activa-
tion of NRF2 contributes to HMOX1 expression in a
manner that is at best additive to the effect of BACH1.
These data suggest that BACH1 inactivation plays a larger
role in the regulation of antioxidant gene expression
than NRF2 activation.
Temporal induction of HMOX1 transcription depends
on BACH1 removal from the enhancers
The observation that expression of HMOX1 is associated
with nuclear efflux of BACH1 rather than with NRF2
activation, suggests that BACH1 removal is the dominant
event regulating HMOX1 induction. To test this hypo-
thesis we followed the temporal dynamics of HMOX1
Relative DNA Enrichment
Relative DNA Enrichment
B
NRF2 Interactions
Kbp 10 1230456789
1 211
21
4
1
2
A
C
BACH1 Interactions
Contro
l
As
+3
&
&
HMOX1 Promoter Position
10 9 8 7 6 5 4 3 2 1
0
10 9 8 7 6 5 4 3 2 1
0
0
1000
2000
3000
4000
5000
HMOX1 Promoter Position
9076
7114
6018
5206
3928
3033
2518
1734
1304
148
0
100
200
300
400
500
600
9076
7114
6018
5206
3928
3033
2518
1734
1304
148
E2
E1 Pr
Figure 2. Identification of NRF2 and BACH1 interactions with core ARE motifs of HMOX1.(A) The position of all 12 putative ARE motifs,
relative to the HMOX1 transcription start site as annotated by the NCBI Homo sapiens Genome Map Viewer, Build 36.2. Open boxes = motif
conforming to the consensus ARE sequence, shaded boxes = imperfect ARE motif. Boxed numerals indicate the number of motif repeats in that
region. Arrows indicate the plus-strand orientation of each ARE motif relative to the consensus ARE sequence (RTGAYnnnGC). Relative DNA
enrichment associated with NRF2 (B) and BACH1 (C) ChIP at each of the HMOX1 ARE-containing sites. Immunoprecipitated DNA was analyzed
for enrichment by qRT–PCR using primers flanking ARE motifs at the positions indicated. Vertical bars and circle symbols graphically depict the
relative magnitude of DNA enrichment at each position. Vertical bars indicate enrichment at positions 3928 bp (E1) and 9069 bp (E2) while
circles correspond with regions where negligible DNA enrichment was observed. Values oriented vertically along the abscissa indicate the position of
each DNA region amplified by qRT–PCR. Amplification was expressed in terms of C
t
as calculated by normalizing the C
t
for each primer in
chromatin immunoprecipitated samples to the C
t
obtained from the respective input DNA and expressed as a percentage relative to the PCR primer
having the maximum enrichment (–9069 bp). Values represent the mean SEM of at least three independent experiments performed in triplicate.
6 Nucleic Acids Research, 2007
transcriptional initiation by ChIP analysis. The time
course of BACH1 and NRF2 binding to the E1
(Figure 4) and E2 (data not shown) enhancer regions of
HMOX1, as well as RNA pol II binding at the proximal
promoter (148), were quantified after treatments with
arsenite, hemin or MG132. These treatments induced very
similar interactions of NRF2 and BACH1 at both of
these enhancer regions, of which the observed interactions
at E1 are representative. As anticipated, arsenite treat-
ment produces a significant loss of BACH1 binding
that is clearly detected 30 min after treatment and reaches
minimal level by 60 min (Figure 4A). Conversely, arsenite
A
HMOX1
DMSO
CHX
Whole Cell Lysates
As
+3
MG132
Hemin
B
Nuclear Lysate
DMSO CHX
b-actin
b-actin
Bach1
Bach1
Nrf2
Nrf2
As
+3
MG132
Hemin
DMSO CHX
Bach1
Nrf2
Cytosolic Lysate
b-actin
C
––+----+
–++--+-+-
–++---++--
––+----+
–++--+-+-
–++---++--
––+----+
–++--+-+-
–++---++--
As
+3
MG132
Hemin
Figure 3. Differential regulation of NRF2 and BACH1 activation.
Immunoblots illustrating differential expression of NRF2 and BACH1
protein following treatment of HaCaT cells with 25 mM arsenite, 5 mM
MG132, 25 mM hemin or MG132 + hemin combined. HaCaT cells were
treated as indicated for 3 h or co-treated with 5 m M cycloheximide
(CHX) following a 30-min CHX pretreatment. (A) Total cellular
proteins (20 mg) from whole cell lysates. (B) Proteins extracts (10 mg)
from nuclear lysates. (C) Proteins extracts (20 mg) from cytosolic
lysates. Blots are representative of 2–3 separate experiments.
Fold change
Nrf2
0
5
10
15
20
25
A
Bach1
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Arsenite
Hemin
Mg132
C
RNA pol II
Minutes
0 30 60 90 120 150 180
0
2
4
6
8
10
12
Fold changeFold change
B
Figure 4. Association of NRF2 and BACH1 DNA binding with
transcriptional activation. Time course of DNA binding by BACH1
(A) NRF2 (B) and RNA polymerase II (C) following treatment with
25 mM arsenite, 5 mM MG132 or 25 mM hemin. ChIP-enriched DNA
was quantified using qRT–PCR with primers flanking the HMOX1
ARE motifs at positions 3992 (NRF2 and BACH1) and 148 (RNA
Pol II) and is expressed as the value for each treatment normalized
to its corresponding input and expressed as fold enrichment relative to
untreated control. Figures represent the results of two independent
experiments performed in triplicate SEM.
Nucleic Acids Research, 2007 7
triggers a prominent increase in NRF2 binding
(Figure 4B), but this increase is delayed by at least
60 min relative to the BACH1 decrease, and is maximal
at 150 min after treatment. Arsenite also produces a
large increase in RNA pol II binding (Figure 4C) that
temporally follows the loss of BACH1 but precedes NRF2
binding by at least 60 min and this coincides with the
relatively low level of NRF2 binding during the period
between 60 and 120 min.
To investigate the individual contributions of BACH1
and NRF2 to HMOX1 induction, cells were treated
with either hemin or MG132 in parallel with arsenite.
Hemin triggers a rapid decrease in BACH1 DNA binding,
mirroring the effect of arsenite and producing
nearly complete BACH1 loss from the promoter 60 min
after treatment (Figure 4A). In comparison to arsenite-
mediated NRF2 binding, hemin treatment is associated
with relatively weak NRF2 binding at time points earlier
than 90 min (Figure 4B) and significantly less than that
observed with arsenite after 120 min. Interestingly, hemin
induces RNA pol II binding that proceeds in a manner
similar in magnitude and time course to that associated
with arsenite treatment (Figure 4C).
In stark contrast to both arsenite and hemin, MG132
has little effect on promoter binding by BACH1
(Figure 4A) and limited binding of NRF2, even after
120 min of treatment (Figure 4B). The level of NRF2
binding elicited by MG132 is of the order of that observed
with hemin and approximately one-half of the level found
with arsenite. Furthermore, MG132 elicits weak RNA pol
II DNA binding relative to the levels associated with
hemin and arsenite treatments, implying much lower levels
of transcription initiation (Figure 4C). Together, these
results strongly suggest that loss of BACH1 from the
HMOX1 promoter is associated with a high level of RNA
polymerase II binding, while NRF2 activation in the
absence of BACH1 removal is associated with weak levels
of gene induction, as inferred from the extent of RNA
pol II binding. Consistent with whole cell lysate
and nuclear extract immunoblots, it is evident that
RNA pol II HMOX1 binding precedes NRF2 activation,
further supporting the conclusion that loss of BACH1-
mediated transcriptional repression is more important
than nuclear accumulation of NRF2 for HMOX1
induction.
To confirm that RNA pol II binding is associated
with gene induction by hemin but not by MG132,
relative steady-state levels of HMOX1 mRNA accumula-
tion were measured by qRT–PCR over the time course
of HMOX1 protein induction. Induction of HMOX1
mRNA expression was differentially regulated by treat-
ment with either hemin or MG132. Despite the absence
of NRF2 activation, hemin triggers induction of
much greater levels of HMOX1 mRNA than MG132
(Figure 5). Given that NRF2 activation is generally
considered a critical event for antioxidant gene
induction, it is surprising that inactivation of BACH1
by hemin produced 5-fold greater levels of HMOX1
induction than were elicited by MG132-mediated NRF2
activation.
