Open Access

Effects of deferoxamine on intrinsic colistin resistance of Proteus mirabilis

  • Authors:
    • Mehmet Erinmez
    • Yasemin Zer
  • View Affiliations

  • Published online on: August 7, 2023     https://doi.org/10.3892/etm.2023.12158
  • Article Number: 459
  • Copyright: © Erinmez et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Proteus mirabilis is a common pathogen, which is responsible for urinary tract infections. Iron is a critical element necessary for both humans and pathogens to maintain their biological functions, and iron limitation via chelator agents may be useful in the treatment of infections. The present study aimed to investigate the synergistic interactions between the iron chelator agent deferoxamine (DFO) and the antibacterial drug colistin. The minimum inhibitory concentration (MIC) values of DFO and colistin for P. mirabilis isolates were determined by broth microdilution. The checkerboard technique was used to examine the potential synergy between DFO and colistin. Furthermore, time‑kill assays were used for the confirmation of synergy detected by the checkerboard assay, as well as for determining bacteriostatic and bactericidal interactions throughout a 24‑h period. As expected, all P. mirabilis isolates were resistant to colistin. DFO did not inhibit P. mirabilis growth when used alone, even at very high doses (10 µg ml‑1). Notably, when in combination with DFO, the MIC values of colistin were markedly reduced, and the checkerboard assay results showed synergy between colistin and DFO for all isolates. In addition, in time‑kill assays, colistin + DFO exhibited synergistic activity against all strains at most time intervals and concentrations tested. Colistin + DFO showed bactericidal activity at colistin concentrations of 1xMIC and 2xMIC, although a degree of re‑growth was observed in one of the strains at 12‑24 h. These findings indicated that DFO has the potential for use as an adjunct to colistin through iron sequestration, thus providing synergistic activity to an antibiotic that would not normally be considered a treatment option against P. mirabilis. In vivo experiments in the future may provide useful information on the efficacy of DFO/colistin since these models effectively reflect physiological parameters.

Introduction

Proteus is a gram-negative bacterium, which is widely distributed in a range of settings, including water sources, soil and sewage, but it is primarily a flora member of the gastrointestinal systems of humans and animals (1). Swarming motility, and the production of urease, hemolysin and numerous fimbrial adhesions are the phenotypic characteristics of this bacterium (2). Proteus mirabilis is the most prevalent cause of human infections among all Proteus species (3). P. mirabilis is the most frequently detected bacteria in long-term urinary catheterization and is a significant cause of complicated urinary tract infections (UTIs), wound infections, gastroenteritis, and, in rare cases, bacteremia (2,3). A distinctive feature of P. mirabilis is that it produces crystalline biofilms, leading to encrusted and clogged catheters in long-term urinary catheterization, which aggravates catheter-associated urinary tract infections (CAUTIs) (4). Therefore, urine retention and reflux, as well as severe bladder distension and pyelonephritis, may develop. Furthermore, crystalline biofilms have been reported to be associated with the persistence of P. mirabilis in the urinary system via shielding it from antimicrobial agents and the host immune mechanisms (5).

Colistin exhibits a broad spectrum of antibacterial activity, mostly against gram-negative bacteria since its antibacterial activity occurs on the outer membrane (6). However, some gram-negative bacteria are naturally resistant to colistin, such as Neisseria meningitidis, Burkholderia species and P. mirabilis (7). Antimicrobial resistance is a major public health issue and antimicrobial resistance in numerous bacterial species has an impact on a number of facets of medical practice, from treatment of infections in primary healthcare to the clinical management of patients with severe diseases in tertiary care (8). The worldwide spread of antibiotic-resistant bacterial strains poses a considerable obstacle to appropriate treatment, as there are few clinically available antibiotics that maintain adequate action against these strains (9).

