Skip to main content

A Zur-mediated transcriptional regulation of the zinc export system in Pseudomonas aeruginosa

Abstract

The control of cellular zinc (Zn) concentrations by dedicated import and export systems is essential for the survival and virulence of Pseudomonas aeruginosa. The transcription of its many Zn transporters is therefore tightly regulated by a known set of transcription factors involved in either the import or the export of Zn. In this work, we show that the Zur protein, a well-known repressor of Zn import, plays a dual role and functions in both import and export processes. In a situation of Zn excess, Zur represses Zn entry, but also activates the transcription of czcR, a positive regulator of the Zn export system. To achieve this, Zur binds at two sites, located by DNA footprinting in the region downstream the czcR transcription start site. In agreement with this regulation, a delay in induction of the efflux system is observed in the absence of Zur and Zn resistance is reduced. The discovery of this regulation highlights a new role of Zur as global regulator of Zn homeostasis in P. aeruginosa disclosing an important link between Zur and zinc export.

Peer Review reports

Introduction

Metal ions are essential to the functioning and survival of all known forms of cellular life. Among all trace elements, zinc (Zn) is of particular importance because it functions as a cofactor of many essential enzymes and is, more generally, needed to sustain protein structure [1,2,3]. Zn is found at the top of the Irving-Williams series (Mg < Ca < Mn < Fe < Co < Ni < Cu > Zn) that ranks metal ions according to their ability to form complexes with proteins [4]. This is why, when present in excess inside the cell, Zn tends to outcompete other metals and cause protein mismetallization, the major cause of its cellular toxicity [5]. The intracellular concentration of Zn has therefore to be tightly and finely controlled by complex homeostasis mechanisms. In bacteria, the cytoplasmic Zn concentration is maintained within the range of 10–3 to 10–4 M, making it the second most abundant metal after iron [6].

In the Gram-negative bacterium Pseudomonas aeruginosa, several systems are involved in Zn homeostasis [7, 8]. Zn import into the periplasm is mediated by TonB-dependent importers such as ZnuD. A P-type ATPase and three ABC transporters help with the uptake of Zn from the periplasm into the cytoplasm. In case of metal excess, Zn is exported by a specific P-type ATPase CadA, a Resistance-Nodulation-Cell Division (RND) efflux pump (CzcCBA), and two cation diffusion facilitators (CDF) that play a minor role in Zn resistance [6, 9]. Control of Zn import and export is extremely important for some bacterial pathogens, especially for their interaction with their host, whether an animal [10] or a plant [11]. P. aeruginosa, for example, is able to grow and thrive in Zn-poor environments such as the bloodstream, where trace metals are chelated to prevent microbial growth [6]. Conversely, in a phagolysosome, high concentrations of Cu and Zn have antimicrobial properties and can eliminate the pathogen [12].

Several regulators of Zn homeostasis have been identified in P. aeruginosa. The major regulator of Zn import is the Zinc Uptake Regulator, hereafter Zur, protein (formerly Np20, [13]), a Fur (Ferric Uptake Regulator)-like protein [14]. In the presence of an excess of cytoplasmic Zn, Zur binds to the metal ion, dimerizes, and, in its dimeric form, binds to specific Zur boxes on the DNA to repress genes involved in Zn import [7, 13]. Zn export is controlled by two other regulators: CadR, a MerR-type regulator that activates the transcription of the P-type ATPase CadA [15], and CzcR, a response regulator (RR) of the two-component system CzcRS, which, when phosphorylated by CzcS, activates the transcription of the efflux pump CzcCBA that expels Zn from the cell [16]. Until now, the regulatory networks controlling Zn import and Zn export have been considered essentially independent from one another and no crosstalk between them has been substantiated.

In this study, we discovered that the Zur protein is needed for the full expression of the Zn efflux system when P. aeruginosa switches from a low-Zn environment to an excess-Zn situation. We show that Zur accelerates czcR transcription, and thus indirectly activates the transcription of Zn efflux systems by binding to two DNA sites in a region upstream the czcR translation start site (ATG). This new function makes Zur the keystone of Zn homeostasis in P. aeruginosa, acting not only as a repressor of Zn import but also as activator of Zn export.

