8
The gallium(III)salicylidene acylhydrazide complex shows synergistic anti-biolm effect and inhibits toxin production by Pseudomonas aeruginosa Olena Rzhepishevska a, , Shoghik Hakobyan a , Barbro Ekstrand-Hammarström b , Yvonne Nygren b , Torbjörn Karlsson a , Anders Bucht b , Mikael Elofsson a , Jean-François Boily a , Madeleine Ramstedt a, ⁎⁎ a Department of Chemistry, Umeå University, 90187 Umeå, Sweden b Swedish Defense Research Institute (FOI), Cementvägen 20, 90182 Umeå, Sweden abstract article info Article history: Received 3 February 2014 Received in revised form 11 April 2014 Accepted 14 April 2014 Available online 23 April 2014 Keywords: Metal complex Gallium Virulence Biolm Bacteria EXAFS Bacterial biolms cause a range of problems in many areas and especially in health care. Biolms are difcult to eradicate with traditional antibiotics and consequently there is a need for alternative ways to prevent and/or remove bacterial biolms. Furthermore, the emergence of antibiotic resistance in bacteria creates a challenge to nd new types of antibiotics with a lower evolutionary pressure for resistance development. One route to develop such drugs is to target the so called virulence factors, i.e. bacterial systems used when bacteria infect a host cell. This study investigates synergy effects between Ga(III) ions, previously reported to suppress biolm formation and growth in bacteria, and salicylidene acylhydrazides (hydrazones) that have been proposed as antivirulence drugs targeting the type three secretion system used by several Gram-negative pathogens, includ- ing Pseudomonas aerugionosa, during bacterial infection of host cells. A library of hydrazones was screened for: Fe(III) binding, enhanced anti-biolm effect with Ga(III) on P. aeruginosa, and low cytotoxicity to mammalian cells. The metal coordination for the most promising ligand, 2-Oxo-2-[N-(2,4,6-trihydroxy-benzylidene)- hydrazino]-acetamide (ME0163) with Ga(III) was investigated using extended X-ray absorption ne structure spectroscopy as well as density functional theory. The results showed that Ga(III) chelates the hydrazone with 5- and 6-membered chelating rings, and that the Ga(III)ME0163 complex enhanced the antibiolm effect of Ga(III) while suppressing the type three secretion system in P. aeruginosa. The latter effect was not observed for the hydrazone alone and was similar for Ga(III)citrate and Ga(III)ME0163 complexes, indicating that the inhibition of virulence was caused by Ga(III). © 2014 The Authors. Published by Elsevier Inc. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/3.0/). 1. Introduction Biolm is a bacterial growth mode characterized by formation of organized cell clusters coated with extracellular polymeric substances. According to National Institute of Health biolms are responsible for up to 80% of all infections, which often become chronic and difcult to eradicate [1]. Extracellular polymeric substances enable bacteria to at- tach to surfaces e.g. prosthetics, catheters, and tissues, and protect them from antibiotics and host immunity. Another explanation for biolm persistence and resistance to antibiotics is the slower metab- olism and distinct gene expression of sessile bacteria in biolms compared to their planktonic, free-swimming, counterparts [24]. Among bacteria that cause biolm-associated infections Pseudomonas aeruginosa is one of the most ubiquitous. P. aeruginosa is an opportunistic human pathogen causing many types of infection e.g. in burn wounds [5], and immunocompromised patients [6,7]. It causes life-threatening pneumonia in cystic brosis patients [8] and is one of the most common pathogens infecting foot ulcers in patients with diabetes mellitus, resulting in frequent foot and leg amputation [9]. Multi-drug resistant P. aeruginosa strains represent a serious problem in healthcare with no effective antibiotic treatments available. Thus, the need for new antimi- crobial drugs and targets for treatment of these multi-resistant strains is urgent and should not be underestimated. Pathogens such as P. aeruginosa, Francisella tularensis, and Mycobac- terium tuberculosis are very sensitive to Fe limitation since Fe plays an important role in the active center of many of their enzymes and respi- ratory proteins [1012]. Consequently, limiting Fe(III) availability is an efcient method used by the human body to control infection [13,14]. In mammalian body uids the level of free iron is very low as it is Journal of Inorganic Biochemistry 138 (2014) 18 Corresponding author. Tel.: +46 90 786 6669; fax: +46 90 786 7655. ⁎⁎ Corresponding author. Tel.: +46 90 786 6328; fax: +46 90 786 7655. E-mail addresses: [email protected] (O. Rzhepishevska), [email protected] (M. Ramstedt). http://dx.doi.org/10.1016/j.jinorgbio.2014.04.009 0162-0134/© 2014 The Authors. Published by Elsevier Inc. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/3.0/). Contents lists available at ScienceDirect Journal of Inorganic Biochemistry journal homepage: www.elsevier.com/locate/jinorgbio

The gallium(III)–salicylidene acylhydrazide complex shows synergistic anti-biofilm effect and inhibits toxin production by Pseudomonas aeruginosa

Embed Size (px)

Citation preview

Page 1: The gallium(III)–salicylidene acylhydrazide complex shows synergistic anti-biofilm effect and inhibits toxin production by Pseudomonas aeruginosa

Journal of Inorganic Biochemistry 138 (2014) 1–8

Contents lists available at ScienceDirect

Journal of Inorganic Biochemistry

j ourna l homepage: www.e lsev ie r .com/ locate / j inorgb io

The gallium(III)–salicylidene acylhydrazide complex shows synergisticanti-biofilm effect and inhibits toxin production byPseudomonas aeruginosa

Olena Rzhepishevska a,⁎, Shoghik Hakobyan a, Barbro Ekstrand-Hammarström b, Yvonne Nygren b,Torbjörn Karlsson a, Anders Bucht b, Mikael Elofsson a, Jean-François Boily a, Madeleine Ramstedt a,⁎⁎a Department of Chemistry, Umeå University, 90187 Umeå, Swedenb Swedish Defense Research Institute (FOI), Cementvägen 20, 90182 Umeå, Sweden

⁎ Corresponding author. Tel.: +46 90 786 6669; fax: +⁎⁎ Corresponding author. Tel.: +46 90 786 6328; fax: +

E-mail addresses: [email protected]@chem.umu.se (M. Ramstedt).

http://dx.doi.org/10.1016/j.jinorgbio.2014.04.0090162-0134/© 2014 The Authors. Published by Elsevier Inc

a b s t r a c t

a r t i c l e i n f o

Article history:Received 3 February 2014Received in revised form 11 April 2014Accepted 14 April 2014Available online 23 April 2014

