Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Modulation of Metabolism and Switching to Biofilm Prevail over Exopolysaccharide Production in the Response of Rhizobium alamii to Cadmium

  • Mathieu Schue,

    Current address: CNRS, UPR 9025, Lab Enzymologie Interfaciale et Physiologie de la Lipolyse, Marseille, France

    Affiliations CEA, Lab Ecol Microbienne Rhizosphere & Environm Extre, iBEB, DSV, Saint-Paul-lez-Durance, France, CNRS, Unite Mixte Rech Biol Vegetale & Microbiol Enviro, UMR 6191, Saint-Paul-lez-Durance, France, Université Aix Marseille, Saint-Paul-lez-Durance, France

  • Agnes Fekete,

    Affiliation Helmholtz-Zentrum Muenchen-German Research Center for Environmental Health, Institute for Ecological Chemistry, Department of BioGeochemistry and Analysis, Neuherberg, Germany

  • Philippe Ortet,

    Affiliations CEA, Lab Ecol Microbienne Rhizosphere & Environm Extre, iBEB, DSV, Saint-Paul-lez-Durance, France, CNRS, Unite Mixte Rech Biol Vegetale & Microbiol Enviro, UMR 6191, Saint-Paul-lez-Durance, France, Université Aix Marseille, Saint-Paul-lez-Durance, France

  • Catherine Brutesco,

    Current address: CEA, DSV, IBEB, Laboratoire de Bioénergétique Cellulaire, Saint-Paul-lez-Durance, France

    Affiliations CEA, Lab Ecol Microbienne Rhizosphere & Environm Extre, iBEB, DSV, Saint-Paul-lez-Durance, France, CNRS, Unite Mixte Rech Biol Vegetale & Microbiol Enviro, UMR 6191, Saint-Paul-lez-Durance, France, Université Aix Marseille, Saint-Paul-lez-Durance, France

  • Thierry Heulin,

    Affiliations CEA, Lab Ecol Microbienne Rhizosphere & Environm Extre, iBEB, DSV, Saint-Paul-lez-Durance, France, CNRS, Unite Mixte Rech Biol Vegetale & Microbiol Enviro, UMR 6191, Saint-Paul-lez-Durance, France, Université Aix Marseille, Saint-Paul-lez-Durance, France

  • Philippe Schmitt-Kopplin,

    Affiliations Helmholtz-Zentrum Muenchen-German Research Center for Environmental Health, Institute for Ecological Chemistry, Department of BioGeochemistry and Analysis, Neuherberg, Germany, Department for Chemical-Technical Analysis Research Center Weihenstephan for Brewing and Food Quality, Technische Universität München, Freising-Weihenstephan, Germany

  • Wafa Achouak,

    Affiliations CEA, Lab Ecol Microbienne Rhizosphere & Environm Extre, iBEB, DSV, Saint-Paul-lez-Durance, France, CNRS, Unite Mixte Rech Biol Vegetale & Microbiol Enviro, UMR 6191, Saint-Paul-lez-Durance, France, Université Aix Marseille, Saint-Paul-lez-Durance, France

  • Catherine Santaella

    catherine.santaella@cea.fr

    Affiliations CEA, Lab Ecol Microbienne Rhizosphere & Environm Extre, iBEB, DSV, Saint-Paul-lez-Durance, France, CNRS, Unite Mixte Rech Biol Vegetale & Microbiol Enviro, UMR 6191, Saint-Paul-lez-Durance, France, Université Aix Marseille, Saint-Paul-lez-Durance, France

Abstract

Heavy metals such as cadmium (Cd2+) affect microbial metabolic processes. Consequently, bacteria adapt by adjusting their cellular machinery. We have investigated the dose-dependent growth effects of Cd2+ on Rhizobium alamii, an exopolysaccharide (EPS)-producing bacterium that forms a biofilm on plant roots. Adsorption isotherms show that the EPS of R. alamii binds cadmium in competition with calcium. A metabonomics approach based on ion cyclotron resonance Fourier transform mass spectrometry has showed that cadmium alters mainly the bacterial metabolism in pathways implying sugars, purine, phosphate, calcium signalling and cell respiration. We determined the influence of EPS on the bacterium response to cadmium, using a mutant of R. alamii impaired in EPS production (MSΔGT). Cadmium dose-dependent effects on the bacterial growth were not significantly different between the R. alamii wild type (wt) and MSΔGT strains. Although cadmium did not modify the quantity of EPS isolated from R. alamii, it triggered the formation of biofilm vs planktonic cells, both by R. alamii wt and by MSΔGT. Thus, it appears that cadmium toxicity could be managed by switching to a biofilm way of life, rather than producing EPS. We conclude that modulations of the bacterial metabolism and switching to biofilms prevails in the adaptation of R. alamii to cadmium. These results are original with regard to the conventional role attributed to EPS in a biofilm matrix, and the bacterial response to cadmium.

Introduction

The exposure of bacterial cells to heavy metals in their environment mediates biological effects, usually through the direct or indirect action of reactive oxygen species [1], [2]. In fact, non-redox-reactive metals, such as cadmium, show a high degree of reactivity towards sulfur, nitrogen and oxygen atoms in biomolecules. Cadmium may bind sulfur in essential enzymes, and alter their functions. Many studies have focused on the molecular mechanism of bacterial cell tolerance to cadmium, mainly for the case of species that are resistant to high metal concentrations, such as Stenotrophomonas [3] or Cupriavidus metallidurans (review in [4]). However, cadmium concentrations and its availability in metal contaminated soils are generally low. At low cadmium concentrations, Dedieu et al. [5] studied the interactions of Sinorhizobium meliloti extracellular compounds on cadmium speciation and availability, and Pagès et al. [6] reported on the completely different adaptation mechanisms of phenotypic variants of Pseudomonas brassicacearum in the presence of cadmium. Varied mechanisms account for cadmium detoxication in bacteria, involving exclusion, binding and sequestration. Cadmium is removed from cells by metal efflux transporters [7], [8], [9], reduced as cadmium sulfide [10], precipitated as insoluble salts [11], immobilized within the cell walls [12], or linked to chelating agents [13], [14]. Cell exudates, such as proteins, siderophores and to a minor extent polysaccharide, play a role in the short-term interaction between Sinorhizobium meliloti and cadmium [5], [15].

