ArticlePDF Available

Abstract and Figures

Zinc is a bivalent cation essential for bacterial growth and metabolism. The human pathogen Neisseria meningitidis expresses a homologue of the Zinc uptake regulator Zur, which has been postulated to repress the putative zinc uptake protein ZnuD. In this study, we elucidated the transcriptome of meningococci in response to zinc by microarrays and quantitative real-time PCR (qRT-PCR). We identified 15 genes that were repressed and two genes that were activated upon zinc addition. All transcription units (genes and operons) harbored a putative Zur binding motif in their promoter regions. A meningococcal Zur binding consensus motif (Zur box) was deduced in silico, which harbors a conserved central palindrome consisting of hexameric inverted repeats separated by three nucleotides (TGTTATDNHATAACA). In vitro binding of recombinant meningococcal Zur to this Zur box was shown for the first time using electrophoretic mobility shift assays. Zur binding to DNA depended specifically on the presence of zinc and was sensitive to mutations in the palindromic sequence. The Zur regulon among genes of unknown function comprised genes involved in zinc uptake, tRNA modification, and ribosomal assembly. In summary, this is the first study of the transcriptional response to zinc in meningococci.
Zur binding to dsDNA fragments in EMSA. Band shifts are indicated with an arrow, unbound dsDNA fragments with an asterisk. (A) Binding of Zur to the dsDNA fragment comprising the nmb0964 Zur box. Addition of different amounts of Zur from 3.75 pmol to 60 pmol led to a band shift of the 34-mer dsDNA fragment. The unlabeled nmb0964 dsDNA fragment in 125-fold excess acted as a specific competitor that abrogated the band shift. (B) Effect of divalent ions and chelators on Zur binding to the nmb0964 dsDNA fragment. Different divalent ions (100 μM) were added to the reactions. Only Zn²⁺ mediated Zur binding to the nmb0964 dsDNA fragment. Addition of the chelator (312.5 μM) EDTA or TPEN reverted the band shift. (C) Mutational analysis of the Zur box. Mutation of conserved nucleotides in the 3′-inverse repeat of the palindrome of the nmb0964 Zur box inhibited the band shift. The four highly conserved nucleotides are displayed in bold. Mutated nucleotides are marked in gray. (D) Binding of Zur to dsDNA fragments comprising the Zur boxes of nmb0546 and nmb1475. Zur led to a band shift of the 34-mer nmb0546 dsDNA fragment and the 51-mer nmb1475 dsDNA fragment covering the respective Zur boxes. Addition of 1,000-fold excess of the respective unlabeled dsDNA fragment led to competition of the shift. (E) Competition studies with the nmb0964 dsDNA fragment. Incubation with a 125-fold excess of the unlabeled nmb0964 and nmb1475 dsDNA fragments abrogated Zur binding; 125-fold excess of unlabeled mutated nmb0964 dsDNA fragments comprising one to four mismatches within the 3′-inverse repeat of the palindrome (panel C) was unable to fully compete the shift.
… 
This content is subject to copyright. Terms and conditions apply.
The Zinc-Responsive Regulon of Neisseria meningitidis Comprises 17
Genes under Control of a Zur Element
Marie-Christin Pawlik, Kerstin Hubert, Biju Joseph, Heike Claus, Christoph Schoen, and Ulrich Vogel
University of Würzburg, Institute for Hygiene and Microbiology, Würzburg, Germany
Zinc is a bivalent cation essential for bacterial growth and metabolism. The human pathogen Neisseria meningitidis expresses a
homologue of the Zinc uptake regulator Zur, which has been postulated to repress the putative zinc uptake protein ZnuD. In this
study, we elucidated the transcriptome of meningococci in response to zinc by microarrays and quantitative real-time PCR
(qRT-PCR). We identified 15 genes that were repressed and two genes that were activated upon zinc addition. All transcription
units (genes and operons) harbored a putative Zur binding motif in their promoter regions. A meningococcal Zur binding con-
sensus motif (Zur box) was deduced in silico, which harbors a conserved central palindrome consisting of hexameric inverted
repeats separated by three nucleotides (TGTTATDNHATAACA). In vitro binding of recombinant meningococcal Zur to this Zur
box was shown for the first time using electrophoretic mobility shift assays. Zur binding to DNA depended specifically on the
presence of zinc and was sensitive to mutations in the palindromic sequence. The Zur regulon among genes of unknown func-
tion comprised genes involved in zinc uptake, tRNA modification, and ribosomal assembly. In summary, this is the first study of
the transcriptional response to zinc in meningococci.
Trace metals, such as zinc, are essential cofactors of many en-
zymes and DNA-binding proteins (40). On the one hand, bac-
teria have to cope with decreased zinc availability during infection
in the host (25), while on the other, at high concentrations, zinc
can lead to the production of toxic reactive oxygen species (49).
Hence, bacteria tightly control metal homeostasis by metalloregu-
latory proteins, which are specialized metal-sensing transcrip-
tional regulators that change their conformation upon binding of
the metal ions (5). Zinc uptake regulators (Zur) belong to the Fur
protein family of transcriptional regulators that also includes Fur,
Mur, and Nur, which are sensors of iron, manganese, and nickel,
respectively (28). Zinc uptake systems and their regulation by Zur
have been characterized for several bacteria such as Escherichia coli
(45), Bacillus subtilis (13), Listeria monocytogenes (7), and Staph-
ylococcus aureus (31). Furthermore, Zur regulons have been de-
scribed for B. subtilis (14), Mycobacterium tuberculosis (32), Yer-
sinia pestis (29), and Corynebacterium glutamicum (54). Panina et
al. identified the promoter binding motifs for Zur for Gammapro-
teobacteria, the Agrobacteria group, the Rhodococcus group, and
Gram-positive bacteria as well as for the streptococcal zinc repres-
sor AdcR using comparative computational analysis (41).
In this study, gene regulation by zinc exposure was analyzed in
the commensal Gram-negative human pathogen Neisseria menin-
gitidis, causing septicemia and meningitis (50). Several genomes
of N. meningitidis have been sequenced (4,43,60), which allowed
for the bioinformatic prediction of a Zur binding motif for Beta-
proteobacteria in the RegPrecise database (39). Until now, there is
experimental evidence for regulation by meningococcal Zur only
for the TonB-dependent outer membrane receptor, ZnuD, which
is involved in zinc acquisition upon zinc limitation (59). Stork et
al. showed that expression of znuC and znuD is enhanced in a zur
knockout mutant and predicted Zur binding motifs upstream of
both genes based on the E. coli Zur binding sequence (59). Kumar
et al. subsequently demonstrated that N. meningitidis znuD is also
regulated by Fur and allows for heme capture on the cell surface
when expressed in E. coli (26). Yet, even for znuD, binding of Zur
to the promoter of Zur-regulated genes has not been demon-
strated by in vitro experiments.
Here, we characterized the Zur regulon of N. meningitidis. The
transcriptomes of strain MC58 grown under high- and low-zinc
conditions were compared. A total of 15 Zur-repressed and two
Zur-activated genes were found. We established a meningococcal
Zur binding motif (Zur box) based on promoter sequences of
these genes. The direct binding of Zur to proposed Zur boxes was
verified by electrophoretic mobility shift assays (EMSAs) and
shown to be zinc dependent. Our results provide the basis for
further studies characterizing the molecular mechanisms of zinc
adaptation in meningococci.
MATERIALS AND METHODS
Strains and mutants. Serogroup B N. meningitidis strain MC58 (36) was
kindly provided by Richard Moxon (Oxford). The zur gene (nmb1266)of
MC58 was replaced with a kanamycin resistance cassette as follows: a
470-bp DNA fragment upstream of zur was amplified by PCR from strain
MC58 with primers MP16 and MP17, introducing restriction sites of XbaI
and EcoRI, respectively. Likewise, primers MP18 and MP19, introducing
restriction sites of EcoRI and XhoI, respectively, were used to amplify 475
bp downstream of zur. Primer sequences are provided in Table 1. The
upstream and downstream fragments were digested with XbaI/EcoRI and
EcoRI/XhoI, respectively. A kanamycin resistance cassette was obtained
from vector pUC4K (GE Healthcare) by EcoRI digestion. The flanking
upstream and downstream fragments and the kanamycin cassette were
cloned into the XbaI/XhoI-digested expression vector pBluescript II SK
(Stratagene). The resulting plasmid pMP5 was transformed into MC58.
Kanamycin-resistant clones were screened by PCR for replacement of zur
by the kanamycin cassette, and zur deletion was confirmed by Southern
Received 20 June 2012 Accepted 20 September 2012
Published ahead of print 5 October 2012
Address correspondence to Ulrich Vogel, uvogel@hygiene.uni-wuerzburg.de.
Copyright © 2012, American Society for Microbiology. All Rights Reserved.
doi:10.1128/JB.01091-12
6594 jb.asm.org Journal of Bacteriology p. 6594 6603 December 2012 Volume 194 Number 23
blotting and sequencing. The resulting zur knockout strain was designated
WUE4812.
Bacterial growth and RNA isolation. Strains MC58 and MC58 zur
were grown overnight at 37°C and 5% CO
2
on GC agar plates (BD).
Colonies of each overnight culture were resuspended in 20 ml of RPMI
medium (Invitrogen) to an optical density at 600 nm (OD
600
) of 0.2 (2
10
8
CFU/ml). The medium was supplemented with 100 M FeCl
3
to
improve bacterial growth. The cultures were incubated at 37°C with shak-
ing at 200 rpm and grown to mid-log phase (OD
600
, 1.0). Both liquid
cultures were then diluted to an OD
600
of 0.2 with RPMI medium plus 100
M FeCl
3
. The MC58 culture was consecutively divided into two parts of
20 ml each. To one culture vial, 0.5 M ZnSO
4
was added to establish a
high-zinc condition according to Stork et al. (59). The other culture and
the MC58 zur culture were further grown without zinc supplementa-
tion. All three 20-ml cultures were grown at 37°C and 200 rpm for 2 h,
resulting in an OD
600
of about 1.0. Bacterial cultures were then mixed with
twice the volume of RNAprotect Bacteria Reagent (Qiagen) to minimize
RNA degradation. Total RNA was isolated using the RNeasy Mini Kit
(Qiagen). Remaining DNA was digested with RNase-free recombinant
DNase I (Invitrogen), and RNA samples were stored at 80°C. RNA
integrity was verified with the Agilent 2100 Bioanalyzer (Agilent Technol-
ogies).