Keywords:Metal complexGalliumVirulenceBiofilmBacteriaEXAFS

Bacterial biofilms cause a range of problems in many areas and especially in health care. Biofilms are difficult toeradicate with traditional antibiotics and consequently there is a need for alternative ways to prevent and/orremove bacterial biofilms. Furthermore, the emergence of antibiotic resistance in bacteria creates a challengeto find new types of antibiotics with a lower evolutionary pressure for resistance development. One route todevelop such drugs is to target the so called virulence factors, i.e. bacterial systems used when bacteria infect ahost cell. This study investigates synergy effects between Ga(III) ions, previously reported to suppress biofilmformation and growth in bacteria, and salicylidene acylhydrazides (hydrazones) that have been proposed asantivirulence drugs targeting the type three secretion system used by several Gram-negative pathogens, includ-ing Pseudomonas aerugionosa, during bacterial infection of host cells. A library of hydrazones was screened for:Fe(III) binding, enhanced anti-biofilm effect with Ga(III) on P. aeruginosa, and low cytotoxicity to mammaliancells. The metal coordination for the most promising ligand, 2-Oxo-2-[N-(2,4,6-trihydroxy-benzylidene)-hydrazino]-acetamide (ME0163) with Ga(III) was investigated using extended X-ray absorption fine structurespectroscopy as well as density functional theory. The results showed that Ga(III) chelates the hydrazone with5- and 6-membered chelating rings, and that the Ga(III)–ME0163 complex enhanced the antibiofilm effect ofGa(III) while suppressing the type three secretion system in P. aeruginosa. The latter effect was not observedfor the hydrazone alone and was similar for Ga(III)–citrate and Ga(III)–ME0163 complexes, indicating that theinhibition of virulence was caused by Ga(III).

© 2014 The Authors. Published by Elsevier Inc. This is an open access article under the CC BY license(http://creativecommons.org/licenses/by/3.0/).

1. Introduction

Biofilm is a bacterial growth mode characterized by formation oforganized cell clusters coated with extracellular polymeric substances.According to National Institute of Health biofilms are responsible forup to 80% of all infections, which often become chronic and difficult toeradicate [1]. Extracellular polymeric substances enable bacteria to at-tach to surfaces e.g. prosthetics, catheters, and tissues, and protectthem from antibiotics and host immunity. Another explanationfor biofilm persistence and resistance to antibiotics is the slowermetab-olism and distinct gene expression of sessile bacteria in biofilmscompared to their planktonic, free-swimming, counterparts [2–4].

46 90 786 7655.46 90 786 7655.(O. Rzhepishevska),

. This is an open access article under

Among bacteria that cause biofilm-associated infections Pseudomonasaeruginosa is one of themost ubiquitous. P. aeruginosa is an opportunistichuman pathogen causingmany types of infection e.g. in burnwounds [5],and immunocompromised patients [6,7]. It causes life-threateningpneumonia in cystic fibrosis patients [8] and is one of the most commonpathogens infecting foot ulcers in patients with diabetes mellitus,resulting in frequent foot and leg amputation [9]. Multi-drug resistantP. aeruginosa strains represent a serious problem in healthcare with noeffective antibiotic treatments available. Thus, the need for new antimi-crobial drugs and targets for treatment of these multi-resistant strains isurgent and should not be underestimated.

Pathogens such as P. aeruginosa, Francisella tularensis, andMycobac-terium tuberculosis are very sensitive to Fe limitation since Fe plays animportant role in the active center of many of their enzymes and respi-ratory proteins [10–12]. Consequently, limiting Fe(III) availability is anefficient method used by the human body to control infection [13,14].In mammalian body fluids the level of free iron is very low as it is

the CC BY license (http://creativecommons.org/licenses/by/3.0/).

Page 2: The gallium(III)–salicylidene acylhydrazide complex shows synergistic anti-biofilm effect and inhibits toxin production by Pseudomonas aeruginosa

2 O. Rzhepishevska et al. / Journal of Inorganic Biochemistry 138 (2014) 1–8

bound to proteins such as transferrin and lactoferrin that have high af-finity to Fe(III) (binding constants log K ~20) [13,15]. The levels offree Fe in different fluids vary between individuals and time pointsbut in a study of 50 healthy adults it was found that the non-proteinbound Fe concentration was between 0 and 18 μM in plasma,0–14 μM for amniotic fluid, 0–11 μM for bronchoalveolar lavage and0–0.59 μM for brain tissue with median values of 1.8 μM, 1.45 μM,0.50 μM and 0.08 μM respectively [16]. Limitation of Fe has beenshown to prevent and disperse biofilms in P. aeruginosa. Bacteria inthe depth of a biofilm are starved due to Fe restriction, and dispersalhas been suggested through induction of twitching motility [17,18].Fe(III) restriction can also be achieved through the addition of Fe(III) an-tagonists. Ga(III) is an Fe(III) mimetic that is proposed to be taken upinto bacterial cells through Fe(III) transport mechanisms such assiderophores and thereafter interfere with metabolic systems requiringFe(III) [19]. Exposure of P. aeruginosa to Ga(III) has been shown to inhib-it bacterial growth and biofilm formation [20–22].

The hydrazone substances, investigated, here have in common achelating motif in the center of the molecule that can bind metal ionssuch as Fe(III) and Ga(III) (Fig. 1). Substances with similar chelatingmotif have been tested as Fe(III) chelators for treatment of Fe(III)overload in patients due to their high binding constants to Fe[23–25]. Additionally, hydrazones were shown to inhibit type IIIsecretion system (T3SS) in a range of Gram-negative bacteria [26].T3SS is a protein complex used for delivery of bacterial toxins intothe host cell, and it is a virulence factor that enables the pathogento infect a host and cause disease. Inhibiting T3SS prevents toxindelivery and, hence, makes bacteria less harmful. The mechanismof action for anti-virulence hydrazones is still to a large extentunknown. Recent studies have shown that these compounds bindto several proteins, suggesting multiple targets in the bacterial cell[26,27]. Fe chelation by hydrazones has been shown to inhibitgrowth and T3SS in sexually transmitted pathogens (Chlamydiatrachomatis and Neisseria gonorrhoeae) [28]. The involvement of Fein the expression of T3SS in these species is unclear while in other spe-cies, Fe depletion stimulates the expression of T3SS through genetic cas-cades [29]. Hydrazones have also been described to restrict Fe frommammalian tumor cells and it was observed that treatment of tumorcells with a combination of hydrazones and Ga(III) enhanced the effect,indicating possible synergies [30–34].