Because the adsorption of cadmium as well as of other metals can be associated with the secretion of exopolysaccharide (EPS) or capsular material [2], [16], [17], EPSs are considered as potential metal transporters in soil [18], [19].

Gram-negative soil bacteria belonging to the commonly named rhizobia are able to produce EPSs with a large diversity of chemical structures [20], [21]. These EPSs are the main contributors in legume-rhizobia interactions, leading to nodulation and nitrogen fixation. Eventually, much of what we know about rhizobia and their EPSs arises from studies of their symbiotic interactions with legume plants, whereas their interactions with non-legumes have been neglected. However, rhizobia associate with the roots of non-legumes such as Arabidopsis thaliana [22], [23], Helianthus annuus [24] and Brassica napus [25]. Rhizobial EPSs also have functions beyond specific recognition in the nodulation process, such as plant growth promotion of non-legumes [24] or evasion from the defense response of plant legumes during crack entry in roots [26]. We should consider communities of rhizobia and their EPSs as integral and functionally important partners of a diverse plant rhizosphere.

For that purpose, we addressed the question of how Rhizobium alamii [27], an EPS-producing bacterium, responds to the toxic impact of cadmium. R. alamii is a rhizobacterium, isolated from the rhizophere of the sunflower, producing a mucoid and an adhesive EPS [24]. This bacterium colonizes the root system of the sunflower, A. thaliana, and rapeseed, and together with its production of EPS in the rhizosphere, it improves the physical structure of the root adherent soil and plant growth under conditions of hydric stress [23], [24]. We studied the mechanisms through which EPS could contribute to the tolerance of R. alamii to cadmium, by using a mutant strain impaired in EPS production (MSΔGT) [23]. We investigated cadmium adsorption on the EPS, in the absence or in competition with calcium. Ion cyclotron resonance Fourier transform mass spectrometry (ICRFT/MS) was used to monitor the metabolic perturbations after exposure of R. alamii cells to a cadmium concentration, which slowed down, but did not inhibit cell growth. We discuss the means by which the bacterial cells adjusted both their entire metabolic processes, and their way of life, to limit cadmium damage. We have for the first time shown that cadmium promotes the formation of a biofilm by R. alamii, and that this change occurs independently of the presence of EPS. These results are original with regard to the conventional role attributed to EPS in a biofilm matrix, and the bacterial response to a heavy metal.

Results and Discussion

Effect of cadmium on R. alamii growth and EPS production

In this study, the minimal inhibitory concentration (MIC) of cadmium (Cd2+) in R. alamii cells was assessed at 44 µM (5 mg.L−1 Cd2+) in a tenfold diluted Tryptic Soy Broth (TSB/10) and as illustrated in Figure 1. The increase of Cd2+ concentration from 1 mg.L−1 to 2 mg.L−1 almost doubled the lag phase. In our hands, cadmium MIC was 53 µM (6 mg.L−1) on R. alamii wt and MSΔGT mutant cells on an agarose solidified TSB/10 medium (Figure 2). 1H NMR spectra of the EPS isolated from the bacteria, cultured in the absence or presence of cadmium, showed that the chemical structure of the EPS was unmodified (A. Heyraud, personal communication). Cadmium concentrations ranging from 0 to 133 µM (0 to 15 mg.L−1 of Cd2+) did not significantly (p>0.5) modify the amount of EPS synthesized by the wt strain in a RCV mineral medium at pH 6.8, supplemented with glucose (Figure S1). The R. alamii behaved like Sinorhizobium meliloti whose EPSs content is unmodified, at pH 7 in response to 10 µM of cadmium nitrate [28].

thumbnail
Figure 1. Growth curves of Rhizobium alamii in ten-fold diluted tryptic soy browth (TSB/10) at 30°C in the presence of cadmium nitrate (0 to 10 mg.mL-1 of Cd2+ (1 to 89 µM).

Dark circle 0 mg.L−1; Open circle 1 mg.L−1; Grey triangle 2 mg.L−1 ; Cross 5 mg.L−1; Lozenge 10 mg.L−1. Mean of 3 replicates ± SD.

https://doi.org/10.1371/journal.pone.0026771.g001

thumbnail
Figure 2. Determination of minimal inhibitory concentrations of cadmium on R. alamii wt and MSΔGT mutant growth on ten-fold diluted tryptic soy browth (TSB/10) solidified with agarose.

Initial bacterial cell cultures in TSB/10 (108 CFU.mL−1) were spotted at dilutions from 10−1 to 10−4. Plates were incubated at 30°C for 96 h.