TABLE 1 Oligonucleotides used in this study for PCR, qRT-PCR, and EMSA
a
Assay and target gene Forward Sequence (5=–3=) Reverse Sequence (5=–3=)
Amplicon
length
(bp)
PCR
zur up (nmb1267) MP16 GCGCGCTCTAGAGATGGCGG
AATACATTTTG
MP17 GCGCGCGAATTCTTCTCTGTT
TATGCCGTCTG
470
zur down
(nmb1264-1265)
MP18 GCGCGCGAATTCATAAGCC
TTTCGAAAGGAGC
MP19 GCGCGCCTCGAGCGTT
TTCACCGATAAGGAAC
475
zur (nmb1266) MP183 GCGCGCTCATGAAAACAAATTT
CAAACAGAAAATTA
MP184 GCGCGCAGATCTCTGCTGACAT
TTTTTACAAATC
469
qRT-PCR
nmb0317 MP141 GCAAATCCCTGAAACTCTACCTCTT MP142 TGATGTTGACGCAGTCTTCATG 72
nmb0525 MP143 CGGCGTTCCGAATACCTTT MP144 GCGTAAATCGCGGCATAGAG 66
nmb0546 MP165 CTGCCGCCGGAATCG MP166 GACCAAAGAGCCGACCACTTC 75
nmb0577 MP179 CGGGCGCGGATAACG MP180 AGGCAGACGCTCCGCATA 61
nmb0586 MP145 GCAACTACCAAATGCAGCTCAA MP146 CAGCAGGGACGGCATTAAAT 69
nmb0587 MP147 GGCTTTGGCACGTCCTCTT MP148 GTGTGCCGAGGGCTTGAA 72
nmb0588 MP149 TCGCCCGCGAAAAAATT MP150 CTTGGGCGAGGTAGGGTTCT 66
nmb0817 MP151 AGGATGTCCTGAATGTACTCGAAAT MP152 CTGTCTCTGTGTCCTCCAACCTAA 80
nmb0819 MP153 CCCTCCTCATCCTCGACACA MP154 CAGCACCTGCGGCAGTAA 69
nmb0820 MP155 CCGCCAAAGCCCTAAACA MP156 GGTTGTCGGGATTTGAACGT 72
nmb0942 MP157 CCGCTGTTTTCGCTGGATA MP158 GTGTTGACGTTGCGCTGTTT 71
nmb0964 MP117 CCGTTCCCCGGTTTTGA MP118 CTGCATCGCCTGCTTTTTC 80
nmb0990 MP171 TGCCATTATCGCGCTTGTC MP172 GCCTGCTGCTTCGCAAAT 81
nmb1475 MP159 TCGGACAAAACTTGGAAATCG MP160 TTTGAAGCGTTCGCCTACGT 68
nmb1497 MP161 TGCCCAACATCCAAGAAATGT MP162 CTGGTTTTAAGGCGGTGTGAA 68
nmb2142 MP163 ACGACGGCGGTCATCTTTAC MP164 CGCCGTATGATGCACCATT 68
EMSA
nmb0546 MP191 CTTTCCAAGATGTTATAATATAACA
TATAATCTAT
MP192 AAATATAGATTATATGTTATAT
TATAACATCTTGG
35
nmb0964
Unaltered MP185 TAAAAAATGTAATGTTATATAATAA
CAAACTTTT
MP186 TTTCAAAAGTTTGTTATTATAT
AACATTACATTT
34
TA/CA MP193 TAAAAAATGTAATGTTATATAACGA
TGAACTTTT
MP194 TTTCAAAAGTTCATCGTTATAT
AACATTACATTT
34
CA MP195 TAAAAAATGTAATGTTATATAATAA
TGAACTTTT
MP196 TTTCAAAAGTTCATTATTATAT
AACATTACATTT
34
TA MP197 TAAAAAATGTAATGTTATATAACGA
CAAACTTTT
MP198 TTTCAAAAGTTTGTCGTTATAT
AACATTACATTT
34
C MP199 TAAAAAATGTAATGTTATATAATAA
TAAACTTTT
MP200 TTTCAAAAGTTTATTATTATAT
AACATTACATTT
34
nmb1475
Short MP187 CGATACCTATTTTGTTATAACATAAC
AAAATCTT
MP188 GTTAAAGATTTTGTTATGTTAT
AACAAAATAGGT
34
Long MP189 TCTTCACACGATACCTATTTTGTTATA
ACATAACAAAATCTTTAACCCACA
MP190 CTCGTGTGGGTTAAAGATTTTG
TTATGTTATAACAAAATAGG
TATCGTGTG
51
a
Underlined bases indicate restriction sites.
Zinc-Responsive Regulon of Meningococci
December 2012 Volume 194 Number 23 jb.asm.org 6595
According to Invitrogen’s data sheet, the RPMI medium used here
does not contain any source of zinc. However, we assumed that this RPMI
contains a minimal concentration of zinc because RPMI distributed by
Sigma was shown to comprise at least 1.69 M zinc (59). We therefore
refer to “low” (RPMI) and “high” (RPMIZnSO
4
) zinc conditions in this
study.
Furthermore, we decided against depletion of zinc, e.g., by an ion
chelator such as N,N,N=,N=-tetrakis-(2-pyridylmethyl)-ethylenediamine
(TPEN), prior to zinc supplementation as conducted in studies of iron
response (18,56) because the probable chelation of ions other than zinc
could have an influence on the general transcriptional response.
cDNA microarray analysis. Microarray analysis was performed using
whole-genome DNA microarrays based on 70-mer oligonucleotides cov-
ering four meningococcal genomes (MC58, Z2491, FAM18, and 14) as
described previously (55). In order to characterize the Zur regulon, we
used a common reference design that allows for future extension of the
data set (65). As outlined in Fig. 1, cDNA obtained from strain MC58
grown with or without zinc supplementation (low- and high-zinc condi-
tions) was hybridized separately along with the common reference, i.e.,
cDNA obtained from strain MC58 zur. MC58 zur was used as the
common reference under the assumption that Zur mostly represses gene
expression. In the zur knockout strain, mRNA of most of the Zur-regu-
lated genes was therefore expected to be present, which is a prerequisite of
valid comparisons. The transcriptome of MC58 grown without zinc sup-
plementation was consecutively indirectly compared to that of MC58
grown in the presence of high zinc.
Ten micrograms of RNA obtained from each culture was transcribed
into cDNA and labeled with either Cy3-dCTP or Cy5-dCTP (GE Health-
care) using SuperScript II (Invitrogen) and random nonamer primer
(Sigma-Aldrich). Remaining RNA was digested with DNase-free RNase
(Roche). The labeled cDNA was purified using illustra AutoSeq G50 col-
umns (GE Healthcare). For each slide, two differentially labeled cDNA
probes were combined, and probes were then hybridized to a prehybrid-
ized microarray slide. Hybridization was carried out in a Tecan
HS4800TM Pro hybridization station. Three biological replicates (inde-
pendent bacterial cultures and RNA isolations) with at least two technical
replicates (independent slides including dye swap) were used for microar-
ray hybridization. The hybridized slides were scanned with the Genepix
4200 scanner, and the raw data were acquired using Genepix Pro 4.0. The
raw gpr files from the microarray slides were processed with the Limma
package (58) implemented in R (47) to identify significantly differentially
regulated genes. Only genes with a false-discovery rate (FDR) of 0.01
and a B statistic value of 0 were considered for further analyses.
qRT-PCR. Quantitative real-time PCR (qRT-PCR) analysis of the
RNA samples described above was performed as described previously
using the StepOnePlus system and the Power SYBR green Master Mix
(53). nmb1567, encoding a putative membrane-associated peptidyl-prolyl
isomerase, which was not differentially regulated in our microarray, was
applied as endogenous control for data normalization. The oligonucleo-
tides used are given in Table 1. The threshold cycle (⌬⌬C
T
) and relative
quantity (RQ) values of qRT-PCR were obtained using the StepOne soft-
ware v2.0. RQ values represent the fold change of expression of the inves-
tigated gene in MC58 grown under low zinc compared to that in MC58
grown under high-zinc conditions. Microarray results were accepted as
differentially regulated only if comparison of the same RNA by qRT-PCR
analysis yielded RQ values of at least 1.5.
Computational analysis of Zur binding sites. Sequences of genes and
upstream regions of the identified candidate genes of strain MC58 were
retrieved from the NeMeSys database (51). Similarity searches were per-
formed using the BLAST program of the National Center for Biotech-
nology Information (NCBI) (http://blast.ncbi.nlm.nih.gov/Blast.cgi).
Operons were predicted using RNAseq data (B. Joseph and C. Schoen,
unpublished data) and DOOR database (35). Annotation data were re-
trieved from the NeMeSys (51) and Uniprot (62) databases. Upstream
sequences were screened for the presence of a Zur box by alignments using
ClustalW in BioEdit 7.0.9 (20) based on the in silico predicted motif for
Betaproteobacteria as implemented in the RegPrecise database (39) and
the hypothetical motif for znuD (nmb0964), which was recently proposed
(59). Furthermore, upstream regions were searched for conserved motifs
with the GIBBS Motif Sampler (61). The meningococcal Zur binding
motif was visualized using WebLogo 3.2 (6).
Expression and purification of recombinant Zur. The Zur protein
was recombinantly expressed for electrophoretic mobility shift assays. For
gene cloning and protein expression, the QIAexpressionist kit (Qiagen)
was used according to the manufacturer’s instructions. Briefly, the entire
coding region of the zur gene without the start and stop codons (469 bp)
was amplified from MC58 using primers MP183 and MP184 harboring a
BspHI and a BglII restriction site, respectively (Table 1). The PCR product
was cloned between the NcoI and BglII sites of pQE-60 (Qiagen) contain-
ing a C-terminal His tag. E. coli M15(pREP4) was transformed with the
recombinant plasmid, and clones were verified by DNA sequencing. Ex-
pression and purification of Zur were performed under native conditions.
Protein expression was induced by 1 mM IPTG (isopropyl--D-thiogalac-
topyranoside) for 4.5 h at 37°C, and cells were harvested, lysed, and bro-
ken by ultrasonic treatment.
For purification, cell extracts were loaded on a HIS GraviTrap nickel
Sepharose column (GE Healthcare) equilibrated with lysis buffer (50 mM
NaH
2
PO
4
, 0.3 M NaCl, 10 mM imidazole [pH 8.0]). The column was
washed twice with wash buffer (50 mM NaH
2
PO
4
, 0.3 M NaCl, 20 mM
imidazole [pH 8.0]) before elution of the His-tagged Zur protein with
elution buffer (50 mM NaH
2
PO
4
, 0.3 M NaCl, 250 mM imidazole [pH
8.0]). The Zur eluate from the nickel Sepharose column was dialyzed
against Tris-borate (TB) buffer (44.5 mM Tris, 44.5 mM boric acid [pH
8.0]) to remove imidazole and nickel. The purity of Zur was verified by
SDS-PAGE with Coomassie blue staining, and its concentration was de-
termined using the Pierce BCA protein assay kit (ThermoScientific). Pro-
tein sample aliquots were stored at 20°C.
Electrophoretic mobility shift assays to visualize DNA-Zur binding.
EMSAs were performed using the digoxigenin (DIG) Gel Shift kit (Roche)
to determine the ability of Zur to interact with short double-stranded
DNA (dsDNA) fragments containing the Zur box. Oligonucleotides were
purchased from Sigma-Aldrich (Table 1) and resuspended in TEN buffer
(10 mM Tris, 1 mM EDTA, 0.1 M NaCl [pH 8.0]). Complementary oli-
gonucleotides in equal concentrations were annealed to yield dsDNA
fragments by heating the oligonucleotide mixture for 10 min at 95°C and
subsequently slowly cooling it down to room temperature. dsDNA frag-
ments were labeled with digoxigenin (DIG Gel Shift kit; Roche). Binding
reactions were performed in a final volume of 10 l with 0.4 ng (15.5 fmol)
FIG 1 Common reference design of microarray comparison used in this
study. cDNAs obtained from MC58 grown without addition of zinc (low zinc)
and from MC58 grown with addition of zinc for 2 h (high zinc) were hybrid-
ized along with cDNA of the MC58 zur knockout strain grown under low-zinc
conditions as a common reference. MC58 zur was used as a common refer-
ence under the assumption that total mRNA of this mutant will contain
mRNA of Zur-repressed genes. The zinc-dependent promoter repression by
Zur is visualized by cartoons using the example of znuD.
Pawlik et al.