Iron is a vital element for growth, and is necessary for the activity of numerous proteins and enzymes participating in various physiological pathways, such as oxygen transportation, gene regulation and nitrogen fixation (10). In the mammalian host, the majority of intracellular iron is stored in ferritin or bound to haem or haem-containing substances, whereas extracellular iron is bound to transferrin, lactoferrin, haemopexin and haptoglobin, making it unavailable for bacterial uptake (11). Bacterial pathogens that have the ability to colonize humans use several strategies to scavenge essential elements, such as iron and zinc; therefore, there is a perpetual competition between bacteria and the host for micronutrients (12). Pathogenic bacteria have evolved multiple iron transportation pathways, intracellular iron stores, redox stress resistance systems and iron-sensitive regulatory sequences to control the expression of genes involved in several cellular activities to counterbalance iron-deficient situations and sustain iron homeostasis (10). A number of pathogenic bacteria use siderophores to overcome the iron-limiting environment in the host. Siderophores are low molecular weight iron-binding substances that are secreted and imported by microorganisms for iron acquisition (13). The iron is released from the siderophore after cellular uptake to aid microbial metabolism and multiplication. During infection with bacterial and fungal pathogens, siderophores are thought to be important virulence components (14). The existence of these regulatory and competing mechanisms underlines the significance of iron in the survival of bacteria.

Given the critical function of iron in the growth and survival of numerous pathogenic bacteria, minimizing the amount of iron available at the infection site may help to improve the treatment outcome. Using chelator compounds that sequester various metals and impede bacterial iron uptake is one of the approaches for attaining iron limitation (8). Iron chelating agents can form complexes with iron in both its ferrous (Fe2+) and ferric (Fe3+) states, although chelators generally exhibit different levels of affinity for the different states (15,16). Deferoxamine (DFO) is a semi-synthetic drug derived from the bacterial siderophore desferrioxamine B, which has been licensed for medical use in the treatment of iron excess in patients (17). DFO also has an impact on the amount of iron available to microorganisms, this feature constitutes the founding of new microbial control strategies, such as novel treatment regimens or even preservation systems (18,19). Considering the ongoing occurrence and spread of antibiotic resistance, and the limited prospect that the current development process will be able to meet the demand for new antimicrobial agents with novel mechanisms of action, it is necessary to investigate the potential of alternatives (8). Iron chelators that have already withstood toxicity and preclinical testing in animals may provide an alternative therapeutic technique in the case of multidrug-resistant bacteria, for which entire classes of antibiotics are no longer considered treatment options (20). Siderophores may also serve as a facilitator for antibiotics across the cell membrane because of increased cell membrane permeability induced by iron deprivation (21). We hypothesize that iron deprivation or interactions with cell membranes caused by DFO or increased siderophore synthesis may cause inhibition or inactivation of proteins and enzymes necessary for critical processes of bacteria, as well as synergy with membrane-active antibiotics, such as colistin.

Materials and methods

Bacterial strains

Clinical P. mirabilis isolates (n=11) recovered from blood culture samples between July and December 2021 in the Research and Application Hospital of Gaziantep University (Gaziantep, Turkey), a tertiary care center, were used in the present study. The isolates were collected from 11 individual patients. Escherichia coli ATCC25922 (American Type Culture Collection) and E. coli NCTC 13846 (National Collection of Type Cultures) strains were provided by the Microbiology Laboratory of Gaziantep University's Medical Faculty Hospital (Gaziantep, Turkey) and employed as quality control in the broth microdilution tests. Prior to testing, the isolates were cultured from frozen stocks onto Columbia Agar with 5% sheep blood (BD Biosciences) and incubated overnight at 35˚C.

Determination of minimum inhibitory concentration (MIC)

MIC values of DFO and colistin against P. mirabilis were determined using the reference broth microdilution method according to the International Organization for Standardization (ISO) standards (ISO 20776-1:2019) (22). Colistin (Biosynth Ltd.) stock solutions were prepared in sterile distilled water (dH2O) and stored in aliquots at -20˚C. DFO (Desferal®; Novartis Corporation) was freshly prepared as a 50 mg/ml stock solution in sterile dH2O prior to the study. Test solutions of DFO and colistin were prepared immediately before use. Briefly, serial two-fold dilutions of DFO and colistin in cation-adjusted Mueller-Hinton Broth (CAMHB; Oxoid; Thermo Fisher Scientific, Inc.) were prepared in a 96-well plate. The inoculum to be tested was prepared from overnight cultures by dilution in MHB to provide a final bacterial density of 5x105 colony-forming unit (CFU)/ml. Colistin was tested over a range from 0.062 to 128 µg/ml, and DFO was tested over a range from 0.062 to 10 mg/ml. Each well was loaded with bacterial suspensions, and the plates were incubated at 35˚C for 24 h. Controls for positive growth and negative sterility were also performed. The lowest concentrations of DFO and colistin with no visible signs of turbidity were defined as MIC values.