https://doi.org/10.1371/journal.pone.0026771.g002

R. alamii metabonomics in response to cadmium

We carried out metabonomic investigations on cells grown up to the end of the exponential phase, with or without 18 µM (2 mg.L−1) Cd2+, in order to determine which metabolism pathways were altered by cadmium in R. alamii. We used a screening method based on ion cyclotron resonance Fourier transform mass spectrometry (ICRFT/MS). This technology allows high-precision measurements to be made of a charged mass, within an error range of only a few parts per million. MassTRIX (http://masstrix.org) allows the compounds, detected with a chemically probable structure, to be assigned in the context of restricted metabolite possibilities for a given organism, using the KEGG pathway database: http://www.genome.jp/kegg/pathway.html [29]. By considering the set of R. alamii metabolites detected in positive and negative ESI of MS, in the experiments with or without cadmium, 1897 putative compounds were common to these conditions, and respectively 936 and 653 probable components were exclusively expressed, in the presence or absence of the metal. Cadmium induced metabolic alterations in major pathways, as summarized in Tables 1 and 2. Cadmium increased the number and level of enzymes and metabolites involved in the sugar metabolism (potentially fructose, glucose, mannose, galactose, cellobiose, inositol, starch and sucrose), phosphorylated intermediates of glycolysis compounds (likely glucose 6-phosphate, fructose 6-phosphate, glucose 1-phosphate), ABC transporters of sugars (potential methyl-galactoside, D-allose, fructose, cellobiose) and the phosphotransferase system which, in bacteria, is the major carbohydrate transport system for incoming sugar substrates, through translocations across the cell membrane. Cadmium ions are admitted into sensitive bacterial cells by the energy-dependent manganese transport systems, where they cause rapid cessation of respiration by binding to sulfhydryl groups in proteins [30], [31]. Glycolysis therefore appeared to be an alternative pathway for energy. Under cadmium stress, hexoses were potentially channelled towards a pentose phosphate pathway. In the presence of Cd2+, the nucleotide metabolism revealed a modulation of the purine metabolism, with an accumulation of purine-based nucleosides (adenosine, deoxyadenosine, guanine and deoxyguanine) at the same time as a decrease in purine-based nucleotides (AMP and dGMP) as illustrated in Figure 3. This result was confirmed by ultra performance liquid chromatographic (UPLC) analysis (Figure 4). The decrease in nucleotide contents and the recovery of pyrimidine and purine compounds could be consistent with an adaptation process, through slowing of cell division in cells likely to adapt to the metal toxicity [32]. In the presence of cadmium, the citric acid cycle and tryptophan metabolism were shut off (Table 1) suggesting the cessation of aerobic sugar respiration and amino acid synthesis. A glutathione precursor, γ-L-Glutamyl-L-cysteine, was identified under cadmium stress. In response to cadmium, Pseudomonas brassicacearum is found to switch from the citric acid cycle to an anaerobic metabolism [6], and the inhibition of protein and glutathione syntheses are described in the response of R. leguminosarum to cadmium [14], [33]. Cadmium activated the biosynthesis of lipopolysaccharide and capsular polysaccharide, which are the outer membrane components of Gram-negative bacteria and are potential metal binding sites [6], [12], [34]. Pyrophosphate was detected following Cd2+ treatment only. Pyrophosphate can form crystals with cadmium [35]. The accumulation of inorganic phosphate is a detoxification mechanism reported in Klebsiella aerogenes [36] and is also involved in the regulation of biofilm formation in Pseudomonas fluorescens [37]. Cadmium is also activated the inositol phosphate metabolism that regulates cytoplasmic Ca2+ and communication, and the cyclic ADP-ribose pathway, which is a secondary messenger for Ca2+ mobilisation [38]. The structural homology between calcium and cadmium ions might account for the activation of pathways needed to maintain calcium homeostasis and signal transduction.

thumbnail
Figure 3. Analysis of purine and pyrimidine metabolism from R. alamii cells grown with no cadmium (white bars) or 2 mg.L-1 of cadmium as cadmium nitrate.

Masses of the purine based cyclic nucleotides were additionally detected from the Cd2+ treated samples. 2,3&3,5-cyclic-AMP from the wt and 2,3&3,5-cyclic-GMP were detectable according to the FTMS measurements and use of MassTrix for data interpretation. Generally, increase in the intensity of the purine based nucleosides and decrease of the detected nucleotides were observed in presence of Cd2+.

https://doi.org/10.1371/journal.pone.0026771.g003

thumbnail
Figure 4. Analysis of nucleotide and nucleoside derivatives.

UPLC chromatograms of cell extracts of R. alamii grown up to the beginning of the stationary phase. a) Blank; b) R. alamii cells grown in the absence of cadmium; c) R. alamii cells grown in the presence of 2 mg.L−1 of cadmium. Dash lines stand for nucleotide standards (AMP,GMP;CMP,UMP,dGTP).

https://doi.org/10.1371/journal.pone.0026771.g004

thumbnail
Table 1. Modification of KEGG pathway metabolites induced by cadmium in Rhizobium alamii.

https://doi.org/10.1371/journal.pone.0026771.t001

thumbnail
Table 2. Main hypotheses in the metabolic pathways altered by cadmium.

https://doi.org/10.1371/journal.pone.0026771.t002

Altogether, these results show that the adaptation of R. alamii cells to cadmium could imply a multifaceted scheme using metal binding, precipitation or export, changes in respiration process, and the retention of cadmium-like metal homeostasis.

Cadmium induces the formation of a biofilm by R. alamii independently of EPS synthesis

R. alamii wt and MSΔGT strains were grown statically in the presence of cadmium (from 0 to 15 mg.L−1 of Cd2+), in mineral RCV medium supplemented with glucose (for composition see materials and methods) to promote EPS synthesis. Figure 5A shows staining of biofilms formed on the tube walls using crystal violet [39], and Figure 5B represents the quantity of planktonic cells monitored by means of optical density measurements, and the quantification of surface-attached cells stained by crystal violet coloration as a function of cadmium concentration. Increasing cadmium concentrations significantly decreased the population of the planktonic cells, and induced the attachment of cells as a biofilm (p<0.05).

thumbnail
Figure 5. Enhancement of biofilm formation in response to cadmium.