Differential gene induction regulated by NRF2 and BACH1
Although NRF2 and BACH1 interact with ARE motifs,
several lines of evidence support that these interactions
with DNA are not identical, thereby suggesting that genes
could be identified that are regulated by one factor but not
by the other (12,29,30). TXNRD1 was identified as one
such gene. ChIP analysis was used to test for NRF2 and
BACH1 binding to the single ARE motif located 9 bp
upstream of the TSS in the TXNRD1 5
0
-flanking region.
As illustrated in the upper panel of Figure 6A, NRF2
binds the ARE motif of the TXNRD1 in ‘control cells’
with an affinity significantly greater than either at
HMOX1 E1 or E2. After arsenite-induced oxidative
stress, this site becomes more strongly bound by NRF2,
reaching a level that is intermediate between NRF2
binding at the HMOX1 E1 and E2 sites after activation,
and 6.5-fold greater than control levels. In contrast,
the lower panel of Figure 6A shows negligible binding
of BACH1 to the TXNRD1 ARE in either control or
arsenite-treated cells. As a consequence of differential
transcription factor binding, we anticipated that MG132,
but not hemin, would elicit NRF2-dependent TXNRD1
induction. To test this prediction, we treated HaCaT cells
with MG132 or hemin and quantified mRNA levels
of HMOX1 and TXNRD1 by qRT–PCR. In untreated
control cells, basal TXNRD1 mRNA levels are 35-fold
greater than the levels of HMOX1 mRNA. Inhibition of
NRF2 activation with CHX significantly inhibited basal
TXNRD1 expression but left HMOX1 expression
unchanged (Figure 6B and C). This pattern of expression
is consistent with functionally active NRF2 residing
basally in the nucleus and rules out the possibility that
this resident nuclear NRF2 contributes to basal HMOX1
expression. The role of activated NRF2 is demonstrated
HMOX1 Expression
Minutes
0 60 120 180
HMOX1 Expression
(% Maximum)
0
20
40
60
80
100
120
Hemin
Mg132
Figure 5. Time course of HMOX1 expression elicited by hemin and
MG132 treatment. The time course of HMOX1 mRNA expression
following treatment with 25 mM hemin or 5 mM MG132. HMOX1
mRNA was determined by quantitative real-time PCR (qRT–PCR),
normalized to b-actin mRNA and expressed as percent maximum
expression. Values represent at least three independent experiments
quantified in triplicate SEM.
8 Nucleic Acids Research, 2007
by the MG132 treatments that trigger maximal TXNRD1
induction but mediate only weak HMOX1 expression
(Figure 6B and C). Pretreatment with CHX prior to
MG132 blocks NRF2 synthesis, returning HMOX1
mRNA to control levels and reducing TXNRD1 mRNA
to the levels of CHX-treated control cells. These expres-
sion data reflect the primary role of NRF2 in basal
and inducible expression of TXNRD1 but not HMOX1.
Hemin treatment, which inactivates BACH1, has no effect
on TXNRD1 expression (Figure 6C), reflecting the fact
that TXNRD1 is regulated independently of BACH1.
In contrast, hemin elicits a significant and nearly maximal
increase in HMOX1 expression (Figure 6B). CHX pre-
treatment predominantly blocks this induction, demon-
strating a requisite role for resident nuclear NRF2 in
HMOX1 induction without necessitating nuclear accumu-
lation of activated NRF2. Combined treatment with
hemin plus MG132 results in induction of TXNRD1
to levels similar to those in cells treated with MG132
alone while the level of HMOX1 induction is similar to
that elicited by hemin (Figure 6B and C). These results
show that TXNRD1 induction is attributable solely
Nrf2
% Enrichment
(relative to E2)
0
20
40
60
80
100
Bach1
Promoter
HMOX1
E2
HMOX1
E1
TXNRD1
0
20
40
60
80
100
Control
Arsenite
A
Treatment
0
20
40
60
80
100
120
Control
Mg132
Hemin
Hemin +
MG132
TXNRD1 Expression
(% Maximum)
C
DMSO
Cycloheximide
Treatment
Control
Mg132
Hemin
Hemin +
MG132
HMOX1 Expression
(% Maximum)
0
20
40
60
80
100
120
B
Figure 6. Differential regulation of HMOX1 and TXNRD1 by BACH1 and NRF2. (A) ChIP analysis of NRF2 and BACH1 interactions with ARE
sites of HMOX1 and TXNRD1 before and after 25 mM arsenite treatment for 3 h. Binding of NRF2 (top panel) and BACH1 (bottom panel)
to HMOX1 E1 and TXNRD1 is expressed relative to binding at the HMOX1 E2 enhancer element. Expression of HMOX1 (B) and TXNRD1
(C) mRNA was measured in HaCaT cells following treatment with 25 mM hemin, 5 mM MG132 or both in the presence (shaded bars) or absence
(filled bars) of 5 m M CHX. Relative mRNA was determined by quantitative real-time PCR (qRT–PCR), normalized to b-actin mRNA and expressed
as percent maximum expression. Values represent at least three independent experiments quantified in triplicate SEM.
Nucleic Acids Research, 2007 9
to activation of NRF2 by proteasome inhibition but that
HMOX1 induction requires inactivation of BACH1.
Thus, BACH1 inactivation is predominantly responsible
for HMOX1 induction and that NRF2 activation contrib-
utes nominally to this process. Interestingly, the combina-
tion of MG132 plus hemin also shows that, when preceded
by CHX treatment, HMOX1 expression remains signifi-
cantly induced despite the lack of NRF2 and hints at the
possible role for BACH1 in repressing the activity
of transcriptional activators in addition to NRF2. These
data establish that inactivation of BACH1 is
more important than activation of NRF2 for induction
of HMOX1 expression and that not all NRF2-regulated
genes respond in this manner, as is the case of TXNRD1.
DISCUSSION
In the present article, we report for the first time the
dynamics associated with inactivation of BACH1 and
NRF2 activation underlying the initiation of endogenous
gene targets. BACH1, which is predominantly localized to
the nucleus of control cells, is exported to the cytosol
within 30 min following arsenite or hemin treatment.
Temporally, inactivation of BACH1 precedes NRF2
activation by at least 30 min, correlating with the period
necessary for de novo synthesis of NRF2, as supported by
the absence of immunoreactive NRF2 when CHX
treatment precedes proteasome inhibition. Consequently,
biological coupling of these transcriptional regulators
insures that the response to stressors is rapid and not
dependent on the delay required for synthesis of high
NRF2 levels.
Of the 12 putative HMOX1 ARE motifs, only two sites,
at 3928 bp (E1) and 8979 bp (E2), are reciprocally
bound by BACH1 and NRF2. Both sites contain multiple
ARE motifs and have been previously recognized as
HMOX1 transcriptional regulators (15). Our data confirm
that these are the only two HMOX1 elements recognized
by NRF2 and BACH1 in vivo. This pattern of binding
contrasts markedly with binding at the TXNRD1 promo-
ter, where NRF2 is capable of significantly binding the
ARE motif but BACH1 is not. Distinct recognition of
ARE motifs by BACH1, compared to NRF2, conveys an
additional level of transcriptional regulation to HMOX1
that is lacking in the regulation of TXNRD1. The fact that
BACH1 and NRF2 differently recognize ARE motifs
presents the possibility that BACH1 regulates an over-
lapping but separate gene battery. NRF2 and BACH1
differ in their affinities for ARE motifs that otherwise
appear to conform equally well to the consensus ARE core
motif; however, this short 10 bp sequence is not the sole
determinant of factor binding. Nucleotides flanking the
ARE core as well as motif multiplicity also contribute
to differential motif recognition by BACH1 and NRF2.