6596 jb.asm.org Journal of Bacteriology
of the labeled dsDNA fragment and 15 pmol (250 ng) of purified Zur
protein (if not stated differently) in the binding buffer [20 mM Tris-HCl
(pH 8.0), 50 mM KCl, 1 mM dithiothreitol (DTT), 5% glycerol, 0.1 g/l
poly(dI-dC), 0.01 g/l poly-L-lysine] modified after the method of Shin
et al. (57). If not stated otherwise, the binding buffer was supplemented
with 100 M ZnSO
4
according to the concentration used by Gaballa and
Helmann for B. subtilis (13). We anticipated that zinc is required for
Zur-DNA interaction as already shown for other bacteria (14,29,32,54,
57) and therefore omitted EDTA from the binding buffer as well as from
the running buffer to avoid chelation of zinc. Reaction mixtures were
incubated at room temperature for 20 min. A native 8% polyacrylamide
gel was prerun at 120 V for 30 min, and samples were then applied to the
gel and separated at 120 V for 110 min at 4°C. After electroblotting onto a
nylon membrane in TB buffer, chemiluminescent detection of the bound
dsDNA fragments was carried out.
CaCl
2
, CoCl
2
, CuSO
4
, FeSO
4
, MgCl
2
, MnSO
4
, and NiSO
4
were pur-
chased from Merck or Sigma-Aldrich and added to the binding buffer at a
final concentration of 100 M. The ion chelators EDTA and TPEN were
used at a final concentration of 312.5 M. For binding competition ex-
periments, 125- to 1,000-fold excess of the unlabeled dsDNA probe was
added to the reactions.
Statistics. The Wilcoxon rank sum test with continuity correction was
applied to analyze the impact of mismatches in the Zur box palindromic
sequence on gene expression.
Microarray data accession number. The microarray data discussed in
this publication have been deposited in NCBI’s Gene Expression Omni-
bus database (12) under accession no. GSE38033.
RESULTS
The aim of this study was to provide a data set of genes regulated
by exposure to zinc. Microarray comparison between MC58
grown with and without zinc supplementation (high- and low-
zinc condition, respectively) was employed using a common ref-
erence design (65) that is schematically represented in Fig. 1.
Growth conditions. Growth of MC58 zur in the chemically
defined medium RPMI supplemented with FeCl
3
was indistin-
guishable from that of the parental wild-type strain (data not
shown). This indicates that deactivation of Zur does not alter
growth in RPMI. Growth conditions for microarray analysis were
validated by qRT-PCR of znuD expression, which previously was
shown to be increased in a zur knockout mutant (59). Strains were
first grown in RPMI medium under low-zinc conditions, followed
by addition of 0.5 M ZnSO
4
to the medium for 1, 2, or 3 h.
After1hofzinc addition, MC58 showed 64.8-fold-reduced
expression levels of znuD by qRT-PCR compared to MC58 zur.
However, znuD expression in MC58 grown under low-zinc con-
ditions still was 27.3-fold reduced compared to MC58 zur, prob-
ably due to intracellular zinc remains of the overnight culture on
GC agar. After2hofzinc addition, compared to the zur knockout,
expression of znuD in MC58 was 48.8-fold reduced, whereas
MC58 grown under low-zinc conditions showed similar expres-
sion. After 3 h, znuD expression of MC58 grown at high zinc was
42.5-fold reduced compared to the zur knockout. Based on these
qRT-PCR results, exposure to zinc for 2 h was selected for mi-
croarray analysis.
Transcriptome analysis of zinc-dependent genes. Gene ex-
pression profiles of strain MC58 grown under low-zinc and high-
zinc conditions were compared by a common reference design
using the MC58 zur knockout strain as the common reference
(Fig. 1). Sixteen genes were upregulated, whereas three genes were
downregulated, in MC58 grown at low zinc compared to high
zinc. The genes of this data set were analyzed for operon struc-
tures. To confirm microarray analysis, we directly compared ex-
pression differences between MC58 (low zinc) and MC58 (high
zinc) by qRT-PCR for the first gene of predicted operon structures
and all single genes. All genes except two were confirmed to be
differentially expressed. Therefore, the regulon verified in this
study comprised 15 genes upregulated and only 2 downregulated
in MC58 grown at low zinc compared to high zinc (Table 2). Nine
genes were organized in four transcriptional units: nmb0317 and
nmb0316 encode a 7-cyano-7-deazaguanine reductase and an in-
tegral membrane protein, respectively; nmb0817 and nmb0818
code for hypothetical proteins that belong to the DUF723 family
and may have a role in DNA-binding (62); nmb0942 and nmb0941
encode paralogues of 50S ribosomal proteins; and nmb0588 and
nmb0587, together with one single gene, nmb0586, encode the
components of the putative ABC transporter for zinc, ZnuCBA
(59). The single genes nmb0819 and nmb0820 both encode pro-
teins that contain putative DNA-binding helix-turn-helix motifs.
Highly differential expression of the TonB-dependent outer
membrane receptor znuD (nmb0964), involved in zinc uptake
(59), represents a positive control for our analysis, as znuD was
used for optimization of our microarray conditions as indicated
above. Expression of another putative TonB-dependent receptor
of yet-unknown function, nmb1497, was also increased at low
zinc. Strongly increased at low zinc were expression of nmb0525
(queC), encoding a zinc-binding 7-cyano-7-deazaguanine syn-
thase, and expression of nmb1475, coding for a conserved hypo-
thetical periplasmic protein with similarities to the acetate kinase
AckA of Bacillus spp. Only two genes were repressed at low zinc:
nmb0546 (adhP), which codes for a zinc-containing alcohol dehy-
drogenase, and nmb0577, which shows similarities to the Haemo-
philus influenzae pfkA, encoding a 6-phosphofructokinase.
In silico prediction of promoter organization and the Zur
box. Zur binding to a palindromic sequence upstream of znuD has
been previously postulated (59). The published sequence resem-
bles the in silico-predicted Zur binding motif for Betaproteobacte-
ria in the RegPrecise database (39). We analyzed the promoter
regions of the confirmed zinc-responsive genes for a conserved
Zur box. Homologies to the postulated Zur binding motif were
found in the upstream regions of all regulated transcription units
(genes or operons). This finding suggests that the approach em-
ployed here at least partially elucidated the Zur regulon and that
zinc treatment of the bacteria did not deregulate genes other than
Zur-controlled ones. The deduced meningococcal Zur binding
motif was visualized as the consensus motif and graphically dis-
played by WebLogo 3.2 (Fig. 2A). It is a 23-bp motif with a central
palindromic part comprising hexameric inverse repeats separated
from each other by three nucleotides, which is also found in the
predicted motif for znuD published by Stork et al. (59). The align-
ment of all putative Zur motifs indicated that the strength of gene
regulation by Zur was dependent on the precision and length of
the palindrome (Fig. 2B). The extent of zinc-dependent gene reg-
ulation (as deduced from RQ values) was reduced if at least one
mismatch occurred in the palindrome of the putative Zur box
(Fig. 2C). Gene expression changes as calculated based on the
absolute ⌬⌬C
T
values obtained by qRT-PCR were significantly
higher in genes with a perfect Zur box than in genes with imperfect
Zur boxes having at least one mismatch (Wilcoxon rank sum test
with continuity correction, P0.01). Extension of the length of
the palindromic sequence may compensate for mismatches in the
central palindrome, e.g., in the case of nmb0317 or nmb0588. The
Zinc-Responsive Regulon of Meningococci
December 2012 Volume 194 Number 23 jb.asm.org 6597
longest palindrome was found for nmb1475 with a decameric in-
verse repeat that probably represents the perfect Zur box, because
this gene showed the highest repression of gene expression upon
zinc exposure (RQ 96.4). We detected two putative Zur boxes
upstream of nmb0586. However, this duplication did not affect
the strength of gene expression.
Verification of Zur binding to predicted DNA motifs by elec-
trophoretic mobility shift assay. To demonstrate that Zur binds
to in silico-predicted DNA motifs, we performed EMSAs with a
His-tagged recombinant Zur protein and synthetic dsDNA frag-
ments. A mobility shift of a 34-bp dsDNA fragment comprising
the nmb0964 Zur box was observed upon incubation with the
recombinant protein (Fig. 3A). The shift of the nmb0964 dsDNA
fragment was competed by 125-fold excess of the unlabeled
nmb0964 dsDNA fragment, which confirms binding specificity
(Fig. 3A). A band shift was observed only in the presence of 100
M ZnSO
4
, whereas CaCl
2
, CoCl
2
, CuSO
4
, FeSO
4
, MgCl
2
,
MnSO
4
, and NiSO
4
did not support binding of Zur to the dsDNA
fragment, indicating that zinc is needed specifically to mediate
Zur-DNA interaction (Fig. 3B). Addition of EDTA and TPEN,
separately, blocked the zinc effect by chelation, underlining the
importance of zinc (Fig. 3B).
Mutations of the four most conserved nucleotides in the 3=-
inverse repeat of the palindrome (ATAACA) (where the con-
served nucleotides are in boldface) of the nmb0964 dsDNA frag-
ment abrogated or reduced Zur binding (Fig. 3C). However,
interference with shifts was less clear upon mutation of marginal
nucleotides or in the case of a single nucleotide mutation. This
indicates that (i) all four conserved bases are important for Zur
binding but the inner bases (TA) of the palindrome are more
essential than the marginal ones (CA) and (ii) only one mutated
nucleotide is not sufficient to completely abrogate the shift.
We further investigated the binding of Zur to the in silico-
predicted Zur motifs of nmb0546, one of the two genes activated in
response to zinc. A clear band shift that could be inhibited by the
unlabeled nmb0546 dsDNA fragment was seen. For reasons un-
known, competition of the nmb0546 shift required higher concen-
trations of the competing DNA than nmb0964. For nmb1475, the
gene with an extended perfect palindrome in the Zur box, no shift
was observed with a 34-mer dsDNA fragment (data not shown).
Upon extending the surrounding region by 8 and 9 bp on each side
of the extended perfect palindrome, a clear shift became also vis-
ible (Fig. 3D). We can only speculate that Zur binding to the
perfect oversized palindrome may be possible only if the DNA
structure is stabilized by extended flanking DNA sequences.
To furthermore affirm binding specificity, we performed
EMSA with the nmb0964 dsDNA fragment and attempted to com-
pete the shift with unlabeled nmb0964 or nmb1475 dsDNA frag-
ments or mutated versions of the nmb0964 sequence. Indeed, the
unlabeled nmb0964 dsDNA fragment itself and the unlabeled
nmb1475 dsDNA fragment comprising the same palindrome ab-
rogated Zur binding, whereas the nmb0964 dsDNA fragment
comprising mismatches within the 3=-inverse repeat palindrome
(ATAACA) was not able to completely compete the shift (Fig. 3E).
In summary, we detected Zur boxes upstream of all genes that
were verified by qRT-PCR to be differentially expressed in re-
sponse to zinc. Our EMSA results prove that Zur binds to the
predicted binding motifs of three selected genes regulated in re-
sponse to zinc. Therefore, we infer that all transcriptional units
with Zur boxes identified here were indeed regulated by Zur. Zinc
was indispensable for interaction of Zur with the DNA. The pal-
indromic part of the motif was essential for the binding as dem-
onstrated by mutational analysis. In vitro binding of dsDNA frag-
TABLE 2 Differentially expressed genes in N. meninigitidis MC58 observed by comparison of low- to high-zinc conditions
a
Locus Gene Predicted function Predicted localization Size (bp)
Differential gene expression
Fold change
by cDNA
microarray
hybridization
RQ by
qRT-PCR
nmb0546 adhP Alcohol dehydrogenase, propanol preferring Cytoplasmic 1,047 4.1 21.7
nmb0577 NosR-related protein Unknown 351 1.8 1.8
nmb0588 znuC ABC transporter, ATP-binding protein Cytoplasmic 756 1.4 4.4
nmb0586 znuA Putative ABC transporter substrate-binding
protein
Cytoplasmic
membrane
915 1.5 6.3
nmb0820 Hypothetical protein Unknown 198 1.6 1.9
nmb1497 Putative TonB-dependent receptor Outer membrane 2,766 1.6 1.8
nmb0818 Hypothetical protein Unknown 411 1.9
nmb0817 Hypothetical protein Unknown 384 2.0 3.1
nmb0587 znuB Putative ABC transporter permease protein Cytoplasmic
membrane
876 2.1 3.3
nmb0942 rpmE2 50S ribosomal protein L31 type B Cytoplasmic 276 2.2 41.6
nmb0819 Hypothetical protein Unknown 393 2.3 1.8
nmb0525 queC 7-Cyano-7-deazaguanine synthase Cytoplasmic 660 2.5 3.9
nmb0941 rpmJ 50S ribosomal protein L36 Cytoplasmic 126 2.7
nmb0317 queF NADPH-dependent 7-cyano-7-deazaguanine
reductase
Cytoplasmic 474 3.4 5.8
nmb1475 Conserved hypothetical periplasmic protein Periplasmic 807 3.6 96.4
nmb0964 znuD TonB-dependent receptor Outer membrane 2,277 4.4 51.1
nmb0316 Conserved hypothetical integral membrane protein Cytoplasmic
membrane
687 4.6
a
Gene information retrieved from NeMeSys (51) and Uniprot (62) databases. Genes are ordered by their differential expression in microarray analysis.