Results

Zur binds to the czcRS promoter

The Zur protein, when concentrations of Zn are sufficient or excessive, binds to a 17 bp DNA Zur box to repress genes involved in Zn import [17]. In a global bioinformatics search on the P. aeruginosa genome, we found, as expected, Zur boxes in the promoter regions of genes involved in the import of Zn, but also, intriguingly, upstream of the czcRS genes [8]. This suggested a possible direct regulatory link, mediated by Zur, between the import and export of Zn. In order to validate the in silico analysis, we purified the Zur protein and performed an electrophoretic mobility shift assay (EMSA) on the czcR upstream region. A clear DNA shift was observed in the presence of Zur (Fig. 1A). The interaction is Zn-dependent as no shift was detected in presence of TPEN, a chelator with high affinity for Zn. DNA footprint analysis revealed two boxes located approximately 40 bps upstream of the translation start codon of czcR (Fig. 1B, 1C and S1A).

Fig. 1
figure 1

Zur binds upstream the czcR translation start codon. A) Electrophoretic Mobility Shift Assay (EMSA) using purified Zur protein and czcR-5’UTR DNA region or znuB ORF DNA region as negative control. 25 ng of DNA per reaction was used. The Zur protein, Zn and TPEN (Zn chelator) concentrations are indicated at the top of the figure. B) DNAse I footprinting of czcR promoter in the absence or presence or of 1 µM Zur protein. DNA fragments were analyzed by capillary electrophoresis. The two Zur boxes are indicated by numbers 1 and 2 in red. The sequencing reaction is visible below the figure (seq2). C) Sequence of the two boxes as determined by the footprinting experiments (Fig. 1B coding strand; Fig S1A template strand)

The transcription start site of czcR was determined by 5’RACE and it was found 300 bps upstream of the czcR translation start, revealing a 292-nucleotide 5’-UTR (Fig. 2A, S1B). Promoter fusions (-gfp) confirmed that a region longer than 292 bps upstream of the ATG is required to sustain czcRS transcription in the presence of Zn (construct #1 and #2 in Fig. 2B).

Fig. 2
figure 2

GFP reporter of the czcR promoters. A) Map of the czcR promoter and 5’-UTR DNA region with the two Zur boxes, the CzcR DNA binding site (according to [18]) and the czcR and czcC transcription start sites (+ 1). The number represents the nucleotide position relative to the translation start of czcR gene. B) GFP promoter fusions #1 and #2 transformed into the wt PAO1 strain. Cultures were incubated at 37 °C with shaking in a microplate reader and fluorescence measurement was performed every 15 min after the addition of 2 mM ZnCl2 (#1 + Zn, #2 + Zn) or without the addition of Zn (#1 and #2). The fluorescence values are normalized with cell density. Standard deviations of the triplicates are indicated

To determine whether the binding at the two Zur boxes was independent, we repeated the band shift assay using different versions of the czcR 5’-UTR DNA region, where box1 (mut1), box 2 (mut2), or both (mut1 + 2) Zur boxes were mutated at key positions (Fig. 3A). A higher shift was obtained when the two Zur boxes were present compared to the presence of only one, and no shift was detected when the two boxes were mutated (Fig. 3B), indicating that the two boxes support the binding of the regulator independently.

Fig. 3
figure 3

Mutations in the Zur boxes and their effects on Zur binding activity. A) Sequences of modified Zur boxes: mut1, mut2 and mut1 + 2 in red. B) Electrophoretic Mobility Shift Assay (EMSA) using purified Zur protein, different czcR 5’-UTR DNA region (czcR) and znuB (negative DNA control). 25 ng of each DNA per reaction was used. The Zur protein and Zn concentrations are indicated at the top of the figure. czcR wt, corresponds to the wt sequence of the czcR 5’-UTR DNA region, czcR mut1 and mut2 possess mutated box 1 or 2, respectively; czcR mut1 + 2 possess the two mutated Zur boxes

Zur is involved in zinc resistance

In order to investigate whether Zur had a role in the regulation of Zn export, we characterized the phenotype of a Δzur mutant. We found that, in the absence of Zur, P. aeruginosa is more sensitive to Zn excess than the parental strain (Fig. 4A). The phenotype is milder than in a ΔczcRS mutant, but more severe than in a ΔcadR mutant [15].

