ArticlePDF AvailableLiterature Review

Competition for zinc binding in the host-pathogen interaction

Frontiers
Frontiers in Cellular and Infection Microbiology
Authors:

Abstract and Figures

Due to its favorable chemical properties, zinc is used as a structural or catalytic cofactor in a very large number of proteins. Despite the apparent abundance of this metal in all cell types, the intracellular pool of loosely bound zinc ions available for biological exchanges is in the picomolar range and nearly all zinc is tightly bound to proteins. In addition, to limit bacterial growth, some zinc-sequestering proteins are produced by eukaryotic hosts in response to infections. Therefore, to grow and multiply in the infected host, bacterial pathogens must produce high affinity zinc importers, such as the ZnuABC transporter which is present in most Gram-negative bacteria. Studies carried in different bacterial species have established that disruption of ZnuABC is usually associated with a remarkable loss of pathogenicity. The critical involvement of zinc in a plethora of metabolic and virulence pathways and the presence of very low number of zinc importers in most bacterial species mark zinc homeostasis as a very promising target for the development of novel antimicrobial strategies.
This content is subject to copyright.
MINI REVIEW ARTICLE
published: 24 December 2013
doi: 10.3389/fcimb.2013.00108
Competition for zinc binding in the host-pathogen
interaction
Mauro Cerasi1, Serena Ammendola 1and Andrea Battistoni1,2*
1Dipartimento di Biologia, Università di Roma Tor Vergata, Rome, Italy
2Istituto Nazionale Biostrutture e Biosistemi, Consorzio Interuniversitario, Rome, Italy
Edited by:
Frédéric J. Veyrier, Institut Pasteur,
France
Reviewed by:
Klaus Hantke, Universität Tübingen,
Germany
Robert D. Perry, University of
Kentucky, USA
*Correspondence:
Andrea Battistoni, Via della Ricerca
Scientifica, Dipartimento di Biologia,
Room 372, Università di Roma Tor
Vergata, Rome 00133, Italy
e-mail: andrea.battistoni@
uniroma2.it
Due to its favorable chemical properties, zinc is used as a structural or catalytic cofactor
in a very large number of proteins. Despite the apparent abundance of this metal in
all cell types, the intracellular pool of loosely bound zinc ions available for biological
exchanges is in the picomolar range and nearly all zinc is tightly bound to proteins.
In addition, to limit bacterial growth, some zinc-sequestering proteins are produced by
eukaryotic hosts in response to infections. Therefore, to grow and multiply in the infected
host, bacterial pathogens must produce high affinity zinc importers, such as the ZnuABC
transporter which is present in most Gram-negative bacteria. Studies carried in different
bacterial species have established that disruption of ZnuABC is usually associated with a
remarkable loss of pathogenicity. The critical involvement of zinc in a plethora of metabolic
and virulence pathways and the presence of very low number of zinc importers in most
bacterial species mark zinc homeostasis as a very promising target for the development
of novel antimicrobial strategies.
Keywords: zinc uptake, ZnuABC, antibacterial therapies, metal cofactor, host-pathogen interaction, Salmonella
enterica, zinc transporter, nutritional immunity
ZINC: CHEMICAL PROPERTIES AND ROLE IN BACTERIAL
PROTEINS
Among transition metals, zinc is likely the one which is used as
a structural or catalytic cofactor in the wider number of pro-
teins. The widespread use of zinc in proteins can be related to
its peculiar chemical properties (Andreini et al., 2008). Unlike the
other biological relevant transition metals (Fe2+,Mn
2+,Cu
2+,
Ni2+) the zinc ion (Zn2+)hasafilleddorbital and, therefore,
it is redox stable. Zinc mainly participates to catalytic reactions
by acting as a Lewis acid able to accept electron pairs or, as an
alternative, by attracting or stabilizing negative charges of the
substrates. Moreover, zinc binding to proteins is facilitated by its
capability to form stable chemical bonds with nitrogen, oxygen
and sulfur atoms and assume different coordination numbers.
As a consequence zinc can be found in a large variety of dis-
tinct chemical environments, which may significantly modulate
its reactivity. However, a potential problem of zinc is that it binds
to proteins stronger than the other divalent metals (Irving and
Williams, 1948) and, therefore, cells maintain the intracellular
pool of “free” metal at very low levels to prevent its unspecific
binding to proteins (Colvin et al., 2010).
Different studies have attempted to measure the amount of
zinc in bacteria. It has been shown that microorganisms have a
remarkable capability to modify their intracellular zinc content
in response to variations in the environmental availability of the
metal (Outten and O’Halloran, 2001; Garmory and Titball, 2004)
and that the total cellular zinc in bacteria growing in rich media is
in the submillimolar range (104M), i.e., a concentration com-
parable to that usually observed in most eukaryotic cells (Eide,
2006). More complex is to obtain a careful evaluation of the intra-
cellular pool of metal ions not tightly bound to proteins. In vitro
studies carried out with purified zinc-responding transcriptional
regulators have initially suggested that cellular “free” zinc levels
are in the femtomolar range, i.e., around 1015 M(Outten and
O’Halloran, 2001). However, recent studies involving protein-
based ratiometric biosensors have established that in vivo the
concentration of intracellular exchangeable zinc is around 20pM,
i.e., 2 ×1011M(Wang et al., 2011). Picomolar values of “free”
zinc have been reported also in several eukaryotic systems (Colvin
et al., 2010).
Interestingly, although the zinc concentration in bacterial cells
is close to that of iron, a significant fraction of iron may be found
in association to proteins such as ferritins, bacterioferritins or
DPS (Andrews et al., 2003), whereas bona fide zinc-storage pro-
teins are present only in a few bacteria (Blindauer et al., 2002).
An experimental attempt to explore the complexity of the bacte-
rial zinc proteome has shown that more than 3% of the proteins
expressed in Escherichia coli contain zinc (Katayama et al., 2002),
whereas bioinformatics investigations have revealed that about
5% of all bacterial proteins contain recognizable zinc-binding
sites (Andreini et al., 2006). This means that an E. coli cell with
about 4300 protein-encoding genes contains more than 200 zinc-
binding proteins. These figures, however, are not sufficient to have
an accurate idea of the actual importance of this metal in the
physiology of a bacterial cell. In fact, in addition to being an essen-
tial cofactor in a large number of enzymes involved in central
metabolic pathways, zinc is bound to several proteins involved
in the management of gene expression, including some riboso-
mal proteins (Hensley et al., 2011), RNA polymerases (Scrutton
et al., 1971), tRNA synthetases (Miller et al., 1991), sigma factor
interacting proteins (Campbell et al., 2007) and zinc responding
transcriptional factors (Chivers, 2007). Moreover, zinc is involved
Frontiers in Cellular and Infection Microbiology www.frontiersin.org December 2013 | Volume 3 | Article 108 |1
CELLULA R AND INFECTION MICROBIOLOG
Y
Cerasi et al. Zinc and bacterial virulence
in other crucial processes, including DNA repair (Kropachev
et al., 2006), response to oxidative stress (Ortiz De Orue Lucana
et al., 2012), antibiotic resistance (Meini et al., 2013)andpro-
duction of virulence-related proteins (Ammendola et al., 2008).
It follows that changes in the intracellular concentrations of zinc
can have pleiotropic effects on the composition of the bacte-
rial proteome, involving changes in the expression and activity
of zinc-containing proteins as well as of proteins which do not
employ this cofactor.
BACTERIAL ZINC UPTAKE SYSTEMS AND RESPONSE TO
ZINC SHORTAGE
Although in bacteria exposed to high levels of zinc the metal
may enter through a large number of unspecific channels, only
a few metal transporters are known to mediate the specific uptake
of zinc (Hantke, 2005)(Figure 1). Some recent studies on the
pneumococcal PsaA protein involved in manganese uptake have
provided interesting hints to understand the mechanisms ensur-
ing specificity in transition metal import (McDevitt et al., 2011;
Counago et al., 2013). PsaA may bind either manganese or other
first-row transition metals, but, due to the propensity of zinc to
form stable complexes with proteins (Irving and Williams, 1948)
it competitively affects Mn2+binding and locks the protein in a
conformation which prevents the entry either of zinc or of man-
ganese. This observation provides an explanation for the ability
of zinc to inhibit pneumococcal growth (McDevitt et al., 2011)
and an elegant example of the strategies used by living cells to
guarantee the correct uptake of specific metal ions (Wal dro n and
Robinson, 2009).
In several Gram-negative bacteria growing in metal replete
conditions, zinc uptake is thought to be primarily mediated by
ZupT, a constitutively expressed low affinity transporter belong-
ing to the ZIP (ZRT-, IRT-like Protein) protein family (Grass
et al., 2002). This metal permease has a broad metal specificity,
but it displays a clear preference for zinc over other divalent met-
als (Grass et al., 2005). ZupT depends on the proton motive force
to energize zinc import (Karlinsey et al., 2010; Taudte and Grass,
2010).