In this study we screened Fe(III) chelating hydrazones, describedearlier [26], in the presence of Ga(III) ions and show that the antibacte-rial and anti-biofilm effects were enhanced. The presence of a Ga(III)–hydrazone complex was evidenced from extended X-ray absorption

NH2

NH

N

OH

OH OHO

O

NH

N

OH

OH OHO

OH

a)

b)

Fig. 1. Chemical structure of two metal chelating hydrazone compounds: a) 2-Oxo-2-[N-(2,4,6-trihydroxy-benzylidene)-hydrazino]-acetamide (ME0163), b) 3-(4-Hydroxy-phenyl)-propionic acid (2,4,6- trihydroxy-benzylidene-)hydrazide (ME0161). A full listof structures for the screened compounds can be found in Dahlgren et al. [26].

fine structure (EXAFS) spectroscopy, which shows that Ga(III) binds tothe chelating motif in the hydrazone.

2. Methods

2.1. Chemicals

Chemicals were purchased from Sigma-Aldrich and used withoutpurification unless otherwise stated. Hydrazones were synthesizedaccording to Dahlgren et al. [26] and Nordfelth et al. [35]. All solu-tions were prepared from deionized and boiled water (resistance =18.2 MΩ) at an ionic strength of 0.1 M NaCl (Merck p.a., dried at 453 K).A 30 mM HCl solution in 0.1 M NaCl was made from concentrated HCl(37% Aldrich) and standardized against tris(hydroxymethyl)aminomethane (Trizma base). A 10 mM NaOH solution in 0.1 mM NaCl(degassed with N2(g)) was made from a 50% NaOH solution and stan-dardized against the standardized HCl solution. A 10.6 mM Ga(NO3)3solution was prepared from Ga(NO3)3xH2O (Aldrich) in 0.1 M NaCl and40.5 mM NaOH, giving Ga(III) in the form of the soluble Ga(OH)4− ion.The exact concentrationofGa(III)was determinedbyusingAtomAbsorp-tion Spectrometry (Perkin Elmer AAS 3110).

2.2. Solubility screening

Hydrazone precipitation in aqueous solutionwas screened to identi-fy which compounds would be soluble during the experimental condi-tions for the study. For this, 100 μM solutions of hydrazones in ISO-SENSITEST medium were incubated at room temperature for 24 h andthe turbidity of the solutions was measured at 600 nm. Screening forFe(III) binding can be found in supporting material.

2.3. Bacterial growth conditions and biofilm experiments

P. aeruginosa, PAO1,was routinely cultured on blood agar plates. ISO-SENSITEST broth and tryptic soy broth (TSB) were used when growingbacteria in liquid cultures. Iron-free medium was prepared accordingto Rzhepishevska et al. [21] with small modifications. Standard ISO-SENSITEST (100%)wasused and the concentrations of the trace elementsolutions were adjusted to the following final concentration: 0.03 mMMgSO4; 0.08 μM ZnCl2; 0.01 μM CuSO4; 0.05 μM MnCl2; 0.25 μM CaCl2.

Minimal inhibitory concentration (MIC) that suppresses 90%of bacterial growth (MIC90) of Ga(III) 2-Oxo-2-[N-(2,4,6-trihydroxy-benzylidene)-hydrazino]-acetamide (Ga(III)–ME0163) was mea-sured as described by Rzhepishevska et al. [21] and inhibition ofP. aeruginosa biofilm formation was assessed by two methods. Methodone: P. aeruginosa PAO1 expressing green fluorescence protein (GFP)was used for screening biofilm inhibition [26]. Bacteria were culturedin multi-well plates with shaking in the presence of Ga–hydrazonecomplexes. Growth was measured as culture absorbance at 600 nm(OD600) after the medium with planktonic bacteria was carefullyremoved. The amount of biofilm was measured as a fluorescent signalfrom GFP at 515 nm. No washing step was included to avoid loss ofloosely attached biofilm. Method two: to confirm screening resultswe used a standard crystal violet staining assay. Briefly, 1 ml of ISO-SENSITEST broth containing Ga(III)–citrate, Ga(III)–ME0163, andME0163, was inoculated with overnight culture of P. aeruginosa PAO1and incubated in a humidified shaker at 37 °C. Crystal violet solution(0.1%) was added to each well of a 24-well plate, incubated for 10 min,rinsed with phosphate buffered saline (PBS) and air dried. Crystal violetwas dissolved in 33% acetic acid and absorbance of the solutions wasmeasured at 595 nm. The biofilm assay was done in 20% ISO-SENSITEST with the addition of tetracycline 10 μM to support the plas-mid for GFP expression.

Page 3: The gallium(III)–salicylidene acylhydrazide complex shows synergistic anti-biofilm effect and inhibits toxin production by Pseudomonas aeruginosa

3O. Rzhepishevska et al. / Journal of Inorganic Biochemistry 138 (2014) 1–8

2.4. Exposure of mammalian cell lines to different hydrazones

To evaluate cytotoxicity and the inflammatory response of mam-malian cells to 4-Methyl-[1,2,3]thiadiazole-5-carboxylic acid (6-bromo-2-hydroxy-3-methoxy-benzylidene)-hydrazide (ME0150),3-(4-Hydroxy-phenyl)-propionic acid (2,4,6-trihydroxy-benzylidene-)hydrazide (ME0161), ME0163 and 3-Methoxy-benzoic acid (3,5-di-tert-butyl-2-hydroxy-benzylidene)-hydrazide (ME0184), we usedthree different cell lines: BEAS-2B (human bronchial epithelial cells),A549 (human alveolar epithelial cells) and L929 (mouse fibroblastcells). BEAS-2B cells were cultured in serum-free bronchial epithelialcell basal medium with supplements (complete medium, BEGMCambrex,Verviers, Belgium). A549 and L929 cells were grown as previ-ously described [21]. Cell viability/cytotoxicity was measured using theAlamarBlue assay (Serotec Scandinavia, Kidlington, United Kingdom)[21]. Stock solutions of ME0150, ME0161, ME0163 and ME0184 werediluted in culture medium to a final concentration of 10 μM, 50 μM,100 μM and 200 μM and added to the cells in 8 replicates. Pro-inflammatory cell response was analyzed by monitoring the releaseof interleukin 8 (IL-8), and interleukin 6 (IL-6) using enzyme-linked im-munosorbent assay (ELISA R&D Systems, Abingdon, United Kingdom)[21].