Pawlik et al.
6598 jb.asm.org Journal of Bacteriology
ments harboring a Zur box was observed for genes irrespective of
their activation or repression by zinc.
DISCUSSION
In the present study, we analyzed the role of the meningococcal
zinc uptake regulator Zur in transcriptional regulation using a
comparative microarray approach. We defined the Zur box for
meningococci and demonstrated in vitro interaction of recombi-
nantly expressed Zur protein with the predicted Zur boxes up-
stream of three selected target genes.
Previous studies of B. subtilis,Y. pestis,M. tuberculosis, and C.
glutamicum applied direct microarray comparisons of a wild-type
strain with a zur knockout mutant to elucidate the Zur regulon
(14,29,32,54). The advantage of our approach of analyzing the
response to zinc was that pleiotropic effects of a constitutive zur
knockout might be avoided. Indeed, the meningococcal Zur regu-
lon deduced from transcriptome data and consecutive Zur box
analysis is remarkably small. We identified 15 genes downregu-
lated and two genes upregulated at high zinc using microarray
analysis and qRT-PCR. All transcriptional units harbored a Zur
box. The entity of 17 regulated genes is comparable to the number
found in microarray analyses of Zur regulons of other bacteria,
i.e., 18 genes and 32 genes upregulated in a zur deletion mutant of
C. glutamicum (54) and M. tuberculosis (32), respectively. Only in
a comparative microarray analysis of a Y. pestis wild-type strain
and its zur knockout mutant, both grown under zinc-rich condi-
tions, were a much higher number of Zur-regulated genes found
(154 genes). However, as only four genes had a Zur box, regula-
tion of most genes presumably has not been a direct result of Zur
binding to promoters but reflects a general alteration of gene reg-
ulation (29). Following several studies of the iron response (18,
56), we could have applied zinc depletion, achieved by addition of
an ion chelator such as TPEN, to compare with a zinc repletion
condition. However, it cannot be excluded that TPEN chelates
ions other than zinc and interferes with the membrane integrity.
Among the 17 genes regulated by zinc exposure, we detected 2
that are upregulated, i.e., nmb0546 and nmb0577. Gene activation
by Zur has been reported in only two other studies. In a microar-
ray analysis of the B. subtilis Zur regulon, two genes were upregu-
lated in the wild type compared to a zur deletion mutant. How-
ever, Zur motifs were not detected in the genes’ regulatory regions
and it was assumed that their upregulation is probably due to
indirect effects of the altered zinc homeostasis (14). This can be
excluded in this study, as we identified Zur boxes in the upstream
regions of both Zur-activated genes. Only in one other study, of
the phytopathogen Xanthomonas campestris, did Zur act as a di-
rect activator of one gene (22).
Based on the prediction of the Zur binding motif for znuD by
Stork et al. (59), we analyzed the promoter regions of all genes
found to be zinc regulated in our study. We established a binding
motif for meningococcal Zur with the ideal palindromic sequence
TGTTATDNHATAACA, which is identical to the Zur binding
motif proposed for znuD (59) and consistent with the palindrome
in the previously bioinformatically predicted motif for the related
Gammaproteobacteria.
We found an ideal palindrome only within the promoters of
the genes nmb0546,nmb0942,znuD, and nmb1475. These genes
showed the highest differences in expression when low- and high-
zinc conditions were compared. Mismatches in the palindrome
reduced gene expression alteration. This finding was confirmed by
EMSA, also in line with previous investigations of the Zur box of
C. glutamicum (54).
The zinc-activated gene nmb0546 was previously shown to be
Hfq repressed in N. meningitidis (37,42). Hfq is an RNA chaper-
one that stabilizes small regulatory RNAs (sRNAs) and mediates
their binding to their mRNA targets, which leads to subsequent
repression of mRNA translation (37). If nmb0546 and nmb0577
were targets of a yet-unknown sRNA transcribed at zinc depletion
FIG 2 Prediction of the putative meningococcal Zur box. (A) Graphical dis-
play of the Zur box for meningococci based on the consensus sequence of
predicted Zur binding sites for all zinc-responsive genes (generated with
WebLogo 3.2). (B) Nucleotide sequence alignment of putative Zur boxes up-
stream of zinc-regulated genes. Distances of the Zur motif to the start codon of
translation and expression differences determined by qRT-PCR are given for
each gene (RQ; low-zinc to high-zinc conditions). The order of motifs is based
on the RQ values. The two genes upregulated by Zur are in bold gray letters.
The strongly conserved palindrome composed of two hexameric inverted re-
peats separated by three nucleotides is indicated (#) and boxed for all motifs.
Bold nucleotides signify complementary nucleotide pairs between the two
sides of the motifs. Symbols: *, an alternative start for the gene is possible; ‡,
two different motifs were found in the same promoter region. (C) Comparison
of the absolute RQ values of gene expression with the number of mismatches in
the palindrome of each Zur box compared to the znuD (nmb0964) palindrome
sequence (TGTTATDNHATAACA). Gene expression changes were signifi-
cantly higher in genes with a perfect Zur box than in genes with at least one
mismatch in the Zur box (Wilcoxon rank sum test with continuity correction,
P0.01).
Zinc-Responsive Regulon of Meningococci
December 2012 Volume 194 Number 23 jb.asm.org 6599
and downregulated by Zur upon zinc repletion, activation of the
genes would occur indirectly due to Zur-mediated downregula-
tion of the putative sRNA, which otherwise leads to degradation of
both mRNAs. However, RNA sequencing needs to be applied to
analyze the regulation of sRNAs by zinc.
Functions of regulated genes. The previously identified me-
ningococcal genes carrying homologues of the E. coli znuCBA
operon encoding an ABC transporter for high-affinity zinc up-
take, i.e., nmb0588-nmb0587-nmb0586, and nmb0964 (znuD),
coding for a TonB-dependent receptor mediating zinc uptake at
low zinc concentration (59), were shown to be repressed at high
zinc in this study. This finding is reminiscent of E. coli (44), M.
tuberculosis (32), B. subtilis (14), C. glutamicum (54), Y. pestis (29),
and Streptomyces coelicolor (57). We demonstrated by EMSA that
meningococcal Zur binds to the in silico predicted Zur box first
proposed by Stork et al. for znuD (59).
The most strongly regulated gene in our study was nmb1475.It
codes for a conserved hypothetical periplasmic protein that shows
34% similarity to the acetate kinase AckA of Bacillus spp. AckA is
an enzyme involved in the conversion of acetate to acetyl coen-
zyme A (acetyl-CoA) (62), and its E. coli homologue was shown to
bind zinc (24). The protein also harbors a conserved domain that
is similar to the CbiK (COG5266) domain of the periplasmic com-
ponent of an ABC-type Co
2
transport system as identified by
NCBI Blast. Thus, NMB1475 may be involved in the uptake of
zinc and/or other transition metals.
It has been suggested that zinc uptake systems are important
for bacterial survival and virulence upon infection, as the access to
FIG 3 Zur binding to dsDNA fragments in EMSA. Band shifts are indicated with an arrow, unbound dsDNA fragments with an asterisk. (A) Binding of Zur to
the dsDNA fragment comprising the nmb0964 Zur box. Addition of different amounts of Zur from 3.75 pmol to 60 pmol led to a band shift of the 34-mer dsDNA
fragment. The unlabeled nmb0964 dsDNA fragment in 125-fold excess acted as a specific competitor that abrogated the band shift. (B) Effect of divalent ions and
chelators on Zur binding to the nmb0964 dsDNA fragment. Different divalent ions (100 M) were added to the reactions. Only Zn
2
mediated Zur binding to
the nmb0964 dsDNA fragment. Addition of the chelator (312.5 M) EDTA or TPEN reverted the band shift. (C) Mutational analysis of the Zur box. Mutation
of conserved nucleotides in the 3=-inverse repeat of the palindrome of the nmb0964 Zur box inhibited the band shift. The four highly conserved nucleotides are
displayed in bold. Mutated nucleotides are marked in gray. (D) Binding of Zur to dsDNA fragments comprising the Zur boxes of nmb0546 and nmb1475. Zur led
to a band shift of the 34-mer nmb0546 dsDNA fragment and the 51-mer nmb1475 dsDNA fragment covering the respective Zur boxes. Addition of 1,000-fold
excess of the respective unlabeled dsDNA fragment led to competition of the shift. (E) Competition studies with the nmb0964 dsDNA fragment. Incubation with
a 125-fold excess of the unlabeled nmb0964 and nmb1475 dsDNA fragments abrogated Zur binding; 125-fold excess of unlabeled mutated nmb0964 dsDNA
fragments comprising one to four mismatches within the 3=-inverse repeat of the palindrome (panel C) was unable to fully compete the shift.
Pawlik et al.
6600 jb.asm.org Journal of Bacteriology
zinc is limited within the human host (21,59). Of note, expression
of NMB0586 (designated MntC by van Alen et al. [63] or ZnuA by
Stork et al. [59]) and NMB1475 was upregulated in biofilms, and
deletion of nmb0586 reduced biofilm formation (63). Extenuated
biofilm formation has previously also been seen upon deletion of
znuA in gonococci (30) and nmb0587 (znuB)inE. coli (19). A
gonococcal znuA (mntC) mutant in vitro also was more sensitive
to hydrogen peroxide than the wild-type strain and showed re-
duced invasion to primary human cervical epithelial cells (64).
Furthermore, znuA contributes to the in vivo pathogenicity of
Salmonella enterica (3). Moreover, ZnuABC and ZnuD were lately
shown to contribute to bacterial adhesion to epithelial cells in E.
coli (15) and N. meningitidis (26), respectively.
Two meningococcal genes, nmb0317 and nmb0525, that were
repressed at high zinc and harbor an upstream Zur box are possi-
bly involved in queuosine biosynthesis as deduced from protein
similarity analysis. nmb0525 encodes the zinc-binding 7-cyano-7-
deazaguanine synthase QueC, and nmb0317 codes for the
7-cyano-7-deazaguanine reductase QueF (62). To date, queuosine
biosynthesis in bacteria is still not fully understood, but the path-
ways have been studied in E. coli and B. subtilis (16,27). Queuosine
is one of the most complex modified nucleosides that are incor-
porated at the wobble position of a subset of tRNAs (10). Such
modification of tRNAs was shown to improve the efficiency and
correctness of translation (9). Dineshkumar et al. demonstrated a
natural defect in queuosine biosynthesis of E. coli and noticed
reduced fitness under nutrient limitation (8). Lack of queuosine
also attenuated Shigella flexneri virulence (9,10). In meningo-
cocci, queuosine biosynthesis seems to be strongly Zur regulated,
since two enzymes of the biosynthesis are upregulated under zinc-
limiting conditions as they occur in the host (25). It will be of
interest for future studies to address the question of whether en-
hanced queuosine modification of tRNAs might be the result of
zinc depletion and contribute to increased expression of virulence
factors and thus support fitness of N. meningitidis upon infection.