A) Staining of Rhizobium alamii biofilms formed by the wild type strain (wt, upper panel) and the MSΔGT mutant impaired in EPS synthesis (lower panel), using a crystal violet assay (O'Toole et al, 1990). The cadmium concentration increases from left to right (0 to 133 µM; 0 to 15 mg.L−1). B) Biofilms of R. alamii wt (white bars) and MSΔGT mutant (dark bars) were quantified using a crystal violet assay (OD 595 nm) after 72 h of static culture at 30°C. Planktonic growth of R. alamii wt (open squares) and MSΔGT mutant (black circles) was measured in the culture supernatant (OD 600 nm). The error bars represent the confidence intervals (95%) of triplicates.

https://doi.org/10.1371/journal.pone.0026771.g005

A biofilm is a lifestyle able to resist various environmental stressors such as antibiotics [40], or metals such as copper, zinc and lead [41], and nickel [42]. In the present study, we reveal for the first time cadmium-induced biofilm formation by R. alamii cells. Interestingly, wt and EPS-mutant strains showed the same change in growth, from free-swimming to biofilm mode, in response to cadmium, suggesting that biofilm formation, rather than EPS, was a remediation to metal toxicity.

Cadmium binds to the EPS of R. alamii

The chemical structure of R. alamii EPS contains carboxylic and hydroxyl functions [43] that may bind metals. The biosorption of metal ions on the EPS of R. alamii has been measured in a MOPS buffer and in a calcium-containing medium, and has been described using the Langmuir and Freundlich isotherms [44]. The parameters calculated from these two models are summarized in Table 3. The Langmuir model, based on the assumption of a monolayer adsorption onto a solid surface with a defined number of identical sites, gave the best fit of equilibrium adsorption data measured in a MOPS buffer. The Freundlich isotherm, which is an empirical model used to describe heterogeneous systems, best fitted the biosorption of cadmium in the presence of an excess of calcium ions. Both models showed that the EPS of R. alamii actually bound cadmium. An apparent distribution coefficient of 3480 L.mg−1 was found at an EPS concentration of 0.125 g.L−1 and a maximum biosorption capacity of 11 mg of cadmium per g of EPS was determined. Both models showed that calcium, an abundant ion in soil solutions, competed with cadmium and reduced its sorption. At the highest EPS concentration tested (1 g.mL−1), the EPS of R. alamii showed the lowest ability to bind Cd2+ (apparent distribution coefficient of 387 L.mg−1 and maximum biosorption capacity of 6 mg of cadmium per g of EPS), showing that polysaccharide chain-chain interactions could preclude the binding of cadmium.

Cadmium and biofilm matrix imaging by confocal laser scanning microscopy

In Figure 6, the structure of cell organization, single cells, microcolonies and biofilms, from R. alamii wt and the MSΔGT mutant, were examined after 3 days and 5 days of incubation at 30°C. The constitutive expression of GFP allowed the bacterial cells to be imaged. EPS was labelled with a fluorescent lectin (left lane only). Figure 6 shows typical pictures of the development of R. alamii wt and MSΔGT mutant cells on a membrane, in the presence (2 mg.L−1) or absence of cadmium. At 3 days of growth in the absence of cadmium, the R. alamii wt and MSΔGT mutant colonized the surface, in the form of microcolonies (Figure 6A), or scattered single cells (Figure 6E), respectively. After 5 days of growth, R. alamii wt and MSΔGT formed spread biofilms with a loose architecture (Figure 6B and 6F). However, cadmium already induced the formation of condensed biofilms of both strains, after 3 days of incubation (Figs. 6C and 6G). After 3 and 5 days, in the presence of cadmium, EPS-producing R. alamii wt cells formed globular biofilms expanded in the z direction (Fig. 6D), whereas the MSΔGT mutant formed broad, flat and dense biofilms (Figure 6H). We conclude that the EPS production of R. alamii enabled the spatial propagation of the cells, and the construction of a 3D-spread biofilm matrix. This observation corroborates the contribution of bacterial exopolymers to the biofilm architecture [45], [46]. The localization of EPS in R. alamii biofilms grown in the presence of cadmium did not occur more frequently than in the absence of the metal, which confirms that R. alamii does not modulate EPS production in response to cadmium toxicity.

thumbnail
Figure 6. Confocal laser-scanning micrographs of Rhizobium alamii wt and MSΔGT mutant.

mg.L−1 of Cd2+ in the form of cadmium nitrate, 3 days (left lane) and 5 days (right lane), Bacteria were grown on polycarbonate membranes in a minimal M9 medium supplemented with 2 g.L−1 of glucose, with 0 or 2 mg.L−1 of cadmium following inoculation of the medium with 2.105 cell.mL−1. The bacterial cells were localized by constitutive expression of GFP (in green); EPS was labeled by binding with fluorescent Alexa660-Concanavalin A (in red) in observations performed at 3 days of growth. 5-day biofilms were unlabeled for EPS. A) R. alamii wt after 3 days of growth in the absence of cadmium nitrate. Projections of Z-sections (1 µm) through 5 µm. The scale bar corresponds to 20 µm. Few cells were attached to the membrane surface as a small core. The EPS can be seen in red, surrounding a single cell. B) After 5 days of growth in the absence of cadmium. Projections of Z-sections (1 µm) through 5 µm. The scale bar corresponds to 50 µm. Bacterial cells colonized the membrane, in the form of a loose spherical biofilm. The EPS can be seen in red in some areas of the biofilm (see picture insert projection of Z-sections through 15 µm, the scale bar corresponds to 20 µm). C) R. alamii wt after 3 days of growth in the presence of cadmium nitrate (Cd2+ 2 mg.L−1). Projections of Z-sections (1 µm) through 5 µm. The scale bar corresponds to 20 µm. The cells grew as a sphere. The EPS was localized inside the biofilm. D) R. alamii wt after 5 days of growth in the presence of cadmium nitrate (Cd2+ 2 mg.L−1). Projections of Z-sections (1 µm) through 15 µm. The scale bar corresponds to 20 µm. The cells were organized in the form of fused, expanded spheres. E) MSΔGT mutant cell development on a membrane surface after 3 days of growth, in the absence of cadmium nitrate. Projection of Z-sections (1 µm) through 6 µm. The scale corresponds to 20 µm. The cells adhered to the membrane surface, although as single cells. F) MSΔGT mutant after 5 days of growth in the absence of cadmium. Projections of Z-sections (1 µm) through 17 µm. The scale bar corresponds to 50 µm. The bacterial cells formed a flat, loose biofilm architecture. G) MSΔGT mutant after 3 days of growth in the presence of cadmium nitrate (Cd2+ 2 mg.L−1). Projection of Z-sections (1 µm) through 5 µm. The scale bar corresponds to 50 µm. Biofilms developed in the yz plane. 2 h. MSΔGT mutant after 5 days of growth in the presence of cadmium nitrate (Cd2+ 2 mg.L−1). Projection of Z-sections (1 µm) through 17 µm. The scale bar corresponds to 50 µm. The cells formed a distributed, dense biofilm.