Fig. 4
figure 4

Zur is involved in CzcR expression. A) Growth curves of P. aeruginosa wt, Δzur, ΔczcRS and ΔcadR mutants in Zn-depleted medium (30 µM TPEN), in 2 mM and 4 mM ZnCl2, as indicated. B) Expression of czcR measured using qRT-PCR. RNAs were extracted at 0, 15 min, 1 h and 5 h after addition of 2 mM ZnCl2 as indicated. Fold expressions are compared to the wt strain at t0 and normalized by oprF expression. Standard deviations of the triplicates are indicated. Statistical analyses were performed according to the Student’s t-test and p-values are given as follows: ≤ 0.05 (*). C) Western blot analysis of total proteins sampled just before (0 h), or 1 h and 5 h after addition of 2 mM Zn of the wt and Δzur mutant, containing the empty plasmid (pE) or the IPTG inducible zur:6His gene (pzur). Blots were decorated with anti-CzcR, anti-6His for Zur detection and anti-Hsp70 as loading control

The effect of the zur deletion on czcR transcription was assessed using qRT-PCR. In the absence of Zur, czcR expression decreased significantly compared to the control (Fig. 4B). The difference was obvious 15 min (400-fold decrease) after Zn addition, still visible after 1 h (230-fold decrease), but no longer present after 5 h of induction. An immunoblot analysis confirmed this expression pattern at the protein level (Fig. 4C). As expected, the delay in czcR expression was abolished when zur:6His (zur containing a C-terminal 6-histidine tag to follow its expression) was overexpressed in trans by the addition of 0.2 mM IPTG (Fig. 4C). It has been shown that the Zn inductions of the CzcCBA efflux pump and the cation diffusion facilitator CzcD were CzcR dependent [7]. In agreement with this, the expression of these two export systems were also reduced in the Δzur mutant (Fig. 5). Conversely, the P-type ATPase CadA, which is known to be overexpressed in the absence of CzcCBA (i.e. ΔczcRS or ΔczcA [15]), showed increased expression in the Δzur mutant (Fig. 5). Altogether, these results strongly suggest that Zur is required to bring about the full expression of czcR and mount the proper response to a Zn excess. Deleting the Zur regulated major Zn import permease ZnuB has no effect on the zur mutant growth phenotype in the presence of Zn (Fig. S2) suggesting a direct effect of Zur on czcR expression. Using transcriptional gfp-fusions of czcC, including wt or mutated Zur boxes, no difference in fluorescence was observed, that Zur affects indirectly the CzcCBA expression, through CzcR (Fig. S3).

Fig. 5
figure 5

Expression of czcC, czcD and cadA measured using qRT-PCR. RNAs were extracted at 0, 15 min, 1 h and 5 h after addition of 2 mM ZnCl2 as indicated. Fold expressions are compared to the wt strain at t0 and normalized by oprF expression. Standard deviations of the triplicates are indicated. Statistical analyses were performed according to the Student’s t-test and p-values are given as follows: ≤ 0.05 (*), ≤ 0.001 (***)

Discussion

Zinc is an essential element whose cellular concentration is tightly regulated [19]. It is vital for cells and must always be present in sufficient amounts, but can become toxic as soon as it exceeds a threshold [20]. In the bacterium P. aeruginosa, eight zinc import and four zinc export systems have been described [7]. In case of zinc excess, the Zur protein represses the expression of many import systems, thus preventing too much Zn from entering the cell. In this work we described a novel key function of the Zur protein in P. aeruginosa, being involved not only in the repression of the import of Zn, but also in the early and fast induction of the CzcCBA Zn export system. Zur is therefore a keystone of Zn homeostasis in P. aeruginosa, integrating both the import and export of the ion.

In some other bacteria, Zur also acts as an activator of zinc efflux genes. In Xanthomonas campestris, for instance, Zur binds a GC-rich sequence overlapping the -35 to -10 region of the XC2976 gene, encoding a CzcD homolog, and directly activates its transcription [21]. In Caulobacter crescentus a Zur box, located just upstream the -35 element of a Zn export system operon allows its expression in the presence of Zn [22]. Another situation has been observed in Streptomyces coelicolor, where Zur binds to a Zur box located just upstream the -35 region of the zitB gene, a czcD homolog. Interestingly, at low Zn concentrations, Zur binds as a dimer to this region, weakly activating zitB transcription. As Zn concentration increases, Zur multimerizes on the promoter and increases zitB transcription [23]. Zur also acts as a positive transcriptional regulator of T6SS4 in Yersinia pseudotuberculosis by binding to the -90 DNA region [24]. This secretion system exports YezP, a Zn binding protein implicated in oxidative stress responses [25]. Other processes, not directly related to metal homeostasis, have also been found to be positively regulated by Zur. This is the case in Streptomyces avermitilis where Zur, by binding to the transcription start site, activates the transcription of the aveR gene, inducing the synthesis of the Avermectin secondary metabolites [26]. In the case of P. aeruginosa, we found that the Zur protein binds in a Zn-dependent manner to two DNA regions (Zur box1 and 2) located 200 bp after the czcR TSS, in the 5’-UTR DNA region. This binding has a positive effect on czcR transcription and, in general, on Zur-mediated Zn resistance.

