You are on page 1of 6

O2 availability impacts iron homeostasis in

Escherichia coli
Nicole A. Beauchenea, Erin L. Metterta, Laura J. Mooreb, Sündüz Keleşc,d, Emily R. Willeya, and Patricia J. Kileya,1
a
Department of Biomolecular Chemistry, University of Wisconsin–Madison, Madison, WI 53706; bDepartment of Chemistry and Biochemistry, Monmouth
College, Monmouth, IL 61462; cDepartment of Statistics, University of Wisconsin–Madison, Madison, WI 53706; and dDepartment of Biostatistics and
Medical Informatics, University of Wisconsin–Madison, Madison, WI 53706

Edited by Carol A. Gross, University of California, San Francisco, CA, and approved October 3, 2017 (received for review April 30, 2017)

The ferric-uptake regulator (Fur) is an Fe2+-responsive transcrip- O2 tension influences the oxidation state of iron, and thereby
tion factor that coordinates iron homeostasis in many bacteria. the means by which it is acquired by cells (1, 3, 16, 17). When O2
Recently, we reported that expression of the Escherichia coli Fur is absent, iron is primarily in its soluble, reduced Fe2+ form and
regulon is also impacted by O2 tension. Here, we show that for can be taken up directly by bacteria via dedicated pathways (e.g.,
most of the Fur regulon, Fur binding and transcriptional repression the FeoABC system). In contrast, when O2 is present, iron is
increase under anaerobic conditions, suggesting that Fur is con- oxidized to an insoluble ferric (Fe3+) state. The uptake of Fe3+
trolled by O2 availability. We found that the intracellular, labile involves the synthesis and secretion of Fe3+-chelating molecules
Fe2+ pool was higher under anaerobic conditions compared with called siderophores (e.g., enterobactin). Transport of the Fe3+-
aerobic conditions, suggesting that higher Fe2+ availability drove bound siderophore into bacterial cells involves specific outer
the formation of more Fe2+-Fur and, accordingly, more DNA bind- membrane receptors, the energy-transducing TonB–ExbB–ExbD
ing. O2 regulation of Fur activity required the anaerobically in-
complex, and ABC-type transporters that ultimately deliver the
Fe3+-siderophore to the cytoplasm, where the iron is reduced
duced FeoABC Fe2+ uptake system, linking increased Fur activity
and released. Given the impact of O2 on the oxidation state of
to ferrous import under iron-sufficient conditions. The increased
iron and the fact that Fur controls expression of both ferrous and
activity of Fur under anaerobic conditions led to a decrease in ex-
ferric iron assimilation pathways, an O2-dependent input into the
pression of ferric import systems. However, the combined positive expression of the Fur regulon would make physiological sense.
regulation of the feoABC operon by ArcA and FNR partially antag- In this study, we show that such an O2-dependent mechanism
onized Fur-mediated repression of feoABC under anaerobic condi- exists by probing how O2 impacts Fur activity. We investigated
tions, allowing ferrous transport to increase even though Fur is the correlation between increased anaerobic Fur DNA binding
more active. This design feature promotes a switch from ferric im- across the genome and increased transcriptional repression
port to the more physiological relevant ferrous iron under anaerobic mediated by Fur. To dissect the mechanisms by which the
conditions. Taken together, we propose that the influence of O2 amount of active Fur is elevated during anaerobiosis, we mea-
availability on the levels of active Fur adds a previously undescribed sured the levels of intracellular chelatable Fe2+ and Fur protein
layer of regulation in maintaining cellular iron homeostasis. in aerobic and anaerobic cells, and we assayed Fur activity in
strains lacking proteins involved in iron uptake or storage. Fi-
ferric uptake regulator | iron homeostasis | anaerobiosis | labile iron pool | nally, we measured the response of Fur to O2 following a shift
protein metallation from anaerobic to aerobic conditions and addressed the influ-
ence of other O2-responsive transcription factors on expression
of genes in the direct Fur regulon. Together, our findings provide
I ron homeostasis requires coordination between iron uptake,
storage, and cofactor synthesis to supply iron for the iron-
containing proteome (1–3). In bacteria, this coordination is or-
insight on the means by which O2-dependent regulation of Fur

chestrated by transcription factors that are predicted to provide a Significance


direct readout of the cellular iron status (4). In particular, the
transcription factor ferric uptake regulator (Fur) plays a leading Our understanding of how cells regulate intracellular iron
role in many bacteria to maintain iron homeostasis by directly pools has been largely shaped by studying cells grown under
aerobic conditions, in which the barrier to iron acquisition is
sensing levels of ferrous iron (Fe2+) and altering gene transcrip-

MICROBIOLOGY
dominated by O2-dependent insolubility. However, less is known
tion accordingly (3, 5). Recently, we reported that transcriptional
about how bacteria meet their iron demands in the O2-limiting or
repression of many genes of the Escherichia coli K12 Fur regulon anaerobic environments that are common to many ecosystems
is enhanced under anaerobic growth conditions (6), suggesting and reflective of the ancient atmosphere of early Earth. Because
that O2 provides an additional input into Fur regulation. the transcription factor ferric-uptake regulator (Fur) plays a central
The regulation of Fur activity by iron has provided the para- role in controlling iron homeostasis in many bacterial species, we
digm for understanding how bacteria maintain appropriate iron use Fur activity as a surrogate for understanding how anaerobi-
levels in vivo. When iron is available, each subunit of a Fur dimer osis alters iron homeostasis. Our finding that levels of active Fur
binds one molecule of Fe2+ ion (7, 8). In this Fe2+-bound form, are increased during anaerobiosis emphasizes fundamental dif-
Fur exhibits increased affinity for specific DNA sequences [re- ferences in bacterial iron requirements between aerobic and an-
ferred to a Fur boxes (9, 10)] and represses transcription of many aerobic growth conditions.
iron assimilation pathways. In contrast, when iron is limiting, Fur
shifts to an iron-free form, thereby inactivating Fur function. The Author contributions: N.A.B. and P.J.K. designed research; N.A.B., E.L.M., L.J.M., and
low micromolar binding affinity of Fe2+ for Fur (11–13) affords a E.R.W. performed research; S.K. contributed new reagents/analytic tools; N.A.B., E.L.M.,
window into the physiologically relevant iron concentration that and P.J.K. analyzed data; and N.A.B., E.L.M., and P.J.K. wrote the paper.