https://doi.org/10.1371/journal.pone.0026771.g006

The view of the role of rhizobia and their EPS being restricted mainly to symbiotic interactions with legumes is insufficient to gain insight into their physiology, and ability to adapt to environmental fluctuations. Altogether, our data has revealed some unique results, related to the metabolic response and lifestyle of R. alamii, and to the function of the EPS in response to cadmium. The belief that EPSs are the main players in metal tolerance of bacteria is persistent in microbiology, mainly in environmental science literature. Our experiments show that the bacterial polysaccharide (EPS) of R. alamii can bind cadmium, although this is not the way that these bacteria use to adapt to the metal toxicity. Changes of lifestyle, from planktonic to biofilm growth, and altering the metabolism, are the means to escape metal toxicity by wt cells as well as by mutant cells that no longer produce EPS. Further studies are being designed to determine the role of R. alamii and its EPS on plant cadmium uptake in the rhizosphere.

Materials and Methods

Bacterial strains and plasmids

Rhizobium alamii wt (described in [24] and avalaible at the Institut Pasteur CNCM I-1809P) and MSΔGT (described in [23] and available upon request), were grown at 30°C in a tenfold diluted tryptic soy broth (DIFCO Laboratories, Detroit, USA), or in a mineral RCV medium [23], at pH 6.8, supplemented with glucose (2 g.L−1) as a carbon source to favour EPS synthesis or in a minimal M9 medium (NaH2PO4 6 g.L−1, KH2PO4 3 g.L−1, NH4Cl 1 g.L−1, NaCl 0.5 g.L−1, MgSO4 1 mM, Thiamine 50 mg.L−1, L-Leucine 0.25 g.L−1, L-Proline 0.25 g.L−1, and glucose 2 g.L−1. Nalidixic acid, kanamycin and tetracycline were respectively used at 50 µg.mL−1, 25 µg.mL−1 and 15 µg.mL−1 for the appropriate antibiotic selection of R. alamii strains.

Rhizobium alamii wild type and its EPS-deficient mutant MSΔGT labelled with green-fluorescent protein (GFP) [23], [47] constitutively expressed the GFP for almost three months in soil, even in the absence of the antibiotic selection pressure.

Cadmium adsorption isotherms

EPS from R. alamii was isolated as described in [23], dialyzed and lyophilized prior to use. Cadmium binding to EPS of R. alamii was performed as described by Nelson et al. [48]. EPS solutions (0.5 to 4 g.L−1, 10 mL) in a MOPS buffer (10−2 M), or CaNO3 (10−2 M), were placed inside dialysis tubing (Cellu-Sep® T3 Membranes, 12-14000 MWCO, Interchim, Montluçon, France) and equilibrated in 10 mL CdNO3 solutions (10−4 to 10−11 M) doped with 109CdCl2 (830 cpm.mL−1, 0.62 nCi final concentration). The pH was adjusted to 6.5±0.2, with NaOH 1N or HCl 1N. Triplicate samples were placed on a rotator at 22°C. Dialysis tubings with no EPS in solution were taken as blanks. Aliquots of the external solutions were taken at 24 and 48 h, and the γ-emissions were measured with a gamma counter.

Tolerance of Rhizobium alamii to cadmium

The wt and MSΔGT strains were grown in a ten-fold diluted tryptic soy broth (TSB/10), or in a RCV supplemented with glucose (2 g.L−1) and various concentrations of cadmium (0 to 133.4 µM; 0 to 15 mg.L−1 of Cd2+). The OD630 nm was measured over time from triplicate samples and reflected planktonic cell concentrations.

The R. alamii wt cells were grown up to an early stationary phase in TSB/10 supplemented with 0 or 2 mg.L−1 of cadmium nitrate. The cells were collected at OD600 nm near to 0.07. Cytosolic cellular extracts were extracted in 50/50 methanol/water, in an ultrasonic bath for 15 min. The pellets were centrifuged at 10000 rpm for 15 min, and the supernatant was analyzed using a Bruker-Daltonics APEXQ 12 Tesla ICR-FT mass spectrometer (Bremen, Germany). The samples were introduced by macrospray infusion, at a flow rate of 120 µL.h−1, ionised in negative and positive electrosprays (ESI), and 256 scans were accumulated over a broadband range of masses (m/z 150–2000). The instrument was externally calibrated on clusters of arginine every measurement day, and the mass spectra were internally calibrated with phtalate diesters in positive ESI, and with fatty acids in negative ESI, thus ensuring a maximum error of 100 ppb. For the ICR-FT MS experiments, three replicates were analyzed for each condition (Figure S2), and the masses common to all three measurements were selected in order to generate a list of potential compounds. From the full set of detected compounds, we focused on those having a suggested attribution for the chemical structure with a difference between the detected and calculated masses of less than ±3 ppm, with detected peaks preferentially confirmed by the presence of 180, 15N or 13C isotopes. The peaks exceeding a threshold signal-to-noise ratio of 3 were exported to peak lists, and were submitted to a metabolite-annotation web interface, MassTRIX (http://masstrix.org). MassTRIX processes the submitted mass peak list by comparing the input experimental masses with all of the compounds recorded in the Kyoto Encyclopedia Genes and Genome (KEGG) chemical compound database, using Rhizobium leguminosarum as the model organism. The MassTRIX annotation of metabolites was used to highlight differences in metabolic pathways between the R. alamii cells incubated with or without Cd2+ 2 mg.L−1. Moreover, the compounds identified by ICR-FT/MS were matched with the KEGG reaction components. A full table of metabolites was added as supporting information Tables S1, S2, S3, S4. The raw data and the position of the potential metabolites in KEGG pathways can be consulted at http://metabolomics.helmholtz-muenchen.de/masstrix2, section “job status”, job access numbers 10031710000016906 and 10031709591316561 for wild type cells [Cd2+] 2 or 0 mg/L in negative mode, and 10031709545915503 and 10031709560715818, for wt cells [Cd2+] 2 or 0 mg/L in positive mode.