What is the molecular action of Zur for the activation of czcR transcription? Recently numerous two-compoent systems response regulators (TCS-RR) binding sites have been detected in the intergenic region between czcR and czcC [27]. Zur could therefore be involved in the competition with the binding of transcriptional repressors in this region. Fine tuning of the efflux system could be essential in environmental conditions not tested in the laboratory. For instance, rapid changes in Zn concentration are known to occur during infection, because of nutritional immunity and metal excess in the phagolysosome [28, 29]. The tight efflux pump control might also be important in the environment, where rapid changes in metal concentrations might also occur. Interestingly, among the Pseudomonas species analyzed for the presence and location of Zur boxes (i.e. P. chlororaphis, P. fluorescens, P. putida, P. stuzeri and P. syringae), only P. aeruginosa possesses boxes upstream of the czcR gene [8]. This characteristic could be related to its ability to proliferate in presence of high metal concentrations. Thus, in P. aeruginosa, Zur appears to be the central cog in Zn homeostasis since it controls both the import and export of this metal. CzcR is also known to be capable of controlling the expression of the gene coding for the OprD porin, involved in imipenem resistance, or the lasI gene involved in quorum sensing [16, 30]. The importance of Zur in the control of czcR expression highlights a novel and important role of Zur in the interaction between Zn homeostasis, virulence factor expression and antibiotic resistance in P. aeruginosa. Deciphering this novel and important regulation occurring by Zur at the czcR gene vicinity is therefore of prime interest.

Materials and methods

Bacterial strains and culture media

The bacterial strains are listed in supplementary table S1. The modified Luria–Bertani medium (M-LB) used for this work was prepared as mentioned previously [15]. When required, antibiotics were added to the medium at the following concentrations: 50 μg/mL tetracycline (Tc, Axxora), 200 μg/mL carbenicillin (Cb, phytotechlab), 50 μg/mL Gentamycin (Gm, AppliChem) for P. aeruginosa or 100 μg/mL ampicillin (Ap, AppliChem) and 15 μg/mL Tc or Gm for E. coli.

Genetic manipulations

The primers used for cloning are detailed in supplementary table S2. Restriction and ligation enzymes (Promega), or fragment insertion using the Gibson assembly Cloning kit (New England Biolabs) were employed according to the supplier’s instructions. Plasmids were transformed into E. coli DH5α by heat shock, verified by sequencing before being electroporated into the P. aeruginosa wild type or Δzur strains [31]. For transcriptional gfp-fusion assays, the two czcR promoter regions (#1 and #2 in Fig. 2A) were obtained by PCR amplification. Fragments were then digested with KpnI and BglII enzymes and ligated into the corresponding sites of the pBBR1-gfp vector. For Zur6HIS overexpression assays, the full zur gene was amplified by PCR with a reverse primer containing the 6xHIS tag. The fragment was digested with the EcoRI and BamHI enzymes and ligated into the pMMB66EH plasmid. The pME6001-czcRS vector was used as a template for the sequencing reactions of the 5’RACE experiments. For this construction, the full czcRS operon and its promoter region were amplified by PCR and cloned into the pME6001 vector with the XhoI and HindIII restriction sites.. For the Δzur mutant, a DNA product, consisting of the two 500 bp amplicons flanking the zur gene fused together, was obtained by megaprimer amplification, digested with EcoRI and BamHI restriction enzymes and ligated into the pME3087 plasmid according to standard procedure [32]. After verification, the resulting plasmids were transformed into the P. aeruginosa wt strain. Merodiploids were selected as described previously [33] and deletions were confirmed by sequencing.

Growth experiments

To monitor the growth of the wt and the mutant strains, overnight cultures were diluted in M-LB medium to an OD600 of 0.05, supplemented with N,N,N′,N′-tetrakis(2-pyridylmethyl)-ethylenediamine (TPEN; Brunschwig) or ZnCl2 (Fluka) as indicated in Figs. 3A and 4. Cultures were incubated at 37 °C with shaking in a microplate reader (BioTek Instruments) and absorbance at 600 nm was measured every 15 min, for 12 h. For strains carrying the pBBR1 plasmid with gfp fusions #1 or #2, overnight cultures were diluted to an OD600 of 0.1 and cultured for 2.5 h in the presence of 30 μM TPEN before adding ZnCl2. Absorbance at 600 nm and fluorescence at 528 nm were then monitored every 15 min. Arbitrary units indicated in the Fig. 2B correspond to the fluorescence values normalized with the cell density.