Several reports indicate that extended sequences flanking
the ARE core are vital for recognition by NRF2 (5,31–33),
though the precise sequence requirements of the extended
ARE motif are not well defined leaving a great deal of
uncertainty when attempting to predict high-affinity
NRF2-binding elements. Similar difficulties exist in
predicting the sequence requirements for high-affinity
BACH1-binding sites. Unlike NRF2 however, detailed
description of BACH1-binding elements is hampered by
the fact that only a few genes are known to be under its
regulatory control (13,34), including NQO1 (10) and
HMOX1 (14). The sequences that BACH1 interacts with
at these genes seems to conform to the core ARE motif
recognized by NRF2; however, only one report has
experimentally examined sequence requirements for
BACH1 binding, the results of which generally support
motif homology with the ARE (29). Binding element
multiplicity may also contribute to differences in DNA-
binding affinity between NRF2 and BACH1. While Nrf2
is capable of binding an individual ARE motif, BACH1
appears to bind poorly to single motifs, preferring to
interact at sites containing multiple motifs. The BTB/POZ
domain of BACH1 is thought to contribute to DNA
binding through oligomerization with neighboring
BACH1 heterodimers to stabilize their interactions with
DNA (12). In this respect, both HMOX1 enhancers
recognized by BACH1 consist of multiple elements while
the TXNRD1 element is a single consensus motif. Thus,
the multiple ARE motifs of HMOX1 might permit
BACH1 oligomerization while TXNRD1, with its single
motif, would not allow formation of stable BACH1
oligomers. Together, nucleotide sequence and motif
multiplicity may account for the observation that the
TXNRD1 promoter is poorly recognized by BACH1.
In control cells, we have noted that the cytosol is
essentially devoid of detectable NRF2 consistent with
reports that NRF2 activity is governed by regulation of
its stability (23,35,36) rather than by being sequestered
in the cytosol (25,37). In the absence of oxidative stress,
KEAP1 mediates Cul3-dependent ubiquitylation of NRF2
that targets it for rapid proteasomal destruction (36).
During oxidative stress several sensitive KEAP1 cysteines
are oxidized resulting a loss of KEAP1-directed protea-
somal degradation of NRF2 and increased proteasomal
degradation of KEAP1 itself (21,38). In the absence of
ongoing KEAP1-directed degradation of NRF2, the
half-life of newly synthesized protein is prolonged permit-
ting rapid nuclear translocation and accumulation with
subsequent antioxidant gene induction (35). Consistent
with a short residence time of NRF2 in the cytosol, we
observe that when NRF2 escapes proteasomal degrada-
tion it accumulates in the nucleus while remaining nearly
undetectable in the cytoplasm. In this respect, we have
observed the persistence of nuclear NRF2 in untreated
control cells indicating that even in the absence of
oxidative stress some NRF2 escapes KEAP1-mediated
degradation and is capable of binding ARE elements and
eliciting gene induction. We propose that the persistence
of resident nuclear NRF2 plays a crucial role in both the
induction of HMOX1 as well as basal TXNRD1 expres-
sion. BACH1 inactivation elicits maximal HMOX1
expression without requiring concurrent NRF2 activation
by permitting resident nuclear NRF2 to bind HMOX1
ARE motifs. In contrast, BACH1 does not inhibit basal
nuclear NRF2 binding to the TXNRD1 promoter and
therefore basal TXNRD1 expression is 20% of maximal
expression, and 35-fold higher than basal HMOX1
10
Nucleic Acids Research, 2007
ARE Motif Position (Kb)
BACH1
BACH1
NRF2 NRF2
NRF2
NRF2
NRF2
BACH1
NRF2
89Kb-10 7 6 5 4 3 2 10
89Kb-10 7 6 5 4 3 2 10
89Kb-10 7 6 5 4 3 2 10
89Kb-10 7 6 5 4 3 2 10
NRF2
BACH1
BACH1
NRF2
Nucleus
K
off
Naïve
A
B
C
D
Hemin
MG132
Arsenite
or
Hemin+
MG132
PolII
PolII
K
on
K
off
K
off
K
on
Nucleus
Nucleus
BACH1
K
off
K
on
K
off
K
on
K
off
K
on
K
off
K
on
K
off
K
on
K
on
NRF2
NRF2
NRF2
NRF2 NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
NRF2
BACH1
BACH1
NRF2
Cytosol
Cytosol
Nucleus
Cytosol
C
y
tosol
Figure 7. Induction of HMOX1 through the interplay of BACH1 and NRF2 with enhancer elements. (A) Binding of BACH1 (red ovals) at the
E1 and E2 enhancer elements of HMOX1 in untreated control cells blocks NRF2 (blue ovals) binding and HMOX1 induction. Competition
Nucleic Acids Research, 2007 11
expression. Loss of basal nuclear NRF2 following CHX
treatment precipitously decreases both hemin-mediated
HMOX1 induction and basal TXNRD1. It is noteworthy
that some HMOX1 expression elicited by hemin inactiva-
tion of BACH1 persists despite the absence of NRF2
protein, suggesting BACH1 may also play a role in the
repression of factor(s) in addition to NRF2. In this regard,
HMOX1 expression is even more pronounced when CHX
treatment precedes co-treatment with hemin plus MG132,
suggesting that proteasomal inhibition may stabilize other
putative transcriptional activator(s).
Without prior inactivation of BACH1, activated NRF2
binds inefficiently to HMOX1 enhancers compared to the
binding elicited by arsenite. Although the level of NRF2
binding associated with MG132 treatment is similar
in extent to that initiated by hemin treatment, NRF2
activation in the absence of BACH1 inactivation results in
significantly lower levels of HMOX1 transcription. This
contrasts profoundly with TXNRD1 expression, where
MG132 strongly stimulates induction and hemin treat-
ment has no effect. Since transcription factor binding to
cognate sites is a stochastic event (39), the massive increase
in nuclear NRF2 concentration associated with its
activation could increase competition with BACH1 for
ARE motifs or sMAF-binding partners. By mass action,
NRF2 activation could establish a new equilibrium
favoring DNA-bound NRF2/sMAF, thereby increasing
the stoichiometric DNA residence of NRF2. Since
BACH1 appears to bind at sites containing multiple
AREs, NRF2 may not displace a large enough fraction of
DNA-bound BACH1 to overcome transcriptional repres-
sion. CHX treatment decreased hemin-mediated HMOX1
induction and basal TXNRD1 expression. When inter-
preted in the light of our ChIP enhancer binding data,
these findings support the conclusion that inactivation of
BACH1 by hemin decreases its DNA-binding affinity
resulting in its generalized removal from ARE motifs and
concomitant loss of transcriptional repression, leaving the
AREs available for binding by NRF2 and possibly other
transcription that mediate gene induction.
Based on the dynamic exchange of BACH1 and NRF2
we propose that in cells naı
¨
ve to oxidative stress, BACH1
is bound to the ARE enhancer motifs preventing
NRF2 from binding and thereby repressing transcription
(Figure 7A). Hemin treatment triggers removal of BACH1
from HMOX1 enhancers thereby allowing NRF2 that is
already present in the nucleus to interact with ARE motifs
and elicit gene induction (Figure 7B). On the other hand,
MG132 triggers NRF2 translocation to the nucleus but
DNA-bound BACH1 blocks NRF2–ARE interactions to
prevent gene induction (Figure 7C). The vast increase
in the presence of nuclear NRF2 permits some increase
in NRF2 binding to HMOX1 AREs but the presence of
BACH1 maintains gene repression. Treatment with either
arsenite or hemin plus MG132 triggers both the removal
of BACH1 and the activation of NRF2, which can now
freely bind vacant enhancer motifs (Figure 7D).