Checkerboard assay

Fractional inhibitory concentration index (FICI) values of combinations of DFO and colistin against a total of 11 P. mirabilis isolates were determined using the checkerboard technique (23,24). In brief, serial two-fold dilutions of the first compound (colistin) were prepared across the columns and the second compound (DFO) dilutions were prepared across the rows of a 96-well plate. Individual wells were inoculated with suspensions of overnight cultures in MHB to provide a final inoculum density of 5x105 CFU/ml. The plates were incubated for 24 h at 35˚C. FIC values were defined by broth microdilution on separate checkerboard panels containing increasing concentrations of DFO (rows G through A, 8 to 512 µg/ml) and colistin (columns 1 through 11, 0.06 to 64 µg/ml). The FIC of both drugs was calculated using the formula: FIC=MIC of Drug in combination/MIC of Drug alone. The FICI was then calculated from the sum of the FIC values using the following formula: FICI=(MIC of colistin in combination/MIC of colistin alone) + (MIC of DFO in combination/MIC of DFO alone). The interpretation of the findings was as follows: Synergy if FICI ≤0.5; no interaction if FICI 0.5-4; and antagonism if FICI >4(23). On an occasion where a MIC for one of the test compounds was off-scale (greater than the highest concentration tested), the MIC was set to the next highest two-fold concentration for calculation of the FIC (e.g. if the highest MIC was tested 32 µg/ml, the FIC was calculated based on a MIC of 64 µg/ml) (25).

Time-kill assay

In accordance with the results from the checkerboard assays, two isolates were randomly selected for further evaluation of the DFO-colistin combination using a time-kill assay. Colistin was tested alone and in combination with DFO for the selected isolates at 0.5X, 1X and 2X the MIC concentrations of colistin. For isolates where the DFO MIC was >512 µg/ml, a DFO concentration of 128 µg/ml (0.25X the highest concentration tested) was used during testing. In preparation for the study, bacterial suspensions were prepared in MHB from freshly-grown blood agar plates, diluted and grown to the log-phase. The turbidity of bacterial cultures was adjusted to form a final inoculum density of 5x105 CFU/ml, as verified by viable count, and added to flasks containing 20 ml MHB. Then, treatments were added to broth cultures to yield desired concentrations. Growth controls without any treatment were also included. Test and control flasks were incubated at 35˚C and viable counts were performed at 0, 1, 3, 6, 9, 12 and 24 h by serial dilution plating. Samples were spread onto Mueller-Hinton Agar (MHA; Difco; BD Biosciences) plates, because swarming ability of Proteus makes colony counting difficult on media other than MHA. All plates were incubated for 20-24 h at 35˚C. Colonies were counted manually, and the CFU/ml was determined from the average count of the duplicate plates, followed by the calculation of the log10 CFU/ml. Antimicrobial activity was calculated for each isolate as the change in the bacterial count of (Δlog10 CFU/ml) obtained in 24 h compared with the count at the start and defined as Δlog10 CFU24. Bactericidal activity was evaluated as a ≥3 log10 decrease in CFU/ml over the time period examined, whereas synergy was considered a ≥2 log10 decrease in CFU/ml for the antibiotic combination in comparison with the most effective monotherapy (26,27). Time-kill studies were performed as three independent replicates and graphs displaying the results were generated using Excel (version 16; Microsoft Corp.).

Results

Broth microdilution and checkerboard method

MICs, detected by broth microdilution method, were 8-16 µg/ml for colistin. However, DFO did not affect bacterial growth even at a concentration of 10 mg/ml (Table I). Because the DFO may easily transfer iron to bacteria with a homologous siderophore receptor and because CAMHB is a rich broth with abundant iron, the outcome was not surprising. Using EUCAST criteria, all P. mirabilis strains were classified as resistant to colistin (28). After combination with DFO, the MIC values of colistin were reduced. The results of a checkerboard assay showed synergy (i.e. FICI ≤0.5) between colistin and DFO for all of the isolates (Table I). No antagonism was observed for the combination.