Concordant with what has been observed in E. coli (17,41), Y.
pestis (29), M. tuberculosis (32), S. coelicolor (57), and B. subtilis
(2), the expression of several genes encoding ribosomal proteins
(r-proteins) was also Zur repressed in N. meningitidis. Several bac-
terial genomes code for two paralogous forms of r-proteins (34).
The first form (C) contains a metal-binding Zn
2
ribbon usu-
ally consisting of four conserved cysteines, whereas in the second
form (C) this zinc ribbon is degenerated (52). For the paralo-
gous pair of the L31 r-protein in B. subtilis, RpmE (C) and YtiA
(C), it was shown that YtiA (C) expression is repressed by Zur
and YtiA liberates RpmE (C) from the ribosome under zinc-
deficient conditions (2,29). Panina et al. identified candidate
binding sites for different zinc repressors upstream of genes cod-
ing for Cparalogues of the r-proteins L31, L33, L36, and S14 of
a range of different bacterial species and confirmed that upon
existence of a CandaCcopy of the r-protein the gene encod-
ing the Ccopy is regulated by a zinc-dependent repressor (41).
The replacement of zinc-containing r-proteins by non-zinc-con-
taining paralogues upon zinc depletion might liberate zinc for
maintaining zinc homeostasis and may enhance bacterial survival
when facing zinc-restrictive conditions in vivo (25,41). A poten-
tial evolutionary explanation for the duplication of r-proteins
with different zinc contents might be that ribosomal assembly in
any case needs to be maintained under zinc-restrictive conditions.
In the meningococcal genome, the genes for the 50S ribosomal
proteins L31 and L36 are duplicated. Based on sequence similarity
to B. subtilis RpmE (C) and YtiA (C), the paralogous pairs of
N. meningitidis r-proteins L31 and L36 are NMB1956 (C)/
NMB0942 (C) and NMB0164 (C)/NMB0941 (C), respec-
tively (34,51). The nmb0942-nmb0941 (rpmEJ) operon was
shown to be repressed at high zinc in this study, and we identified
a Zur box upstream of nmb0942. Therefore, we assume that also in
meningococci in the absence of zinc the enzyme lacking the zinc
ribbon takes over the function of the Cprotein.
Comparison to other regulons. In a study by Wu et al., the
manganese-responsive PerR regulon has been described (64).
Gonococcal PerR is a member of the Fur family and shows 96%
protein identity to meningococcal Zur. Wu et al. conducted a
microarray comparison of a Neisseria gonorrhoeae wild-type strain
versus its perR knockout mutant. They found 11 genes upregu-
lated and 1 gene downregulated in the perR mutant. Except for
three genes, all genes do have meningococcal homologues with
93% to 98% protein identity. All of those homologues also have
been identified in our microarray analysis of the zinc-responsive
regulon of Neisseria meningitidis:nmb0586-nmb0588 (ng0170-
ng0168), nmb0941-nmb0942 (ng0931-ng0930), nmb0964 (ng1205),
nmb1475 (ng1049), nmb1497 (ng0952), and nmb0546 (ng1442)(51,
64). This suggests that the meningococcal zinc-dependent Zur
regulation is related to the gonococcal manganese-dependent
PerR regulation.
Several of the Zur-regulated proteins were also regulated dif-
ferentially in a microarray study that compared transcriptional
profiles of N. meningitidis grown with different host iron binding
proteins (i.e., hemoglobin, transferrin, and lactoferrin) as the sole
iron source (23). The genes nmb0941,nmb0942,nmb1475, and
nmb1497 were downregulated, and nmb0546 was upregulated
upon exposure to lactoferrin (compared to hemoglobin or trans-
ferrin) (23). Lactoferrin, which is present in secretions and on
mucosal surfaces of the human host, can be used by meningococci
employing a lactoferrin receptor to catch the iron from lactoferrin
(23,38). Therefore, regulation of these genes upon lactoferrin
exposure may additionally be accomplished by a second regulator
that senses iron. This was shown recently for znuD, whose expres-
sion also was iron induced (26). Besides its upstream Zur binding
site, znuD also harbors a Fur binding site where Fur in vitro binds
to, independently of Zur (26). However, we did not detect any Fur
box upstream of nmb0942-nmb0941,nmb1475,nmb1497, and
nmb0546, nor has znuD been regulated upon lactoferrin exposure
in the study by Jordan and Saunders (23). Hence, we favor a sec-
ond hypothesis to explain the overlap of the zinc and lactoferrin
regulons. As lactoferrin has been reported to also loosely bind zinc
(1), meningococci might use it as an additional source of zinc.
Zinc could then act as a cofactor for Zur, leading to Zur-mediated
regulation of the genes mentioned above as also seen in this study.
The transcriptomic response of N. meningitidis to whole-blood
exposure was recently recorded using microarrays (11). Interest-
ingly, several of the genes deregulated by this approach were also
found to be changed in their expression upon zinc exposure. This
finding suggests that zinc depletion upon transfer of bacteria from
liquid culture to whole blood is responsible for a part of the tran-
scriptional changes observed. Furthermore, this finding also rein-
forces that changes of zinc concentration mimic an important
environmental signal encountered by bacteria during pathogen-
host interaction.
In summary, we elucidated the transcriptional adaptation of N.
Zinc-Responsive Regulon of Meningococci
December 2012 Volume 194 Number 23 jb.asm.org 6601
meningitidis to zinc using strain MC58. The regulon is assumed to
cover at least a significant portion of the Zur regulon, as all tran-
scriptional units (genes/operons) were preceded by promoters
harboring a Zur box. The functionality of representative motifs
was confirmed by EMSA. It will be of interest to investigate the
concerted action of genes derepressed at low-zinc conditions in
vivo, as low zinc is encountered by the bacteria upon infection
(25).
ACKNOWLEDGMENTS
The study was funded by a grant to U.V. provided by the German Federal
Ministry of Education and Research (reference number 0315434) via the
ERA-NET PathoGenoMics program (2nd call).
REFERENCES
1. Ainscough EW, Brodie AM, Plowman JE. 1980. Zinc transport by lac-
toferrin in human milk. Am. J. Clin. Nutr. 33:1314–1315.
2. Akanuma G, Nanamiya H, Natori Y, Nomura N, Kawamura F. 2006.
Liberation of zinc-containing L31 (RpmE) from ribosomes by its paralo-
gous gene product, YtiA, in Bacillus subtilis. J. Bacteriol. 188:2715–2720.
3. Ammendola S, et al. 2007. High-affinity Zn2uptake system ZnuABC is
required for bacterial zinc homeostasis in intracellular environments and
contributes to the virulence of Salmonella enterica. Infect. Immun. 75:
5867–5876.
4. Bentley SD, et al. 2007. Meningococcal genetic variation mechanisms
viewed through comparative analysis of serogroup C strain FAM18. PLoS
Genet. 3:e23. doi:10.1371/journal.pgen.0030023.
5. Chen PR, He C. 2008. Selective recognition of metal ions by metalloregu-
latory proteins. Curr. Opin. Chem. Biol. 12:214–221.
6. Crooks GE, Hon G, Chandonia JM, Brenner SE. 2004. WebLogo: a
sequence logo generator. Genome Res. 14:1188–1190.
7. Dalet K, Gouin E, Cenatiempo Y, Cossart P, Hechard Y. 1999. Char-
acterisation of a new operon encoding a Zur-like protein and an associated
ABC zinc permease in Listeria monocytogenes. FEMS Microbiol. Lett.
174:111–116.
8. Dineshkumar TK, Thanedar S, Subbulakshmi C, Varshney U. 2002. An
unexpected absence of queuosine modification in the tRNAs of an Esch-
erichia coli B strain. Microbiology 148:3779–3787.
9. Durand JM, Bjork GR. 2003. Putrescine or a combination of methionine
and arginine restores virulence gene expression in a tRNA modification-
deficient mutant of Shigella flexneri: a possible role in adaptation of viru-
lence. Mol. Microbiol. 47:519–527.
10. Durand JM, Dagberg B, Uhlin BE, Bjork GR. 2000. Transfer RNA
modification, temperature and DNA superhelicity have a common target
in the regulatory network of the virulence of Shigella flexneri: the expres-
sion of the virF gene. Mol. Microbiol. 35:924–935.
11. Echenique-Rivera H, et al. 2011. Transcriptome analysis of Neisseria
meningitidis in human whole blood and mutagenesis studies identify vir-
ulence factors involved in blood survival. PLoS Pathog. 7:e1002027. doi:
10.1371/journal.ppat.1002027.
12. Edgar R, Domrachev M, Lash AE. 2002. Gene Expression Omnibus:
NCBI gene expression and hybridization array data repository. Nucleic
Acids Res. 30:207–210.
13. Gaballa A, Helmann JD. 1998. Identification of a zinc-specific metallo-
regulatory protein, Zur, controlling zinc transport operons in Bacillus
subtilis. J. Bacteriol. 180:5815–5821.
14. Gaballa A, Wang T, Ye RW, Helmann JD. 2002. Functional analysis of
the Bacillus subtilis Zur regulon. J. Bacteriol. 184:6508 6514.
15. Gabbianelli R, et al. 2011. Role of ZnuABC and ZinT in Escherichia coli
O157:H7 zinc acquisition and interaction with epithelial cells. BMC Mi-
crobiol. 11:36. doi:10.1186/1471-2180-11-36.
16. Gaur R, Varshney U. 2005. Genetic analysis identifies a function for the
queC (ybaX) gene product at an initial step in the queuosine biosynthetic
pathway in Escherichia coli. J. Bacteriol. 187:6893–6901.
17. Graham AI, et al. 2009. Severe zinc depletion of Escherichia coli: roles for
high affinity zinc binding by ZinT, zinc transport and zinc-independent
proteins. J. Biol. Chem. 284:18377–18389.
18. Grifantini R, et al. 2003. Identification of iron-activated and -repressed
Fur-dependent genes by transcriptome analysis of Neisseria meningitidis
group B. Proc. Natl. Acad. Sci. U. S. A. 100:9542–9547.
19. Gunasekera TS, Herre AH, Crowder MW. 2009. Absence of ZnuABC-
mediated zinc uptake affects virulence-associated phenotypes of uro-
pathogenic Escherichia coli CFT073 under Zn(II)-depleted conditions.
FEMS Microbiol. Lett. 300:36 41.
20. Hall TA. 1999. BioEdit: a user-friendly biological sequence alignment
editor and analysis program for Windows 95/98/NT. Nucleic Acids Symp.
41:95–98.
21. Hantke K. 2001. Bacterial zinc transporters and regulators. Biometals
14:239–249.
22. Huang DL, et al. 2008. The Zur of Xanthomonas campestris functions as
a repressor and an activator of putative zinc homeostasis genes via recog-
nizing two distinct sequences within its target promoters. Nucleic Acids
Res. 36:4295–4309.
23. Jordan PW, Saunders NJ. 2009. Host iron binding proteins acting as
niche indicators for Neisseria meningitidis. PLoS One 4:e5198. doi:
10.1371/journal.pone.0005198.
24. Katayama A, Tsujii A, Wada A, Nishino T, Ishihama A. 2002. System-
atic search for zinc-binding proteins in Escherichia coli. Eur. J. Biochem.
269:2403–2413.