The analyses of pyrimidine and purine metabolites, with or without cadmium culture conditions, were performed by ULPC-MS on three independent replicates, which were characterised by similar profiles, indicating a high reproducibility of the three experiments (Figure S3 and Figure S4).

Adhesion of bacterial cells

The R. alamii wt and MSΔGT mutant were statically grown to a stationary phase, in polypropylene tubes containing RCV-medium, supplemented with 2 g.L−1 glucose and cadmium nitrate (0 to 133.4 µM; 0 to 15 mg.L−1 of Cd2+). Bacterial adhesion to solid surfaces and the formation of biofilms were monitored using a crystal violet staining-based protocol adapted from [39]. Breafly, the planktonic bacteria were removed from the tubes. Washing was performed with sterile water. After careful remove of the water, staining was performed with 2 mL of 1% crystal violet solution, 15 min at room temperature. The crystal violet solution was gently removed and successive washings were performed with water. Each tube was inverted and gently tapped on paper towels to remove any excess liquid and allow tubes to air-dry. 2 mL of 100% ethanol was added to each tube and OD595 was measured in a cuvette on a spectrophotometer.

Microscopy

Biofilms of R. alamii wt and MSΔGT mutant were statically grown on a polycarbonate membrane (Millipore 0.2 µm) incubated in a modified M9 medium (Na2HPO4 6 g.L−1, KH2PO4 3 g.L−1, NH4Cl 1 g.L−1, NaCl 0.5 g.L−1, MgSO4 1 mM, thiamine 50 mg.L−1, L-Leucine 0.25 g.L−1, L-Proline 0.25 g.L−1 and glucose 20 g.L−1), supplemented with 0 or 2 mg.mL−1 of cadmium nitrate. 30 mL of M9 medium were inoculated with 2.105 cells of R. alamii wt or MSΔGT mutant, grown to half of the exponential phase in TSB/10, and washed with sterile ultrapure water. EPS labelling with concanavalinA and confocal laser scanning microscopy were performed on an Olympus confocal microscope, as described in Santaella et al. [23].

Data treatment

Statistical analyses (ANOVA) were made using version XV of the Statgraphics Centrion software.

Supporting Information

Figure S1.

Amount of EPS isolated from 5-day old R. alamii cultures (5 mL) in RCV medium (white bars) and in M9 modified medium (grey bars) supplemented with glucose (2 g.L−1). Mean of 3 replicates ± SD.

https://doi.org/10.1371/journal.pone.0026771.s001

(DOC)

Figure S2.

Examples of off-line FTMS measurements.

https://doi.org/10.1371/journal.pone.0026771.s002

(DOC)

Figure S3.

Examples of UPLC chromatographs of cell extracts of R. alamii grown without cadmium to the beginning of the stationary phase.UPLC-PDA, HSS column. Eluent: A: 10 mM ammonium acetate pH 6.4, B: 20% 10 mM ammonium acetate in acetonitrile. Flow: 0.9 mL.min−1 (pressure drop 600–900 bar). Column temperature: 40°C. Injection: 5 µL, partial loop with needle overfill. Wavelength detection 256 nm. Illustration of the reproducibilty of the analyses from three repetitions of independent cell cultures and extracts.

https://doi.org/10.1371/journal.pone.0026771.s003

(DOC)

Figure S4.

Example of UPLC chromatographs of cell extracts of R. alamii grown with 2 mg.L−1 of cadmium to the beginning of the stationary phase. Analysis of three independent cell cultures and extracts.

https://doi.org/10.1371/journal.pone.0026771.s004

(DOC)

Table S1.

Potential metabolites detected by ICR-FT MS negative ESI in R. alamii grown with no cadmium nitrate.

https://doi.org/10.1371/journal.pone.0026771.s005

(XLS)

Table S2.

Potential metabolites detected by ICR-FT MS positive ESI in R. alamii grown with no cadmium nitrate.

https://doi.org/10.1371/journal.pone.0026771.s006

(XLS)

Table S3.

Potential metabolites detected by ICR-FT MS negative ESI in R. alamii grown with 2 mg.L−1 of cadmium nitrate.

https://doi.org/10.1371/journal.pone.0026771.s007

(XLS)

Table S4.

Potential metabolites detected by ICR-FT MS positive ESI in R. alamii grown with 2 mg.L−1 of cadmium nitrate.

https://doi.org/10.1371/journal.pone.0026771.s008

(XLS)

Acknowledgments

We wish to thank Dr. Alain Heyraud (CERMAV, CNRS, Grenoble, FR) for providing us with the NMR spectra of the EPS. We thank Dr Jérôme Balesdent for constructive discussions and his assistance with the radionuclide manipulations.