Table I

Antimicrobial activity of colistin with DFO against PM strains.

Table I

Antimicrobial activity of colistin with DFO against PM strains.

 MIC value, µg/ml 
IsolateColistinDFOColistin MIC in combination with DFOFICIInterpretation
PM116>5120.50.06Synergy
PM216>51240.5Synergy
PM316>5120.50.06Synergy
PM416>51210.12Synergy
PM516>5120.50.06Synergy
PM616>5120.50.06Synergy
PM716>51240.5Synergy
PM816>51240.5Synergy
PM916>5120.50.06Synergy
PM108>5120.50.09Synergy
PM1116>5120.50.06Synergy

[i] PM, Proteus mirabilis; MIC, minimum inhibitory concentration; DFO, deferoxamine; FICI, fractional inhibitory concentration index.

Time-kill method

The two randomly selected P. mirabilis isolates (PM2 and PM7) were used in the time-kill assays. The time-kill profiles of isolates and changes in log10 CFU/ml from the initial inoculum to 24 h are shown in Fig. 1. DFO monotherapy (512 µg/ml) and colistin monotherapy at a concentration of 0.5xMIC produced little or no bacterial killing at any timepoint, with bacterial growth close to control values across the 24 h. Colistin monotherapy at concentrations of 1xMIC and 2xMIC showed bacteriostatic activity maintained for 9 h, with subsequent re-growth in varying amounts. Colistin + DFO showed synergistic activity against all strains at most time intervals and concentrations tested. Colistin + DFO showed bactericidal activity at colistin concentrations of 1xMIC and 2xMIC although a degree of re-growth was observed in isolate PM7 at 12-24 h (Fig. 1).

Discussion

A limitation of the present study is that there were no animal experiments conducted, in which the physiological environment could be better evaluated, since they were not included in the initial research design and ethics committee approvals. Excess iron has been shown to aggravate the situation in various infections, including tuberculosis, malaria, invasive bacterial infections, cystitis, keratitis and wound infections (29-31). P. mirabilis is a significant causative agent of UTIs, especially CAUTIs. Bacterial capability to import Fe3+, Zn2+ and Ni2+, and export Cu+ is essential for efficacious colonization of the urinary system (32). In the course of a UTI, it has been shown that iron overload in the bladder or in the urine can exacerbate inflammation and escalate urothelial cell death (33). Urine samples from patients with acute UTIs have markedly higher iron levels and sloughed epithelial cells than urine samples from the healthy population (32). In addition, it has been reported that, for infants, having higher basal iron serum concentrations or increased iron supplementation is associated with an increased risk of UTIs (34-36). Similarly, iron concentrations in the urine of postmenopausal women who are prone to recurrent and chronic UTIs have been shown to be higher than those in the general postmenopausal female population (37-39). Increased ambient iron has also been reported to lead to enhanced growth and a significant increase in the intracellular bacterial load of E. coli, another frequent causative agent of UTIs, in bladder epithelial cells (33). Given that high iron levels enhance bacterial colonization, infection development and infection chronicity in specific systems, limiting ambient iron appears to be an appropriate strategy to battle pathogenic microorganisms.

Sequestration of iron by chelation may be a beneficial adjunct for the treatment of infections, given the relationship between iron excess or dietary iron supplementation and infection (40). For a number of years, iron chelators have been used to treat iron excess conditions, and their pharmacological and safety profiles have been widely investigated (41). Depletion of iron through synthetic iron chelators can be effective in inhibiting bacterial growth (29). DFO, as an iron chelator, aids the host's intrinsic iron-withholding systems and appears as a promising treatment option for local infections (21). Iron chelation with DFO has been reported to improve host cell survival, reduce bacterial proliferation in urothelial cells and reduce autophagy (33). Furthermore, DFO has shown antibacterial activity as a single agent, and can inhibit the growth of Acinetobacter baumannii, Pseudomonas aeruginosa and Staphylococcus aureus at varying concentrations (21). Furthermore, impaired iron acquisition can reduce the urinary tract colonization of P. mirabilis (42). To the best of our knowledge, the present study is the first investigation to determine DFO MIC values against P. mirabilis isolates using the standard broth microdilution method. However, the results revealed that DFO does not have the potential to be an antibacterial agent on its own.