25. Kehl-Fie TE, Skaar EP. 2010. Nutritional immunity beyond iron: a role
for manganese and zinc. Curr. Opin. Chem. Biol. 14:218–224.
26. Kumar P, Sannigrahi S, Tzeng YL. 2012. The Neisseria meningitidis
ZnuD zinc receptor contributes to interactions with epithelial cells and
supports heme utilization when expressed in Escherichia coli. Infect. Im-
mun. 80:657–667.
27. Lee BW, Van Lanen SG, Iwata-Reuyl D. 2007. Mechanistic studies of
Bacillus subtilis QueF, the nitrile oxidoreductase involved in queuosine
biosynthesis. Biochemistry 46:12844 –12854.
28. Lee JW, Helmann JD. 2007. Functional specialization within the Fur
family of metalloregulators. Biometals 20:485–499.
29. Li Y, et al. 2009. Characterization of Zur-dependent genes and direct Zur
targets in Yersinia pestis. BMC Microbiol. 9:128. doi:10.1186/1471-2180-
9-128.
30. Lim KH, et al. 2008. Metal binding specificity of the MntABC permease of
Neisseria gonorrhoeae and its influence on bacterial growth and interac-
tion with cervical epithelial cells. Infect. Immun. 76:3569–3576.
31. Lindsay JA, Foster SJ. 2001. zur: a Zn(2)-responsive regulatory element
of Staphylococcus aureus. Microbiology 147:1259–1266.
32. Maciag A, et al. 2007. Global analysis of the Mycobacterium tuberculosis
Zur (FurB) regulon. J. Bacteriol. 189:730–740.
33. Reference deleted.
34. Makarova KS, Ponomarev VA, Koonin EV. 2001. Two C or not two C:
recurrent disruption of Zn-ribbons, gene duplication, lineage-specific
gene loss, and horizontal gene transfer in evolution of bacterial ribosomal
proteins. Genome Biol. 2:research 0033–research0033.14. doi:10.1186/gb-
2001-2-9-research0033.
35. Mao F, Dam P, Chou J, Olman V, Xu Y. 2009. DOOR: a database for
prokaryotic operons. Nucleic Acids Res. 37:D459–D463.
36. McGuinness BT, et al. 1991. Point mutation in meningococcal porA gene
associated with increased endemic disease. Lancet 337:514–517.
37. Mellin JR, et al. 2010. Role of Hfq in iron-dependent and -independent
gene regulation in Neisseria meningitidis. Microbiology 156:2316–2326.
38. Morgenthau A, Livingstone M, Adamiak P, Schryvers AB. 2012. The
role of lactoferrin binding protein B in mediating protection against hu-
man lactoferricin. Biochem. Cell Biol. 90:417–423.
39. Novichkov PS, et al. 2010. RegPrecise: a database of curated genomic
inferences of transcriptional regulatory interactions in prokaryotes. Nu-
cleic Acids Res. 38:D111–D118.
40. O’Halloran TV. 1993. Transition metals in control of gene expression.
Science 261:715–725.
41. Panina EM, Mironov AA, Gelfand MS. 2003. Comparative genomics of
bacterial zinc regulons: enhanced ion transport, pathogenesis, and rear-
rangement of ribosomal proteins. Proc. Natl. Acad. Sci. U. S. A. 100:9912–
9917.
42. Pannekoek Y, et al. 2009. Molecular characterization and identification
of proteins regulated by Hfq in Neisseria meningitidis. FEMS Microbiol.
Lett. 294:216–224.
43. Parkhill J, et al. 2000. Complete DNA sequence of a serogroup A strain of
Neisseria meningitidis Z2491. Nature 404:502–506.
44. Patzer SI, Hantke K. 2000. The zinc-responsive regulator Zur and its
control of the znu gene cluster encoding the ZnuABC zinc uptake system
in Escherichia coli. J. Biol. Chem. 275:24321–24332.
Pawlik et al.
6602 jb.asm.org Journal of Bacteriology
45. Patzer SI, Hantke K. 1998. The ZnuABC high-affinity zinc uptake system
and its regulator Zur in Escherichia coli. Mol. Microbiol. 28:1199–1210.
46. Reference deleted.
47. R Development Core Team. 2008. R: a language and environment for sta-
tistical computing. R Foundation for Statistical Computing, Vienna, Austria.
48. Reference deleted.
49. Reyes-Caballero H, Campanello GC, Giedroc DP. 2011. Metalloregula-
tory proteins: metal selectivity and allosteric switching. Biophys. Chem.
156:103–114.
50. Rosenstein NE, Perkins BA, Stephens DS, Popovic T, Hughes JM. 2001.
Meningococcal disease. N. Engl. J. Med. 344:1378–1388.
51. Rusniok C, et al. 2009. NeMeSys: a biological resource for narrowing the
gap between sequence and function in the human pathogen Neisseria
meningitidis. Genome Biol. 10:R110. doi:10.1186/gb-2009-10-10-r110.
52. Sankaran B, et al. 2009. Zinc-independent folate biosynthesis: genetic,
biochemical, and structural investigations reveal new metal dependence
for GTP cyclohydrolase IB. J. Bacteriol. 191:6936 6949.
53. Schielke S, et al. 2011. Characterization of FarR as a highly specialized,
growth phase-dependent transcriptional regulator in Neisseria meningi-
tidis. Int. J. Med. Microbiol. 301:325–333.
54. Schröder J, Jochmann N, Rodionov DA, Tauch A. 2010. The Zur
regulon of Corynebacterium glutamicum ATCC 13032. BMC Genomics
11:12. doi:10.1186/1471-2164-11-12.
55. Schwarz R, et al. 2010. Evaluation of one- and two-color gene expression
arrays for microbial comparative genome hybridization analyses in rou-
tine applications. J. Clin. Microbiol. 48:3105–3110.
56. Shaik YB, et al. 2007. Expression of the iron-activated nspA and secY
genes in Neisseria meningitidis group B by Fur-dependent and -indepen-
dent mechanisms. J. Bacteriol. 189:663–669.
57. Shin JH, Oh SY, Kim SJ, Roe JH. 2007. The zinc-responsive regulator
Zur controls a zinc uptake system and some ribosomal proteins in Strep-
tomyces coelicolor A3(2). J. Bacteriol. 189:4070 4077.
58. Smyth G. 2005. Limma: linear models for microarray data, p 397– 420. In
Gentleman R, Carey V, Dudoit SR, Irizarry R, Huber W (ed), Bioinfor-
matics and computational biology solutions using R and Bioconductor.
Springer, New York, NY.
59. Stork M, et al. 2010. An outer membrane receptor of Neisseria meningi-
tidis involved in zinc acquisition with vaccine potential. PLoS Pathog.
6:e1000969. doi:10.1371/journal.ppat.1000969.
60. Tettelin H, et al. 2000. Complete genome sequence of Neisseria menin-
gitidis serogroup B strain MC58. Science 287:1809–1815.
61. Thompson WA, Newberg LA, Conlan S, McCue LA, Lawrence CE.
2007. The Gibbs centroid sampler. Nucleic Acids Res. 35:W232–W237.
62. UniProt Consortium. 2012. Reorganizing the protein space at the Uni-
versal Protein Resource (UniProt). Nucleic Acids Res. 40:D71–D75.
63. van Alen T, et al. 2010. Comparative proteomic analysis of biofilm and
planktonic cells of Neisseria meningitidis. Proteomics 10:4512–4521.
64. Wu HJ, et al. 2006. PerR controls Mn-dependent resistance to oxidative
stress in Neisseria gonorrhoeae. Mol. Microbiol. 60:401–416.
65. Yang YH, Speed T. 2002. Design issues for cDNA microarray experi-
ments. Nat. Rev. Genet. 3:579–588.
Zinc-Responsive Regulon of Meningococci
December 2012 Volume 194 Number 23 jb.asm.org 6603
... Transition metals, such as zinc, manganese, and copper, are essential to many host processes, including oxidative stress resistance (Girotto et al., 2014;Jarosz et al., 2017;Ganini et al., 2018), cell signaling and metabolism (Maares et al., 2018), immune modulation (Hu Frisk et al., 2017), post-translational modifications (Braiterman et al., 2015;Tidball et al., 2015), and structural maintenance and enzymatic processing (Tidball et al., 2015;Ganini et al., 2018). These metals play similar roles in pathogens including Neisseria meningitidis (Persson et al., 2001;Pawlik et al., 2012;Hecel et al., 2018) and Escherichia coli (Kaur et al., 2017). ...
... This import system may be required for growth and pathogenesis in N. meningitidis much like it is in N. gonorrhoeae. Pawlik et al., 2012;Jean et al., 2016) nk, (not known) indicates the affinity for that ligand is not known. These proteins have been shown to be regulated by or interact with multiple metals. ...
... Pathogenic Neisseriae express two well-characterized metal dependent regulators: Fur and Zur (Zur has also previously been called PerR) Pawlik et al., 2012;Jean, 2015;Jean et al., 2016). In N. gonorrhoeae, Zur is hypothesized to repress zinc and manganese import genes in the presence of these metals and to de-repress zinc and manganese import genes in the absence of these metals ( Table 1) (Chen and Morse, 2001;Wu et al., 2006). ...
Article
Full-text available
Neisseria gonorrhoeae and Neisseria meningitidis are human-specific pathogens in the Neisseriaceae family that can cause devastating diseases. Although both species inhabit mucosal surfaces, they cause dramatically different diseases. Despite this, they have evolved similar mechanisms to survive and thrive in a metal-restricted host. The human host restricts, or overloads, the bacterial metal nutrient supply within host cell niches to limit pathogenesis and disease progression. Thus, the pathogenic Neisseria require appropriate metal homeostasis mechanisms to acclimate to such a hostile and ever-changing host environment. This review discusses the mechanisms by which the host allocates and alters zinc, manganese, and copper levels and the ability of the pathogenic Neisseria to sense and respond to such alterations. This review will also discuss integrated metal homeostasis in N. gonorrhoeae and the significance of investigating metal interplay.
... In conditions of high cytoplasmic zinc concentration, Zur binds to zinc. This increases the affinity of Zur for a palindromic DNA motif called the Zur box, which is found in the promoter region of Zur-regulated genes, thereby repressing their transcription (35). Conversely, when cytoplasmic zinc is low, zinc ions dissociate from Zur, which loses its affinity to DNA, leading to de-repression of gene expression. ...
... Four ORFs were more abundant in all three comparisons. Three of these ORFs are known to be zinc-repressed and regulated by Zur: the periplasmic zinc shuttle protein ZnuA (NGO_0168), a zinc-regulated hypothetical periplasmic protein (NGO_1049), and a zinc-independent ribosomal subunit RpmE2 (NGO_0930) (35,36). One gene was reduced in abundance in all of the aforementioned comparisons, AdhP (NGO_1442), which has been previously described as induced under high zinc concentrations and in the presence of Zur (35,36). ...
... Three of these ORFs are known to be zinc-repressed and regulated by Zur: the periplasmic zinc shuttle protein ZnuA (NGO_0168), a zinc-regulated hypothetical periplasmic protein (NGO_1049), and a zinc-independent ribosomal subunit RpmE2 (NGO_0930) (35,36). One gene was reduced in abundance in all of the aforementioned comparisons, AdhP (NGO_1442), which has been previously described as induced under high zinc concentrations and in the presence of Zur (35,36). Transcripts for an additional six ORFs were higher with adherence under zinc sequestered conditions (adh seq versus sus seq) and with zinc sequestration when adherent (adh seq versus adh excess), but not with zinc sequestration in suspension (sus seq versus sus excess). ...