Fur responds to in cells. Indeed, many studies suggest that under The authors declare no conflict of interest.
aerobic growth conditions, the intracellular iron pool available for This article is a PNAS Direct Submission.
Fur binding is in the low micromolar range (14, 15). However, our Published under the PNAS license.
recent finding that Fur-mediated repression is enhanced under 1
To whom correspondence should be addressed. Email: pjkiley@wisc.edu.
anaerobic conditions (6) raises the question of whether the in- This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
tracellular iron pool is altered during anaerobiosis. 1073/pnas.1707189114/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1707189114 PNAS | November 14, 2017 | vol. 114 | no. 46 | 12261–12266


occurs, which involves the up-regulation of Fe2+ uptake under A Fur Binding Site Is Sufficient to Confer Regulation by O2. To test if
anaerobic conditions. the presence of a Fur DNA binding site is sufficient to confer O2-
dependent regulation as suggested by our genome-wide analysis,
Results we engineered a synthetic Fur-dependent promoter (PfepBS), in
Fur DNA Occupancy Increases During Anaerobiosis. A previous genome- which the tac promoter (Ptac) contains the Fur binding site from
wide study of Fur–DNA interactions using ChIP-sequencing PfepB (which drives expression of FepB, an enterobactin binding
(ChIP-seq) revealed that under anaerobic conditions, Fur bound protein). This binding site is predicted to bind two Fur dimers
more regions in the E. coli K12 genome than under aerobic growth (20, 21) and was positioned in Ptac to maintain the same spacing
conditions (6). This finding led us to hypothesize that during an- (−19 to +21 bp) with respect to the promoter −35 hexamer from
aerobiosis, an increase in the amount of active Fe2+-Fur could PfepB. In a strain containing PfepBS driving expression of a chro-
explain Fur binding to DNA sites that are more varied from the mosomal lacZ reporter gene, β-galactosidase activity was signif-
consensus Fur binding motif (6) than those occupied by Fur in icantly decreased when Fur was present (Fig. S1), indicating that
both the presence and absence of O2. Here, we build upon these Fur mediates repression of PfepBS in vivo. Furthermore, Fur-
results to address whether the O2-dependent changes in expres- dependent repression was increased threefold under anaerobic
sion observed for the Fur regulon (6) can be attributed to changes growth conditions compared with aerobic growth conditions
in Fur DNA occupancy genome-wide. (Fig. 2), demonstrating a direct effect of O2 availability on Fur-
By quantifying the ratio of Fur enrichment for 74 of the dependent regulation of PfepBS. Taken together, the enhance-
highest intensity iron-dependent ChIP-seq peaks (6) using the ment in Fur occupancy at genomic sites, along with the increase
statistics-based differential binding algorithm DBChIP (18), we in Fur repression under anaerobic conditions, provides evidence
found that, overall, the ratio of anaerobic to aerobic Fur ChIP- for a direct role of Fur in the global response to anaerobiosis.
seq signal increased, ranging from ∼1.5- to 10-fold, with a me-
dian of threefold (Tables S1 and S2). The increase in the ratio of Additional Transcription Factors Can Tailor O2-Mediated Expression of
anaerobic to aerobic Fur ChIP-seq signal at each site suggested Some Members of the Fur Regulon. For a few members of the Fur
an increase in Fe2+-Fur binding at these sites. regulon, we hypothesized from previous results that the impact
The anaerobic increase in genome-scale Fur DNA binding of O2 on their transcription might also require the O2-regulated
appears to be transcriptionally relevant since, for most of the Fur transcription factors FNR and ArcA (22, 23). Therefore, we
regulon, the ratio of anaerobic to aerobic Chip-seq signal cor- measured expression of lacZ fusions to PexbB, Pfiu, PcirA, or PfeoA
related positively with the fold increase in Fur-mediated re- in strains lacking one or more of these transcription factors. In
pression calculated from anaerobic and aerobic transcriptomic the absence of any mutations, expression of PexbB, Pfiu, and PcirA
data, respectively (Table S1). For example, Fur binding to the decreased under anaerobic conditions (Fig. 3 A–C), in agree-
promoter regions of exbB and tonB (both involved in Fe3+- ment with transcriptomic data (6). For PexbB, we found that de-
siderophore transport) and bfd (involved in iron storage/release letion of fur increased expression to comparable levels under
from bacterioferritin) increased 6.2-, 5.2-, and 3.9-fold under both aerobic and anaerobic conditions, indicating Fur-mediated
anaerobic conditions, respectively (Fig. 1 and Table S1). This repression was sufficient to explain its decreased expression
corresponded to a 6.0-, 3.3-, and 7.0-fold respective increase in under anaerobic conditions (Fig. 3A). Deletion of fnr had no
Fur-mediated repression of these genes under anaerobic condi- effect on PexbB expression, despite a reported FNR binding site
tions compared with aerobic conditions (Table S1). In instances (22). In contrast, for Pfiu and PcirA, we found that decreased
where the ChIP-seq ratio measured for Fur occupancy in an- expression of these promoters under anaerobic conditions was
aerobic and aerobic cells was closer to 1, smaller O2-dependent only partially dependent on Fur (Fig. 3 B and C). For both
changes in gene expression were detected (e.g., entC and fhuE, promoters, we found that ArcA also contributes to anaerobic
involved in Fe3+-siderophore transport; ryhB, encodes a regula- repression (Fig. 3 B and C), in agreement with ArcA binding
tory small RNA) (Table S1). This latter group of promoters these promoter regions in vivo (23). This additional role of ArcA
represents some of the more strongly repressed Fur genes (6, likely explains why the increase in Fur binding to Pfiu and PcirA
19), and the lower ratio likely stems from nearly comparable Fur anaerobically did not correlate well with the fold increase in Fur-
binding under both aerobic and anaerobic conditions. In sum- mediated repression (Table S1).
mary, these data indicate that under anaerobic conditions, Fur Similar analysis of the promoter driving transcription of feoABC,
binding is increased at many promoters, which can lead to encoding a ferrous uptake system, revealed several distinctive fea-
comparable changes in Fur-mediated gene repression. tures (Fig. 3D). First, in contrast to the ferric uptake systems, PfeoA
is poorly expressed aerobically even in the absence of Fur. Second,
unlike the ferric uptake systems (e.g., cirA, exbB, fiu), expression of
PfeoA increased a small amount under anaerobic conditions com-
pared with aerobic conditions, despite increased anaerobic Fe2+-
Fur levels. Third, we found that this increased anaerobic expression
of PfeoA required both of the anaerobic transcription factors FNR
and ArcA, as suggested by genome-wide and other data (22–
24). This positive effect of both ArcA and FNR appears to limit
the negative effect of Fur under anaerobic conditions. Thus,
FNR and ArcA produce an expression pattern for PfeoA under
anaerobic conditions opposite to the expression pattern of the
ferric transport systems. Taken together, these data support the
notion that Fur plays a central role in reprogramming gene ex-
pression profiles in response to O2 because of changes in levels of
Fig. 1. Occupancy of Fur binding regions increases during anaerobiosis. Fur
Fe2+-Fur available for DNA binding and by Fur acting in concert
ChIP-seq peak regions for exbB, tonB, and bfd from cells grown under an-
with other transcription factors.
aerobic (blue) or aerobic (red) conditions in MOPS minimal glucose media
containing 10 μM FeSO4, were plotted from data reported by Beauchene The Labile Fe2+ Pool Increases During Anaerobiosis. To determine
et al. (6). Shown are three replicates for each growth condition. The x axis the basis for increased levels of Fe2+-Fur under anaerobic con-
indicates the genomic position of the ChIP-seq peak, and the y axis indicates ditions, we measured the levels of both Fur protein and iron.
the read count after each dataset was normalized to 20 million reads. The Western blot analysis revealed that Fur protein levels are com-
ratio of the anaerobic ChIP signal to aerobic ChIP signal for all Fur ChIP-seq parable under aerobic (22 μM) and anaerobic (17 μM) growth
peaks is presented in Tables S1 and S2. conditions (Fig. S2). To test if increased intracellular Fe2+