Immunoblot analyses

Immunoblot analyses were performed as previously described [30]. Briefly, overnight cultures were diluted to an OD 0.1 and incubated 2.5 h in M-LB supplemented with 30 μM TPEN. Isopropyl-1-thio-D-galactopyranoside (IPTG, Sigma Aldrich) was added at a final concentration of 0.2 mM to the cultures of strains complemented with zur:6His on the pMMB66EH or the empty plasmid. 1 ml was collected and centrifuged immediately prior to ZnCl2 addition (t0) and after several time points, as specified in Fig. 4C. Pellets were solubilized in the appropriate volume of 2 × β-mercaptoethanol gel-loading buffer (an OD600 of 1 gives 0.175 mg/ml of protein) to obtain a final concentration of 2 mg/ml and loaded onto SDS PAGE, using 4–12% precast gels (Thermofisher Scientific). Transfers were performed with iBlot2 transfer stacks (Invitrogen) and nitrocellulose membranes were incubated with the anti-CzcR [34], anti-penta HIS (Invitrogen) and anti-HSP70 [30] antibodies. Blots were revealed by chemiluminescence using the Amersham Imager 680 System.

EMSA

The ability of Zur to bind the 5’-UTR DNA region of czcR, was determined by electrophoretic mobility shift assays (EMSA). The znuB gene, the 5’-UTR of czcR, wt or containing the mutated zur box-es (#mut1, #mut2 or #mut 1 + 2), were amplified by PCR from either the P. aeruginosa wt genomic DNA or from the in vitro synthetized mutated DNA (GeneArt, Thermo Fisher Scientific, Fig. 3A). Primers used for PCR amplification are indicated in supplementary table S2. Amplicons were then purified on agarose gel. Binding assays were performed as previously described [7]. Results were analyzed by staining the gel with 0.1% ethidium bromide and revealing with UV light using a NuGenius instrument.

DNAse I Footprinting assays

The czcR promoter was amplified by PCR from the wild type P. aeruginosa genomic DNA with 5’ Fluorescein (6FAM) primer as indicated in supplementary table S 2. DNA, with the plus or minus strand labeled independently, were gel purified. 50 ng of DNA fragments were mixed with or without 1 µM Zur in 40 μL of EMSA binding buffer [7] supplemented with 5 μM zinc. The reaction was incubated for 30 min at room temperature, followed by partial DNAse I digestion as described previously [15]. The sequence of the pME6001-czcRS plasmid, determined with the same FAM-labeled primers and the Thermo Sequenase Dye kit (Thermofisher Scientific) was used to align the peaks. All fragments were analyzed by capillary sequencing (Microsynth AG, Switzerland) using ILS600 as a size standard, then peaks were assessed using PeakScanner2 software (Thermofisher Scientific).

RNA extraction

Overnight cultures of the P. aeruginosa wt or mutant strains were diluted to an OD600 of 0.1 in M-LB containing 30 μM TPEN and cultured for 2.5 h before being induced with 2 mM ZnCl2. For RT-qPCR, 0.5 ml of culture was mixed with 1 ml of RNA protect (Qiagen) immediately prior to metal addition (t0) and after several time points as indicated in the figures. The P. aeruginosa wt culture intended for the 5’RACE experiment was collected after 5 h of 2 mM ZnCl2 induction. Total RNA extractions were performed as mentioned in previous publications [34]. Briefly, pellets were resuspended in 100 μL Tris–EDTA buffer supplemented with 5 mg/mL Lysozyme (Fluka) and incubated at room temperature for 10 min. The following steps were carried out with the RNeasy mini kit (Qiagen), according to the manufacturer’s directives. Residual genomic DNA was removed by treating with 10 units/sample of RNAse-free RQ1 DNAse (Promega) for 1 h at 37 °C followed by phenol–chloroform extraction and ethanol precipitation. Total RNA was then resuspended in RNAse-free water.