2.5. EXAFS spectroscopy data collection and analysis

Samples for EXAFS were prepared by adding a high pH 10.6 mMGa(OH)4− solution and 10 mM NaOH solution into a solution ofhydrazone in 0.1 M NaCl under N2 atmosphere and stirring to give afinal concentration of 1.74–1.95 mM (ratio Ga:ME0163 1.06:1). ThepH was adjusted using 30 mM HCl, which gave solutions with [Ga(III)]= 2.07 mM and pH = 10.6 (N1), [Ga(III)] = 1.89 mM and pH = 8.9(N2), and [Ga(III)] = 1.84 mM and pH = 7.9 (N3). The samples wereequilibrated for five days before EXAFS analysis. An aqueous paste ofamorphous Ga(OH)3(s) [36], an aqueous solutions of Ga(III)–citrate[Ga(C6H5O7)23−] [37], andGa(OH)4− (10.6mMGa; pH 11.3)were select-ed as EXAFS references. More detailed EXAFS method description anddata analysis can be found in supporting material.

2.6. Density functional theory (DFT) calculations

A geometry optimization of the neutrally-charged Ga(OH)21+–

H3L1−·H2O complex (where L−4 is the completely deprotonatedME0163 ligand and Ga(III) is in 5- and 6-membered chelate coordina-tion) was carried out at the B3LYP/6-31G level of theory [38]. A watermolecule was added to this complex to complete the octahedral coordi-nation shell of Ga(III). Calculations were carried out using Gaussian 09[39].

2.7. Induction and analysis of type three secretion system (T3SS)

T3SSwas induced by lowcalciumgrowth conditions (TSB brothwithEGTA and 5 mM MgCl2) as described by Lee et al. [40]. P. aeruginosaPAO1 was grown to OD600 of 1.9, bacteria were pelleted and culturesupernatant (1.8 ml) was used for analysis of the secreted proteins.Secreted proteins were precipitated from supernatant by 10% trichloro-acetic acid, washed with acetone and re-suspended in 100 μl loadingbuffer for electrophoresis [33,41,42]. At the same time, aliquots of theculture containing both bacteria and the supernatant (100 μl) were col-lected and mixed with the loading buffer to analyze total (expressedand secreted) protein. Protein fractions were separated on NuPAGEready-made gradient gels (Novex, Life Technologies) and ExoS protein,a marker of T3SS expression and secretion, was detected usingWesternblot with ExoS antibody. Total protein fraction was additionally ana-lyzed using FliC (H7) antibody as a control of protein expression inthe presence of Ga(III) complexes at 50 μM (1:1 complex). All experi-mental conditions (cultures, optical density measurements and

secreted protein analysis) were performed in triplicates. Total proteinwas analyzed in duplicates.

3. Results and discussion

3.1. Hydrazone Fe(III) binding ability

A hydrazone collection of 54 compounds, obtained by statistical mo-lecular design and subsequent synthesis [26], was screened with excesshydrazone to assess Fe(III) binding (Table S2 in supporting material).For most hydrazones, Fe(III) complexes were formed to different ex-tents. For example, in the 1:2 Fe:hydrazone mixture used, ME0161gave a 1:1 ratio between free ligand and complex indicating possibleformation of a 1:1 Fe:hydrazone complex. However, no Fe(III) complexwas detected for the hydrazoneME0163 under the experimental condi-tions used.

3.2. Antibiofilm and antibacterial effects of Ga(III)–hydrazone complexes

Since complex formationwith Fe(III) varied for different hydrazones(Table S2) it could be expected to vary also for Ga(III). Hence in the nextstep the hydrazone library was screened for the most efficient Ga(III)complexwith respect to anti-biofilm effects on P. aeruginosa. First, com-pounds with good solubility in aqueous media were selected as we ob-served that some hydrazones precipitated in these conditions.Thereafter, 1:1 Ga(III):hydrazone mixtures at different concentrationswere studied for effect on growth and biofilm formation of P. aeruginosa(Fig. 2 and Fig. S1). A solution with Ga(III)–citrate complexes was usedas a reference since it has previously been reported to have both antibac-terial and antibiofilm properties [21]. A large inhibition of planktonicgrowth was seen for ME0150 and ME0184 at concentrations higherthan 6 μM (Fig. 2). An anti-biofilm effect could be observed for most ofthe tested compounds already at 3 μM of Ga(III)–hydrazone althoughthe effect was similar to that obtained with Ga(III)–citrate. ME0161and ME0163 showed the largest antibiofilm effect at several differentconcentrations, displayed an enhanced antibiofilm effect compared tothe Ga(III)–citrate control, and had little effect on growth (Figs. 1 and2). The increased activity of ME0163 with Ga(III) suggested that itcould formmetal complexes even if an Fe(III) complexwas not detectedin the aforementioned Fe(III) binding assay.

3.3. Cytotoxic effects on human cells

For any application in medicine it is important that the substance isnot toxic to mammalian cells. The hydrazones alone have previouslybeen screened for toxicity to HeLa cells [26], but we were interested tolook into the toxicity of the hydrazones with and without Ga(III) toepithelial cells (that could be expected to become exposed to thesesubstances were they to be used topically or on medical devices).Consequently, we studied cytotoxicity and inflammatory response onhuman bronchial epithelial cells (BEAS-2B), alveolar epithelial cells(A549) and mouse fibroblast cells (L929 — a common reference cellline). CompoundsME0150 andME0184 that had a large effect on bacte-rial growthwere also toxic tomammalian cells (Fig. S2) andwere there-fore excluded from further investigations. However, compoundsME0161 and ME0163 in the presence of gallium were not found to becytotoxic (Fig. S3) nor produce inflammatory responses when testedon bronchial or alveolar cells (data not shown). No toxicity to bronchialepithelial cells was seen even in the absence of serum, which is an im-portant observation since serum has been reported to protect cellsfrom antibiotic toxicity by metal ions in vitro [43].