12262 | www.pnas.org/cgi/doi/10.1073/pnas.1707189114 Beauchene et al.


composition (rich versus minimal) influences Fur-mediated re-
pression under aerobic or anaerobic conditions. Fur-mediated
repression was assayed using the synthetic PfebBS-lacZ fusion.
We did not find any changes in O2-dependent regulation of Fur
activity when the media were varied from our standard 3-(N-
morpholino)propanesulfonic acid (MOPS) minimal media (Fig.
S3). We also considered the possibility that decreased repression
by Fur under aerobic conditions could be explained if iron up-
take was less efficient due to the decreased solubility of iron in
the presence of O2. However, increasing external iron levels 10-
fold resulted in no change in the levels of Fur-mediated re-
pression under aerobic conditions, suggesting that iron avail-
ability was not limiting (Fig. S3). In contrast, under anaerobic
conditions, repression by Fur was increased even further, sug-
gesting that iron transport could be contributing to Fur regula-
tion under anaerobic conditions.
Fig. 2. Fur-dependent repression of a synthetic promoter containing a Fur Since the FeoABC system has a major role in anaerobic Fe2+
DNA binding site is influenced by O2 and the FeoABC ferrous uptake sys- uptake (17, 33, 34), we investigated whether anaerobic regulation
tem. Fur-mediated repression of a chromosomal Ptac-lacZ fusion bearing the of Fur was disrupted in a strain lacking feoB. Indeed, we found
E. coli fepB Fur binding site (PfepBS) is shown in wild-type, Δfur, ΔfeoB, ΔfecC, that elimination of the FeoB iron transporter caused a threefold
ΔentA, and ΔftnA Δbfr Δdps strains grown under aerobic (gray bars) or defect in PfepBS repression by Fur under anaerobic conditions (Fig.
anaerobic (white bars) conditions in MOPS minimal glucose media containing 2). Furthermore, O2-dependent regulation of Fur required FeoB
1.0 μM FeSO4. Fold repression was calculated by dividing the average promoter since Fur repressed PfepBS to a similar extent under aerobic and
activity in the Δfur strain by the average promoter activity in each of the other
anaerobic conditions in the ΔfeoB mutant (Fig. 2). The effect of
strains tested (at least two replicates of each; promoter activity was determined
using β-galactosidase assays, as shown in Fig. S1), and error bars represent the
feoB on Fur activity did not appear to reflect a general defect in
propagation of errors calculated according to the method of Ku (45). iron-containing transcription factors since we did not find any
change in activity of two Fe-S cluster-containing transcription
factors (FNR and IscR) when similarly assayed under anaerobic
availability could explain the increase in active Fe2+-Fur protein conditions (Fig. S4). These data suggest that FeoABC contributes
levels under anaerobic conditions, we used chelator-assisted elec- to the O2-dependent regulation of Fur by increasing the anaerobic
tron paramagnetic resonance (EPR) spectroscopy (25) to measure labile iron pool available for Fur binding.
the intracellular chelatable Fe2+ pool. This labile pool is estimated
to be 1.0% of cellular iron (14) and is defined as iron not tightly
bound to cellular proteins, and possibly associated with metabolites,
such as glutathione, citrate, or phosphorylated sugar compounds
(25–27). We found that the amount of chelatable Fe2+ was ap-
proximately sevenfold higher in anaerobic cells compared with
aerobic cells (Fig. 4 A and B). When total cellular iron levels (which
encompass both Fe2+ and Fe3+ in proteins and the labile iron pool)
were measured by inductively coupled plasma mass spectrometry,
there was only a slight increase in the total iron content of anaer-
obically grown cells (Fig. 4C). These results indicate that while total
cellular iron in aerobically and anaerobically grown cells is com-
parable, the labile iron pool (that which is accessible for Fur
binding) is higher in the absence of O2.
To provide further support for an elevated labile Fe2+ pool in
anaerobic cells, we assayed the sensitivity of cells to hydrogen
peroxide (H2O2). H2O2 is reduced by Fe2+, resulting in the