Author Contributions

Conceived and designed the experiments: CS MS WA. Performed the experiments: MS CS AF CB. Analyzed the data: CS MS PO AF PS-K WA TH. Contributed reagents/materials/analysis tools: PO. Wrote the paper: CS WA.

References

  1. 1. Gracy RW, Talent JM, Kong Y, Conrad CC (1999) Reactive oxygen species: the unavoidable environmental insult? Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis 428: 17–22.
  2. 2. Ma Z, Jacobsen FE, Giedroc DP (2010) Coordination Chemistry of Bacterial Metal Transport and Sensing. Chem Rev 109: 4644–ì4681.
  3. 3. Pagès D, Rose J, Conrod S, Cuine S, Carrier P, et al. (2008) Heavy Metal Tolerance in Stenotrophomonas maltophilia. PLoS ONE 3: e1539.
  4. 4. von Rozycki T, Nies DH (2009) Cupriavidus metallidurans: evolution of a metal-resistant bacterium. Antonie Van Leeuwenhoek 96: 115–139.
  5. 5. Dedieu K, Iuranova T, Slaveykova VI (2006) Do Exudates Affect Cadmium Speciation and Bioavailability to the Rhizobacterium Sinorhizobium meliloti? Environmental Chemistry 3: 424–427.
  6. 6. Pagès D, Sanchez L, Conrod S, Gidrol X, Fekete A, et al. (2007) Exploration of intraclonal adaptation mechanisms of Pseudomonas brassicacearum facing cadmium toxicity. Environmental Microbiology 9: 2820–2835.
  7. 7. Anton A, Grosse C, Reissmann J, Pribyl T, Nies DH (1999) CzcD is a heavy metal ion transporter involved in regulation of heavy metal resistance in Ralstonia sp. strain CH34. J Bacteriol 181: 6876–6881.
  8. 8. Nies DH (2003) Efflux-mediated heavy metal resistance in prokaryotes. FEMS Microbiol Rev 27: 313–339.
  9. 9. Rodrigue A, Effantin G, Mandrand-Berthelot MA (2005) Identification of rcnA (yohM), a nickel and cobalt resistance gene in Escherichia coli. J Bacteriol 187: 2912–2916.
  10. 10. Holmes JD, Richardson DJ, Saed S, Evans-Gowing R, Russell DA, et al. (1997) Cadmium-specific formation of metal sulfide ‘Q-particles’ by Klebsiella pneumoniae. Microbiology 143: 2521–2530.
  11. 11. Blake RC, Choate DM, Bardhan S, Revis N, Barton LL, et al. (1993) Chemical transformation of toxic metals by a Pseudomonas strain from a toxic waste site. 12: 1365–1376.
  12. 12. Volesky B (1987) Biosorbents for metal recovery. Trends in Biotechnology 5: 96–101.
  13. 13. Higham DP, Sadler PJ, Scawen MD (1984) Cadmium-Resistant Pseudomonas putida Synthesizes Novel Cadmium Proteins. Science 225: 1043–1046.
  14. 14. Figueira EM, Gusmão Lima AI, Pereira SIA (2005) Cadmium tolerance plasticity in Rhizobium leguminosarum bv. viciae: glutathione as a detoxifying agent. Canadian Journal of Microbiology 51: 7–14.
  15. 15. Slaveykova VI, Nalini P, Karine D, Toescher D (2010) Role of extracellular compounds in Cd-sequestration relative to Cd uptake by bacterium Sinorhizobium meliloti. Environmental Pollution 158: 2561–2565.
  16. 16. Ron EZ, Minz D, Finkelstein NP, Rosenberg E (1992) Interactions of bacteria with cadmium. Biodegradation 3: 161–170.
  17. 17. Lamelas C, Benedetti M, Wilkinson KJ, Slaveykova VI (2006) Characterization of H+ and Cd2+ binding properties of the bacterial exopolysaccharides. Chemosphere 65: 1362–1370.
  18. 18. Chen JH, Czajka DR, Lion LW, Shuler ML, Ghiorse WC (1995) Trace metal mobilization in soil by bacterial polymers. Environ Health Perspect 103: 53–58.
  19. 19. Czajka DR, Lion LW, Shuler ML, Ghiorse WC (1997) Evaluation of the utility of bacterial extracellular polymers for treatment of metal-contaminated soils: Polymer persistence, mobility, and the influence of lead. Water Research 31: 2827–2839.
  20. 20. Hebbar KP, Gueniot B, Heyraud A, Colin-Morel P, Heulin T, et al. (1992) Characterization of exopolysaccharides produced by rhizobacteria. Applied Microbiology and Biotechnology 38: 248–253.
  21. 21. Skorupska A, Janczarek M, Marczak M, Mazur A, Król J (2006) Rhizobial exopolysaccharides: genetic control and symbiotic functions. Microbial Cell Factories 5: 1–19.
  22. 22. Fujishige NA, Kapadia NN, Hirsch AM (2006) A feeling for the micro-organism: structure on a small scale. Biofilms on plant roots. Botanical Journal of the Linnean Society - Wiley Online Library 150: 79–88.
  23. 23. Santaella C, Schue M, Berge O, Heulin T, Achouak W (2008) The exopolysaccharide of Rhizobium sp. YAS34 is not necessary for biofilm formation on Arabidopsis thaliana and Brassica napus roots but contributes to root colonization. Environ Microbiol 10: 2150–2163.
  24. 24. Alami Y, Achouak W, Marol C, Heulin T (2000) Rhizosphere soil aggregation and plant growth promotion of sunflowers by an exopolysaccharide-producing Rhizobium sp. strain isolated from sunflower roots. Appl Environ Microbiol 66: 3393–3398.