The response to zinc paucity is controlled through the coordi-
nated expression of a set of genes regulated by the transcriptional
factor Zur, which may bind two or more zinc ions, depending on
the species (Outten et al., 2001; Lucarelli et al., 2007; Shin et al.,
2011). One atom of zinc serves a structural role, whereas the other
atom(s) favors the folding of a DNA-binding domain enabling
the protein to tightly bind to a consensus sequence located in
the promoter of said genes. In contrast, when the intracellular
zinc content decreases, zinc-devoid Zur is no longer able to sta-
bly interact with DNA and to repress transcription. The number
of known Zur–regulated genes changes in different bacteria, but
in all species they include a small operon encoding for the com-
ponents of ZnuABC, a high affinity zinc importer of the ABC
family, and one or more genes encoding for paralogs of zinc-
containing ribosomal proteins (Panina, 2003; Graham et al., 2009;
Li et al., 2009; Lim et al., 2013). The ZnuABC uptake system is
composed of three proteins: the ZnuB channel, the ZnuC ATPase
component which provides the energy necessary for ion trans-
port through the inner membrane, and ZnuA, a soluble protein
which captures Zn(II) in the periplasm with high efficiency and
delivers it to ZnuB (Patzer and Hantke, 1998). In some bacteria
there is also an accessory component of the ZnuABC transporter,
ZinT, which is known to form a complex with ZnuA in pres-
ence of zinc and is thought to enhance ZnuA ability to recruit
zinc (Petrarca et al., 2010; Ilari et al., 2013). A similar zinc uptake
system (AdcABC) can be found in pneumococci and some other
Gram-positive bacteria, where the lipoprotein AdcA is charac-
terized by two structural domains that show clear sequence and
structural homology with ZnuA and ZinT, respectively (Dintilhac
et al., 1997; Panina et al., 2003).
An interesting facet of the Zur-mediated response to zinc
shortage is the substitution of ribosomal proteins containing zinc
with homologous proteins lacking the zinc-binding motif. This
change in ribosomal structure reduces the metal requirements
of bacterial cells, as the majority of intracellular zinc is thought
to be associated to ribosomes (Hensley et al., 2011). Moreover,
the production of zinc-independent ribosomal proteins may be
useful to mobilize a relevant amount of metal from pre-existing
ribosomes and facilitates the adaptation to zinc-limiting condi-
tions (Gabriel and Helmann, 2009). From this point of view, the
ribosome may be described as a zinc storage protein complex.
Additional paralogs of zinc-containing proteins have been identi-
fied in several bacteria (Haas et al., 2009) and include a homolog
of the transcriptional factor DksA which is involved in the con-
trol of the bacterial response to stress and starvation (Blaby-Haas
et al., 2011).
The outer membrane of Gram-negative bacteria allows the
passive diffusion of low molecular weight molecules. However, a
mechanism of nutrient uptake solely based on diffusion may be
hardly able to ensure the adequate absorption of elements which
are poorly available in the environment. In recent years, an outer
membrane TonB-dependent receptor involved in zinc uptake has
been identified in Neisseria meningitidis and some other Gram-
negative bacteria (Stork et al., 2010). This protein, denominated
ZnuD, mediates either zinc or heme uptake and is regulated either
by Zur or by Fur (Kumar et al., 2012; Pawlik et al., 2012). The
pneumococcal surface protein PhtD has been proposed to play a
functionally similar role in favoring zinc uptake through AdcAII
(Loisel et al., 2011).Nooutermembranezincreceptorshave
been so far identified in Enterobacteria or in other Gram-negative
bacteria. However, it has been observed that apo-ZinT can be
extruded outside the cell, suggesting that it could have some role
in the acquisition of zinc from the environment (Ho et al., 2008;
Gabbianelli et al., 2011).
It should also be noted that a few bacterial species have been
shown to express more than one high affinity zinc uptake sys-
tems. This is the case of Listeria monocytogenes which expresses
two ABC-type zinc importers (ZnuABC and ZurAM), both con-
tributing to full virulence (Corbett et al., 2012) and of non-
typeable Haemophilus influenzae, where the zinc binding system
ZevAB facilitates growth in zinc-limiting conditions and lung
colonization in infected mice (Rosadini et al., 2011). Similarly,
disruption of znuA in Pseudomonas aeruginosa results in a very
limited growth defect under zinc-limiting conditions (Ellison
et al., 2013), possibly due to the expression of a zinc-importing
P-t yp e ATPase (Lewinson et al., 2009).
Frontiers in Cellular and Infection Microbiology www.frontiersin.org December 2013 | Volume 3 | Article 108 |2
Cerasi et al. Zinc and bacterial virulence
FIGURE 1 | Schematic diagram of transporters involved in zinc uptake.
The bacterial outer membrane is thought to be permeable to hydrophilic
solutes of <600 dalton (Nikaido and Vaara, 1985) and, therefore, zinc
concentration in the periplasmic space is largely dependent on zinc availability
in the environment. Under zinc replete conditions (left), the metal is imported
through low affinity import systems, such as ZupT, and Zur inhibits the
expression of the importer ZnuABC. Under conditions of zinc shortage (right),
apoZur is unable to bind DNA and the high affinity zinc importer ZnuABC is
expressed. Neisseria meningitidis expresses a Zur-regulated TonB-dependent
outer membrane protein, ZnuD, involved in zinc uptake.
ZINC IN THE HOST-PATHOGEN INTERACTION
Whereas the competition for iron acquisition has been recognized
as a key element of the host pathogen interaction for a long time,
only in recent years the efficient uptake of other divalent metals
has emerged to play a comparable role (Kehl-Fie and Skaar, 2010).
In particular the importance of zinc has become clear through
a series of investigations which have established that deletion of
the znuABC genes not only decrease bacterial ability to growth
in in vitro environments poor of this metal, but also dramati-
cally affects their pathogenicity. Bacterial pathogens which have
been shown to critically depend on ZnuABC to infect their hosts
include Acinetobacter baumannii (Hood et al., 2012), Brucella
abortus (Kim et al., 2004; Yang et al., 2006), Campylobacter
jejuni (Davis et al., 2009), pathogenic E. coli strains (Sabri et al.,
2009; Gabbianelli et al., 2011), H. ducreyi (Lewis et al., 1999),
Moraxella catarrhalis (Murphy et al., 2013), Pasteurella multo-
cida (Garrido et al., 2003), Salmonella enterica (Campoy et al.,
2002; Ammendola et al., 2007)andYersinia ruckeri (Dahiya
and Stevenson, 2010). In contrast, while being required for zinc
uptake in vitro,ZnuABCdoesnotcontributetoY. pes t is virulence
(Desrosiers et al., 2010) and provides only a limited advantage
to Proteus mirabilis during urinary tract infections (Nielubowicz
et al., 2010). It is not yet clear whether these bacteria possess addi-
tional zinc importers or if they show limited zinc requirements
during infections.
Probably, the previous underestimation of the importance of
zinc in the interaction between bacteria and their hosts can be
largely attributed to the apparent abundance of this element in
all tissues. In fact, high levels of zinc are present either within
cells or in the plasma, where most of the metal is loosely associ-
ated to proteins (Zalewski et al., 2006) and, therefore, potentially
available for invading microorganisms. However, it should be
noted that a typical feature of the early response to the infec-
tion is the rapid fall of plasma zinc concentration, accompanied
by zinc accumulation in the liver. Redistribution of zinc among
the various tissues is regulated by a lipopolysaccharide-induced
cytokines cascade (with IL-6 playing a central role) which stim-
ulates increased synthesis of acute phase proteins, such as met-
allothionein, and the hepatic uptake of the metal through the
induction of the solute carrier 39 (SLC39) protein ZIP14 (Liuzzi,
2005). In view of the importance of the ZnuABC transporter in
bacterial zinc uptake during infections, this feature of the acute
phase response appears as an adaptive mechanism intended to
deprive pathogens of zinc.
The role of the ZnuABC transporter has been investigated in
details in S. enterica serovar Typhimurium (S. Typhimurium).
Expression of znuABC is repressed in S. Typhimurium culti-
vated in synthetic media containing zinc concentrations as low
as 1 µM. In contrast, the znuABC operon is strongly induced in
bacteria recovered from the spleens of infected mice or from cul-
tured epithelial or macrophagic cells (Ammendola et al., 2007).
These observations suggest that zinc bound to proteins is not eas-
ily available for invading bacteria and that ZnuABC is required
to have rapid access to the pool of “free” zinc. More recently,
it has been shown that during gut infections ZnuABC confers
resistance to the antimicrobial protein calprotectin (Liu et al.,
2012). Calprotectin is a neutrophilic protein of the S100 fam-
ily of calcium binding proteins, that is abundantly released at
sites of infection to control the multiplication of pathogens by
the sequestration of zinc and manganese (Kehl-Fie and Skaar,
2010). In support to the experimental evidence that calprotectin
starves bacteria for metal ions, structural studies have confirmed
that calprotectin possesses two distinct binding sites for transition
metals, one of which is specific for zinc and the other one may
Frontiers in Cellular and Infection Microbiology www.frontiersin.org December 2013 | Volume 3 | Article 108 |3
Cerasi et al. Zinc and bacterial virulence
accommodate either zinc or manganese (Brunjes Brophy et al.,
2013; Damo et al., 2013). Bacteria expressing ZnuABC are able to
resist to such antimicrobial strategy and this favor their growth
over competing microbes in the inflamed gut (Gielda and Dirita,
2012; Liu et al., 2012). Taken together, these studies suggest that
zinc acquisition through the ZnuABC transporter is essential for
the colonization of Salmonella in mice, and provide a parallelism
between the mechanisms of iron and zinc sequestration in the
host-pathogen relationships. It is worth noting that calprotectin
is not the unique S100 protein involved in zinc sequestration. In
fact, a comparable function has been proposed for the antibacte-
rial protein psoriasin (S100A7), which protects human skin from
E. coli infections (Glaser et al., 2005).
Although, the above mentioned studies have suggested that
zinc sequestration is a strategy widely used by vertebrates to con-
trol microbial infections, a few recent observations have revealed
an alternative way to use zinc in host defense. In fact, it has been
shown that human macrophages control mycobacteria by elevat-
ing zinc levels in the bacteria-containing phagosomes (Botella
et al., 2011). This process is dependent on reactive oxygen species
generated by the phagocytic NADPH oxidase and involves the
mobilization of zinc from intracellular stores. Mycobacterial resis-
tance to zinc intoxication in macrophages relies on their ability to
induce the expression of heavy metal efflux P-type ATPases, which
prevent the intracellular accumulation of zinc at toxic levels.