MICROBIOLOGY
formation of hydroxyl radicals via the Fenton reaction (28).
Hydroxyl radicals are highly detrimental to the integrity of cel-
lular components, such as DNA (29, 30), and can lead to cell
death. We determined the sensitivity of aerobic and anaerobic
cells to H2O2 using a strain lacking RecA, which is defective in
DNA repair and thus more sensitive to H2O2 compared with a
wild-type strain (31). Anaerobically grown cells showed 10-fold
more sensitivity to killing by H2O2 than aerobically grown cells
(Fig. 4D), suggesting that Fe2+ is more readily available in an-
aerobic cells to promote hydroxyl radicals and DNA damage.
These results agree with previous findings that anaerobic cells
are highly sensitive to H2O2 (31, 32).
Taken together, these data show that increased levels of Fe2+-
Fur under anaerobic conditions are not due to elevated Fur
protein levels, but rather to an increase in the labile Fe2+ pool.
Consequently, these higher Fe2+ levels would allow for increased Fig. 3. Fur and other O2-regulated transcription factors tailor the expres-
formation of Fe2+-Fur, and thus explain increased Fur DNA sion of iron-uptake systems. Strains bearing lacZ fusions to PexbB (A), Pfiu (B),
binding and transcriptional repression during anaerobiosis. PcirA (C), and PfeoA (D) were assayed for β-galactosidase activity under aerobic
(gray) or anaerobic (white) growth conditions. Deletion of the relevant
Regulation of the Labile Iron Pool by O2. To determine how the transcription factor in the reporter strain is indicated by (−). Cultures were
labile iron pool is increased under anaerobic conditions, we grown in MOPS minimal glucose media containing 1.0 μM FeSO4. Error bars
asked whether the source of iron (Fe2+ or Fe3+) or media represent the SE from triplicate experiments.

Beauchene et al. PNAS | November 14, 2017 | vol. 114 | no. 46 | 12263
changes in environmental O2, conditions in which E. coli would
experience in its natural habitat.
Discussion
Our data provide insight into how O2 availability alters iron
homeostasis in the facultative bacterium E. coli. Overall, we
propose that an increase in the intracellular labile Fe2+ pool,
mediated by the FeoABC ferrous transport system, leads to an
increase in Fur-mediated repression under anaerobic conditions.
The rise in anaerobic Fe2+ availability increases the population
of Fur protein that is bound to Fe2+ and, accordingly, drives Fur
DNA binding and transcriptional repression. The global change
Fig. 4. Intracellular labile Fe2+ levels increase during anaerobiosis. EPR spectra in expression of the Fur regulon under iron-sufficient anaerobic
showing the signal at g = 4.3, normalized for OD600 from aerobic or anaerobic conditions promotes a shift in expression from Fe3+ to Fe2+
cells grown in MOPS minimal glucose media containing 1.0 μM FeSO4. Aerobic transport systems, allowing E. coli to use the most physiologically
(red) or anaerobic (blue) cells were treated with desferoxamine mesylate (DFO) relevant form of iron.
or were left untreated (black). (A) Four to five replicates of each experiment
are shown. (B) Chelatable Fe levels in aerobic (gray) or anaerobic cells (white) The Labile Iron Pool Increases Under Anaerobic Conditions. It is
calculated from EPR spectra. (C) Inductively coupled plasma mass spectrometry was generally considered that the affinity of a metal sensor for its
used to measure whole-cell Fe levels in three biological replicates of aerobic (gray) metal, and the subsequent homeostatic control the regulator
or anaerobic cells (white), which were grown in MOPS minimal glucose media exerts on its regulon, defines intracellular free metal concen-
containing 1.0 μM FeSO4 and normalized for cell pellet (milligrams). (D) Viable cell trations (37–39). Thus, before this work, one would expect a
counts (assayed in triplicate) for aerobic (gray) or anaerobic cells (white) grown in priori that the intracellular free (labile) Fe2+ pool would be
MOPS minimal glucose media containing 1.0 μM FeSO4, with and without treat-
comparable in aerobically and anaerobically grown cells. Despite
ment of 2.5 mM H2O2 for 15 min. For B, C, and D, error bars represent the SE.
this expectation, we found that intracellular chelatable iron lev-
els increased nearly sevenfold (∼26 μM in aerobic cells and
We also tested whether other iron transport systems affected ∼177 μM in anaerobic cells) under anaerobic conditions. Pre-
Fur activity (3). Enterobactin-mediated Fe3+ uptake was disrupted vious studies have shown that the FeoABC transporter, which is
widely distributed in bacteria (17), contributes to anaerobic iron
by deletion of entA. The inability to synthesize the enterobactin
uptake (33, 34). Therefore, our finding that FeoABC was re-
siderophore caused a twofold loss in the ability of Fur to repress
quired for increased Fur activity under anaerobic conditions
PfepBS under both aerobic and anaerobic conditions (Fig. 2).
provided a link between increased iron import under anaerobic
However, the EntA− mutant retained O2-dependent regulation of conditions and the anaerobic labile Fe2+ pool. The ability of
Fur, presumably because of the continued function and regulated FeoABC to mediate an increase in the iron pool under anaerobic
expression of FeoABC (Fig. 2). As a control, we showed that Fur- conditions appears to be due, in part, to its unique mechanism of
mediated repression was not affected in a strain lacking fecC, in- transcriptional regulation (discussed below).
volved in transport of ferric citrate, which was not added to our
media (Fig. 2). Fur Fe2+ Occupancy Varies with O2 Tension. Since the affinity of Fur
Finally, we tested whether iron storage proteins affected the for Fe2+ is estimated at 1.2–55 μM (11–13), and the cellular Fur
amount of iron available for Fur binding by measuring Fur-
mediated repression in iron storage-deficient strains. The only
known mechanisms of iron storage in E. coli require O2 or H2O2
(35), suggesting that less iron might be stored anaerobically, and
potentially contributing to the increase in the labile iron pool.
However, elimination of iron storage proteins by deletion of
ftnA, bfr, and dps did not alter the level of Fur-mediated re-
pression under either aerobic or anaerobic conditions (Fig. 2).