3.4. Antibiofilm and antibacterial effects of the complex Ga(III)–ME0163

After the screening we focused on ME0163 with Ga(III) as it hadbetter anti-biofilm properties compared to Ga(III)–citrate, exhibited

Page 4: The gallium(III)–salicylidene acylhydrazide complex shows synergistic anti-biofilm effect and inhibits toxin production by Pseudomonas aeruginosa

Fig. 2. The combination effect of Ga(III) and different hydrazones (ratio 1:1, full names inTable S1) on (from left) biofilm formation (red bars) and growth of PAO1 in solution (bluebars) tested at a concentration of a) 25 μM, b) 12.5 μM, c) 6.12 μM, and d) 3 μM in 20% ISO-SENSITEST. Biofilmwas judgedbyfluorescence of greenfluorescence protein expressed bybacteria. “Control” is no Ga(III) or hydrazone and “Ga cit” corresponds to samples withGa(III)–citrate (previously shown to inhibit both growth and biofilm formation [21]).(For interpretation of the references to color in this figure legend, the reader is referredto the web version of this article.)

a)

b)

c)

Fig. 3. a) Biofilm measured by crystal violet staining in 20% medium (from left): bluecontrol, green ME0163, red Ga(III)–citrate, orange Ga(III)–ME0163, y-axis shows ratio tobaseline control (no bacteria). b) Growth curves of PAO1 in the presence of Ga(III)–citrate(red circles), ME0163 (inverted green triangles) and Ga(III)–ME0163 (orange triangles),control (blue squares), empty symbols 12.5 μM and filled symbols 50 μM. All in 20% ISO-SENSITEST. c). Growth curves of PAO1 in 100% ISO-SENSITEST (empty symbols) and che-lated, iron-free ISO-SENSITEST (filled symbols) at 50 μM of Ga(III)–citrate (red circles),Ga(III)–ME0163 (orange triangles), ME0163 (inverted green triangles), control (bluesquares). (For interpretation of the references to color in this figure legend, the reader isreferred to the web version of this article.)

4 O. Rzhepishevska et al. / Journal of Inorganic Biochemistry 138 (2014) 1–8

no cytotoxic effect on mammalian cells, and since the solubility ofME0163 enabled detailed thermodynamic studies of protonation andmetal binding (reported in Hakobyan et al. [44]). First, the data fromthe biofilm screeningwas confirmed using a secondmethod, i.e. the tra-ditional crystal violet assay. Both biofilm assays showed that Ga(III)–ME0163 inhibited biofilm formation more efficiently at low concentra-tions than Ga(III)–citrate did, while ME0163 alone had very little effecton biofilm formation (Fig. 3a). Bacterial growth curves showed thatGa(III)–ME0163 inhibited growth of P. aeruginosa more efficientlythan Ga(III)–citrate at a concentration of 50 μM (Fig. 3b) while at 12.5μM the inhibition was comparable between the two Ga(III) complexes.These results are consistent with previous studies of Ga(III) showingthat lower concentrations of Ga(III) are needed to prevent biofilmformation than to inhibit growth [21]. Thus, the results indicate thatwhen used together with ME0163, the anti-biofilm action of Ga(III) isachieved at lower concentrations. At higher concentrations, e.g. 50 μM,the difference in biofilm inhibition becomes masked by growthinhibition.

Since both compounds are described as Fe antagonists and themedi-umdiluted to 20%will have a low amount of Fe, we also investigated the

antibacterial effect in standard (100%) medium, that contains ~6–7 μMFe, and standard medium treated with Chelex-100, a resin that bindsfree Fe. In standard ISO-SENSITEST MIC90 for both Ga(III)–ME0163 andGa(III)–citrate was 200 μM, but in Fe–free ISO-SENSITEST, MIC90 was100 μM for Ga(III)–ME0163 and 200 μM Ga(III)–citrate. In Fe-free

Page 5: The gallium(III)–salicylidene acylhydrazide complex shows synergistic anti-biofilm effect and inhibits toxin production by Pseudomonas aeruginosa

5O. Rzhepishevska et al. / Journal of Inorganic Biochemistry 138 (2014) 1–8

conditions 50 μM of ME0163 alone (Fig. 3c) inhibited growth ofP. aeruginosa to some extent, while it had no effect on growth in 20%ISO-SENSITEST (Fig. 3b). In both Fe restricted conditions (Fig. 3band c) Ga(III)–ME0163 suppressed bacterial growth, indicating a syner-gistic inhibitory effect between Ga(III) ions and ME0163, especially inthe Chelex-treated medium where Ga(III)–citrate had almost no effecton growth. An explanation for this lack of inhibition fromGa(III)–citratecould be that metabolic pathways for bacteria growing in chelatedmedia may differ. Together with iron, Chelex-100 also binds organiccompounds and concentrations of certain amino acids can be decreasedup to 80% in growth media due to chelation [45]. This may cause bacte-ria to use another energy source (e.g. glucose) and express different en-zymes. The alternative enzymes may be less affected by Ga(III) therebydecreasing bacterial sensitivity to gallium [21].

The synergistic or additive effect of a Ga(III)–ME0163 complex couldbe accounted for in several ways: i) the presence of Ga(III) and forma-tion of a Ga(III) complex solubilizes Ga(III) and/or the hydrazone to alarger extent or thereby increase the active concentration in solutionfor both substances, ii) thehydrazone andGa(III) both act on Fe(III) sen-sitive systems (e.g. enzymes) and therefore the effect is enhancedwhenboth substances are present, or iii) the complex has an increased effectcaused by differences in uptake or target. The enhanced synergy effectin Fe-free conditions suggests that the first explanation is unlikely asthe solubility should be similar or higher in the presence of Fe(III)since complexation of ME0163 with Fe(III) as well as Ga(III) wouldshift the solubility equilibrium towards soluble species. The secondand third explanation seems more probable. Explanation ii) is possiblesince Fe(III) sensitive enzymes could be rescued at some base level ofFe(III) in the medium, thus suggesting a reason for the difference be-tween 20%and 100%media, however, this does not clarify the differencebetween Ga complexes. The difference observed between Ga(III)–cit-rate and Ga(III)–ME0163 in chelated medium could be accounted forby hypothesis iii) and/or that the bacteria change carbon source inChelex-treated medium. We showed previously that P. aeruginosagrown on glucose expresses approximately 4 times less pyoverdinethan P. aeruginosa grown on Casamino acids [21]. Consequently, ifP. aeruginosa loses the potent siderophore pyoverdine and free Fe(III)is sequestered by ME0163, it may experience iron starvation. Whilethe mechanism of this process is outside the scope of this article, it isclear that ME0163 is a successful ligand for Ga(III) potentiating its anti-bacterial and anti-biofilm effects in both types of medium, and that theeffect on biofilm occurs at lower concentrations than the effect on bac-terial growth. Furthermore, considering that the 20% medium had freeFe levels corresponding to those measured in biological fluids, we can

b)a)

Fig. 4. a) Speciation diagramcalculated from formation constants [44] for EXAFS sampleN2 contcomplex showing distances from the DFT calculations.

assume that the complex would still be effective in most biologicalfluids.