Structurally homologous zinc efflux pumps have been identi-
fied in a large number of bacteria, including E. coli,wherethe
P-type zinc exporter ZntA has been proved to be critical for
zinc tolerance (Beard et al., 1997; Rensing et al., 1997)andfor
the maintaining of appropriate levels of intracellular zinc (Wang
et al., 2012). Whereas mycobacteria lacking the efflux pump CtpC
or E. coli cells devoid of ZntA display a reduced ability to survive
in human macrophages, disruption of CtpC does not affect the
ability of M. tuberculosis to infect mice (Botella et al., 2011).
To add to the complexity, a mobilization of zinc in the oppo-
site direction to that found in response to mycobacteria has
been observed in murine macrophages infected with the fungus
Histoplasma capsulatum (Subramanian Vignesh et al., 2013). In
this case, phagosomes are deprived of zinc and the metal accu-
mulates in the Golgi or in the cytoplasm, in association to met-
allothioneins. Further studies are needed to understand whether
these different responses depend on the cross-talk between each
specific microorganism and phagocytic cells, or whether the abil-
ity to poison bacteria through an excess of zinc is a prerogative of
human macrophages.
ZINC HOMEOSTASIS AS A TARGET FOR ANTIBACTERIAL
THERAPIES
Different studies have proposed that ABC transporters could be
effective targets for the development of novel antibacterial drugs
or vaccines (Garmory and Titball, 2004; Counago et al., 2012). In
this view, ZnuABC appears as a particularly promising candidate.
Whereas a large number of pathogens produce multiple metal
import systems that enable the uptake of different iron forms
(reduced or oxidized, “free or bound to proteins or to heme),
in most bacteria there is only one high affinity zinc importer.
Moreover, it has been shown that Salmonella strains lacking the
whole znuABC operon display the same dramatic loss of virulence
of strains producing ZnuB and ZnuC, but lacking ZnuA (Petrarca
et al., 2010). This finding indicates that the ability of ZnuA to
effectively compete for zinc binding with other periplasmic pro-
teins is critical to ensure zinc import in the cytoplasm (Berducci
et al., 2004) and suggests that drugs targeting the soluble com-
ponent of the transporter could block zinc import. This is a very
interesting possibility because ZnuA is a suitable substrate for bio-
chemical and structural studies (Banerjee et al., 2003; Chandra
et al., 2007; Wei et al., 2007; Yatsunyk et al., 2008; Ilari et al., 2011;
Castelli et al., 2013). It is worth noting that several molecules
able to interfere with zinc uptake in Candida albicans have been
recently identified through the screening of a small-molecule
library of 2000 compounds (Simm et al., 2011), thus providing
a proof of concept that it is possible to pharmacologically target
zinc homeostasis in pathogenic microorganisms.
Bacterial mutant strains unable to produce the ZnuABC trans-
porter are also putative candidate for the development of live–
attenuated vaccines. It has been shown that a S.Typhimurium
znuABC strain is able to induce short lasting infections in mice
which induce a solid and durable immune-based protection
against virulent strains of S. Ty phi mu r ium ( Pasquali et al., 2008;
Pesciaroli et al., 2011). The same strain proved to be attenuated,
immunogenic and protective also in pigs (Gradassi et al., 2013;
Pesciaroli et al., 2013). Similarly, it has been shown that vacci-
nation with Brucella znuA mutants protects mice from Brucella
infections (Yang et al., 2006; Clapp et al., 2011). In addition, it has
been shown that ZnuD is able to elicit the production of antibod-
ies triggering complement-mediated killing of several Neisseria
meningitidis serogroup B strains, suggesting that it is a promising
candidate for the generation of an effective vaccine (Hubert et al.,
2013). Taken together, these studies suggest that zinc homeosta-
sis offers interesting options to generate vaccines against different
pathogenic bacteria.
REFERENCES
Ammendola, S., Pasquali, P., Pacello, F., Rotilio, G., Castor, M., Libby, S. J., et al.
(2008). Regulatory and structural differences in the Cu,Zn-superoxide dismu-
tases of Salmonella enterica and their significance for virulence. J. Biol. Chem.
283, 13688–13699. doi: 10.1074/jbc.M710499200
Ammendola, S., Pasquali, P., Pistoia, C., Petrucci, P., Petrarca, P., Rotilio, G.,
et al. (2007). High-affinity Zn2+ uptake system ZnuABC is required for bac-
terial zinc homeostasis in intracellular environments and contributes to the
virulence of Salmonella enterica. Infect. Immun. 75, 5867–5876. doi: 10.1128/
IAI.00559-07
Andreini, C., Banci, L., Bertini, I., and Rosato, A. (2006). Zinc through the three
domains of life. J. Proteome Res. 5, 3173–3178. doi: 10.1021/pr0603699
Andreini, C., Bertini, I., Cavallaro, G., Holliday, G. L., and Thornton, J. M. (2008).
Metal ions in biological catalysis: from enzyme databases to general principles.
J. Biol. Inorg. Chem. 13, 1205–1218. doi: 10.1007/s00775-008-0404-5
Andrews, S. C., Robinson, A. K., and Rodríguez-Quiñones, F. (2003). Bacterial
iron homeostasis. FEMS Microbiol. Rev. 27, 215–237. doi: 10.1016/S0168-
6445(03)00055-X
Banerjee, S., Wei, B., Bhattacharyya-Pakrasi, M., Pakrasi, H. B., and Smith, T.
J. (2003). Structural determinants of metal specificity in the zinc transport
protein ZnuA from synechocystis 6803. J. Mol. Biol. 333, 1061–1069. doi:
10.1016/j.jmb.2003.09.008
Beard, S. J., Hashim, R., Membrillo-Hernandez, J., Hughes, M. N., and Poole, R.
K. (1997). Zinc(II) tolerance in Escherichia coli K-12: evidence that the zntA
gene (o732) encodes a cation transport ATPase. Mol. Microbiol. 25, 883–891.
doi: 10.1111/j.1365-2958.1997.mmi518.x
Frontiers in Cellular and Infection Microbiology www.frontiersin.org December 2013 | Volume 3 | Article 108 |4
Cerasi et al. Zinc and bacterial virulence
Berducci, G., Mazzetti, A. P., Rotilio, G., and Battistoni, A. (2004). Periplasmic com-
petition for zinc uptake between the metallochaperone ZnuA and Cu, Zn super-
oxide dismutase. FEBS Lett. 569, 289–292. doi: 10.1016/j.febslet.2004.06.008
Blaby-Haas, C. E., Furman, R., Rodionov, D. A., Artsimovitch, I., and De
Crécy-Lagard, V. (2011). Role of a Zn-independent DksA in Zn homeosta-
sis and stringent response. Mol. Microbiol. 79, 700–715. doi: 10.1111/j.1365-
2958.2010.07475.x
Blindauer, C. A., Harrison, M. D., Robinson, A. K., Parkinson, J. A., Bowness,
P. W., Sadler, P. J., et al. (2002). Multiple bacteria encode metallothioneins
and SmtA-like zinc fingers. Mol. Microbiol. 45, 1421–1432. doi: 10.1046/j.1365-
2958.2002.03109.x
Botella, H., Peyron, P., Levillain, F., Poincloux, R., Poquet, Y., Brandli, I.,
et al. (2011). Mycobacterial p(1)-type ATPases mediate resistance to zinc
poisoning in human macrophages. Cell Host Microbe 10, 248–259. doi:
10.1016/j.chom.2011.08.006
Brunjes Brophy, M., Nakashige, T. G., Gaillard, A., and Nolan, E. M. (2013).
Contributions of the S100A9 C-terminal tail to high-affinity Mn(II) chela-
tion by the host-defense protein human calprotectin. J. Am. Chem. Soc. 135,
17804–17817. doi: 10.1021/ja407147d
Campbell, E. A., Greenwell, R., Anthony, J. R., Wang, S., Lim, L., Das,
K., et al. (2007). A conserved structural module regulates transcriptional
responses to diverse stress signals in bacteria. Mol. Cell 27, 793–805. doi:
10.1016/j.molcel.2007.07.009
Campoy, S., Jara, M., Busquets, N., Perez De Rozas, A. M., Badiola, I., and Barbe,
J. (2002). Role of the high-affinity zinc uptake znuABC system in Salmonella
enterica serovar typhimurium virulence. Infect. Immun. 70, 4721–4725. doi:
10.1128/IAI.70.8.4721-4725.2002
Castelli, S., Stella, L., Petrarca, P., Battistoni, A., Desideri, A., and Falconi, M.