Fur Activity Responds to a Shift in O2 Availability. Adapting to changes in


O2 tension is key for most organisms. Given the significant differ-
ences in the bioavailability of iron in the presence and absence of O2,
we wanted to test if the sudden introduction of O2 to anaerobic cells
altered Fur activity. To do this, we assayed β-galactosidase from lacZ
fusions to either our synthetic promoter (PfepBS) or the fepA promoter
(PfepA), which drives expression of the FepA Fe3+-enterobactin
transporter. Upon exposure to either O2 or α,α′-dipyridyl, a known
cell-permeable iron chelator, we observed an increase in PfepBS
and PfepA expression (Fig. 5 and Fig. S5). Although the rate of
increase in expression of these promoters after exposure to O2 was
slower than with α,α′-dipyridyl, de-repression of these promoters
occurred on a similar time scale.
To determine how other genes that are directly repressed by
Fur respond to O2 exposure, we analyzed time series tran-
Fig. 5. Anaerobic Fur activity decreases upon exposure to O2. A strain
scriptomic data from RNA isolated from E. coli at 0.5, 1, 2, 5,
containing PfepBS-lacZ was grown anaerobically and was then either shifted
and 10 min after a shift from anaerobic to aerobic conditions to aerobic growth conditions (●) or treated with an iron chelator (200 μM
[first reported by von Wulffen et al. (36)]. Similar to what we α,α′-dipyridyl; ▲) at time 0. Samples were removed and assayed for
observed for PfepBS and PfepA, expression of these genes showed β-galactosidase at various time points. Untreated cells (♦) are indicated.
an immediate response to the addition of O2, with almost all Cultures were grown in MOPS minimal glucose media containing 1.0 μM
RNA levels increasing by 2 min (Table S3). Thus, these data FeSO4. Error bars represent the SE from three experiments. (Inset) Expanded
indicate that expression of the Fur regulon responds to sudden view of O2- and α,α′-dipyridyl–treated cells from time 0–20 min.

12264 | www.pnas.org/cgi/doi/10.1073/pnas.1707189114 Beauchene et al.


concentration is comparable between aerobic (∼22 μM) and conditions in which their substrate (Fe2+ or Fe3+) is most likely
anaerobic (∼17 μM) cells, our data suggest that the fraction of to be found. In most cases, Fur-mediated repression was suffi-
Fur bound by Fe2+ is impacted by O2-dependent changes in Fe2+ cient to explain decreased expression of Fe3+ uptake systems
levels. We propose that under aerobic conditions even when iron under anaerobic conditions. However, for a few Fe3+ uptake
levels are sufficient, Fur is not completely occupied with Fe2+, systems (e.g., fiu, cirA), decreased expression under anaerobic
whereas under anaerobic conditions, Fur-Fe2+ occupancy increases conditions required repression by both ArcA and Fur. In con-
due to the increase in the labile iron pool (Fig. 6). It was also trast, anaerobic expression of the FeoABC Fe2+ uptake system
surprising that the decrease in iron available for Fur binding ob- is under positive control of ArcA and FNR. This regulatory
served with the ΔfeoB mutant, did not impact gene regulation mechanism prevents additional Fur-mediated repression of PfeoA
mediated by the transcription factors FNR and IscR that ligate Fe-S (unlike that of the ferric uptake systems) and explains how ex-
clusters. This result could imply that the affinity of iron for Fe-S pression increases under anaerobic conditions even when Fur is
cluster biogenesis machinery is much stronger than the binding more active. By wiring feoABC expression in this manner, Fe2+
affinity to Fur, creating a hierarchy of iron usage in the cell. can enter cells through a transport system whose expression is
not strictly controlled by Fur under anaerobic growth conditions.
The Strength of Fur DNA Recognition Sites May also Contribute to O2 In turn, anaerobic up-regulation of feoABC facilitates the in-
Regulation of the Fur Regulon. The effect of O2 on Fur-mediated crease in the iron pool and recalibrates iron homeostasis.
repression varied for individual members of the Fur regulon. It might be counterintuitive that even small changes in Fur-
Some genes exhibited only small differences in Fur-mediated mediated repression between aerobic and anaerobic conditions
regulation between aerobic and anaerobic conditions, whereas could be physiologically significant. However, it is important to
others showed a much stronger dependence on anaerobic con- note that under iron-sufficient conditions, the low expression
ditions to observe regulation by Fur. To explain these findings, levels of Fur-repressed iron uptake systems are sufficient to drive
we propose that the increase in Fe2+-Fur levels under anaerobic iron uptake. Iron uptake systems only become further induced
conditions allows binding to Fur sites of a wider range of affin- when external iron becomes scarce. Thus, it is logical to assume
ities. In contrast, under aerobic conditions, the reduced levels of that even small changes in levels of these transport systems will
Fe2+-Fur bias binding to stronger affinity Fur sites. In support of affect iron uptake rates.
this model, bioinformatic analysis of sites bound by Fur only In summary, our data show that O2 has a direct effect on levels
under anaerobic conditions was much further from consensus of active Fur in vivo and that several factors contribute to the
than that of sites bound by Fur both in the presence and absence expression profile of the Fur regulon in response to O2. The
of O2 (6). Although, there is no resource to compare Fur af- expression output of each gene is a summation of Fur activity,
which depends on the extent of Fur Fe2+ occupancy, the Fur
finities obtained under similar solution conditions for individual
DNA binding sequence, and coordinated regulation by other
sites across the regulon, information theory predictions of Fur
transcription factors. Our data highlight the complexity of co-
binding site strength suggest that many of the DNA sites, which
ordinating a central cellular activity like iron homeostasis to
bound Fur well both in the presence and absence of O2 (e.g., multiple environmental cues. This fine-tuning of transcriptional
ryhB, entC, cirA, yjjZ, fepA), should be high-affinity sites (21). regulation ensures that the necessary gene products are available
Thus, the strength of the Fur DNA binding site likely contributes for cellular function. Finally, we predict that the differential ef-
to how much of an effect O2 has on an individual regulon fect of O2 on iron homeostasis observed in this study is likely to
member. For the Fur homolog Zur, the affinity of the transcrip- be conserved in other Fur-containing facultative bacteria since it
tion factor for its DNA sites also correlates to the fold repression reflects an adaptation to the intrinsic sensitivity of Fe2+ to oxi-
(40). Further, it has been shown that stepwise metallation of the dation in aerobic environments.
Zur Zn binding sites fine-tunes the degree of Zur repression of its
target promoters (41, 42). Methods
Detailed materials and methods can be found in SI Methods.
Physiological Significance of O2 Regulation of Fur. Our data reveal
that the O2-dependent changes in levels of active Fur play a role Differential Fur DNA Binding Analyses. Differential binding of Fur between
in tailoring expression of iron uptake systems to the growth aerobic and anaerobic conditions was assessed using the algorithm DBChIP
(18). This analysis was performed on 74 previously reported iron-dependent
Fur ChIP-seq peaks from anaerobically grown E. coli with read counts >4,000 at
the peak summit (6). Read counts of each dataset were normalized to 20 mil-