  25. 25. Trinick MJH, P. A. (1995) Formation of nodular structures on the non-legumes Brassica napus, B. campestris, B. juncea and Arabidopsis thaliana with Bradyrhizobium and Rhizobium isolated from Parasponia spp. or legumes grown in tropical soils. Plant and soil 172: 207–219.
  26. 26. Morgante C, Castro S, Fabra A (2007) Role of rhizobial EPS in the evasion of peanut defense response during the crack-entry infection process. Soil Biology and Biochemistry 39: 1222–1225.
  27. 27. Berge O, Lodhi A, Brandelet G, Santaella C, Roncato M-A, et al. (2009) Rhizobium alamii sp. nov., an exopolysaccharide-producing species isolated from legume and non-legume rhizospheres. International journal of systematic and evolutionary microbiology 59: 367–372.
  28. 28. Slaveykova VI, Parthasarathy N, Dedieu K, Toeschera D (2010) Role of extracellular compounds in Cd-sequestration relative to Cd uptake by bacterium Sinorhizobium meliloti. Environmental Pollution 158: 2561–2565.
  29. 29. Suhre K, Schmitt-Kopplin P (2008) MassTRIX: mass translator into pathways. Nucl Acids Res 36: 481–484.
  30. 30. Foster TJ (1983) Plasmid-determined resistance to antimicrobial drugs and toxic metal ions in bacteria. Microbiol Rev 47: 361–409.
  31. 31. Pacheco CC, Passos JF, Castro AR, Moradas-Ferreira P, De Marco P (2008) Role of respiration and glutathione in cadmium-induced oxidative stress in Escherichia coli K-12. Arch Microbiol 189: 271–278.
  32. 32. Mitra RS, Gray RH, Chin B, Bernstein IA (1975) Molecular mechanisms of accommodation in Escherichia coli to toxic levels of Cd2+. – Mitra et al. 121 (3): 1180 – The Journal of Bacteriology 121: 1180–1188.
  33. 33. Lima AIG, Corticeiro SC, Figueira E (2006) Glutathione-mediated cadmium sequestration in Rhizobium leguminosarum. Enzyme and Microbial Technology 39: 763–769.
  34. 34. Pereira SI, Lima AI, Figueira EM (2006) Screening Possible Mechanisms Mediating Cadmium Resistance in Rhizobium leguminosarum bv. viciae Isolated from Contaminated Portuguese Soils. Microbial Ecology 52: 176–186.
  35. 35. Bouchelaghem W, Benloucif MR, Mayoufi M, Benmoussa A, Schenk KJ (2006) Reinvestigation of cadmium diphosphate. Acta Cryst E62: i99–i102.
  36. 36. Aiking H, Stijnman A, van Garderen C, van Heerikhuizen H, van 't Riet J (1984) Inorganic phosphate accumulation and cadmium detoxification in Klebsiella aerogenes NCTC 418 growing in continuous culture. Appl Environ Microbiol 47: 374–377.
  37. 37. Monds RD, Newell PD, Gross RH, O'Toole GA (2007) Phosphate-dependent modulation of c-di-GMP levels regulates Pseudomonas fluorescens Pf0-1 biofilm formation by controlling secretion of the adhesin LapA. Issue Molecular Microbiology Molecular Microbiology 63: 656–ì679.
  38. 38. Karasawa T, Takasawa S, Yamakawa K, Yonekura H, Okamoto H, et al. (1995) NAD(+)-glycohydrolase from Streptococcus pyogenes shows cyclic ADP-ribose forming activity. FEMS Microbiol Lett 130: 201–204.
  39. 39. O'Toole GA, Kolter R (1998) Initiation of biofilm formation in Pseudomonas fluorescens WCS365 proceeds via multiple, convergent signalling pathways: a genetic analysis. Mol Microbiol 28: 449–461.
  40. 40. Whiteley M, Bangera MG, Bumgarner RE, Parsek MR, Teitzel GM, et al. (2001) Gene expression in Pseudomonas aeruginosa biofilms. Nature 413: 860–864.
  41. 41. Teitzel GM, Parsek MR (2003) Heavy Metal Resistance of Biofilm and Planktonic Pseudomonas aeruginosa. Applied and Environmental Microbiology 69: 2313–2320.
  42. 42. Perrin C, Briandet R, Jubelin G, Lejeune P, Mandrand-Berthelot MA, et al. (2009) Nickel promotes biofilm formation by Escherichia coli K-12 strains that produce curli. Appl Environ Microbiol 75: 1723–1733.
  43. 43. Villain-Simonnet A, Milas M, Rinaudo M (2000) A new bacterial polysaccharide (YAS34). I. Characterization of the conformations and conformational transition. International Journal of Biological Macromolecules 27: 65–75.
  44. 44. Zhou W, Wang J, Shena B, Houb Y, Zhang W (2009) Biosorption of copper(II) and cadmium(II) by a novel exopolysaccharide secreted from deep-sea mesophilic bacterium. Colloids and Surfaces B: Biointerfaces 72: 295–302.
  45. 45. Vu B, Chen M, Crawford RJ, Ivanova EP (2009) Bacterial extracellular polysaccharides involved in biofilm formation. Molecules 14: 2535–2554.
  46. 46. Flemming HC, Wingender J (2010) The biofilm matrix. Nat Rev Microbiol 8: 623–633.
  47. 47. Vicré M, Santaella C, Blanchet S, Gateau A, Driouich A (2005) Root Border-Like Cells of Arabidopsis. Microscopical Characterization and Role in the Interaction with Rhizobacteria. Plant Physiol 138: 998–1008.
  48. 48. Nelson YM, Lo W, Lion LW, Shuler ML, Ghiorse WC (1995) Lead distribution in a simulated aquatic environment: effects of bacterial biofilms and iron oxide. Water Research 29: 1934–1944.