Though not all ARE-containing genes are regulated
by BACH1, whenever BACH1 interacts with ARE motifs
it contributes an important regulatory dimension that
provides a more complex response to environmental
stress. Reciprocal ARE binding of BACH1 for NRF2
imparts an increased level of complexity to the ARE-
regulated genes producing distinct patterns of gene
expression patterns, as shown here for HMOX1 and
THXRD1. The differential interaction of BACH1 and
NRF2 with various forms of the ARE motif suggests
the possibility that BACH1 may regulate a distinct but
overlapping battery of genes. In this case, genes specifi-
cally repressed by BACH1 could be globally induced in
response to redox stress.
SUPPLEMENTARY DATA
Supplementary Data are available at NAR Online.
ACKNOWLEDGEMENTS
We thank Dr N. Fusenig (Division of Differentiation
and Carcinogenesis in Vitro, German Cancer Research
Center, Heidelberg, Germany) for a gift of HaCaT
keratinocytes. This research was supported by NIEHS
grants R01 ES10807, The NIEHS Center for Environ-
mental Genetics grant P30 ES06096 and the NIEHS
Superfund Basic Research Program grant P42 ES04908.
J.F.R. is a Postdoctoral Trainee partly supported by
NIEHS T32 ES07250, Environmental Carcinogenesis and
Mutagenesis Training Grant. Funding to pay the Open
Access publication charges for this article was provided
by NIEHS grants RO1 ES10807.
Conflict of interest statement. None declared.
REFERENCES
1. Spuches,A.M., Kruszyna,H.G., Rich,A.M. and Wilcox,D.E. (2005)
Thermodynamics of the As(III)-thiol interaction: arsenite and
monomethylarsenite complexes with glutathione, dihydrolipoic acid,
and other thiol ligands. Inorg. Chem., 44, 2964–2972.
2. Shi,H., Shi,X. and Liu,K.J. (2004) Oxidative mechanism of arsenic
toxicity and carcinogenesis. Mol. Cell Biochem., 255, 67–78.
3. Kann,S., Estes,C., Reichard,J.F., Huang,M.Y., Sartor,M.A.,
Schwemberger,S., Chen,Y., Dalton,T.P., Shertzer,H.G. et al. (2005)
Butylhydroquinone protects cells genetically deficient in glutathione
biosynthesis from arsenite-induced apoptosis without significantly
changing their prooxidant status. Toxicol. Sci., 87, 365–384.
4. Itoh,K., Chiba,T., Takahashi,S., Ishii,T., Igarashi,K., Katoh,Y.,
Oyake,T., Hayashi,N., Satoh,K. et al. (1997) An Nrf2/small Maf
heterodimer mediates the induction of phase II detoxifying enzyme
genes through antioxidant response elements. Biochem. Biophys.
Res. Commun., 236, 313–322.
5. Wasserman,W.W. and Fahl,W.E. (1997) Functional antioxidant
responsive elements. Proc. Natl Acad. Sci. USA, 94, 5361–5366.
for ARE-binding elements between activated NRF2 and DNA-bound BACH1 is indicated by hypothetical rate constants representing stochastic
binding (k
on
) and dissociation (k
off
) of NRF2. (B) Inactivation of BACH1 by hemin results in its removal from ARE motifs and elimination from the
nucleus. Consequently, NRF2 can interact with exposed ARE enhancers to recruit RNA pol II (green oval) leading to high-level HMOX1 induction.
(C) DNA-bound BACH1 blocks DNA binding of NRF2 despite its nuclear accumulation and prevents efficient HMOX1 induction. (D) Treatment
with arsenite or co-treatment with hemin + MG132 results in BACH1 inactivation, nuclear accumulation and ARE binding of NRF2, and high-level
HMOX1 induction.
12 Nucleic Acids Research, 2007
6. Deppmann,C.D., Alvania,R.S. and Taparowsky,E.J. (2006) Cross-
species annotation of basic leucine zipper factor interactions: insight
into the evolution of closed interaction networks. Mol. Biol. Evol.,
23, 1480–1492.
7. Motohashi,H., O’Connor,T., Katsuoka,F., Engel,J.D. and
Yamamoto,M. (2002) Integration and diversity of the regulatory
network composed of Maf and CNC families of transcription
factors. Gene, 294, 1–12.
8. Venugopal,R. and Jaiswal,A.K. (1996) Nrf1 and Nrf2 positively and
c-Fos and Fra1 negatively regulate the human antioxidant response
element-mediated expression of NAD(P)H:quinone oxidoreductase1
gene. Proc. Natl Acad. Sci. USA, 93, 14960–14965.
9. Sankaranarayanan,K. and Jaiswal,A.K. (2004) Nrf3 negatively
regulates antioxidant-response element-mediated expression and
antioxidant induction of NAD(P)H:quinone oxidoreductase1 gene.
J. Biol. Chem., 279 , 50810–50817.
10. Dhakshinamoorthy,S., Jain,A.K., Bloom,D.A. and Jaiswal,A.K.
(2005) Bach1 competes with Nrf2 leading to negative regulation of
the antioxidant response element (ARE)-mediated
NAD(P)H:quinone oxidoreductase 1 gene expression and induction
in response to antioxidants. J. Biol. Chem., 280, 16891–16900.
11. Amoutzias,G., Veron,A., Weiner,A., Robinson-Rechavi,M.,
Bornberg-Bauer,E., Oliver,S. and Robertson,D. (2006) One billion
years of bZIP transcription factor evolution: conservation and
change in dimerization, and DNA-binding site specificity.
Mol. Biol. Evol., 24, 827–835.
12. Igarashi,K., Hoshino,H., Muto,A., Suwabe,N., Nishikawa,S.,
Nakauchi,H. and Yamamoto,M. (1998) Multivalent DNA binding
complex generated by small Maf and Bach1 as a possible
biochemical basis for beta-globin locus control region complex.
J. Biol. Chem., 273 , 11783–11790.
13. Oyake,T., Itoh,K., Motohashi,H., Hayashi,N., Hoshino,H.,
Nishizawa,M., Yamamoto,M. and Igarashi,K. (1996) Bach proteins
belong to a novel family of BTB-basic leucine zipper transcription
factors that interact with MafK and regulate transcription through
the NF-E2 site. Mol. Cell. Biol., 16, 6083–6095.
14. Sun,J., Hoshino,H., Takaku,K., Nakajima,O., Muto,A., Suzuki,H.,
Tashiro,S., Takahashi,S., Shibahara,S. et al. (2002) Hemoprotein
Bach1 regulates enhancer availability of heme oxygenase-1 gene.
EMBO J., 21, 5216–5224.
15. Sun,J., Brand,M., Zenke,Y., Tashiro,S., Groudine,M. and
Igarashi,K. (2004) Heme regulates the dynamic exchange of Bach1
and NF-E2-related factors in the Maf transcription factor network.
Proc. Natl Acad. Sci. USA, 101, 1461–1466.
16. Shan,Y., Lambrecht,R.W., Ghaziani,T., Donohue,S.E. and
Bonkovsky,H.L. (2004) Role of Bach-1 in regulation of heme
oxygenase-1 in human liver cells: insights from studies with small
interfering RNAS. J. Biol. Chem., 279, 51769–51774.
17. Ishikawa,M., Numazawa,S. and Yoshida,T. (2005) Redox regula-
tion of the transcriptional repressor Bach1. Free Radic. Biol. Med.,
38, 1344–1352.
18. Dohi,Y., Alam,J., Yoshizumi,M., Sun,J. and Igarashi,K. (2006)
Heme Oxygenase-1 Gene Enhancer Manifests Silencing Activity in a
Chromatin Environment Prior to Oxidative Stress. Antioxid. Redox.
Signal., 8, 60–67.
19. Igarashi,K. and Sun,J. (2006) The heme-Bach1 pathway in the
regulation of oxidative stress response and erythroid differentiation.
Antioxid. Redox. Signal., 8, 107–118.
20. Dinkova-Kostova,A.T., Holtzclaw,W.D., Cole,R.N., Itoh,K.,
Wakabayashi,N., Katoh,Y., Yamamoto,M. and Talalay,P. (2002)
Direct evidence that sulfhydryl groups of Keap1 are the sensors
regulating induction of phase 2 enzymes that protect against
carcinogens and oxidants. Proc. Natl Acad. Sci. USA, 99,
11908–11913.