An overabundance of iron is hazardous to the host, not just because it enhances bacterial growth, but also because it induces increased inflammatory activity and epithelial cell stress (33,43). Furthermore, local iron deprivation may have an effect on the host immune system response by reducing local reactive oxygen species (ROS) formation (29). It has been demonstrated that the cumulation of iron causes an elevated inflammatory response in local bladder tissue, as well as increased bacterial colonization, whereas these findings can be reversed by a low-iron diet (44). During the course of infection, unbound labile iron accumulates in the plasma and induces a variety of responses, including elevated cell proliferation rates, persistent suppression of cell proliferation, and apoptotic or necrotic cell death via ROS. Therefore, pharmacological regulation of iron by chelation therapy is crucial for achieving a balance between inhibiting cell damage and supplying the cellular demands (29). Although the most concrete indicator of the success of medical treatment against infections is the discontinuation of bacterial growth, since limiting ambient iron reduces inflammation, iron chelators combined with antibiotics can reduce patient complaints and symptoms.

Medical device-associated infections are frequently related to indwelling objects, such as urinary catheters and prosthetic joints, and are responsible for ~50% of all nosocomial infections (45). These objects provide an abiotic surface and facilitate the formation of biofilms (46). Biofilm production is an important virulence factor in the emergence of CAUTIs (42). Biofilm-related infections are notoriously difficult to treat due to the adjustments enabled by biofilm development, which result in increased antibiotic resistance as well as increased resistance to host immune mechanisms (46). Higher levels of iron are required for the production of biofilms than for bacterial proliferation (47,48). A recent study showed that iron can trigger biofilm formation in P. mirabilis (42). Similarly, biofilm production is decreased when lactoferrin, a physiological iron chelator, is added to P. aeruginosa cultures (49). A previous study also observed that DFO can improve the ability of tobramycin to dissolve formed biofilms grown on human airway epithelial cells (50). Iron chelation may be useful in infection treatment due to the crucial function of iron in biofilm development and bacterial pathogenicity (29). Inhibition of bacterial iron acquisition via catheters made of iron-scavenging materials or coated with chelators may decrease biofilm development and further infection, but this approach has yet to be tested in a clinical setting (29,42). However, generally, it has been assumed that the local administration of iron-chelating agents represents a safe pharmacological therapy with a low risk of side effects (29). Iron chelation may be employed as a prophylactic strategy to minimize medical device-associated infections, especially for certain systems, such as the urinary system.

Bacteria sense an iron-deficient environment and react correspondingly by upregulating iron-acquisition pathways as well as virulence genes (51). A β-barrel receptor on the outer membrane of gram-negative bacteria recognizes iron-bound siderophores. The iron-bound siderophore is translocated into the periplasm after ligand engagement causes a conformational alteration (29). An ATP-binding cassette transporter in the inner membrane mediates conformational alterations in transportation into the cytoplasm and iron reduction. This complex will then attach to certain receptor proteins on the bacterial cell surface, allowing it to be taken through active transport (29). It is hypothesized that iron limitation conditions may result in increased production of siderophores, which are specific molecules for transporting iron. Siderophore secretion has the physical outcome of allowing molecules to diffuse away from producers while possibly preventing diffused molecules from returning producer cells. Diffusion can still result in significant siderophore loss, putting bacterial fitness at risk (52). In the present study, in addition to discovering a synergistic interaction between DFO and colistin in all P. mirabilis isolates using the checkerboard method, the MIC values of the DFO and colistin combination were in the susceptible ranges of colistin against Enterobacterales for the vast majority of the isolates, according to EUCAST guidelines (28). However, in vivo experiments would be necessary to confirm the in vitro findings. According to data not shown, no significant synergy was detected in E. coli, K. pneumoniae, P. aeruginosa and A. baumannii isolates in the subsequent investigation to assess whether DFO can exhibit synergistic effects when used with colistin against other gram-negative bacteria. From this point of view, we hypothesize that the ability of all bacteria to produce siderophores and their responses to iron restriction is not the same, and the synergy of DFO and colistin detected in the present study is based on a specific mechanism. Effective siderophore production was previously thought to be a feature of aerobic gram-negative bacteria (53). However, it has been discovered that some gram-positive bacteria may generate siderophores (54). The identification of bacterial species that lack effective siderophore production will become more relevant if iron chelators are combined with antimicrobial treatment.