Article
Full-text available
Neisseria gonorrhoeae (Gc) must overcome limitation of metals such as zinc to colonize mucosal surfaces in its obligate human host. While the zinc-binding nutritional immunity proteins calprotectin (S100A8/A9) and psoriasin (S100A7) are abundant in human cervicovaginal lavage fluid, Gc possesses TonB-dependent transporters TdfH and TdfJ that bind and extract zinc from the human version of these proteins, respectively. Here we investigated the contribution of zinc acquisition to Gc infection of epithelial cells of the female genital tract. We found that TdfH and TdfJ were dispensable for survival of strain FA1090 Gc that were associated with Ect1 human immortalized epithelial cells, when zinc was limited by calprotectin and psoriasin. In contrast, suspension-grown bacteria declined in viability under the same conditions. Exposure to murine calprotectin, which Gc cannot use as a zinc source, similarly reduced survival of suspension-grown Gc, but not Ect1-associated Gc. We ruled out epithelial cells as a contributor to the enhanced growth of cell-associated Gc under zinc limitation. Instead, we found that attachment to glass was sufficient to enhance bacterial growth when zinc was sequestered. We compared the transcriptional profiles of WT Gc adherent to glass coverslips or in suspension, when zinc was sequestered with murine calprotectin or provided in excess, from which we identified open reading frames that were increased by zinc sequestration in adherent Gc. One of these, ZnuA, was necessary but not sufficient for survival of Gc under zinc-limiting conditions. These results show that adherence protects Gc from zinc-dependent growth restriction by host nutritional immunity proteins.
... Pathogenic Neisseria encode the transcriptional regulators Fur and Zur, which mediate changes in expression of metal homeostasis genes in response to intracellular levels of iron [13] and zinc [14], respectively (Fig 1). When intracellular iron concentration is high, Fur binds iron and dimerizes, increasing its affinity for palindromic A/T rich sequences found in the promoters of iron-responsive genes to repress their transcription. ...
... Similarly, Zur binds zinc and recognizes a different consensus DNA sequence in Zur-regulated genes. Genes in the fur and zur regulons encode proteins involved in metal acquisition, metal transport, and metabolism, and as-yet uncharacterized proteins [13,14]. Fur can also enhance gene expression, potentially by opening DNA for binding of RNA polymerase or other regulators [15]. ...
... Additionally, genes, namely, queC and queF are also Zur-regulated, these encode for the cytoplasmic enzymes involved in the quenosine production, which is basically a modified nucleoside and is often present at the first/wobble position of the tRNA anticodon. The expression of these genes increases under low-Zn conditions, as reported in N. meningitidis [113]. ...
... Number of total genes regulated by Zur vary for different bacterial species, while in C. glutamicum [121] it is only 9, but for N. meningitidis [113] and M. tuberculosis [79], it is 17 and 30, respectively. Additionally, in certain cases, the Zur regulon is extremely large, that is, up to 154 and 121 genes present in Y pestis and S. suis serotype 2 strain [77,122], respectively. ...
Article
Full-text available
Zinc (Zn) is the quintessential d block metal, needed for survival in all living organisms. While Zn is an essential element, its excess is deleterious, therefore, maintenance of its intracellular concentrations is needed for survival. The living organisms, during the course of evolution, developed proteins that can track the limitation or excess of necessary metal ions, thus providing survival benefits under variable environmental conditions. Zinc uptake regulator (Zur) is a regulatory transcriptional factor of the FUR superfamily of proteins, abundant among the bacterial species and known for its intracellular Zn sensing ability. In this study, we highlight the roles played by Zur in maintaining the Zn levels in various bacterial species as well as the fact that in recent years Zur has emerged not only as a Zn homeostatic regulator but also as a protein involved directly or indirectly in virulence of some pathogens. This functional aspect of Zur could be exploited in the ventures for the identification of newer antimicrobial targets. Despite extensive research on Zur, the insights into its overall regulon and its moonlighting functions in various pathogens yet remain to be explored. Here in this review, we aim to summarise the disparate functional aspects of Zur proteins present in various bacterial species.
... In addition to its classical role as a repressor, Zur has been reported to activate zinc exporter genes in Xanthomonas campestris and S. coelicolor [11,41]. Furthermore, Zur works as an activator of the zinc-containing alcohol dehydrogenases nmb0546 from Neisseria meningitidis and adhA (dip2114) from Corynebacterium diphtheriae as well as for the transcription of the NosR-related protein (nmb0577) in N. meningitidis [42,43]. Therefore, as their Fur and PerR paralogs, Zur can work in different ways. ...
... Since zinc deficiency seems to stimulate the σ R regulatory system in S. coelicolor, this finding unveils a connection between thiol-oxidizing stress and zinc depletion. Regulation of ribosomal proteins by Zur has also been reported in M. tuberculosis [49], B. subtilis [50], E. coli [51], Yersinia pestis [25] and N. meningitidis [43]. Other Zur targets, less frequent but also present in several phyla are metallochaperones [35,52], proteins involved in the synthesis of cobalamin and tetrapyrrole [35,53], oxidoreductases [35,54], as well as genes involved in pathogenesis and virulence of species as different as Xantomonas spp. ...
Article
Proteins belonging to the FUR (ferric uptake regulator) family are the cornerstone of metalloregulation in most prokaryotes. Although numerous reviews have been devoted to these proteins, these reports are mainly focused on the Fur paralog that gives name to the family. In the last years, the increasing knowledge on the other, less ubiquitous members of this family has evidenced their importance in bacterial metabolism. As the Fur paralog, the major regulator of iron homeostasis, Zur, Irr, BosR and PerR are tightly related to stress defenses and host-pathogen interaction being in many cases essential for virulence. Furthermore, the Nur and Mur paralogs largely contribute to control nickel and manganese homeostasis, which are cofactors of pivotal proteins for host colonization and bacterial redox homeostasis. The present review highlights the main features of FUR proteins that differ to the canonical Fur paralog either in the coregulatory metal, such as Zur, Nur and Mur, or in the action mechanism to control target genes, such as PerR, Irr and BosR.
... Queuosine plays an essential role in tRNA modifications, and the gut microbiome is thought to be a major provider of this micronutrient to the host [22,23]. However, it was shown to be upregulated under Zn limiting conditions, so again increased FCP may potentially be responsible for its increased abundance in active colitis patients [24,25]. ...
Article
Full-text available
Purpose Chronic granulomatous disorder (CGD) is a primary immunodeficiency which is frequently complicated by inflammatory colitis and is associated with systemic inflammation. Herein, we aimed to investigate the role of the microbiome in the pathogenesis of colitis and systemic inflammation. Methods We performed 16S rDNA sequencing on mucosal biopsy samples from each segment of 10 CGD patients’ colons and conducted compositional and functional pathway prediction analyses. Results The microbiota in samples from colitis patients demonstrated reduced taxonomic alpha-diversity compared to unaffected patients, even in apparently normal bowel segments. Functional pathway richness was similar between the colitic and non-colitic mucosa, although metabolic pathways involved in butyrate biosynthesis or utilization were enriched in patients with colitis and correlated positively with fecal calprotectin levels. One patient with very severe colitis was dominated by Enterococcus spp., while among other patients Bacteroides spp. abundance correlated with colitis severity measured by fecal calprotectin and an endoscopic severity score. In contrast, Blautia abundance is associated with low severity scores and mucosal health. Several taxa and functional pathways correlated with concentrations of inflammatory cytokines in blood but not with colitis severity. Notably, dividing patients into “high” and “low” systemic inflammation groups demonstrated clearer separation than on the basis of colitis status in beta-diversity analyses. Conclusion The microbiome is abnormal in CGD-associated colitis and altered functional characteristics probably contribute to pathogenesis. Furthermore, the relationship between the mucosal microbiome and systemic inflammation, independent of colitis status, implies that the microbiome in CGD can influence the inflammatory phenotype of the condition.
Article
Deazaguanine modifications play multifaceted roles in the molecular biology of DNA and tRNA, shaping diverse yet essential biological processes, including the nuanced fine-tuning of translation efficiency and the intricate modulation of codon-anticodon interactions. Beyond their roles in translation, deazaguanine modifications contribute to cellular stress resistance, self-nonself discrimination mechanisms, and host evasion defenses, directly modulating the adaptability of living organisms. Deazaguanine moieties extend beyond nucleic acid modifications, manifesting in the structural diversity of biologically active natural products. Their roles in fundamental cellular processes and their presence in biologically active natural products underscore their versatility and pivotal contributions to the intricate web of molecular interactions within living organisms. Here, we discuss the current understanding of the biosynthesis and multifaceted functions of deazaguanines, shedding light on their diverse and dynamic roles in the molecular landscape of life.
Article
Full-text available
Zur is a Fur-family metalloregulator that is widely used to control zinc homeostasis in bacteria. In Streptomyces coelicolor, Zur (ScZur) acts as both a repressor for zinc uptake (znuA) gene and an activator for zinc exporter (zitB) gene. Previous structural studies revealed three zinc ions specifically bound per ScZur monomer; a structural one to allow dimeric architecture and two regulatory ones for DNA-binding activity. In this study, we present evidence that Zur contains a fourth specific zinc-binding site with a key histidine residue (H36), widely conserved among actinobacteria, for regulatory function. Biochemical, genetic, and calorimetric data revealed that H36 is critical for hexameric binding of Zur to the zitB zurbox and further binding to its upstream region required for full activation. A comprehensive thermodynamic model demonstrated that the DNA-binding affinity of Zur to both znuA and zitB zurboxes is remarkably enhanced upon saturation of all three regulatory zinc sites. The model also predicts that the strong coupling between zinc binding and DNA binding equilibria of Zur drives a biphasic activation of the zitB gene in response to a wide concentration change of zinc. Similar mechanisms may be pertinent to other metalloproteins, expanding their response spectrum through binding multiple regulatory metals.
Article
Full-text available
Chromobacterium violaceum is a ubiquitous environmental bacterium that causes sporadic life-threatening infections in humans. How C. violaceum acquires zinc to colonize environmental and host niches is unknown. In this work, we demonstrated that C. violaceum employs the zinc uptake system ZnuABC to overcome zinc limitation in the host, ensuring the zinc supply for several physiological demands. Our data indicated that the C. violaceum ZnuABC transporter is encoded in a zur -CV_RS15045-CV_RS15040- znuCBA operon. This operon was repressed by the zinc uptake regulator Zur and derepressed in the presence of the host protein calprotectin (CP) and the synthetic metal chelator EDTA. A Δ znuCBA mutant strain showed impaired growth under these zinc-chelated conditions. Moreover, the deletion of znuCBA provoked a reduction in violacein production, swimming motility, biofilm formation, and bacterial competition. Remarkably, the Δ znuCBA mutant strain was highly attenuated for virulence in an in vivo mouse infection model and showed a low capacity to colonize the liver, grow in the presence of CP, and resist neutrophil killing. Overall, our findings demonstrate that ZnuABC is essential for C. violaceum virulence, contributing to subvert the zinc-based host nutritional immunity.
Article
Full-text available
Vibrio cholerae is the causative agent of cholera, a notorious diarrheal disease that is typically transmitted via contaminated drinking water. The current pandemic agent, the El Tor biotype, has undergone several genetic changes that include horizontal acquisition of two genomic islands (VSP-I and VSP-II). VSP presence strongly correlates with pandemicity; however, the contribution of these islands to V . cholerae ’s life cycle, particularly the 26-kb VSP-II, remains poorly understood. VSP-II-encoded genes are not expressed under standard laboratory conditions, suggesting that their induction requires an unknown signal from the host or environment. One signal that bacteria encounter under both host and environmental conditions is metal limitation. While studying V . cholerae ’s zinc-starvation response in vitro , we noticed that a mutant constitutively expressing zinc starvation genes (Δ zur ) congregates at the bottom of a culture tube when grown in a nutrient-poor medium. Using transposon mutagenesis, we found that flagellar motility, chemotaxis, and VSP-II encoded genes were required for congregation. The VSP-II genes encode an AraC-like transcriptional activator (VerA) and a methyl-accepting chemotaxis protein (AerB). Using RNA-seq and lacZ transcriptional reporters, we show that VerA is a novel Zur target and an activator of the nearby AerB chemoreceptor. AerB interfaces with the chemotaxis system to drive oxygen-dependent congregation and energy taxis. Importantly, this work suggests a functional link between VSP-II, zinc-starved environments, and energy taxis, yielding insights into the role of VSP-II in a metal-limited host or aquatic reservoir.