21. Zhang,D.D. and Hannink,M. (2003) Distinct cysteine residues in
Keap1 are required for Keap1-dependent ubiquitination of Nrf2
and for stabilization of Nrf2 by chemopreventive agents and
oxidative stress. Mol. Cell Biol., 23, 8137–8151.
22. Hong,F., Freeman,M.L. and Liebler,D.C. (2005) Identification of
sensor cysteines in human Keap1 modified by the cancer
chemopreventive agent sulforaphane. Chem. Res. Toxicol., 18,
1917–1926.
23. He,X., Chen,M.G., Lin,G.X. and Ma,Q. (2006) Arsenic induces
NAD(P)H-quinone oxidoreductase I by disrupting the Nrf2 x
Keap1 x Cul3 complex and recruiting Nrf2 x Maf to the antioxidant
response element enhancer. J. Biol. Chem., 281, 23620–23631.
24. Boukamp,P., Petrussevska,R.T., Breitkreutz,D., Hornung,J.,
Markham,A. and Fusenig,N.E. (1988) Normal keratinization in a
spontaneously immortalized aneuploid human keratinocyte cell line.
J. Cell Biol., 106, 761–771.
25. Itoh,K., Wakabayashi,N., Katoh,Y., Ishii,T., Igarashi,K.,
Engel,J.D. and Yamamoto,M. (1999) Keap1 represses nuclear
activation of antioxidant responsive elements by Nrf2 through
binding to the amino-terminal Neh2 domain. Genes Dev. , 13, 76–86.
26. Nguyen,T., Sherratt,P.J., Huang,H.C., Yang,C.S. and Pickett,C.B.
(2003) Increased protein stability as a mechanism that enhances
Nrf2-mediated transcriptional activation of the antioxidant response
element. Degradation of Nrf2 by the 26 S proteasome. J. Biol.
Chem., 278, 4536–4541.
27. Ogawa,K., Sun,J., Taketani,S., Nakajima,O., Nishitani,C., Sassa,S.,
Hayashi,N., Yamamoto,M., Shibahara,S. et al. (2001) Heme
mediates derepression of Maf recognition element through direct
binding to transcription repressor Bach1. EMBO J., 20, 2835–2843.
28. Suzuki,H., Tashiro,S., Hira,S., Sun,J., Yamazaki,C., Zenke,Y.,
Ikeda-Saito,M., Yoshida,M. and Igarashi,K. (2004) Heme regulates
gene expression by triggering Crm1-dependent nuclear export of
Bach1. EMBO J., 23, 2544–2553.
29. Kanezaki,R., Toki,T., Yokoyama,M., Yomogida,K., Sugiyama,K.,
Yamamoto,M., Igarashi,K. and Ito,E. (2001) Transcription factor
BACH1 is recruited to the nucleus by its novel alternative spliced
isoform. J. Biol. Chem., 276, 7278–7284.
30. Sakurai,A., Nishimoto,M., Himeno,S., Imura,N., Tsujimoto,M.,
Kunimoto,M. and Hara,S. (2005) Transcriptional regulation of
thioredoxin reductase 1 expression by cadmium in vascular
endothelial cells: role of NF-E2-related factor-2. J. Cell Physiol,
203, 529–537.
31. Rushmore,T.H., Morton,M.R. and Pickett,C.B. (1991) The anti-
oxidant responsive element. Activation by oxidative stress and
identification of the DNA consensus sequence required for
functional activity. J. Biol. Chem., 266, 11632–11639.
32. Erickson,A.M., Nevarea,Z., Gipp,J.J. and Mulcahy,R.T. (2002)
Identification of a variant antioxidant response element in the
promoter of the human glutamate-cysteine ligase modifier subunit
gene. Revision of the ARE consensus sequence. J. Biol. Chem., 277,
30730–30737.
33. Nioi,P., McMahon,M., Itoh,K., Yamamoto,M. and Hayes,J.D.
(2003) Identification of a novel Nrf2-regulated antioxidant response
element (ARE) in the mouse NAD(P)H:quinone oxidoreductase
1 gene: reassessment of the ARE consensus sequence. Biochem. J.,
374, 337–348.
34. Tahara,T., Sun,J., Igarashi,K. and Taketani,S. (2004) Heme-
dependent up-regulation of the alpha-globin gene expression by
transcriptional repressor Bach1 in erythroid cells. Biochem. Biophys.
Res. Commun., 324, 77–85.
35. McMahon,M., Itoh,K., Yamamoto,M. and Hayes,J.D. (2003)
Keap1-dependent proteasomal degradation of transcription factor
Nrf2 contributes to the negative regulation of antioxidant
response element-driven gene expression. J. Biol. Chem., 278,
21592–21600.
36. Furukawa,M. and Xiong,Y. (2005) BTB protein Keap1 targets
antioxidant transcription factor Nrf2 for ubiquitination by the
Cullin 3-Roc1 ligase. Mol. Cell Biol., 25, 162–171.
37. Zipper,L.M. and Mulcahy,R.T. (2002) The Keap1 BTB/POZ
dimerization function is required to sequester Nrf2 in cytoplasm.
J. Biol. Chem., 277 , 36544–36552.
38. Zhang,D.D., Lo,S.C., Sun,Z., Habib,G.M., Lieberman,M.W. and
Hannink,M. (2005) Ubiquitination of Keap1, a BTB-Kelch
substrate adaptor protein for Cul3, targets Keap1 for degradation
by a proteasome-independent pathway. J. Biol. Chem., 280,
30091–30099.
39. Cranz,S., Berger,C., Baici,A., Jelesarov,I. and Bosshard,H.R. (2004)
Monomeric and dimeric bZIP transcription factor GCN4
bind at the same rate to their target DNA site. Biochemistry, 43,
718–727.
Nucleic Acids Research, 2007 13

Supplementary resource (1)

Data
October 2007
John Reichard · Gregory T. Motz · Alvaro Puga
... In the case of HMOX1, it is important to note that the ability of NRF2 to upregulate the expression of this gene has been shown to be dependent on the inhibition of the repressive action of BACH1 [37]. The latter protein essentially blocks the access of NRF2 to AREs upstream of the HMOX1 promoter. ...
... The latter protein essentially blocks the access of NRF2 to AREs upstream of the HMOX1 promoter. For example, the proteasome inhibitor MG132 causes NRF2 accumulation but not BACH1 inhibition, and therefore does not stimulate a marked increase in HMOX1 expression, in contrast to pro-oxidants and other types of NRF2 activators that can inactivate BACH1 and upregulate HMOX1 [37]. Another important consideration in using HMOX1 mRNA as a marker of NRF2 signalling is its distinct kinetics of upregulation and return to baseline level in comparison to many other NRF2 regulated genes. ...
Article
Full-text available
The cytoprotective transcription factor NRF2 regulates the expression of several hundred genes in mammalian cells and is a promising therapeutic target in a number of diseases associated with oxidative stress and inflammation. Hence, an ability to monitor basal and inducible NRF2 signalling is vital for mechanistic understanding in translational studies. Due to some caveats related to the direct measurement of NRF2 levels, the modulation of NRF2 activity is typically determined by measuring changes in the expression of one or more of its target genes and/or the associated protein products. However, there is a lack of consensus regarding the most relevant set of these genes/proteins that best represents NRF2 activity across cell types and species. We present the findings of a comprehensive literature search that according to stringent criteria identifies GCLC, GCLM, HMOX1, NQO1, SRXN1 and TXNRD1 as a robust panel of markers that are directly regulated by NRF2 in multiple cell and tissue types. We assess the relevance of these markers in clinically accessible biofluids and highlight future challenges in the development and use of NRF2 biomarkers in humans.