(2013). Zinc ion coordination as a modulating factor of the ZnuA histidine-
rich loop flexibility: a molecular modeling and fluorescence spectroscopy study.
Biochem. Biophys. Res. Commun. 430, 769–773. doi: 10.1016/j.bbrc.2012.11.073
Chandra, B. R., Yogavel, M., and Sharma, A. (2007). Structural analysis of
ABC-family periplasmic zinc binding protein provides new insights into
mechanism of ligand uptake and release. J. Mol. Biol. 367, 970–982. doi:
10.1016/j.jmb.2007.01.041
Chivers, P. T. (2007). A galvanizing story–protein stability and zinc homeostasis.
J. Bacteriol. 189, 2953–2954. doi: 10.1128/JB.00173-07
Clapp, B., Skyberg, J. A., Yang, X., Thornburg, T., Walters, N., and Pascual, D.
W. (2011). Protective live oral brucellosis vaccines stimulate Th1 and th17 cell
responses. Infect. Immun. 79, 4165–4174. doi: 10.1128/IAI.05080-11
Colvin, R. A., Holmes, W. R., Fontaine, C. P., and Maret, W. (2010). Cytosolic zinc
buffering and muffling: their role in intracellular zinc homeostasis. Metallomics
2, 306–317. doi: 10.1039/b926662c
Corbett, D., Wang, J., Schuler, S., Lopez-Castejon, G., Glenn, S., Brough, D., et al.
(2012). Two zinc uptake systems contribute to the full virulence of Listeria
monocytogenes during growth in vitro and in vivo.Infect. Immun. 80, 14–21.
doi: 10.1128/IAI.05904-11
Counago, R. M., McDevitt, C. A., Ween, M. P., and Kobe, B. (2012). Prokaryotic
substrate-binding proteins as targets for antimicrobial therapies. Curr. Dru g
Ta r ge t s 13, 1400–1410. doi: 10.2174/138945012803530170
Counago,R.M.,Ween,M.P.,Begg,S.L.,Bajaj,M.,Zuegg,J.,OMara,M.L.,etal.
(2013). Imperfect coordination chemistry facilitates metal ion release in the Psa
permease. Nat. Chem. Biol.
Dahiya, I., and Stevenson, R. M. (2010). The ZnuABC operon is important for
Yersinia ruckeri infections of rainbow trout, Oncorhynchus mykiss (Walbaum).
J. Fish Dis. 33, 331–340. doi: 10.1111/j.1365-2761.2009.01125.x
Damo, S. M., Kehl-Fie, T. E., Sugitani, N., Holt, M. E., Rathi, S., Murphy, W. J.,
et al. (2013). Molecular basis for manganese sequestration by calprotectin and
roles in the innate immune response to invading bacterial pathogens. Proc. Natl.
Acad. Sci. U.S.A. 110, 3841–3846. doi: 10.1073/pnas.1220341110
Davis, L. M., Kakuda, T., and Dirita, V. J. (2009). A Campylobacter jejuni znuA
orthologue is essential for growth in low-zinc environments and chick colo-
nization. J. Bacteriol. 191, 1631–1640. doi: 10.1128/JB.01394-08
Desrosiers, D. C., Bearden, S. W., Mier, I. Jr., Abney, J., Paulley, J. T., Fetherston, J.
D., et al. (2010). Znu is the predominant zinc importer in Yersinia pestis during
in vitro growth but is not essential for virulence. Infect. Immun. 78, 5163–5177.
doi: 10.1128/IAI.00732-10
Dintilhac, A., Alloing, G., Granadel, C., and Claverys, J. P. (1997). Competence
and virulence of Streptococcus pneumoniae: Adc and PsaA mutants exhibit
a requirement for Zn and Mn resulting from inactivation of putative
ABC metal permeases. Mol. Microbiol. 25, 727–739. doi: 10.1046/j.1365-
2958.1997.5111879.x
Eide, D. J. (2006). Zinc transporters and the cellular trafficking of zinc. Biochim.
Biophys. Acta 1763, 711–722. doi: 10.1016/j.bbamcr.2006.03.005
Ellison, M. L., Farrow, J. M., Parrish, W., Danell, A. S., and Pesci, E. C. (2013).
The Transcriptional regulator Np20 Is the Zinc uptake regulator in Pseudomonas
aeruginosa. PLoS ONE 8:e75389. doi: 10.1371/journal.pone.0075389
Gabbianelli, R., Scotti, R., Ammendola, S., Petrarca, P., Nicolini, L., and Battistoni,
A. (2011). Role of ZnuABC and ZinT in Escherichia coli O157:H7 zinc
acquisition and interaction with epithelial cells. BMC Microbiol. 11:36. doi:
10.1186/1471-2180-11-36
Gabriel, S. E., and Helmann, J. D. (2009). Contributions of Zur-controlled ribo-
somal proteins to growth under zinc starvation conditions. J. Bacteriol. 191,
6116–6122. doi: 10.1128/JB.00802-09
Garmory, H. S., and Titball, R. W. (2004). ATP-binding cassettet ransporters are tar-
gets for the development of antibacterial vaccines and therapies. Infect. Immun.
72, 6757–6763. doi: 10.1128/IAI.72.12.6757-6763.2004
Garrido, M. E., Bosch, M., Medina, R., Llagostera, M., Perez De Rozas, A.
M., Badiola, I., et al. (2003). The high-affinity zinc-uptake system znu-
ACB is under control of the iron-uptake regulator (fur) gene in the ani-
mal pathogenPast eurella m ulto cid a.FEMS Microbiol. Lett. 221, 31–37. doi:
10.1016/S0378-1097(03)00131-9
Gielda, L. M., and Dirita, V. J. (2012). Zinc competition among the intestinal
microbiota. MBio 3, e00171–00112. doi: 10.1128/mBio.00171-12
Glaser, R., Harder, J., Lange, H., Bartels, J., Christophers, E., and Schroder, J. M.
(2005). Antimicrobial psoriasin (S100A7) protects human skin from Escherichia
coli infection. Nat. Immunol. 6, 57–64. doi: 10.1038/ni1142
Gradassi, M., Pesciaroli, M., Martinelli, N., Ruggeri, J., Petrucci, P., Hassan, W.
H., et al. (2013). Attenuated Salmonella enterica serovar Typhimurium
lacking the ZnuABC transporter: an efficacious orally-administered
mucosal vaccine against salmonellosis in pigs. Vaccine 31, 3695–3701.
doi: 10.1016/j.vaccine.2013.05.105
Graham, A. I., Hunt, S., Stokes, S. L., Bramall, N., Bunch, J., Cox, A. G., et al.
(2009). Severe zinc depletion of Escherichia coli: roles for high affinity zinc bind-
ing by ZinT, zinc transport and zinc-independent proteins. J. Biol. Chem. 284,
18377–18389. doi: 10.1074/jbc.M109.001503
Grass, G., Franke, S., Taudte, N., Nies, D. H., Kucharski, L. M., Maguire, M.
E., et al. (2005). The metal permease ZupT from Escherichia coli is a trans-
porter with a broad substrate spectrum. J. Bacteriol. 187, 1604–1611. doi:
10.1128/JB.187.5.1604-1611.2005
Grass, G., Wong, M. D., Rosen, B. P., Smith, R. L., and Rensing, C. (2002). ZupT
is a Zn(II) uptake system in Escherichia coli.J. Bacteriol. 184, 864–866. doi:
10.1128/JB.184.3.864-866.2002
Haas, C. E., Rodionov, D. A., Kropat, J., Malasarn, D., Merchant, S. S., and De
Crecy-Lagard, V. (2009). A subset of the diverse COG0523 family of putative
metal chaperones is linked to zinc homeostasis in all kingdoms of life. BMC
Genomics 10:470. doi: 10.1186/1471-2164-10-470
Hantke, K. (2005). Bacterial zinc uptake and regulators. Curr. Opin. Microbiol. 8,
196–202. doi: 10.1016/j.mib.2005.02.001
Hensley, M. P., Tierney, D. L., and Crowder, M. W. (2011). Zn(II) bind-
ing to Escherichia coli 70S ribosomes. Biochemistry 50, 9937–9939. doi:
10.1021/bi200619w
Ho, T. D., Davis, B. M., Ritchie, J. M., and Waldor, M. K. (2008). Type
2 secretion promotes enterohemorrhagic Escherichia coli adherence and
intestinal colonization. Infect. Immun. 76, 1858–1865. doi: 10.1128/IAI.
01688-07
Hood, M. I., Mortensen, B. L., Moore, J. L., Zhang, Y., Kehl-Fie, T. E., Sugitani, N.,
et al. (2012). Identification of an Acinetobacter baumannii zinc acquisition sys-
tem that facilitates resistance to calprotectin-mediated zinc sequestration. PLoS
Path og. 8:e1003068. doi: 10.1371/journal.ppat.1003068
Hubert, K., Devos, N., Mordhorst, I., Tans, C., Baudoux, G., Feron, C.,
et al. (2013). ZnuD, a potential candidate for a simple and universal
Neisseria meningitidis vaccine. Infect. Immun. 81, 1915–1927. doi: 10.1128/IAI.