3.5. EXAFS analysis and DFT calculations of complex

The biological effects observed did not necessarily have to be a resultof a Ga(III)–ME0163 complex but could also have been a combined ef-fect of the two substances acting alone on the bacterial cell, especiallysince no Fe(III)–ME0163 complex was detected. However, a parallelstudy [44] showed strong complex formation at physiological pH(Fig. 4a), with hydrazone binding constants to Ga(III) similar to that ofEDTA. While those experiments were carried out under more diluteconditions than those considered in this work, the thermodynamicmodel generated in that study ought to be applicable to a wide rangeof concentrations. As concentrated systems can generate multinuclearspecies and/or solid phases that do not occur under more dilute sys-tems, we made use of EXAFS and DFT calculations to confirm that i)Ga(III)–ME0163 complexation was effectively achieved and ii) to pro-vide complementary characterization of the bonding environment ofGa(III) and ME0163.

EXAFS data for a sample at pH 10.6 (N1) was successfully modeledwith only a first coordination shell consisting of 3.9 O/N at a distanceof 1. 84 Å in accordance with the Ga–O distances in Ga(OH)4− (Fig. 5;Table S3). Thus, at this pH there was no or only minor complex forma-tion between Ga(III) and the hydrazone ligand and Ga(III) exists as asoluble hydroxo-complex, in agreement with the thermodynamicmodel in Hakobyan et al. [44] (Fig. 4a). For samples at pH 8.9 and 7.9(N2 and N3) a model with 1 O/N distance in the first shell resulted ina poor fit indicating a distorted first shell or that several Ga(III) speciesare present in the samples. The fit was significantly improved by intro-ducing a second Ga–O/N distance and the two obtained distances(1.83–1.85 Å and 1.95 Å) were in accordance with Ga–O distances infour-coordinated Ga(OH)4− and in six-coordinated Ga(NO)3(aq). Thecoordination number (CN) obtained for the longer Ga–O/N path washigher in both of these samples suggesting that 6-coordinated Ga(III)species dominated (Table S3) and illustrating a large change in Ga(III)speciation as pH is decreased. Furthermore, at the lower–pH samplesGa–C/N interactions were detected in the second coordination shellshowing that Ga(III) forms inner-sphere complexes with the hydrazoneligand. The Ga–C/N distances obtained (2.95–2.96 Å) are fairly close toGa–C distances in Ga(III)–citrate (2.73–2.93 Å), where Ga(III) is com-plexed by both 5- and 6-membered chelate rings [37], thus in a coordi-nationmotif similar to that expected for the Ga(III)–hydrazone complex(based on crystal structure of a Cu–hydrazone complex [46]) and,

aining1.89mMGa(III); b) geometry of a representative neutrally-chargedGa(III)–ME0163

Page 6: The gallium(III)–salicylidene acylhydrazide complex shows synergistic anti-biofilm effect and inhibits toxin production by Pseudomonas aeruginosa

2 4 6 8 10 12

k3(k

)

k (Å-1)

I

II

III

IV

V

VI

a)

0 1 2 3 4 5 6F

ou

rier

Tra

nsf

rom

Mag

nit

ud

eR (Å)

I

IV

II

V

III

VI

b)

χ

Fig. 5. a) k3-weighted EXAFS spectra (solid lines) with fit results (dotted lines) and b) corresponding Fourier transforms (solid lines) with fit results (dotted lines) for the Ga-hydrazonesamples compared with references; (I) Ga(OH)4− (II) N1; 2.07 mMGa(III) and pH 10.6, (III) N2; 1.89mMGa(III) and pH 8.9 (IV) N3; 1.84mMGa(III) and pH 7.9 (V) Ga(OH)3(s) [36] and,(VI) Ga(III)–citrate [37]. Vertical dashed lines highlight the peak position of the second oscillation in k-space for Ga(OH)4− and peak positions of the second coordination shells in the FT forGa(III)–citrate and Ga(OH)3(s).

6 O. Rzhepishevska et al. / Journal of Inorganic Biochemistry 138 (2014) 1–8

again, in accordance with Hakobyan et al. [44]. Since the only possi-ble ligands available to coordinate Ga(III) in the EXAFS solutions wereH2O, OH− and ME0163, these data clearly indicate a direct interactionbetween Ga(III) and the hydrazone in the lower–pH samples.

In an effort to constrain our understanding of this complex further,DFT calculations were performed. We chose to study the neutrally-charged Ga(OH)21+–H3L1−·H2O complex (where L−4 is the completelydeprotonated ME0163 ligand and Ga(III) is in 5- and 6-membered che-late coordination, Fig. 4b), well noting that species of different Ga(III)-hydrolysis and ligand-protonation states can very well give rise to dif-ferent sets of atomic distances. The values for this neutrally-chargedspecies are therefore only representative of a range of values that canbe experimentally detected in a mixture of species. The resulting com-plex had a first coordination shell consisting of six Ga–O/N distancesin the 1.85–2.11 Å range, with an average at 2.00 Å, thus comparableto EXAFS results. A second Ga–N distance lies at 2.90 Å, and a secondshell Ga–C distances in the 2.84–3.46 range, again comparable withEXAFS results.

In addition to the Ga(III)–ME0163 complex, EXAFS also providedevidence forGa–Ga interactionswith distances that are highly compara-ble to those occurring in Ga(OH)3(s) (Table S3). Formation constantsfrom Hakobyan et al. [44] suggest that small amounts of GaOOH(s)should be formed at pH lower than 8.5 at these Ga(III) concentrations(Fig. 4a). However, it is possible that an amorphous Ga(OH)3(s) phasewas formed with a different solubility to the crystalline GaOOH(s)used in speciation determinations in Hakobyan et al. [44]. The EXAFSsamples were analyzed five days after sample preparation whereasthe equilibrium data in Hakobyan et al. were determined at shorterequilibrium times (4 h). Furthermore, the concentrations in the EXAFSsamples were higher (1.8 mM vs. 50 μM). Thus, it is possible that anamorphous solid phase could have been formed after 5 days at thehigher concentrations in the EXAFS samples even if it did not form dur-ing titrations at 50 μM described in Hakobyan et al. [44].