RT-qPCR

Quantitative RT-PCR were performed as previously described [35]. Briefly 500 ng of total RNA were reverse transcribed using random primers (Promega) with the ImProm-II reverse transcriptase (Promega) according to the manufacturer’s instructions. Quantitative PCR was performed on a tenfold dilution of resulting cDNA and using the SYBR Select Master Mix (Applied Biosystem), with the primers listed in table S2. Results were analyzed as formerly mentioned [36] and standardized with oprF.

5’RACE

The transcription starts of the czcR and czcC genes, were determined using the Rapid amplification of cDNA-5’ends (5’RACE) kit (Roche). Briefly, 5 μg of total RNA from the P. aeruginosa wt were reverse transcribed using the specific primers sp1R for czcR mRNA or sp1C for czcC mRNA. Resulting cDNAs were purified with the Wizard SV Gel and PCR Clean-Up System (Promega) and tailing with dATP and TdT. Tailed-cDNA strands were then amplified with the dT-Anchor primer (Roche) and the specific primer sp2R or sp2C (for czcR cDNA and czcC cDNA respectively). A 20-fold dilution of the first PCR was used as template for a second PCR with primers Anchor (Roche) and sp3R for czcR cDNA or sp2C for czcC cDNA. PCR products were then purified by agarose gel electrophoresis and subcloned into a pCR2.1 vector with the TA Cloning kit (Life Technologies) according to the supplier’s instructions. Plasmids isolated from E. coli clones were analyzed by sequencing. The transcription starts of czcR or czcC indicated in Fig. 2A correspond to the first nucleotide following the polyA tail.

Experimental relevance and statistical data

Regarding the graph representations, the mean values of at least three independent experiments are indicated in the figures, along with the corresponding standard deviations. When specified, statistical analyses were performed according to the Student’s t-test and significance p-values were set at p ≤ 0.05(*) or p ≤ 0.001 (***). For other figures, data points represent the mean of triplicate values.

Availability of data and materials

Datasets generated and analyzed during this study are included in this published article and its supplementary information files.

References

  1. Auld DS. The ins and outs of biological zinc sites. Biometals. 2009;22:141–8.

    Article  CAS  Google Scholar 

  2. Auld, D.S. (2013) In Kretsinger, R. H., Uversky, V. N. and Permyakov, E. A. (eds.), Encyclopedia of Metalloproteins. Springer New York, New York, NY, pp. 2554–2559.

  3. Vallee BL, Auld DS. Zinc coordination, function, and structure of zinc enzymes and other proteins. Biochemistry. 1990;29:5647–59.

    Article  CAS  Google Scholar 

  4. Irving, H. and Williams, R.J.P. (1953) The stability of transition-metal complexes. J Chem Soc (Resumed), 3192–3210.

  5. Foster AW, Osman D, Robinson NJ. Metal preferences and metallation. J Biol Chem. 2014;289:28095–103.

    Article  CAS  Google Scholar 

  6. Chandrangsu P, Rensing C, Helmann JD. Metal homeostasis and resistance in bacteria. Nat Rev Microbiol. 2017;15:338–50.

    Article  CAS  Google Scholar 

  7. Ducret V, Abdou M, Goncalves Milho C, Leoni S, Martin-Pelaud O, Sandoz A, Segovia Campos I, Tercier-Waeber ML, Valentini M, Perron K. Global analysis of the zinc homeostasis network in pseudomonas aeruginosa and its gene expression dynamics. Front Microbiol. 2021;12: 739988.

    Article  Google Scholar 

  8. Ducret, V., Gonzalez, D. and Perron, K. (2022) Zinc homeostasis in Pseudomonas. Biometals.

  9. Gonzalez MR, Ducret V, Leoni S, Perron K. Pseudomonas aeruginosa zinc homeostasis: Key issues for an opportunistic pathogen. Biochim Biophys Acta Gene Regul Mech. 2019;1862:722–33.

    Article  CAS  Google Scholar 

  10. Lonergan ZR, Skaar EP. Nutrient zinc at the host-pathogen interface. Trends Biochem Sci. 2019;44:1041–56.

    Article  CAS  Google Scholar 

  11. Cabot C, Martos S, Llugany M, Gallego B, Tolra R, Poschenrieder C. A role for zinc in plant defense against pathogens and herbivores. Front Plant Sci. 2019;10:1171.

    Article  Google Scholar 

  12. Kehl-Fie TE, Skaar EP. Nutritional immunity beyond iron: a role for manganese and zinc. Curr Opin Chem Biol. 2010;14:218–24.