... After oxidative or electrophilic stress, the NRF2-sMAFG/K heterodimer binds at Enhancer 1 (−4 kb) and Enhancer 2 (−9 kb) of the HMOX1 gene (Chorley et al., 2012), inducing its transactivation (Reichard et al., 2007). In the H 2 O 2 -primed cells, we found a clear enrichment of NRF2 in Enhancer 2. Similarly, Tu et al. (2017) described the ETM of the TNFSF10 gene, induced by PMA/I, associated with chromatin structure changes and the enrichment of transcription factor ETS-1 at the gene enhancer. ...
... Since BACH1 has been evidenced as a MAFK-dependent HMOX1 repressor (Reichard et al., 2007;Sun et al., 2002), evaluating its enrichment at the HMOX1 enhancers in H 2 O 2 -primed cells becomes relevant. However, other ETM epigenetic mechanisms could be studied in this ETM model, such as H3K4me2 enrichment at the core-promoter or, as reported in other ETM models, the H3ace and H4ace enrichment at the core-promoter and enhancers (Wong et al., 2014). ...
Article
Maternal obesity (MO) is a significant cause of increased cardiometabolic risk in offspring, who present endothelial dysfunction at birth. Alterations in physiologic and cellular redox status are strongly associated with altered gene regulation in arterial endothelium. However, specific mechanisms by which the pro‐oxidant fetal environment in MO could modulate the vascular gene expression and function during the offspring's postnatal life are elusive. We tested if oxidative stress could reprogram the antioxidant‐coding gene's response to a pro‐oxidant challenge through an epigenetic transcriptional memory (ETM) mechanism. A pro‐oxidant double‐hit protocol was applied to human umbilical artery endothelial cells (HUAECs) and EA.hy 926 endothelial cell lines. The ETM acquisition in the HMOX1 gene was analyzed by RT‐qPCR. HMOX1 mRNA decay was evaluated by Actinomycin‐D treatment and RT‐qPCR. To assess the chromatin accessibility and the enrichment of NRF2, RNAP2, and phosphorylation at serin‐5 of RNAP2, at HMOX1 gene regulatory regions, were used DNase HS‐qPCR and ChIP‐qPCR assays, respectively. The CpG methylation pattern at the HMOX1 core promoter was analyzed by DNA bisulfite conversion and Sanger sequencing. Data were analyzed using two‐way ANOVA, and p < 0.05 was statistically significant. Using a pro‐oxidant double‐hit protocol, we found that the Heme Oxygenase gene ( HMOX1 ) presents an ETM response associated with changes in the chromatin structure at the promoter and gene regulatory regions. The ETM response was characterized by a paused‐RNA Polymerase 2 and NRF2 enrichment at the transcription start site and Enhancer 2 of the HMOX1 gene, respectively. Changes in DNA methylation pattern at the HMOX1 promoter were not a hallmark of this oxidative stress‐induced ETM. These data suggest that a pro‐oxidant milieu could trigger an ETM at the vascular level, indicating a potential epigenetic mechanism involved in the increased cardiovascular risk in the offspring of women with obesity.
... The bZip protein BACH1 represses the HMOX1 gene, which encodes HO-1. This property is the main element that inhibits NRF2 activation of HMOX1 transcription [53]. To effectively activate HMOX1, BACH1 must first be deactivated. ...
Article
Full-text available
The phrase “Let food be thy medicine…” means that food can be a form of medicine and medicine can be a form of food; in other words, that the diet we eat can have a significant impact on our health and well-being. Today, this phrase is gaining prominence as more and more scientific evi-dence suggests that one’s diet can help prevent and treat disease. A diet rich in fruits, vegetables, whole grains, and lean protein can help reduce the risk of heart disease, cancer, diabetes, and other health problems and, on the other hand, a diet rich in processed foods, added sugars, and satu-rated fats can increase the risk of the same diseases. Electrophilic compounds in the diet can have a significant impact on our health, and they are molecules that covalently modify cysteine residues present in the thiol-rich Keap1 protein. These compounds bind to Keap1 and activate NRF2, which promotes its translocation to the nucleus and its binding to DNA in the ARE region, triggering the antioxidant response and protecting against oxidative stress. These compounds include poly-phenols and flavonoids that are nucleophilic but are converted to electrophilic quinones by met-abolic enzymes such as polyphenol oxidases (PPOs) and sulfur compounds present in foods such as the Brassica genus (broccoli, cauliflower, cabbage, Brussel sprouts, etc.) and garlic. This review summarizes our current knowledge on this subject.
... Under normal conditions, BACH1 represses HMOX1 transcription by binding to the HMOX1 enhancer with small MAF proteins. Stimulation with heme or arsenite induces HMOX1 transcription by inactivating BACH1 and promoting NRF2 interaction with the HMOX1 promoter [30,31]. Additionally, PERK, STAT3 and HIF-1a have been found to promote the expression of HMOX1 [32][33][34][35]. ...
Article
Full-text available
Background Sanguinarine chloride (S.C) is a benzophenanthrine alkaloid derived from the root of sanguinaria canadensis and other poppy-fumaria species. Studies have reported that S.C exhibits antioxidant, anti-inflammatory, proapoptotic, and growth inhibitory effects, which contribute to its anti-cancer properties. Recent studies suggested that the antitumor effect of S.C through inducing ferroptosis in some cancers. Nevertheless, the precise mechanism underlying the regulation of ferroptosis by S.C remains poorly understood. Methods A small molecule library was constructed based on FDA and CFDA approved small molecular drugs. CCK-8 assay was applied to evaluate the effects of the small molecule compound on tumor cell viability. Prostate cancer cells were treated with S.C and then the cell viability and migration ability were assessed using CCK8, colony formation and wound healing assay. Reactive oxygen species (ROS) and iron accumulation were quantified through flow cytometry analysis. The levels of malondialdehyde (MDA) and total glutathione (GSH) were measured using commercially available kits. RNA-seq analysis was performed to identify differentially expressed genes (DEGs) among the treatment groups. Western blotting and qPCR were utilized to investigate the expression of relevant proteins and genes. In vivo experiments employed a xenograft mice model to evaluate the anti-cancer efficacy of S.C. Results Our study demonstrated that S.C effectively inhibited the viability of various prostate cancer cells. Notably, S.C exhibited the ability to enhance the cytotoxicity of docetaxel in DU145 cells. We found that S.C-induced cell death partially relied on the induction of ferroptosis, which was mediated through up-regulation of HMOX1 protein. Additionally, our investigation revealed that S.C treatment decreased the stability of BACH1 protein, which contributed to HMOX1expression. We further identified that S.C-induced ROS caused BACH1 instability by suppressing USP47expression. Moreover, In DU145 xenograft model, we found S.C significantly inhibited prostate cancer growth, highlighting its potential as a therapeutic strategy. Collectively, these findings provide evidence that S.C could induce regulated cell death (RCD) in prostate cancer cells and effectively inhibit tumor growth via triggering ferroptosis. This study provides evidence that S.C effectively suppresses tumor progression and induces ferroptosis in prostate cancer cells by targeting ROS/USP47/BACH1/HMOX1 axis. Conclusion This study provides evidence that S.C effectively suppresses tumor progression and induces ferroptosis in prostate cancer cells by targeting the ROS/USP47/BACH1/HMOX1 axis. These findings offer novel insights into the underlying mechanism by which S.C inhibits the progression of prostate cancer. Furthermore, leveraging the potential of S.C in targeting ferroptosis may present a new therapeutic opportunity for prostate cancer. This study found that S.C induces ferroptosis by targeting the ROS/USP47/BACH1/HMOX1 axis in prostate cancer cells. Graphical Abstract
... Subsequently, BACH1 molecules translocate to the cytoplasm, undergo ubiquitination, and are subsequently degraded by the proteasome. This process allows Maf to form a homodimer with Nrf2 through the same ARE region, leading to transcriptional activation (Reichard et al., 2007). Activation of Nrf2 requires its partner Kelch-like ECH-associated protein 1 (Keap1). ...