Data
Full-text available
The mission of UniProt is to support biological research by providing a freely accessible, stable, comprehensive, fully classified, richly and accurately annotated protein sequence knowledgebase, with extensive cross-references and querying interfaces. UniProt is comprised of four major components, each optimized for different uses: the UniProt Archive, the UniProt Knowledgebase, the UniProt Reference Clusters and the UniProt Metagenomic and Environmental Sequence Database. A key development at UniProt is the provision of complete, reference and representative proteomes. UniProt is updated and distributed every 4 weeks and can be accessed online for searches or download at http://www.uniprot.org.
Article
Full-text available
Summary The Gene Expression Omnibus (GEO) project was initiated at NCBI in 1999 in response to the growing demand for a public repository for data generated from high-throughput microarray experiments. GEO has a flexible and open design that allows the submission, storage, and retrieval of many types of data sets, such as those from high-throughput gene expression, genomic hybridization, and antibody array experiments. GEO was never intended to replace lab-specific gene expression databases or laboratory information management systems (LIMS), both of which usually cater to a particular type of data set and analytical method. Rather, GEO complements these resources by acting as a central, molecular abundance-data distribution hub. GEO is available on the World Wide Web at http://www.ncbi.nih.gov/geo (http://www.ncbi.nih.gov/geo).
Article
The mission of UniProt is to support biological research by providing a freely accessible, stable, comprehensive, fully classified, richly and accurately annotated protein sequence knowledgebase, with extensive cross-references and querying interfaces. UniProt is comprised of four major components, each optimized for different uses: the UniProt Archive, the UniProt Knowledgebase, the UniProt Reference Clusters and the UniProt Metagenomic and Environmental Sequence Database. A key development at UniProt is the provision of complete, reference and representative proteomes. UniProt is updated and distributed every 4 weeks and can be accessed online for searches or download at http://www.uniprot.org.
Article
Background: Ribosomal proteins are encoded in all genomes of cellular life forms and are, generally, well conserved during evolution. In prokaryotes, the genes for most ribosomal proteins are clustered in several highly conserved operons, which ensures efficient co-regulation of their expression. Duplications of ribosomal-protein genes are infrequent, and given their coordinated expression and functioning, it is generally assumed that ribosomal-protein genes are unlikely to undergo horizontal transfer. However, with the accumulation of numerous complete genome sequences of prokaryotes, several paralogous pairs of ribosomal protein genes have been identified. Here we analyze all such cases and attempt to reconstruct the evolutionary history of these ribosomal proteins. Results: Complete bacterial genomes were searched for duplications of ribosomal proteins. Ribosomal proteins L36, L33, L31, S14 are each duplicated in several bacterial genomes and ribosomal proteins L11, L28, L7/L12, S1, S15, S18 are so far duplicated in only one genome each. Sequence analysis of the four ribosomal proteins, for which paralogs were detected in several genomes, two of the ribosomal proteins duplicated in one genome (L28 and S18), and the ribosomal protein L32 showed that each of them comes in two distinct versions. One form contains a predicted metal-binding Zn-ribbon that consists of four conserved cysteines (in some cases replaced by histidines), whereas, in the second form, these metal-chelating residues are completely or partially replaced. Typically, genomes containing paralogous genes for these ribosomal proteins encode both versions, designated C+ and C-, respectively. Analysis of phylogenetic trees for these seven ribosomal proteins, combined with comparison of genomic contexts for the respective genes, indicates that in most, if not all cases, their evolution involved a duplication of the ancestral C+ form early in bacterial evolution, with subsequent alternative loss of the C+ and C- forms in different lineages. Additionally, evidence was obtained for a role of horizontal gene transfer in the evolution of these ribosomal proteins, with multiple cases of gene displacement 'in situ', that is, without a change of the gene order in the recipient genome. Conclusions: A more complex picture of evolution of bacterial ribosomal proteins than previously suspected is emerging from these results, with major contributions of lineage-specific gene loss and horizontal gene transfer. The recurrent theme of emergence and disruption of Zn-ribbons in bacterial ribosomal proteins awaits a functional interpretation.
Article
The 2,272,351–base pair genome of Neisseria meningitidis strain MC58 (serogroup B), a causative agent of meningitis and septicemia, contains 2158 predicted coding regions, 1158 (53.7%) of which were assigned a biological role. Three major islands of horizontal DNA transfer were identified; two of these contain genes encoding proteins involved in pathogenicity, and the third island contains coding sequences only for hypothetical proteins. Insights into the commensal and virulence behavior of N. meningitidis can be gleaned from the genome, in which sequences for structural proteins of the pilus are clustered and several coding regions unique to serogroup B capsular polysaccharide synthesis can be identified. Finally, N. meningitidis contains more genes that undergo phase variation than any pathogen studied to date, a mechanism that controls their expression and contributes to the evasion of the host immune system.
Data
The mission of UniProt is to provide the scientific community with a comprehensive, high-quality and freely accessible resource of protein sequence and functional information that is essential for modern biological research. UniProt is produced by the UniProt Consortium which consists of groups from the European Bioinformatics Institute, the Protein Information Resource and the Swiss Institute of Bioinformatics. The core activities include manual curation of protein sequences assisted by computa-tional analysis, sequence archiving, a user-friendly UniProt website and the provision of additional value-added information through cross-references to other databases. UniProt is comprised of four major components, each optimized for different uses: the UniProt Archive, the UniProt Knowledge-base, the UniProt Reference Clusters and the Uni-Prot Metagenomic and Environmental Sequence Database. One of the key achievements of the UniProt consortium in 2008 is the completion of the first draft of the complete human proteome in UniProtKB/Swiss-Prot. This manually annotated representation of all currently known human protein-coding genes was made available in UniProt release 14.0 with 20 325 entries. UniProt is updated and distributed every three weeks and can be accessed online for searches or downloaded at www.uniprot.org. INTRODUCTION
Data
In previous studies it has been established that resis-tance to superoxide by Neisseria gonorrhoeae is dependent on the accumulation of Mn(II) ions involv-ing the ABC transporter, MntABC. A mutant strain lacking the periplasmic binding protein component (MntC) of this transport system is hypersensitive to killing by superoxide anion. In this study the mntC mutant was found to be more sensitive to H 2 O 2 killing than the wild-type. Analysis of regulation of MntC expression revealed that it was de-repressed under low Mn(II) conditions. The N. gonorrhoeae mntABC locus lacks the mntR repressor typically found asso-ciated with this locus in other organisms. A search for a candidate regulator of mntABC expression revealed a homologue of PerR, a Mn-dependent per-oxide-responsive regulator found in Gram-positive organisms. A perR mutant expressed more MntC pro-tein than wild-type, and expression was independent of Mn(II), consistent with a role for PerR as a repressor of mntABC expression. The PerR regulon of N. gonorrhoeae was defined by microarray analysis and includes ribosomal proteins, TonB-dependent receptors and an alcohol dehydrogenase. Both the mntC and perR mutants had reduced intracellular sur-vival in a human cervical epithelial cell model.
Data
mntABC from Neisseria gonorrhoeae encodes an ABC permease which includes a periplasmic divalent cation binding receptor protein of the cluster IX family, encoded by mntC. Analysis of an mntC mutant showed that growth of N. gonorrhoeae could be stimulated by addition of either manganese(II) or zinc(II) ions, suggesting that the MntABC system could transport both ions. In contrast, growth of the mntAB mutant in liquid culture was possible only when the medium was supplemented with an antioxidant such as mannitol, consistent with the view that ion transport via MntABC is essential for protection of N. gonorrhoeae against oxidative stress. Using recombinant MntC, we determined that MntC binds Zn 2 and Mn 2 with almost equal affinity (dissociation constant of 0.1 M). Competition assays with the metallochromic zinc indicator 4-(2-pyridy-lazo)resorcinol showed that MntC binds Mn 2 and Zn 2 at the same binding site. Analysis of the N. gonorrhoeae genome showed that MntC is the only Mn/Zn metal binding receptor protein cluster IX in this bacterium, in contrast to the situation in many other bacteria which have systems with dedicated Mn and Zn binding proteins as part of distinctive ABC cassette permeases. Both the mntC and mntAB mutants had reduced intracellular survival in a human cervical epithelial cell model and showed reduced ability to form a biofilm. These data suggest that the MntABC transporter is of importance for survival of Neisseria gonorrhoeae in the human host. Neisseria gonorrhoeae is a mucosal pathogen often associated with the genitourinary tract, and it is the etiological agent of the sexually transmitted disease gonorrhea (31). Gonorrhea is often characterized by a localized inflammatory response in-volving inflamed urogenital tissues and activated polymorpho-nuclear leukocytes (PMNs) (26). This innate immune response includes production of superoxide radical (O 2), hydrogen peroxide (H 2 O 2), and reactive nitrogen species. The accumu-lation of reactive oxygen species around the site of gonococcal infection can exert a cytotoxic effect, which results from oxi-dant interaction with proteins, lipids, and nucleic acids (18, 32). Despite the constant environmental stress elicited by O 2 and H 2 O 2 , gonococci are routinely isolated from PMN-laden pu-rulent exudates. Furthermore, studies suggest that gonococci not only survive within PMNs but may also replicate within these cells (27, 29, 36). The ability of gonococci to survive in environments high in reactive oxygen species suggests that this bacterium has an efficient antioxidant defense system. Previous studies show that the accumulation of manganese (Mn), via the Mn transporter, MntABC, is important for
Article
Full expression of the virulence genes of Shigella flexneri requires the presence of two modified nucleosides in the tRNA [queuosine, Q34, present in the wobble position (position 34) and 2-methylthio-N6-isopentenyladenosine (ms2i6A37, adjacent to and 3′ of the anticodon)]. The synthesis of these two nucleosides depends on the products of the tgt and miaA genes respectively. We have shown that the intracellular concentration of the virulence-related transcriptional regulator VirF is reduced in the absence of either of these modified nucleosides. The intracellular concentration of VirF is correlated with the expression of the virulence genes. Overproduction of VirF in the tgt and the miaA mutants suppressed the less virulent (tgt) or the avirulent (miaA) phenotypes respectively, caused by the tRNA modification deficiency. This suggests that the primary result of undermodification of the tRNA is a poor translation of virF mRNA and not of any other mRNA whose product acts downstream of the action of VirF. Shigella showed no virulence phenotypes at 30°C, but forced synthesis of VirF at 30°C induced the virulence phenotype at this low temperature. In addition, removal of the known gene silencer H-NS by a mutation in its structural gene hns increased the synthesis of VirF at low temperature and thus induced a virulent phenotype at 30°C. Conversely, decreased expression of VirF at 37°C induced by the addition of novobiocin, a known inhibitor of gyrase, led to an avirulent phenotype. We conclude that tRNA modification, temperature and superhelicity have the same target – the expression of VirF – to influence the expression of the central regulatory gene virB and thereby the virulence of Shigella. These results further strengthen the suggestion that the concentration of VirF is the critical factor in the regulation of virulence in Shigella. In addition, they emphasize the role of the bacterial translational machinery in the regulation of the expression of virulence genes which appears here quantitatively as important as the well-established regulation on the transcriptional level.