MICROBIOLOGY
lion reads. The fold change between aerobic and anaerobic Fur DNA occu-
pancy, reported here as the ChIP ratio, was calculated by dividing the average
ChIP signal from three anaerobic replicates by that of three aerobic replicates.
Fur was annotated as differentially bound if the adjusted P value of the fold
change was <0.05.

Growth Conditions. For the β-galactosidase assays (Figs. 2, 3, and 5 and Figs.
S1 and S3–S5) and H2O2 sensitivity assays (Fig. 4D), strains were grown at
37 °C in MOPS minimal 0.2% glucose media (43) containing either 1.0 μM or
10 μM FeSO4 under aerobic conditions (by shaking) or anaerobic conditions
(using filled screw-capped tubes) as described previously (44). In Fig. S3,
different media and iron sources are indicated in the figure legend.
β-galactosidase assays were performed as described (6).
Fig. 6. O2-dependent regulation of the Fur regulon. How O2 influences the For the in vivo O2 shift experiments (Fig. 5 and Fig. S5), strains bearing
Fe2+ occupancy of Fur under aerobic and anaerobic growth conditions is PfepA-lacZ or PfepBS-lacZ were grown at 37 °C in MOPS minimal 0.2% glucose
depicted. In the absence of O2, increased intracellular labile Fe2+ levels, media containing 1.0 μM FeSO4 by sparging with an anaerobic gas mix of 95%
which are mediated by the FeoABC system, promote Fur Fe2+ occupancy, and N2 and 5% CO2. At time 0 min, which corresponded to an OD600 of 0.1, cultures
thus lead to increased repression of Fur-dependent promoters. In the pres- were either shifted to aerobic conditions by switching the gas mix to 70% N2,
ence of O2, cells shift to using Fe3+ uptake systems and intracellular labile 5% CO2, and 25% O2, or were treated with α,α′-dipyridyl (final concentration of
Fe2+ levels are lower compared with those in anaerobic cells. As a result, 200 μM). Samples were removed at various time points for measurement of
levels of Fe2+-Fur are decreased, leading to partial occupancy and less re- OD600 and β-galactosidase activity as described by Beauchene et al. (6).
pression of promoters containing less conserved Fur binding sites. However, For EPR spectroscopy and Western blot analysis, cultures were grown at
promoters containing more conserved Fur binding sites remain repressed. 37 °C in MOPS minimal glucose media containing 1.0 μM FeSO4 to an OD600