3.6. Gallium complexes inhibit the expression of virulence factors

To study the induction and activity of T3SS in the presence ofME0163 and Ga(III)–ME0163 we measured expression and secretionof ExoS, a toxin secreted by the T3SS. Cultures of equal optical densityi.e. equal number of bacterial cells were used for all experimental condi-tions. ExoS expression and secretion in the presence of ME0163 weresimilar to control (Fig. 6a and b) which is in agreement with data ob-tained with Yersinia pseudotuberculosis [26]. In contrast, both Ga(III)–citrate and Ga(III)–ME0163 expression and secretion of ExoS dramati-cally dropped to low or undetectable levels (Fig. 6a and b). Our conclu-sion is that the T3SS is not completely shut down but stronglysuppressed in the presence of Ga(III) complexes. Previously, we haveobserved an altered protein expression pattern in P. aeruginosa in thepresence of Ga(III) [21]. However, this is the first report on suppressionof such an important virulence factor as T3SS by Ga(III) complexes. Thesuppression is obviously selective as the levels of FliC protein in the samesampleswere not changed. Selective inhibition of T3SS implies that eventhough bacteria are able tomultiply at a given concentration of the com-pound, their ability to damage the host cells is much weakened throughthe presence of Ga(III).

3.7. Conclusions

In this study we have shown that Ga(III) forms a complex withthe hydrazone ME0163 exhibiting two chelating rings. This Ga(III)–ME0163 complex was found to have an enhanced anti-biofilm effectin low Fe(III) conditions but this effect disappeared in 100% ISO-SENSITEST medium where Fe(III) was presumably present at highenough levels to counteract the Fe antagonist effect. While ME0163alone did not have a large effect on the T3SS in P. aeruginosa, weobserved that Ga(III)–ME0163 and Ga(III)–citrate suppressed expression

Page 7: The gallium(III)–salicylidene acylhydrazide complex shows synergistic anti-biofilm effect and inhibits toxin production by Pseudomonas aeruginosa

Fig. 6. Expression and activity of T3SS are specifically inhibited in the presence of50 μM of Ga(III) complexes; a) presence of ExoS effector in the secreted protein fraction;,b) presence of ExoS effector in the total protein fraction, c) presence of flagellar structuralprotein FliC in the total protein fraction. For all the experimental conditions aliquots forprotein analysis were taken at equal culture density to ensure the equal number of cellsin the sample.

7O. Rzhepishevska et al. / Journal of Inorganic Biochemistry 138 (2014) 1–8

and secretion of the T3SS effector protein ExoS. To the best of our knowl-edge, this work is the first to present Ga(III) complexes that suppress theT3SS in bacteria. Thus, this study opens up for new approaches to targetvirulence in the form of T3SS in bacteria.

AbbreviationsME0163 2-Oxo-2-[N-(2,4,6-trihydroxy-benzylidene)-hydrazino]-

acetamideEXAFS extended X-ray absorption fine structureDFT density functional theoryT3SS type III secretion systemMIC90 minimal inhibitory concentration that suppresses 90% of bac-

terial growthGFP green fluorescence proteinME0150 4-Methyl-[1,2,3]thiadiazole-5-carboxylic acid (6-bromo-2-

hydroxy-3-methoxy-benzylidene)-hydrazideME0161 3-(4-Hydroxy-phenyl)-propionic acid (2,4,6-trihydroxy-

benzylidene-)hydrazideME0184 3-Methoxy-benzoic acid (3,5-di-tert-butyl-2-hydroxy-

benzylidene)-hydrazideIL-8 interleukin 8

IL-6 interleukin 6ELISA enzyme-linked immunosorbent assayTSB tryptic soy brothEGTA ethylene glycol tetraacetic acidOD600 optical density at 600 nmCN coordination number

Acknowledgments

The Swedish Research Council is acknowledged for funding (#2011-3504 for M. Ramstedt; #2012-2976 for J.-F. Boily). ExoS antibody was akind gift from Åke Forsberg's group at Umeå University, Sweden, andFliC antibody was kindly provided by the group of Bernt Eric Uhlin atUmeå University, Sweden. Portions of this research were carried out atbeamline I811, MAX-lab synchrotron radiation source, Lund University,Sweden. Funding for the beamline I811 project was kindly provided bythe Swedish Research Council and the Knut och Alice WallenbergsStiftelse. Stefan Carlsson and the rest of the staff at beamline I811 aregratefully acknowledged for their help and advice. AlionaZbirenko is ac-knowledged for art work in the TOC graphic.

Appendix A. Supplementary data

Fe(III) binding assay, tablewith full names of ligands, detailed exper-imental description of EXAFS analysis, table with distances from EXAFSanalysis of the Ga(III)–ME0163 complex, as well as additional figureswith results from growth and biofilm formation assay, and mammaliancell viability tests are all available free of charge via the internet. Supple-mentary data associatedwith this article can be found in the online ver-sion, at http://dx.doi.org/10.1016/j.jinorgbio.2014.04.009. These datainclude MOL files and InChiKeys of the most important compounds de-scribed in this article.

References

[1] D. Davies, Nat. Rev. Drug Discov. 2 (2003) 114–122.[2] D.J. Evans, D.G. Allison, M.R. Brown, P. Gilbert, J. Antimicrob. Agents Chemother. 26

(1990) 473–478.[3] H. Anwar, J.L. Strap, K. Chen, J.W. Costerton, Antimicrob. Agents Chemother. 36

(1992) 1208–1214.[4] M. Alhede, T. Bjarnsholt, M. Givskov, Adv. Appl. Microbiol. 86 (2014) 1–40.[5] H.K. Estahbanati, P.P. Kashani, F. Ghanaatpisheh, Burns 28 (2002) 340–348.[6] R.T. Sadikot, T.S. Blackwell, J.W. Christman, A.S. Prince, Am. J. Respir. Crit. Care Med.

171 (2005) 1209–1223.[7] S.C. Schimpff, W.H. Greene, V.M. Young, P.H. Wiernik, J. Infect. Dis. 130S (1974)

S24–S31.[8] N. Høiby, O. Ciofu, T. Bjarnsholt, Future Microbiol 5 (2010) 1663–1674.[9] Z. Aziz, W.K. Lin, A. Nather, C.Y. Huak, Diabet. Foot Ankle 2 (2011), http://dx.doi.org/

10.3402/dfa.v2i0.7463.[10] S.C. Andrews, A.K. Robinson, F. Rodríguez-Quiñones, FEMSMicrobiol. Rev. 27 (2003)

215–237.[11] H. Lindgren, M. Honn, E. Salomonsson, K. Kuoppa, Å. Forsberg, A. Sjöstedt, Infect.