01312-12
Ilari, A., Alaleona, F., Petrarca, P., Battistoni, A., and Chiancone, E. (2011). The X-
ray structure of the zinc transporter ZnuA from Salmonella enterica discloses
a unique triad of zinc-coordinating histidines. J. Mol. Biol. 409, 630–641. doi:
10.1016/j.jmb.2011.04.036
Frontiers in Cellular and Infection Microbiology www.frontiersin.org December 2013 | Volume 3 | Article 108 |5
Cerasi et al. Zinc and bacterial virulence
Ilari, A., Alaleona, F., Tria, G., Petrarca, P., Battistoni, A., Zamparelli, C., et al.
(2013). The Salmonella enterica ZinT structure, zinc affinity and interac-
tion with the high-affinity uptake protein ZnuA provide insight into the
management of periplasmic zinc. Biochim. Biophys. Acta 1840, 535–544. doi:
10.1016/j.bbagen.2013.10.010
Irving, H., and Williams, R. J. P. (1948). Order of stability of metal complexes.
Nature 162, 746–747. doi: 10.1038/162746a0
Karlinsey, J. E., Maguire, M. E., Becker, L. A., Crouch, M. L., and Fang, F. C.
(2010). The phage shock protein PspA facilitates divalent metal transport and is
required for virulence of Salmonella enterica sv. Typhimurium. Mol. Microbiol.
78, 669–685. doi: 10.1111/j.1365-2958.2010.07357.x
Katayama, A., Tsujii, A., Wada, A., Nishino, T., and Ishihama, A. (2002). Systematic
search for zinc-binding proteins in Escherichia coli.Eur. J. Biochem. 269,
2403–2413. doi: 10.1046/j.1432-1033.2002.02900.x
Kehl-Fie, T. E., and Skaar, E. P. (2010). Nutritional immunity beyond iron:
a role for manganese and zinc. Curr. Opin. Chem. Biol. 14, 218–224. doi:
10.1016/j.cbpa.2009.11.008
Kim, S., Watanabe, K., Shirahata, T., and Watarai, M. (2004). Zinc uptake sys-
tem (znuA locus) of Brucella abortus is essential for intracellular survival and
virulence in mice. J. Vet. Med. Sci. 66, 1059–1063. doi: 10.1292/jvms.66.1059
Kropachev, K. Y., Zharkov, D. O., and Grollman, A. P. (2006). Catalytic mechanism
of Escherichia coli endonuclease VIII: roles of the intercalation loop and the zinc
finger. Biochemistry 45, 12039–12049. doi: 10.1021/bi060663e
Kumar, P., Sannigrahi, S., and Tzeng, Y. L. (2012). The Neisseria meningitidis ZnuD
zinc receptor contributes to interactions with epithelial cells and supports heme
utilization when expressed in Escherichia coli.Infect. Immun. 80, 657–667. doi:
10.1128/IAI.05208-11
Lewinson, O., Lee, A. T., and Rees, D. C. (2009). A P-type ATPase importer that
discriminates between essential and toxic transition metals. Proc. Natl. Acad.
Sci. U.S.A. 106, 4677–4682. doi: 10.1073/pnas.0900666106
Lewis, D. A., Klesney-Tait, J., Lumbley, S. R., Ward, C. K., Latimer, J. L., Ison, C.
A., et al. (1999). Identification of the znuA-encoded periplasmic zinc transport
protein of Haemophilus ducreyi.Infect Immun 67, 5060–5068.
Li, Y., Qiu, Y., Gao, H., Guo, Z., Han, Y., Song, 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
Lim, C. K., Hassan, K. A., Penesyan, A., Loper, J. E., and Paulsen, I. T. (2013). The
effect of zinc limitation on the transcriptome of Pseudomonas protegens Pf-5.
Environ. Microbiol. 15, 702–715. doi: 10.1111/j.1462-2920.2012.02849.x
Liu, J. Z., Jellbauer, S., Poe, A. J., Ton, V., Pesciaroli, M., Kehl-Fie, T. E., et al.
(2012). Zinc sequestration by the neutrophil protein calprotectin enhances
Salmonella growth in the inflamed gut. Cell Host Microbe 11, 227–239. doi:
10.1016/j.chom.2012.01.017
Liuzzi, J. P. (2005). Interleukin-6 regulates the zinc transporter Zip14 in liver and
contributes to the hypozincemia of the acute-phase response. Proc. Natl. Acad.
Sci. U.S.A. 102, 6843–6848. doi: 10.1073/pnas.0502257102
Loisel, E., Chimalapati, S., Bougault, C., Imberty, A., Gallet, B., Di Guilmi, A.
M., et al. (2011). Biochemical characterization of the histidine triad protein
PhtD as a cell surface zinc-binding protein of pneumococcus. Biochemistry 50,
3551–3558. doi: 10.1021/bi200012f
Lucarelli, D., Russo, S., Garman, E., Milano, A., Meyer-Klaucke, W., and Pohl,
E. (2007). Crystal Structure and function of the zinc uptake regulator FurB
from Mycobacterium tuberculosis.J. of Biol. Chem. 282, 9914–9922. doi:
10.1074/jbc.M609974200
McDevitt, C. A., Ogunniyi, A. D., Valkov, E., Lawrence, M. C., Kobe, B., McEwan,
A. G., et al. (2011). A molecular mechanism for bacterial susceptibility to zinc.
PLoS Pathog. 7:e1002357. doi: 10.1371/journal.ppat.1002357
Meini, M. R., Gonzalez, L. J., and Vila, A. J. (2013). Antibiotic resistance in Zn(II)-
deficient environments: metallo-beta-lactamase activation in the periplasm.
Future Microbiol. 8, 947–979. doi: 10.2217/fmb.13.34
Miller, W. T., Hill, K.a.W., and Schimmel, P. (1991). Evidence for a cysteine-
histidine box” metal-binding site in an Escherichia coli aminoacyl-tRNA syn-
thetase. Biochemistry 30, 6970–6976. doi: 10.1021/bi00242a023
Murphy, T. F., Brauer, A. L., Kirkham, C., Johnson, A., Koszelak-Rosenblum, M.,
and Malkowski, M. G. (2013). Role of the zinc uptake ABC transporter of
Moraxella catarrhalis in persistence in the respiratory tract. Infect. Immun. 81,
3406–3413. doi: 10.1128/IAI.00589-13
Nielubowicz, G. R., Smith, S. N., and Mobley, H. L. (2010). Zinc uptake con-
tributes to motility and provides a competitive advantage to Proteus mirabilis
during experimental urinary tract infection. Infect. Immun. 78, 2823–2833. doi:
10.1128/IAI.01220-09
Nikaido, H., and Vaara, M. (1985). Molecular basis of bacterial outer membrane
permeability. Microbiol. Rev. 49, 1–32.
Ortiz De Orue Lucana, D., Wedderhoff, I., and Groves, M. R. (2012). ROS-
mediated signalling in bacteria: zinc-containing Cys-X-X-Cys redox cen-
tres and iron-based oxidative stress. J. Signal Transduct. 2012, 605905. doi:
10.1155/2012/605905
Outten, C. E., and O’Halloran, T. V. (2001). Femtomolar sensitivity of metal-
loregulatory proteins controlling zinc homeostasis. Science 292, 2488–2492. doi:
10.1126/science.1060331
Outten, C. E., Tobin, D. A., Penner-Hahn, J. E., and O’Halloran, T. V. (2001).
Characterization of the metal receptor sites in Escherichia coli Zur, an ultra-
sensitive zinc(II) metalloregulatory protein. Biochemistry 40, 10417–10423. doi:
10.1021/bi0155448
Panina, E. M. (2003). Comparative genomics of bacterial zinc regulons: enhanced
ion transport, pathogenesis, and rearrangement of ribosomal proteins. Proc.
Natl. Acad. Sci. U.S.A. 100, 9912–9917. doi: 10.1073/pnas.1733691100
Panina, E. M., Mironov, A. A., and Gelfand, M. S. (2003). Comparative genomics
of bacterial zinc regulons: enhanced ion transport, pathogenesis, and rearrange-
ment of ribosomal proteins. Proc. Natl. Acad. Sci. U.S.A. 100, 9912–9917. doi:
10.1073/pnas.1733691100
Pasquali, P., Ammendola, S., Pistoia, C., Petrucci, P., Tarantino, M., Valente, C.,
et al. (2008). Attenuated Salmonella enterica serovar Typhimurium lacking
the ZnuABC transporter confers immune-based protection against challenge
infections in mice. Vaccine 26, 3421–3426. doi: 10.1016/j.vaccine.2008.04.036
Patzer, S. I., and Hantke, K. (1998). The ZnuABC high-affinity zinc uptake sys-
tem and its regulator Zur in Escherichia coli.Mol. Microbiol. 28, 1199–1210. doi:
10.1046/j.1365-2958.1998.00883.x
Pawlik, M. C., Hubert, K., Joseph, B., Claus, H., Schoen, C., and Vogel, U. (2012).
The zinc-responsive regulon of Neisseria meningitidis comprises 17 genes
under control of a Zur element. J. Bacteriol. 194, 6594–6603. doi: 10.1128/JB.
01091-12
Pesciaroli, M., Aloisio, F., Ammendola, S., Pistoia, C., Petrucci, P., Tarantino,
M., et al. (2011). An attenuated Salmonella enterica serovar Typhimurium
strain lacking the ZnuABC transporter induces protection in a mouse
intestinal model of Salmonella infection. Vaccine 29, 1783–1790. doi:
10.1016/j.vaccine.2010.12.111
Pesciaroli, M., Gradassi, M., Martinelli, N., Ruggeri, J., Pistoia, C., Raffatellu,
M., et al. (2013). Salmonella Typhimurium lacking the Znuabc trans-
porter is attenuated and immunogenic in pigs. Vaccine 31, 2868–2873. doi:
10.1016/j.vaccine.2013.04.032
Petrarca, P., Ammendola, S., Pasquali, P., and Battistoni, A. (2010). The Zur-
regulated ZinT protein is an auxiliary component of the high-affinity ZnuABC
zinc transporter that facilitates metal recruitment during severe zinc shortage.