Beauchene et al. PNAS | November 14, 2017 | vol. 114 | no. 46 | 12265
of ∼0.4 (Lambda 25 UV/Vis Spectrophotometer; PerkinElmer) by sparging H2O2 Sensitivity Assays. Cells were treated with 2.5 mM H2O2 (equilibrated in
with either an aerobic (70% N2, 5% CO2, 25% O2) or anaerobic (95% N2, 5% an anaerobic chamber for anaerobic cultures) for 15 min at 37 °C (31) and
CO2) gas mix (44). then immediately diluted 1,000-fold into fresh MOPS glucose minimal media
containing 1.0 μM FeSO4. Treated and nontreated cultures were plated on
Chelator-Assisted EPR Spectroscopy. A MG1655 culture volume of 450 mL TYE agar plates to determine the number of viable cells. Anaerobic condi-
(OD600 of ∼0.4) was pelleted, resuspended in 10 mL of fresh media con- tions were maintained by use of a Coy anaerobic chamber and GasPak
taining either 20 mM membrane-permeable iron chelator desferoxamine anaerobic jar.
mesylate (DFO; Sigma Aldrich) and 10 mM membrane-impermeable iron
chelator diethylenetriaminepentaacetic acid (DTPA; Sigma Aldrich) or just
ACKNOWLEDGMENTS. We thank Martin Shafer (Wisconsin State Laboratory
10 mM DTPA (to remove contaminating iron) as described previously (25). of Hygiene) for assistance with whole-cell element analyses and Monika
After a 15-min incubation at 37 °C, cells were washed twice with 5 mL of Ivancic and Bob Shanks (Department of Chemistry EPR facilities, University of
20 mM Tris·HCl (pH 7.4) and resuspended in 350 μL of 20 mM Tris·HCl (pH 7.4) Wisconsin) for their assistance with EPR instrumentation. We thank Jim Imlay
and 30% glycerol. A portion of these cells (200 μL) was transferred to EPR (University of Illinois) and Thomas Brunold (University of Wisconsin) for
tubes (4-mm Thin Wall Quartz EPR sample tube with a length of 250 mm; helpful discussions on labile iron measurements. We also thank Dan Park for
Wilmad Lab Glass) and frozen at −80 °C. The remaining cells were diluted to the gift of Fur protein, and Kevin Myers for data analysis. This work was
measure the final OD600 (Lambda 25 UV/Vis Spectrophotometer; PerkinElmer). funded by Grants GM045844 and GM115894 from the NIH (to P.J.K.). N.A.B.
All manipulations were completed in a Coy anaerobic chamber. Four to five was supported by University of Wisconsin–Madison NIH Chemistry Biology
replicates of each sample were analyzed. Interface Training Grant T32GM008505.