Immun. 79 (2011) 1218–1224.[12] J.J. De Voss, K. Rutter, B.G. Schroeder, C.E. Barry, J. Bacteriol. 181 (1999) 4443–4451.[13] J.J. Bullen, H.J. Rogers, P.B. Spalding, C.G. Ward, FEMS Immunol. Med. Microbiol. 43

(2005) 325–330.[14] C. Ratledge, L.G. Dover, Annu. Rev. Microbiol. 54 (2000) 881–941.[15] C. Wandersman, P. Delepelaire, Annu. Rev. Microbiol. 58 (2004) 611–647.[16] P. Paffetti, S. Perrone, M. Longini, A. Ferrari, D. Tanganelli, B. Marzocchi, G.

Buonocore, Biol. Trace Elem. Res. 112 (2006) 221–232.[17] E. Banin, K.M. Brady, E.P. Greenberg, Appl. Environ.Microbiol. 72 (2006) 2064–2069.[18] P.K. Singh, M.R. Parsek, E.P. Greenberg, M.J. Welsh, Nature 417 (2002) 552–555.[19] A.B. Kelson, M. Carnevali, V. Truong-Le, Curr. Opin. Pharmacol. 13 (2013) 707–716.[20] Y. Kaneko, M. Thoendel, O. Olakanmi, B. Britigan, P. Singh, J. Clin. Invest. (2007)

877–888.[21] O. Rzhepishevska, B. Ekstrand-Hammarstrom, M. Popp, E. Bjorn, A. Bucht, A.

Sjostedt, H. Antti, M. Ramstedt, Antimicrob. Agents Chemother. 55 (2011) 5568–5580.

[22] E. Banin, A. Lozinski, K.M. Brady, E. Berenshtein, P.W. Butterfield, M. Moshe, M.Chevion, E.P. Greenberg, Proc. Natl. Acad. Sci. U. S. A. 105 (2008) 16761–16766.

[23] D. Richardson, L. Vitolo, G. Hefter, P. May, B. Clare, J. Webb, P. Wilairat, Inorg. Chim.Acta 170 (1990) 165–170.

[24] L. Vitolo, G. Hefter, B. Clare, J. Webb, Inorg. Chim. Acta 170 (1990) 171–176.[25] Z. Kovacevic, Y. Yu, D.R. Richardson, Chem. Res. Toxicol. 24 (2011) 279–282.

Page 8: The gallium(III)–salicylidene acylhydrazide complex shows synergistic anti-biofilm effect and inhibits toxin production by Pseudomonas aeruginosa

8 O. Rzhepishevska et al. / Journal of Inorganic Biochemistry 138 (2014) 1–8

[26] M. Dahlgren, C. Zetterstrom, A. Gylfe, A. Linusson, M. Elofsson, Bioorg. Med. Chem.18 (2010) 2686–2703.

[27] D. Wang, C. Zetterstrom, M. Gabrielsen, K. Beckham, J. Tree, S. Macdonald, O. Byron,T. Mitchell, D. Gally, P. Herzyk, A. Mahajan, H. Uvell, R. Burchmore, B. Smith, M.Elofsson, A. Roe, J. Biol. Chem. 286 (2011) 29922–29931.

[28] H. Chu, A. Slepenkin, M. Elofsson, P. Keyser, L.M. de la Maza, E.M. Peterson, Int. J.Antimicrob. Agents 36 (2010) 145–150.

[29] T.J. Brickman, C.A. Cummings, S.Y. Liew, D.A. Relman, S.K. Armstrong, J. Bacteriol. 193(2011) 4798–4812.

[30] J. Gao, D.R. Richardson, Blood 98 (2001) 842–850.[31] P.V. Bernhardt, P.C. Sharpe,M. Islam, D.B. Lovejoy, D.S. Kalinowski, D.R. Richardson, J.

Med. Chem. 52 (2009) 407–415.[32] D.R. Richardson, E.H. Tran, P. Ponka, Blood 86 (1995) 4295–4306.[33] D.R. Richardson, Antimicrob. Agents Chemother. 41 (1997) 2061–2063.[34] A. Despaigne, G. Parrilha, J. Izidoro, P. da Costa, R. dos Santos, O. Piro, E. Castellano,

W. Rocha, H. Beraldo, Eur. J. Med. Chem. 50 (2012) 163–172.[35] R. Nordfelth, A. Kauppi, H. Norberg, H. Wolf-Watz, M. Elofsson, Infect. Immun. 73

(2005) 3104–3114.[36] P. Persson, K. Zivkovic, S. Sjöberg, Langmuir 22 (2006) 2096–2104.[37] M. Clausen, L.-O. Öhman, P. Persson, J. Inorg. Biochem. 99 (2005) 716–726.[38] J.E. Delbene, W.B. Person, K. Szczepaniak, J. Phys. Chem. 99 (1995) 10705–

10707.

[39] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, G.Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H.P.Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M. Hada, M. Ehara, K.Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T.Vreven, J.A. Montgomery, Jr., J.E. Peralta, F. Ogliaro, M. Bearpark, J.J. Heyd, E. Brothers,K.N. Kudin, V.N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C.Burant, S.S. Iyengar, J. Tomasi, M. Cossi, N. Rega, N.J. Millam, M. Klene, J.E. Knox, J.B.Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J.Austin, R. Cammi, C. Pomelli, J.W.Ochterski, R.L.Martin, K.Morokuma, V.G. Zakrzewski,G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, Ö. Farkas, J.B.Foresman, J.V. Ortiz, J. Cioslowski, D.J. Fox, Gaussian, Inc., Wallingford CT, 2009.

[40] V.T. Lee, R.S. Smith, B. Tummler, S. Lory, Infect. Immun. 73 (2005) 1695–1705.[41] G. Jackson, Polyhedron 9 (1990) 163–170.[42] A. Evers, R. Hancock, A. Martell, R. Motekaitis, Inorg. Chem. 28 (1989) 2189–2195.[43] M. Ramstedt, B. Ekstrand-Hammarstrom, A.V. Shchukarev, A. Bucht, L. Osterlund, M.

Welch, W.T.S. Huck, Biomaterials 30 (2009) 1524–1531.[44] S. Hakobyan, J.-F. Boily, M. Ramstedt, J. Inorg. Biochem. (2014), http://dx.doi.org/10.

1016/j.jinorgbio.2014.04.012 in press.[45] U. Carpentieri, J. Myers, C.W. Daeschner, M.E. Haggard, J. Biochem. Biophys. Methods

14 (1987) 93–100.[46] L. Koh, O. Kon, K. Loh, Y. Long, J. Ranford, A. Tan, Y. Tjan, J. Inorg. Biochem. 72 (1998)

155–162.