J. Bacteriol. 192, 1553–1564. doi: 10.1128/JB.01310-09
Rensing, C., Mitra, B., and Rosen, B. P. (1997). The zntA gene of Escherichia coli
encodes a Zn(II)-translocating P-type ATPase. Proc. Natl. Acad. Sci. U.S.A. 94,
14326–14331. doi: 10.1073/pnas.94.26.14326
Rosadini, C. V., Gawronski, J. D., Raimunda, D., Arguello, J. M., and Akerley, B.
J. (2011). A novel zinc binding system, ZevAB, is critical for survival of nonty-
peable Haemophilus influenzae in a murine lung infection model. Infect . Immun.
79, 3366–3376. doi: 10.1128/IAI.05135-11
Sabri, M., Houle, S., and Dozois, C. M. (2009). Roles of the extraintestinal
pathogenic Escherichia coli ZnuACB and ZupT zinc transporters during urinary
tract infection. Infect. Immun. 77, 1155–1164. doi: 10.1128/IAI.01082-08
Scrutton, M. C., Wu, C. W., and Goldthwait, D. A. (1971). The presence and pos-
sible role of zinc in RNA polymerase obtained from Escherichia coli.Proc. Natl.
Acad. Sci. U.S.A. 68, 2497–2501. doi: 10.1073/pnas.68.10.2497
Shin, J. H., Jung, H. J., An, Y. J., Cho, Y. B., Cha, S. S., and Roe, J. H. (2011). Graded
expression of zinc-responsive genes through two regulatory zinc-binding sites in
Zur. Proc. Natl. Acad.Sc i. U.S.A. 108, 5045–5050. doi: 10.1073/pnas.1017744108
Simm, C., Luan, C. H., Weiss, E., and O’Halloran, T. (2011). High-throughput
screen for identifying small molecules that target fungal zinc homeostasis. PLoS
ONE 6:e25136. doi: 10.1371/journal.pone.0025136
Stork, M., Bos, M. P., Jongerius, I., De Kok, N., Schilders, I., Weynants, V. E., et al.
(2010). An outer membrane receptor of Neisseria meningitidis involved in zinc
acquisition with vaccine potential. PLoS Pathog. 6:e1000969. doi: 10.1371/jour-
nal.ppat.1000969
Frontiers in Cellular and Infection Microbiology www.frontiersin.org December 2013 | Volume 3 | Article 108 |6
Cerasi et al. Zinc and bacterial virulence
Subramanian Vignesh, K., Landero Figueroa, J. A., Porollo, A., Caruso, J.
A., and Deepe, G. S. Jr. (2013). Granulocyte macrophage-colony stim-
ulating factor induced zn sequestration enhances macrophage superox-
ide and limits intracellular pathogen survival. Immunity 39, 697–710. doi:
10.1016/j.immuni.2013.09.006
Taudte, N., and Grass, G. (2010). Point mutations change specificity and kinet-
ics of metal uptake by ZupT from Escherichia coli.Biometals 23, 643–656. doi:
10.1007/s10534-010-9319-z
Waldron, K. J., and Robinson, N. J. (2009). How do bacterial cells ensure that
metalloproteins get the correct metal? Nat. Rev. Microbiol. 7, 25–35. doi:
10.1038/nrmicro2057
Wang, D., Hosteen, O., and Fierke, C. A. (2012). ZntR-mediated transcription
of zntA responds to nanomolar intracellular free zinc. J. Inorg. Biochem. 111,
173–181. doi: 10.1016/j.jinorgbio.2012.02.008
Wang, D., Hurst, T. K., Thompson, R. B., and Fierke, C. A. (2011). Genetically
encoded ratiometric biosensors to measure intracellular exchangeable zinc in
Escherichia coli.J. Biomed. Opt. 16, 087011. doi: 10.1117/1.3613926
Wei, B., Randich, A. M., Bhattacharyya-Pakrasi, M., Pakrasi, H. B., and Smith, T. J.
(2007). Possible regulatory role for the histidine-rich loop in the zinc transport
protein, ZnuA. Biochemistry 46, 8734–8743. doi: 10.1021/bi700763w
Yang, X., Becker, T., Walters, N., and Pascual, D. W. (2006). Deletion of znuA viru-
lence factor attenuates Brucella abortus and confers protection against wild-type
challenge. Infect. Immun. 74, 3874–3879. doi: 10.1128/IAI.01957-05
Yatsunyk, L. A., Easton, J. A., Kim, L. R., Sugarbaker, S. A., Bennett, B., Breece, R.
M., et al. (2008). Structure and metal binding properties of ZnuA, a periplasmic
zinc transporter from Escherichia coli.J. Biol. Inorg. Chem. 13, 271–288. doi:
10.1007/s00775-007-0320-0
Zalewski, P., Truong-Tran, A., Lincoln, S., Ward, D., Shankar, A., Coyle, P.,
et al. (2006). Use of a zinc fluorophore to measure labile pools of zinc in
body fluids and cell-conditioned media. BioTechniques 40, 509–520. doi:
10.2144/06404RR02
Conflict of Interest Statement: The authors declare that the research was con-
ducted in the absence of any commercial or financial relationships that could be
construed as a potential conflict of interest.
Received: 01 October 2013; accepted: 11 December 2013; published online: 24
December 2013.
Citation: Cerasi M, Ammendola S and Battistoni A (2013) Competition for zinc bind-
ing in the host-pathogen interaction. Front. Cell. Infect. Microbiol. 3:108. doi: 10.3389/
fcimb.2013.00108
This article was submitted to the journal Frontiers in Cellular and Infection
Microbiology.
Copyright © 2013 Cerasi, Ammendola and Battistoni. This is an open-access
article distributed under the terms of the Creative Commons Attribution License
(CC BY). The use, distribution or reproduction in other forums is permitted, pro-
vided the original author(s) or licensor are credited and that the original publi-
cation in this journal is cited, in accordance with accepted academic practice. No
use, distribution or reproduction is permitted which does not comply with these
terms.
Frontiers in Cellular and Infection Microbiology www.frontiersin.org December 2013 | Volume 3 | Article 108 |7
... Pathogenic bacteria could exploit the interaction with Zn 2+ because Zn 2+ role in organism is difficult to overestimate 20 . Zn 2+ is an essential part of the immune system of mammalians, in particular in action against bacterial infections 18,21,22 . Zn 2+ affects multiple aspects of immunity -both innate and adaptive 23 , Zn 2+ deficiency in aging is involved in the shift of immune cells balance 24 . ...
... Zinc is a major element necessary for the function of all cells and is the second most abundant transition metal in humans. Apparently, it plays crucial roles in many facets of the immune system 18,[21][22][23] . Zinc is also essential for the growth of pathogenic microorganisms and is involved in the regulation of various virulence factors. ...
Article
Full-text available
Proteorhodopsins (PRs), bacterial light-driven outward proton pumps comprise the first discovered and largest family of rhodopsins, they play a significant role in life on the Earth. A big remaining mystery was that up-to-date there was no described bacterial rhodopsins pumping protons at acidic pH despite the fact that bacteria live in different pH environment. Here we describe conceptually new bacterial rhodopsins which are operating as outward proton pumps at acidic pH. A comprehensive function-structure study of a representative of a new clade of proton pumping rhodopsins which we name “mirror proteorhodopsins”, from Sphingomonas paucimobilis (SpaR) shows cavity/gate architecture of the proton translocation pathway rather resembling channelrhodopsins than the known rhodopsin proton pumps. Another unique property of mirror proteorhodopsins is that proton pumping is inhibited by a millimolar concentration of zinc. We also show that mirror proteorhodopsins are extensively represented in opportunistic multidrug resistant human pathogens, plant growth-promoting and zinc solubilizing bacteria. They may be of optogenetic interest.
... This homeostasis is crucial not only for the expression of metallozymes (Hood and Skaar, 2012) and other proteins related to bacterial metabolism but also for the adequate expression of virulent factors to cause infection (Waldron and Robinson, 2009;Vickers, 2017). Among micronutrients like copper, zinc, manganese, and iron, maintaining the intracellular and extracellular levels of zinc is of utmost importance for the structural and catalytic regulation of various bacterial proteins involved in processes like DNA replication and oxidative stress response (Cerasi et al., 2013). This regulation is achieved by zinc efflux and influx transporters like P1Btype ATPase ZntA and ZnuABC in E. coli, respectively (Hazan et al., 2001;Wei and Fu, 2006;Romiguier and Roux, 2017). ...