Article
Full-text available
Ferroptosis is a type of programmed cell death that pathogens can leverage to enhance their replication, transmission, and pathogenicity. Hosts typically combat pathogenic infections by utilizing oxidative stress as a defense mechanism. Nonetheless, some pathogens can trigger considerable oxidative stress while infecting, inducing an intense inflammatory response in the host’s immune system and activating cell death. The process of ferroptosis is closely linked to oxidative stress, with their interaction exerting a substantial impact on the outcome of infectious diseases. This article presents an overview of the interrelated mechanisms of both Ferroptosis and oxidative stress in infectious diseases, identifying potential targets for treating such diseases in the context of their interaction.
... Unlike the Nrf2-sMaf complexes, the BACH1-sMaf primarily acts as a transcriptional repressor (Oyake et al., 1996;Liu et al., 2022). Previous literature has indicated that BACH1-sMaf represses NQO1 and HMOX-1 gene expression in a variety of cell lines (Sun et al., 2002;Dhakshinamoorthy et al., 2005;Reichard et al., 2007). Using siRNA to silence BACH1, HMOX-1 expression is strongly upregulated in HaCaT cells and Huh-7 hepatocytes (Shan et al., 2004;Casares et al., 2020). ...
Article
Full-text available
Intracerebral hemorrhage (ICH) is a subtype of stroke with a high mortality rate. Oxidative stress cascades play an important role in brain injury after ICH. Cannabidiol, a major non-psychotropic phytocannabinoids, has drawn increasing interest in recent years as a potential therapeutic intervention for various neuropsychiatric disorders. Here we provide a comprehensive review of the potential therapeutic effects of cannabidiol in countering oxidative stress resulting from ICH. The review elaborates on the various sources of oxidative stress post-ICH, including mitochondrial dysfunction, excitotoxicity, iron toxicity, inflammation, and also highlights cannabidiol’s ability to inhibit ROS/RNS generation from these sources. The article also delves into cannabidiol’s role in promoting ROS/RNS scavenging through the Nrf2/ARE pathway, detailing both extranuclear and intranuclear regulatory mechanisms. Overall, the review underscores cannabidiol’s promising antioxidant effects in the context of ICH and suggests its potential as a therapeutic option.
Article
Full-text available
With the rapid development of new generations of antitumor therapies, the average survival time of cancer patients is expected to be continuously prolonged. However, these therapies often lead to cardiotoxicity, resulting in a growing number of tumor survivors with cardiovascular disease. Therefore, a new interdisciplinary subspecialty called “cardio-oncology” has emerged, aiming to detect and treat cardiovascular diseases associated with tumors and antitumor therapies. Recent studies have highlighted the role of ferroptosis in both cardiovascular and neoplastic diseases. The balance between intracellular oxidative stress and antioxidant defense is crucial in regulating ferroptosis. Tumor cells can evade ferroptosis by upregulating multiple antioxidant defense pathways, while many antitumor therapies rely on downregulating antioxidant defense and promoting ferroptosis in cancer cells. Unfortunately, these ferroptosis-inducing antitumor therapies often lack tissue specificity and can also cause injury to the heart, resulting in ferroptosis-induced cardiotoxicity. A range of cardioprotective agents exert cardioprotective effects by inhibiting ferroptosis. However, these cardioprotective agents might diminish the efficacy of antitumor treatment due to their antiferroptotic effects. Most current research on ferroptosis only focuses on either tumor treatment or heart protection but rarely considers both in concert. Therefore, further research is needed to study how to protect the heart during antitumor therapies by regulating ferroptosis. In this review, we summarized the role of ferroptosis in the treatment of neoplastic diseases and cardiovascular diseases and also attempted to propose further research directions for ferroptosis in the field of cardio-oncology.
Article
Full-text available
We have characterized further the antioxidant responsive element (ARE) identified in the 5'-flanking region of the rat glutathione S-transferase Ya subunit gene and the NAD(P)H:quinone reductase gene by mutational and deletion analyses. Our data suggest that the sequence, 5'-puGTGACNNNGC-3' 3'-pyCACTGNNNCG-5' where N is any nucleotide, represents the core sequence of the ARE required for transcriptional activation by phenolic antioxidants and metabolizable planar aromatic compounds (e.g. beta-naphthoflavone and 3-methylcholanthrene). We also have found that the ARE is responsive to hydrogen peroxide and phenolic antioxidants that undergo redox cycling. These latter data suggest that the ARE is responsive to reactive oxygen species and thus may represent part of a signal transduction pathway that allow eukaryotic cells to sense and respond to oxidative stress.
Article
Full-text available
In contrast to mouse epidermal cells, human skin keratinocytes are rather resistant to transformation in vitro. Immortalization has been achieved by SV40 but has resulted in cell lines with altered differentiation. We have established a spontaneously transformed human epithelial cell line from adult skin, which maintains full epidermal differentiation capacity. This HaCaT cell line is obviously immortal (greater than 140 passages), has a transformed phenotype in vitro (clonogenic on plastic and in agar) but remains nontumorigenic. Despite the altered and unlimited growth potential, HaCaT cells, similar to normal keratinocytes, reform an orderly structured and differentiated epidermal tissue when transplanted onto nude mice. Differentiation-specific keratins (Nos. 1 and 10) and other markers (involucrin and filaggrin) are expressed and regularly located. Thus, HaCaT is the first permanent epithelial cell line from adult human skin that exhibits normal differentiation and provides a promising tool for studying regulation of keratinization in human cells. On karyotyping this line is aneuploid (initially hypodiploid) with unique stable marker chromosomes indicating monoclonal origin. The identity of the HaCaT line with the tissue of origin was proven by DNA fingerprinting using hypervariable minisatellite probes. This is the first demonstration that the DNA fingerprint pattern is unaffected by long-term cultivation, transformation, and multiple chromosomal alterations, thereby offering a unique possibility for unequivocal identification of human cell lines. The characteristics of the HaCaT cell line clearly document that spontaneous transformation of human adult keratinocytes can occur in vitro and is associated with sequential chromosomal alterations, though not obligatorily linked to major defects in differentiation.
Article
Full-text available
Members of the small Maf family (MafK, MafF, and MafG) are basic region leucine zipper (bZip) proteins that can function as transcriptional activators or repressors. The dimer compositions of their DNA binding forms determine whether the small Maf family proteins activate or repress transcription. Using a yeast two-hybrid screen with a GAL4-MafK fusion protein, we have identified two novel bZip transcription factors, Bach1 and Bach2, as heterodimerization partners of MafK. In addition to a Cap'n'collar-type bZip domain, these Bach proteins possess a BTB domain which is a protein interaction motif; Bach1 and Bach2 show significant similarity to each other in these regions but are otherwise divergent. Whereas expression of Bach1 appears ubiquitous, that of Bach2 is restricted to monocytes and neuronal cells. Bach proteins bind in vitro to NF-E2 binding sites, recognition elements for the hematopoietic transcription factor NF-E2, by forming heterodimers with MafK. Furthermore, a DNA binding complex that contained MafK as well as Bach2 or a protein related closely to Bach2 was found to be present in mouse brain cells. Bach1 and Bach2 function as transcription repressors in transfection assays using fibroblast cells, but they function as a transcriptional activator and repressor, respectively, in cultured erythroid cells. The results suggest that members of the Bach family play important roles in coordinating transcription activation and repression by MafK.
Article
Full-text available
Twenty-four base pairs of the human antioxidant response element (hARE) are required for high basal transcription of the NAD(P)H:quinone oxidoreductase1 (NQO1) gene and its induction in response to xenobiotics and antioxidants. hARE is a unique cis-element that contains one perfect and one imperfect AP1 element arranged as inverse repeats separated by 3 bp, followed by a "GC" box. We report here that Jun, Fos, Fra, and Nrf nuclear transcription factors bind to the hARE. Overexpression of cDNA derived combinations of the nuclear proteins Jun and Fos or Jun and Fra1 repressed hARE-mediated chloramphenicol acetyltransferase (CAT) gene expression in transfected human hepatoblastoma (Hep-G2) cells. Further experiments suggested that this repression was due to overexpression of c-Fos and Fra1, but not due to Jun proteins. The Jun (c-Jun, Jun-B, and Jun-D) proteins in all the possible combinations were more or less ineffective in repression or upregulation of hARE-mediated gene expression. Interestingly, overexpression of Nrf1 and Nrf2 individually in Hep-G2 and monkey kidney (COS1) cells significantly increased CAT gene expression from reporter plasmid hARE-thymidine kinase-CAT in transfected cells that were inducible by beta-naphthoflavone and teri-butyl hydroquinone. These results indicated that hARE-mediated expression of the NQO1 gene and its induction by xenobiotics and antioxidants are mediated by Nrf1 and Nrf2. The hARE-mediated basal expression, however, is repressed by overexpression of c-Fos and Fra1.