1. Braun V, Hantke K (2011) Recent insights into iron import by bacteria. Curr Opin 27. Varghese S, Wu A, Park S, Imlay KR, Imlay JA (2007) Submicromolar hydrogen per-
Chem Biol 15:328–334. oxide disrupts the ability of Fur protein to control free-iron levels in Escherichia coli.
2. Frawley ER, et al. (2013) Iron and citrate export by a major facilitator superfamily Mol Microbiol 64:822–830.
pump regulates metabolism and stress resistance in Salmonella typhimurium. Proc 28. Walling C (1975) Fenton’s reagent revisited. Acc Chem Res 8:125–131.
Natl Acad Sci USA 110:12054–12059. 29. Imlay JA, Chin SM, Linn S (1988) Toxic DNA damage by hydrogen peroxide through
3. Andrews S, et al. (2013) Control of iron metabolism in bacteria. Met Ions Life Sci 12: the Fenton reaction in vivo and in vitro. Science 240:640–642.
203–239. 30. Imlay JA (2013) The molecular mechanisms and physiological consequences of oxi-
4. Fleischhacker AS, Kiley PJ (2011) Iron-containing transcription factors and their roles dative stress: Lessons from a model bacterium. Nat Rev Microbiol 11:443–454.
as sensors. Curr Opin Chem Biol 15:335–341. 31. Imlay JA, Linn S (1986) Bimodal pattern of killing of DNA-repair-defective or anoxi-
5. McHugh JP, et al. (2003) Global iron-dependent gene regulation in Escherichia coli. A cally grown Escherichia coli by hydrogen peroxide. J Bacteriol 166:519–527.
new mechanism for iron homeostasis. J Biol Chem 278:29478–29486. 32. Keyer K, Gort AS, Imlay JA (1995) Superoxide and the production of oxidative DNA
6. Beauchene NA, et al. (2015) Impact of anaerobiosis on expression of the iron- damage. J Bacteriol 177:6782–6790.
33. Lodge JS, Emery T (1984) Anaerobic iron uptake by Escherichia coli. J Bacteriol 160:
responsive Fur and RyhB regulons. MBio 6:e01947–e15.
801–804.
7. Lee JW, Helmann JD (2007) Functional specialization within the Fur family of met-
34. Kammler M, Schön C, Hantke K (1993) Characterization of the ferrous iron uptake
alloregulators. Biometals 20:485–499.
system of Escherichia coli. J Bacteriol 175:6212–6219.
8. Fillat MF (2014) The FUR (ferric uptake regulator) superfamily: Diversity and versatility
35. Bradley JM, Moore GR, Le Brun NE (2014) Mechanisms of iron mineralization in fer-
of key transcriptional regulators. Arch Biochem Biophys 546:41–52.
ritins: One size does not fit all. J Biol Inorg Chem 19:775–785.
9. Baichoo N, Helmann JD (2002) Recognition of DNA by Fur: A reinterpretation of the
36. von Wulffen J, Sawodny O, Feuer R; RecogNice-Team (2016) Transition of an anaer-
Fur box consensus sequence. J Bacteriol 184:5826–5832.
obic Escherichia coli culture to aerobiosis: Balancing mRNA and protein levels in a
10. Lavrrar JL, McIntosh MA (2003) Architecture of a fur binding site: A comparative
demand-directed dynamic flux balance analysis. PLoS One 11:e0158711.
analysis. J Bacteriol 185:2194–2202. 37. Waldron KJ, Robinson NJ (2009) How do bacterial cells ensure that metalloproteins
11. Bagg A, Neilands JB (1987) Ferric uptake regulation protein acts as a repressor, em- get the correct metal? Nat Rev Microbiol 7:25–35.
ploying iron (II) as a cofactor to bind the operator of an iron transport operon in 38. Waldron KJ, Rutherford JC, Ford D, Robinson NJ (2009) Metalloproteins and metal
Escherichia coli. Biochemistry 26:5471–5477. sensing. Nature 460:823–830.
12. Hamed MY (1993) Binding of the ferric uptake regulation repressor protein (Fur) to 39. Reyes-Caballero H, Campanello GC, Giedroc DP (2011) Metalloregulatory proteins:
Mn(II), Fe(II), Co(II), and Cu(II) ions as co-repressors: Electronic absorption, equilibrium, Metal selectivity and allosteric switching. Biophys Chem 156:103–114.
and 57Fe Mössbauer studies. J Inorg Biochem 50:193–210. 40. Gilston BA, et al. (2014) Structural and mechanistic basis of zinc regulation across the
13. Mills SA, Marletta MA (2005) Metal binding characteristics and role of iron oxidation E. coli Zur regulon. PLoS Biol 12:e1001987.
in the ferric uptake regulator from Escherichia coli. Biochemistry 44:13553–13559. 41. Ma Z, Gabriel SE, Helmann JD (2011) Sequential binding and sensing of Zn(II) by
14. Keyer K, Imlay JA (1996) Superoxide accelerates DNA damage by elevating free-iron Bacillus subtilis Zur. Nucleic Acids Res 39:9130–9138.
levels. Proc Natl Acad Sci USA 93:13635–13640. 42. Shin JH, et al. (2011) Graded expression of zinc-responsive genes through two reg-
15. Jacques JF, et al. (2006) RyhB small RNA modulates the free intracellular iron pool and ulatory zinc-binding sites in Zur. Proc Natl Acad Sci USA 108:5045–5050.
is essential for normal growth during iron limitation in Escherichia coli. Mol Microbiol 43. Blattner FR, et al. (1997) The complete genome sequence of Escherichia coli K-12.
62:1181–1190. Science 277:1453–1462.
16. Kosman DJ (2013) Iron metabolism in aerobes: Managing ferric iron hydrolysis and 44. Sutton VR, Kiley PJ (2003) Techniques for studying the oxygen-sensitive transcription
ferrous iron autoxidation. Coord Chem Rev 257:210–217. factor FNR from Escherichia coli. Methods Enzymol 370:300–312.
17. Lau CK, Krewulak KD, Vogel HJ (2016) Bacterial ferrous iron transport: The Feo sys- 45. Ku HH (1966) Notes on the use of propagation of error formulas. J Res Natl Bur Stand
tem. FEMS Microbiol Rev 40:273–298. Sect C 70C:263–273.
18. Liang K, Keles S (2012) Detecting differential binding of transcription factors with 46. Kang Y, Weber KD, Qiu Y, Kiley PJ, Blattner FR (2005) Genome-wide expression
ChIP-seq. Bioinformatics 28:121–122. analysis indicates that FNR of Escherichia coli K-12 regulates a large number of genes
19. Seo SW, et al. (2014) Deciphering Fur transcriptional regulatory network highlights its of unknown function. J Bacteriol 187:1135–1160.
47. Baba T, et al. (2006) Construction of Escherichia coli K-12 in-frame, single-gene
complex role beyond iron metabolism in Escherichia coli. Nat Commun 5:4910.
knockout mutants: The Keio collection. Mol Syst Biol 2:2006.0008.
20. Brickman TJ, Ozenberger BA, McIntosh MA (1990) Regulation of divergent tran-
48. Datsenko KA, Wanner BL (2000) One-step inactivation of chromosomal genes in
scription from the iron-responsive fepB-entC promoter-operator regions in Escherichia
Escherichia coli K-12 using PCR products. Proc Natl Acad Sci USA 97:6640–6645.
coli. J Mol Biol 212:669–682.
49. Sutton VR, Mettert EL, Beinert H, Kiley PJ (2004) Kinetic analysis of the oxidative
21. Chen Z, et al. (2007) Discovery of Fur binding site clusters in Escherichia coli by in-
conversion of the [4Fe-4S]2+ cluster of FNR to a [2Fe-2S]2+ cluster. J Bacteriol 186:
formation theory models. Nucleic Acids Res 35:6762–6777.
8018–8025.
22. Myers KS, et al. (2013) Genome-scale analysis of Escherichia coli FNR reveals complex
50. Imlay JA, Fridovich I (1991) Assay of metabolic superoxide production in Escherichia
features of transcription factor binding. PLoS Genet 9:e1003565. coli. J Biol Chem 266:6957–6965.
23. Park DM, Akhtar MS, Ansari AZ, Landick R, Kiley PJ (2013) The bacterial response 51. Schwartz CJ, et al. (2001) IscR, an Fe-S cluster-containing transcription factor, re-
regulator ArcA uses a diverse binding site architecture to regulate carbon oxidation presses expression of Escherichia coli genes encoding Fe-S cluster assembly proteins.
globally. PLoS Genet 9:e1003839. Proc Natl Acad Sci USA 98:14895–14900.
24. Carpenter C, Payne SM (2014) Regulation of iron transport systems in Enter- 52. Mettert EL, Outten FW, Wanta B, Kiley PJ (2008) The impact of O(2) on the Fe-S cluster
obacteriaceae in response to oxygen and iron availability. J Inorg Biochem 133: biogenesis requirements of Escherichia coli FNR. J Mol Biol 384:798–811.
110–117. 53. Giel JL, et al. (2013) Regulation of iron-sulphur cluster homeostasis through tran-
25. Woodmansee AN, Imlay JA (2002) Quantitation of intracellular free iron by electron scriptional control of the Isc pathway by [2Fe-2S]-IscR in Escherichia coli. Mol
paramagnetic resonance spectroscopy. Methods Enzymol 349:3–9. Microbiol 87:478–492.
26. Hider RC, Kong X (2013) Iron speciation in the cytosol: An overview. Dalton Trans 42: 54. Keseler IM, et al. (2011) EcoCyc: A comprehensive database of Escherichia coli biology.
3220–3229. Nucleic Acids Res 39:D583–D590.

12266 | www.pnas.org/cgi/doi/10.1073/pnas.1707189114 Beauchene et al.

You might also like