Article
Full-text available
Prokaryotic deacetylases are classified into nicotinamide adenine dinucleotide (NAD⁺)-dependent sirtuins and Zn²⁺-dependent deacetylases. NAD⁺ is a coenzyme for redox reactions, thus serving as an essential component for energy metabolism. The NAD⁺-dependent deacetylase domain is quite conserved and well characterized across bacterial species like CobB in Escherichia coli and Salmonella, Rv1151c in Mycobacterium, and SirtN in Bacillus subtilis. E. coli CobB is the only bacterial deacetylase with a known crystal structure (PDB ID: 1S5P), which has 91% sequence similarity with Salmonella CobB (SeCobB). Salmonella encodes two CobB isoforms, SeCobBS and SeCobBL, with a difference of 37 amino acids in its N-terminal domain (NTD). The hydrophobic nature of NTD leads to the stable oligomerization of SeCobBL. The homology modeling-based predicted structure of SeCobB showed the presence of a zinc-binding motif of unknown function. Tryptophan fluorescence quenching induced by ZnCl2 showed that Zn²⁺ has a weak interaction with SeCobBS but higher binding affinity toward SeCobBL, which clearly demonstrated the crucial role of NTD in Zn²⁺ binding. In the presence of Zn²⁺, both isoforms had significantly reduced thermal stability, and a greater effect was observed on SeCobBL. Dynamic light scattering (DLS) studies reflected a ninefold increase in the scattering intensity of SeCobBL upon ZnCl2 addition in contrast to an ∼onefold change in the case of SeCobBS, indicating that the Zn²⁺ interaction leads to the formation of large particles of SeCobBL. An in vitro lysine deacetylase assay showed that SeCobB deacetylated mammalian histones, which can be inhibited in the presence of 0.25–1.00 mM ZnCl2. Taken together, our data conclusively showed that Zn²⁺ strongly binds to SeCobBL through the NTD that drastically alters its stability, oligomeric status, and enzymatic activity in vitro.
... The latter calibrate intracellular Zn concentrations by binding and shuttling Zn into cellular organelles (4,6,7). Thus, the microbe and host battle for Zn to survive (8)(9)(10)(11)(12)(13)(14). ...
Article
Full-text available
Histoplasma capsulatum is a dimorphic fungal pathogen acquired via inhalation of soil-resident spores. Upon exposure to mammalian body temperatures, these fungal elements transform into yeasts that reside primarily within phagocytes. Macrophages (MΦ) provide a permissive environment for fungal replication until T cell-dependent immunity is engaged. MΦ activated by granulocyte macrophage colony stimulating factor (GM-CSF) induces metallothioneins (MTs) that bind zinc (Zn) and deprive yeast cells of labile Zn, thereby disabling fungal growth. Prior work demonstrated that the zinc transporter, ZRT2 , was important for fungal survival in vivo . Hence, we constructed a yeast cell reporter strain that expresses green fluorescent protein (GFP) under control of the ZRT2 zinc-regulated promoter. This reporter accurately responds to a medium devoid of Zn. ZRT2 expression increased in GM-CSF, but not interferon-γ, stimulated MΦ. To examine the in vivo response, we infected mice with a reporter yeast strain and assessed ZRT2 expression at 0, 3, 7, and 14 days post-infection (dpi). ZRT2 expression minimally increased at 3 dpi and peaked at 7 dpi, corresponding with the onset of adaptive immunity. We discovered that the major MΦ populations that restrict Zn from the fungus are interstitial MΦ and exudate MΦ. Neutralizing GM-CSF blunted the control of infection but unexpectedly increased ZRT2 expression. This increase was dependent on another cytokine that activates MΦ to control H. capsulatum replication, M-CSF. These findings illustrate the reporter’s ability to sense Zn in vitro and in vivo and correlate ZRT2 expression with GM-CSF and M-CSF activation of MΦ. IMPORTANCE Phagocytes use an arsenal of defenses to control the replication of Histoplasma yeasts, one of which is the limitation of trace metals. On the other hand, H. capsulatum combats metal restriction by upregulating metal importers such as the Zn importer ZRT2 . This transporter contributes to H. capsulatum pathogenesis upon activation of adaptive immunity. We constructed a fluorescent ZRT2 transcriptional reporter to probe H. capsulatum Zn sensing during infection and exposed the role for M-CSF activation of macrophages when GM-CSF is absent. These data highlight the ways in which fungal pathogens sense metal deprivation in vivo and reveal the potential of metal-sensing reporters. The work adds a new dimension to study how intracellular pathogens sense and respond to the changing environments of the host.
... Interestingly, many annotation terms ranked by both databases were related to metal binding in the molecular function category. Indeed, many host-pathogen interactions for bacteria need co-factors for catalytic activity (Cerasi et al., 2013;Khan et al., 2020;Shah et al., 2015). In the E.coli list of the PredPrIn interactions, there was also great participation of terms related to kinase activities. ...
Article
Full-text available
Semantic web standards have shown importance in the last 20 years in promoting data formalization and interlinking between the existing knowledge graphs. In this context, several ontologies and data integration initiatives have emerged in recent years for the biological area, such as the broadly used Gene Ontology that contains metadata to annotate gene function and subcellular location. Another important subject in the biological area is protein–protein interactions (PPIs) which have applications like protein function inference. Current PPI databases have heterogeneous exportation methods that challenge their integration and analysis. Presently, several initiatives of ontologies covering some concepts of the PPI domain are available to promote interoperability across datasets. However, the efforts to stimulate guidelines for automatic semantic data integration and analysis for PPIs in these datasets are limited. Here, we present PPIntegrator, a system that semantically describes data related to protein interactions. We also introduce an enrichment pipeline to generate, predict and validate new potential host–pathogen datasets by transitivity analysis. PPIntegrator contains a data preparation module to organize data from three reference databases and a triplification and data fusion module to describe the provenance information and results. This work provides an overview of the PPIntegrator system applied to integrate and compare host–pathogen PPI datasets from four bacterial species using our proposed transitivity analysis pipeline. We also demonstrated some critical queries to analyze this kind of data and highlight the importance and usage of the semantic data generated by our system. Availability and implementation https://github.com/YasCoMa/ppintegrator, https://github.com/YasCoMa/ppi_validation_process and https://github.com/YasCoMa/predprin.
... Intake of the whole grains in the diet reduces the mortality from cardiovascular diseases [60]. Zinc is a cofactor for various commensal bacterial proteins playing a role in the metabolism [61]. Deficiency of zinc leads to higher β-diversity [62]. ...
Article
Full-text available
The human gut is colonized by a variety of microorganisms especially bacteria. There are multiple evidences that gut microflora dysbiosis is a novel risk factor for development of various intestine-related diseases such as irritable bowel syndrome and inflammatory bowel disease as well as nonintestinal diseases including obesity, type II diabetes, and cardiovascular diseases. A mutual relationship among the host’s immune system and the metabolites produced by the gut microflora, including trimethylamine N-oxide (TMAO), short-chain fatty acids (SCFAs), and bile acids, is present. Alterations in the host-microbial interaction lead to impaired homeostasis and thus contribute towards the activation of several pathways that causes progression of cardiovascular diseases. This review summarizes the role of gut microflora dysbiosis in the development and progression of atherosclerosis, coronary artery disease, and hypertension. Dysbiosis has been implicated in the pathogenesis of atherosclerosis by TLR activation, intracellular Ca2+ release, FXR-induced signalling, and decreased removal of cholesterol from peripheral macrophages, while in hypertension the mechanism involved is prolonged haemodynamic effects of angiotensin-II and oxLDL-induced hypertension. In fact, CVDs are the leading cause of mortality across the globe; thus, targeting the gut microflora in the treatment of these diseases along with the conventional therapy can markedly reduce the cardiovascular disease burden. The gut microbiota-targeted treatment including prebiotics, probiotics, and postbiotics can be therapeutically beneficial. In future, the heart-gut axis can be presented as a novel and clinically relevant area for research.
Article
Full-text available
The rapid surge in antibiotic resistance to pathogens has emerged as a grave threat to public health, globally. This multiple drug resistance (MDR) is directly linked to the high rates of morbidity and mortality worldwide due to untreated microbial infections. Therefore, it is inevitable to identify some novel, efficient, and comparatively safer antimicrobial agents to rescue the declining health index. In this regard, nanomaterials with modified structure, size, and infinity have risen as the sole source to tackle the MDR either through ameliorating the efficacy of existing drugs or by triggering entirely new bactericidal mechanisms. Out of all the nanomaterials, metals, and metal oxide nanoparticles with biopolymer-induced reduction have fetched the attention of global researchers due to their significant and promising pathogen-killing ability without any hint of resistance. The current review covers the updated molecular modes of resistance development in Gram-positive and Gram-negative bacteria, comprehensively. This review also highlighted the detailed mode of action of various metallic nanoparticles (silver, zinc, copper, titanium, and cobalt) against MDR pathogens. Moreover, this review article thoroughly discussed the correlation between the mechanisms of resistance and alternative NPs bactericidal modes for better understanding for the readers. Last but not least, toxicity analysis is also explained for safe further use. Graphical Abstract
Preprint
Full-text available
Limiting the availability of transition metals at infection sites serves as a critical defense mechanism employed by the innate immune system to combat microbial infections. Pseudomonas aeruginosa exhibits a remarkable ability to thrive in zinc-deficient environments, which is facilitated by intricate cellular responses governed by numerous genes regulated by the zinc-responsive transcription factor Zur. Many of these genes have unknown functions, including those within the predicted PA2911-PA2914 and PA4063-PA4066 operons. A bioinformatic analysis revealed that PA2911-PA2914 comprises a TonB-dependent outer membrane receptor and an inner membrane ABC-permease responsible for importing metal-chelating molecules, whereas PA4063-PA4066 contains genes encoding a MacB transporter, likely involved in the export of large molecules. Molecular genetics and biochemical experiments, feeding assays, and intracellular metal content measurements demonstrated that PA2911-PA2914 and PA4063-PA4066 are engaged in the import and export of the pyochelin-cobalt complex, respectively. Notably, cobalt can reduce zinc demand and promote the growth of P. aeruginosa strains unable to import zinc, highlighting pyochelin-mediated cobalt import as a novel bacterial strategy to counteract zinc deficiency. These results unveil an unexpected role for pyochelin in zinc homeostasis and challenge the traditional view of this metallophore exclusively as an iron transporter.