    Article  CAS  Google Scholar 

  13. Ellison ML, Farrow JM 3rd, Parrish W, Danell AS, Pesci EC. The transcriptional regulator Np20 is the zinc uptake regulator in Pseudomonas aeruginosa. PLoS ONE. 2013;8: e75389.

    Article  CAS  Google Scholar 

  14. Fillat MF. The FUR (ferric uptake regulator) superfamily: diversity and versatility of key transcriptional regulators. Arch Biochem Biophys. 2014;546:41–52.

    Article  CAS  Google Scholar 

  15. Ducret V, Gonzalez MR, Leoni S, Valentini M, Perron K. The CzcCBA Efflux system requires the CadA P-type ATPase for timely expression upon zinc excess in pseudomonas aeruginosa. Front Microbiol. 2020;11:911.

    Article  Google Scholar 

  16. Perron K, Caille O, Rossier C, Van Delden C, Dumas JL, Kohler T. CzcR-CzcS, a two-component system involved in heavy metal and carbapenem resistance in Pseudomonas aeruginosa. J Biol Chem. 2004;279:8761–8.

    Article  CAS  Google Scholar 

  17. Pederick VG, Eijkelkamp BA, Begg SL, Ween MP, McAllister LJ, Paton JC, McDevitt CA. ZnuA and zinc homeostasis in Pseudomonas aeruginosa. Sci Rep. 2015;5:13139.

    Article  CAS  Google Scholar 

  18. Fan K, Cao Q, Lan L. Genome-wide mapping reveals complex regulatory activities of BfmR in Pseudomonas aeruginosa. Microorganisms. 2021;9(3):485.

    Article  CAS  Google Scholar 

  19. Yin, S., Duan, M., Fang, B., Zhao, G., Leng, X. and Zhang, T. (2022) Zinc homeostasis and regulation: Zinc transmembrane transport through transporters. Crit Rev Food Sci Nutr, 1–11.

  20. Xia P, Lian S, Wu Y, Yan L, Quan G, Zhu G. Zinc is an important inter-kingdom signal between the host and microbe. Vet Res. 2021;52:39.

    Article  CAS  Google Scholar 

  21. Huang DL, Tang DJ, Liao Q, Li HC, Chen Q, He YQ, Feng JX, Jiang BL, Lu GT, Chen B, et al. The Zur of Xanthomonas campestris functions as a repressor and an activator of putative zinc homeostasis genes via recognizing two distinct sequences within its target promoters. Nucleic Acids Res. 2008;36:4295–309.

    Article  CAS  Google Scholar 

  22. RR Mazzon VS Braz JF Silva Neto da do Valle Marques, M. 2014 Analysis of the Caulobacter crescentus Zur regulon reveals novel insights in zinc acquisition by TonB-dependent outer membrane proteins BMC Genomics 15 734

  23. Choi SH, Lee KL, Shin JH, Cho YB, Cha SS, Roe JH. Zinc-dependent regulation of zinc import and export genes by Zur. Nat Commun. 2017;8:15812.

    Article  Google Scholar 

  24. Cai R, Gao F, Pan J, Hao X, Yu Z, Qu Y, Li J, Wang D, Wang Y, Shen X, et al. The transcriptional regulator Zur regulates the expression of ZnuABC and T6SS4 in response to stresses in Yersinia pseudotuberculosis. Microbiol Res. 2021;249: 126787.

    Article  CAS  Google Scholar 

  25. Wang T, Si M, Song Y, Zhu W, Gao F, Wang Y, Zhang L, Zhang W, Wei G, Luo ZQ, et al. Type VI secretion system Transports Zn2+ to combat multiple stresses and host immunity. PLoS Pathog. 2015;11: e1005020.

    Article  Google Scholar 

  26. Lyu M, Cheng Y, Dai Y, Wen Y, Song Y, Li J, Chen Z. Zinc-responsive regulator Zur regulates zinc homeostasis, secondary metabolism, and morphological differentiation in streptomyces avermitilis. Appl Environ Microbiol. 2022;88: e0027822.

    Article  Google Scholar 

  27. Trouillon J, Imbert L, Villard A-M, Vernet T, Attrée I, Elsen S. Determination of the two-component systems regulatory network reveals core and accessory regulations across Pseudomonas aeruginosa lineages. Nucleic Acids Res. 2021;49:11476–90.

    Article  CAS  Google Scholar 

  28. Murdoch CC, Skaar EP. Nutritional immunity: the battle for nutrient metals at the host–pathogen interface. Nat Rev Microbiol. 2022;20(11):657–70.