Article
Full-text available
Exposure of human and rodent cells to a wide variety of chemoprotective compounds confers resistance against a broad set of carcinogens. For a subset of the chemoprotective compounds, protection is generated by an increase in the abundance of protective enzymes like glutathione S-transferases (GST). Antioxidant responsive elements (AREs) mediate the transcriptional induction of a battery of genes which comprise much of this chemoprotective response system. Past studies identified a necessary ARE "core" sequence of RTGACnnnGC, but this sequence alone is insufficient to mediate induction. In this study, the additional sequences necessary to define a sufficient, functional ARE are identified through systematic mutational analysis of the murine GST Ya ARE. Introduction of the newly identified necessary nucleotides into the regions flanking a nonresponsive, ARE-like, GST-Mu promoter sequence produced an inducible element. A screen of the GenBank database with the newly identified ARE consensus identified 16 genes which contained the functional ARE consensus sequence in their promoters. Included within this group was an ARE sequence from the murine ferritin-L promoter that mediated induction when tested. In an electrophoretic mobility-shift assay, the ferritin-L ARE was bound by ARE-binding protein 1, a protein previously identified as the likely mediator of the chemoprotective response. A three-level ARE classification system is presented to account for the distinct induction strengths observed in our mutagenesis studies. A model of the ARE as a composite regulatory site, where multiple transcription factors interact, is presented to account for the complex characteristics of ARE-mediated chemoprotective gene expression.
Article
Full-text available
The human beta-globin locus control region (LCR) is required to properly regulate chromatin domain opening, replication timing, and globin gene activation. The LCR contains multiple NF-E2 sites (Maf recognition elements, MAREs) that allow the binding of various basic leucine zipper (bZip) proteins like p45 NF-E2, Nrf1, Nrf2, Bach1, and Bach2, in some cases as obligate heterodimers with a small Maf protein. In addition to the bZip domain, the Bach proteins bear a BTB/POZ domain, which has been implicated in the regulation of chromatin structure. We show here that Bach1 is highly expressed in hematopoietic cells and constitutes one of the two MARE-binding activities in murine erythroleukemic (MEL) cells. We further demonstrate that Bach1/MafK heterodimers interact with each other through the BTB domain, generating a multimeric and multivalent DNA binding complex. These results strongly implicate Bach1/MafK heterodimer as an architectural transcription factor that mediates interactions among multiple MAREs. Such a factor could then provide a model for assembly of the theoretical beta-globin LCR "holocomplex. " Other BTB domain proteins have already been demonstrated to be involved in remodeling chromatin, and thus this class of proteins likely promote the formation of nucleoprotein complexes required to establish the architecture of regulatory domains.
Article
Full-text available
Transcription factor Nrf2 is essential for the antioxidant responsive element (ARE)-mediated induction of phase II detoxifying and oxidative stress enzyme genes. Detailed analysis of differential Nrf2 activity displayed in transfected cell lines ultimately led to the identification of a new protein, which we named Keap1, that suppresses Nrf2 transcriptional activity by specific binding to its evolutionarily conserved amino-terminal regulatory domain. The closest homolog of Keap1 is a Drosophila actin-binding protein called Kelch, implying that Keap1 might be a Nrf2 cytoplasmic effector. We then showed that electrophilic agents antagonize Keap1 inhibition of Nrf2 activity in vivo, allowing Nrf2 to traverse from the cytoplasm to the nucleus and potentiate the ARE response. We postulate that Keap1 and Nrf2 constitute a crucial cellular sensor for oxidative stress, and together mediate a key step in the signaling pathway that leads to transcriptional activation by this novel Nrf2 nuclear shuttling mechanism. The activation of Nrf2 leads in turn to the induction of phase II enzyme and antioxidative stress genes in response to electrophiles and reactive oxygen species.
Article
Full-text available
The transcription factor Bach1 is a member of a novel family of broad complex, tramtrack,bric-a-brac/poxvirus and zinc finger (BTB/POZ) basic region leucine zipper factors. Bach1 forms a heterodimer with MafK, a member of the small Maf protein family (MafF, MafG, and MafK), which recognizes the NF-E2/Maf recognition element, a cis-regulatory motif containing a 12-O-tetradecanoylphorbol-13-acetate-responsive element. Here we describe the gene structure of human BACH1, including a newly identified promoter and an alternatively RNA-spliced truncated form of BACH1, designated BACH1t, abundantly transcribed in human testis. The alternate splicing originated from the usage of a novel exon located 5.6 kilobase pairs downstream of the exon encoding the leucine zipper domain, and produced a protein that contained the conserved BTB/POZ, Cap'n collar, and basic region domains, but lacked the leucine zipper domain essential for NF-E2/Maf recognition element binding. Subcellular localization studies using green fluorescent protein as a reporter showed that full-length BACH1 localized to the cytoplasm, whereas BACH1t accumulated in the nucleus. Interestingly, coexpression of BACH1 and BACH1t demonstrated interaction between the molecules and the induction of nuclear import of BACH1. These results suggested that BACH1t recruits BACH1 to the nucleus through BTB domain-mediated interaction.
Article
Full-text available
Heme controls expression of genes involved in the synthesis of globins and heme. The mammalian transcription factor Bach1 functions as a repressor of the Maf recognition element (MARE) by forming antagonizing hetero-oligomers with the small Maf family proteins. We show here that heme binds specifically to Bach1 and regulates its DNA-binding activity. Deletion studies demonstrated that a heme-binding region of Bach1 is confined within its C-terminal region that possesses four dipeptide cysteine-proline (CP) motifs. Mutations in all of the CP motifs of Bach1 abolished its interaction with heme. The DNA-binding activity of Bach1 as a MafK hetero-oligomer was markedly inhibited by heme in gel mobility shift assays. The repressor activity of Bach1 was lost upon addition of hemin in transfected cells. These results suggest that increased levels of heme inactivate the repressor Bach1, resulting in induction of a host of genes with MARES:
Article
The induction of phase II detoxifying enzymes is an important defense mechanism against intake of xenobiotics. While this group of enzymes is believed to be under the transcriptional control of antioxidant response elements (AREs), this contention is experimentally unconfirmed. Since the ARE resembles the binding sequence of erythroid transcription factor NF-E2, we investigated the possibility that the phase II enzyme genes might be regulated by transcription factors that also bind to the NF-E2 sequence. The expression profiles of a number of transcription factors suggest that an Nrf2/small Maf heterodimer is the most likely candidate to fulfill this rolein vivo.To directly test these questions, we disrupted the murinenrf2 genein vivo.While the expression of phase II enzymes (e.g., glutathione S-transferase and NAD(P)H: quinone oxidoreductase) was markedly induced by a phenolic antioxidantin vivoin both wild type and heterozygous mutant mice, the induction was largely eliminated in the liver and intestine of homozygousnrf2-mutant mice. Nrf2 was found to bind to the ARE with high affinity only as a heterodimer with a small Maf protein, suggesting that Nrf2/small Maf activates gene expression directly through the ARE. These results demonstrate that Nrf2 is essential for the transcriptional induction of phase II enzymes and the presence of a coordinate transcriptional regulatory mechanism for phase II enzyme genes. Thenrf2-deficient mice may prove to be a very useful model for thein vivoanalysis of chemical carcinogenesis and resistance to anti-cancer drugs.