Article
Zinc serves critical catalytic, regulatory, and structural roles. Hosts and their resident gut microbiota both require zinc, leading to competition, where a balance must be maintained. This systematic review examined evidence on dietary zinc and physiological status (zinc deficiency or high zinc/zinc overload) effects on gut microbiota. This review was conducted according to PRISMA (Preferred Reporting Items for Systematic reviews and Meta-Analyses) guidelines and registered in PROSPERO (CRD42021250566). PubMed, Web of Science, and Scopus databases were searched for in vivo (animal) studies, resulting in eight selected studies. Study quality limitations were evaluated using the SYRCLE risk of bias tool and according to ARRIVE guidelines. The results demonstrated that zinc deficiency led to inconsistent changes in α-diversity and short-chain fatty acid production but led to alterations in bacterial taxa with functions in carbohydrate metabolism, glycan metabolism, and intestinal mucin degradation. High dietary zinc/zinc overload generally resulted in either unchanged or decreased α-diversity, decreased short-chain fatty acid production, and increased bacterial metal resistance and antibiotic resistance genes. Additional studies in human and animal models are needed to further understand zinc physiological status effects on the intestinal microbiome and clarify the applicability of utilizing the gut microbiome as a potential zinc status biomarker.
Article
Full-text available
Chronic Obstructive Pulmonary Disease (COPD) has significantly contributed to global mortality, with three million deaths reported annually. This impact is expected to increase over the next 40 years, with approximately 5 million people predicted to succumb to COPD-related deaths annually. Immune mechanisms driving disease progression have not been fully elucidated. Airway microbiota have been implicated. However, it is still unclear how changes in the airway microbiome drive persistent immune activation and consequent lung damage. Mechanisms mediating microbiome-immune crosstalk in the airways remain unclear. In this review, we examine how dysbiosis mediates airway inflammation in COPD. We give a detailed account of how airway commensal bacteria interact with the mucosal innate and adaptive immune system to regulate immune responses in healthy or diseased airways. Immune-phenotyping airway microbiota could advance COPD immunotherapeutics and identify key open questions that future research must address to further such translation.
Article
Full-text available
Macrophages possess numerous mechanisms to combat microbial invasion, including sequestration of essential nutrients, like zinc (Zn). The pleiotropic cytokine granulocyte macrophage-colony stimulating factor (GM-CSF) enhances antimicrobial defenses against intracellular pathogens such as Histoplasma capsulatum, but its mode of action remains elusive. We have found that GM-CSF-activated infected macrophages sequestered labile Zn by inducing binding to metallothioneins (MTs) in a STAT3 and STAT5 transcription-factor-dependent manner. GM-CSF upregulated expression of Zn exporters, Slc30a4 and Slc30a7; the metal was shuttled away from phagosomes and into the Golgi apparatus. This distinctive Zn sequestration strategy elevated phagosomal H(+) channel function and triggered reactive oxygen species generation by NADPH oxidase. Consequently, H. capsulatum was selectively deprived of Zn, thereby halting replication and fostering fungal clearance. GM-CSF mediated Zn sequestration via MTs in vitro and in vivo in mice and in human macrophages. These findings illuminate a GM-CSF-induced Zn-sequestration network that drives phagocyte antimicrobial effector function.
Article
Full-text available
Zinc is essential for all bacteria, but excess amounts of the metal can have toxic effects. To address this, bacteria have developed tightly regulated zinc uptake systems, such as the ZnuABC zinc transporter which is regulated by the Fur-like zinc uptake regulator (Zur). In Pseudomonas aeruginosa, a Zur protein has yet to be identified experimentally, however, sequence alignment revealed that the zinc-responsive transcriptional regulator Np20, encoded by np20 (PA5499), shares high sequence identity with Zur found in other bacteria. In this study, we set out to determine whether Np20 was functioning as Zur in P. aeruginosa. Using RT-PCR, we determined that np20 (hereafter known as zur) formed a polycistronic operon with znuC and znuB. Mutant strains, lacking the putative znuA, znuB, or znuC genes were found to grow poorly in zinc deplete conditions as compared to wild-type strain PAO1. Intracellular zinc concentrations in strain PAO-Zur (Δzur) were found to be higher than those for strain PAO1, further implicating the zur as the zinc uptake regulator. Reporter gene fusions and real time RT-PCR revealed that transcription of znuA was repressed in a zinc-dependent manner in strain PAO1, however zinc-dependent transcriptional repression was alleviated in strain PAO-Zur, suggesting that the P. aeruginosa Zur homolog (ZurPA) directly regulates expression of znuA. Electrophoretic mobility shift assays also revealed that recombinant ZurPA specifically binds to the promoter region of znuA and does not bind in the presence of the zinc chelator N,N',N-tetrakis(2-pyridylmethyl) ethylenediamine (TPEN). Taken together, these data support the notion that Np20 is the P. aeruginosa Zur, which regulates the transcription of the genes encoding the high affinity ZnuABC zinc transport system.
Article
Human calprotectin (CP) is an antimicrobial protein that coordinates Mn(II) with high affinity in a Ca(II)-dependent manner at an unusual histidine-rich site (site 2) formed at the S100A8/S100A9 dimer interface. We present a 16-member CP mutant family where mutations in the S100A9 C-terminal tail (residues 96-114) are employed to evaluate the contributions of this region, which houses three histidines and four acidic residues, to Mn(II) coordination at site 2. The results from analytical size-exclusion chromatography, Mn(II) competition titrations, and electron paramagnetic resonance spectroscopy establish that the C-terminal tail is essential for high-affinity Mn(II) coordination by CP in solution. The studies indicate that His103 and His105 (HXH motif) of the tail complete the Mn(II) coordination sphere in solution, affording an unprecedented biological His6 site. These solution studies are in agreement with a Mn(II)-CP crystal structure reported recently (Damo, S. M.; et al. Proc. Natl. Acad. Sci. U.S.A. 2013, 110, 3841). Remarkably high-affinity Mn(II) binding is retained when either H103 or H105 are mutated to Ala, when the HXH motif is shifted from positions 103-105 to 104-106, and when the human tail is substituted by the C-terminal tail of murine S100A9. Nevertheless, antibacterial activity assays employing human CP mutants reveal that the native disposition of His residues is important for conferring growth inhibition against Escherichia coli and Staphylococcus aureus. Within the S100 family, the S100A8/S100A9 heterooligomer is essential for providing high-affinity Mn(II) binding; the S100A7, S100A9(C3S), S100A12, and S100B homodimers do not exhibit such Mn(II)-binding capacity.
Article
The relative stability of divalent first-row transition metal ion complexes, as defined by the Irving-Williams series, poses a fundamental chemical challenge for selectivity in bacterial metal ion acquisition. Here we show that although the substrate-binding protein of Streptococcus pneumoniae, PsaA, is finely attuned to bind its physiological substrate manganese, it can also bind a broad range of other divalent transition metal cations. By combining high-resolution structural data, metal-binding assays and mutational analyses, we show that the inability of open-state PsaA to satisfy the preferred coordination chemistry of manganese enables the protein to undergo the conformational changes required for cargo release to the Psa permease. This is specific for manganese ions, whereas zinc ions remain bound to PsaA. Collectively, these findings suggest a new ligand binding and release mechanism for PsaA and related substrate-binding proteins that facilitate specificity for divalent cations during competition from zinc ions, which are more abundant in biological systems.
Article
In Gram-negative bacteria the ZnuABC transporter ensures adequate zinc import in Zn(II)-poor environments, like those encountered by pathogens within the infected host. Recently, the metal-binding protein ZinT was suggested to operate as an accessory component of ZnuABC in periplasmic zinc recruitment. Since ZinT is known to form a ZinT-ZnuA complex in the presence of Zn(II) it was proposed to transfer Zn(II) to ZnuA. The present work was undertaken to test this claim. ZinT and its structural relationship with ZnuA have been characterized by multiple biophysical techniques (X-ray crystallography, SAXS, analytical ultracentrifugation, fluorescence spectroscopy). The metal-free and metal-bound crystal structures of Salmonella enterica ZinT show one Zn(II) binding site and limited structural changes upon metal removal. Spectroscopic titrations with Zn(II) yield a KD value 22 ± 2 nM for ZinT, while those with ZnuA point to one high affinity (KD<20 nM) and one low affinity Zn(II) binding site (KD in the micromolar range). Sedimentation velocity experiments established that Zn(II)-bound ZinT interacts with ZnuA, whereas apo-ZinT does not. The model of the ZinT-ZnuA complex derived from Small Angle X-ray Scattering experiments points to a disposition that favors metal transfer as the metal binding cavities of the two proteins face each other. ZinT acts as a Zn(II)-buffering protein that delivers Zn(II) to ZnuA. Knowledge of the ZinT-ZnuA relationship is crucial for understanding bacterial Zn(II) uptake.