    Article  CAS  Google Scholar 

  29. Djoko KY, Ong CL, Walker MJ, McEwan AG. the role of copper and zinc toxicity in innate immune defense against bacterial pathogens. J Biol Chem. 2015;290:18954–61.

    Article  CAS  Google Scholar 

  30. Dieppois G, Ducret V, Caille O, Perron K. The transcriptional regulator CzcR modulates antibiotic resistance and quorum sensing in Pseudomonas aeruginosa. PLoS ONE. 2012;7: e38148.

    Article  CAS  Google Scholar 

  31. Choi KH, Kumar A, Schweizer HP. A 10-min method for preparation of highly electrocompetent Pseudomonas aeruginosa cells: application for DNA fragment transfer between chromosomes and plasmid transformation. J Microbiol Methods. 2006;64:391–7.

    Article  CAS  Google Scholar 

  32. Sambrook, J., and D. W. Russell. . (2001. ) Molecular cloning: a laboratory manual 3rd ed.Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.

  33. Gonzalez MR, Ducret V, Leoni S, Fleuchot B, Jafari P, Raffoul W, Applegate LA, Que YA, Perron K. Transcriptome analysis of pseudomonas aeruginosa cultured in human burn wound exudates. Front Cell Infect Microbiol. 2018;8:39.

    Article  Google Scholar 

  34. Ducret V, Gonzalez MR, Scrignari T, Perron K. OprD repression upon metal treatment requires the RNA Chaperone Hfq in Pseudomonas aeruginosa. Genes (Basel). 2016;7(10):82.

    Article  Google Scholar 

  35. Marguerettaz M, Dieppois G, Que Y-A, Ducret V, Zuchuat S, Perron K. Sputum containing zinc enhances carbapenem resistance, biofilm formation and virulence of Pseudomonas aeruginosa. Microb Pathog. 2014;77:36–41.

    Article  CAS  Google Scholar 

  36. Schmittgen TD, Livak KJ. Analyzing real-time PCR data by the comparative C(T) method. Nat Protoc. 2008;3:1101–8.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

Not applicable

Funding

This work was supported by the Swiss National Science Foundation grant 31003A_179336 for K.P., Eccellenza grant PCEFP3_203343 for M.V and Ambizione grant PZ00P3_180142 for D.G.

Author information

Authors and Affiliations

Authors

Contributions

V.D., D.G., S.L., M.V., and K.P. conceived and planned the experiments. V.D., S.L., and K.P. carried out the experiments. V.D., D.G., S.L., M.V., and K.P. contributed to the interpretation of the results. V.D., D.G., M.V., and K.P wrote the manuscript.

Corresponding author

Correspondence to Karl Perron.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no conflict of interest.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1:

Figure S1. A) DNAse I footprinting of czcR promoter (template strand) in the absence or presence or of 1 µM Zur protein. DNA fragments were analyzed by capillary electrophoresis. The two Zur boxes are indicated by the numbers 1 and 2 in red. The sequencing reaction is visible below the figure (seq1) and the sequence with the two boxes indicated in red is shown below the figure. B) Transcription start site (+1) of czcR and czcC determined by 5’RACE. The CzcR box, according to (1) is indicated in the box and the -35, -10 sequences were determined in silico using BPROM program (2).

Additional file 2:

Figure S2. Growth curves of the wt PAO1 strain, the Δzur mutant and the ΔzurΔznuB double mutant without or in presence of 2 mM ZnCl2 (Zn). Standard deviations of the triplicates are indicated.

Additional file 3:

Figure S3. GFP promoter fusions containing 542 bp upstream of the czcC DNA regionincluding either the wt or the mutated Zur boxes (mut1+2). Fusions were transformed into the wt PAO1 strain. Cultures were incubated at 37°C with shaking in a microplate reader and fluorescence measurement was performed every 15 min after the addition of 2mM ZnCl2, or without the addition of Zn, as indicated. The fluorescence values are normalized with cell density. Standard deviations of the triplicates are indicated.

Additional file 4:

Table S1. Strains and plasmids used in this study

Additional file 5:

Table S2. Primers used in this study

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ducret, V., Gonzalez, D., Leoni, S. et al. A Zur-mediated transcriptional regulation of the zinc export system in Pseudomonas aeruginosa. BMC Microbiol 23, 6 (2023). https://doi.org/10.1186/s12866-022-02750-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12866-022-02750-4

Keywords