Colistin is a polymyxin antibiotic with rapid bactericidal activity against gram-negative bacteria (55). Similarly, the bactericidal action of the DFO-colistin combination (colistin 1xMIC + DFO, colistin 2xMIC+ DFO) was evident from the onset of the present study, according to our time-kill assay results. As a result, it was hypothesized that several intrinsic mechanisms related to colistin resistance of P. mirabilis were rapidly eliminated via the addition of DFO. Furthermore, the fact that the bactericidal effect of the DFO-colistin combination was maintained for ≥12 h in the present experiments indicated that it could be beneficial without the requirement for frequent dose repetitions since colistin has a long half-life (14.4 h) (56). This insight can be beneficial for optimizing doses and improving treatment approaches when combined with the pharmacokinetic profile of the targeted patient population. Conformational changes in the outer membrane of the bacteria during both increased secretion and uptake of siderophores may be responsible for vulnerabilities against colistin activity. Similarly, vancomycin, which is not preferred in P. aeruginosa infections due to its low gram-negative activity, has been reported to be effective when used with the DFO-gallium complex due to disturbance of the outer membrane via a combination of electrostatic and hydrophobic interactions coupled with selective binding and resulting in enhanced permeability of vancomycin as a result of the membrane damage (57). Under low iron concentrations, several physiological changes may occur in the bacterial pathogens, including a shift to a planktonic state (58,59). Bacteria in a planktonic state are known to be more susceptible to certain antimicrobials, suggesting a potential mechanism of iron chelation-induced sensitization to antimicrobials (8). Because of the increased permeability induced by iron deprivation, siderophores may potentially serve as a facilitator for colistin across the cell membrane. The use of antivirulence compounds combined with antibiotics may be a promising approach for virulence attenuation and pathogen elimination (60). In addition to binding iron, DFO may also have an affinity for zinc. The ability to sequestrate zinc is thought to be responsible for making metallo-β lactamase producers susceptible to β-lactams (61). Similarly, deprivation of iron reduces the activity of key proteins and enzymes, such as cytochromes, which are examples of iron-dependent proteins that are crucial for energy metabolism and ribonucleotide reductase, which is involved in DNA synthesis. If any of these are disrupted, the multiplication of the microorganism may be halted (21). Briefly, inhibition of bacterial growth, reducing biofilm formation, membrane disruption, inactivation of specific enzymes or proteins that are essential for cell replication, and reducing oxidant stress on host cells may constitute potential pathways for the enhanced bactericidal effect of colistin via iron chelation. We are currently investigating whether the synergistic interaction discovered between DFO and colistin extends to other antibiotics. Identifying patient groups and explaining the mechanisms underlying the synergistic interaction will be aided by determining the interactions between different drug groups and iron chelators.

In conclusion, the present study suggested that DFO has the potential for use as an adjunct to colistin through iron sequestration, thus providing synergistic activity to an antibiotic that is not normally considered a treatment option against P. mirabilis. In vivo experiments will provide useful information on DFO-colistin efficacy, since these models are better in terms of reflecting physiological conditions such as metal ion levels. Also, in vivo models account for parameters, such as compound biodistribution, pH and the presence of host factors. Because CAMHB is a rich broth with substantially higher iron, carbon sources and other cofactors than the levels in the human body, the synergy that was detected in vitro may be greater in vivo.

Acknowledgements

Not applicable.

Funding

Funding: This work was supported by the Scientific Research Programme of Gaziantep University (grant no. TF.UT.21.28).

Availability of data and materials

The datasets used and/or analyzed during the current study are available from the corresponding author on reasonable request.

Authors' contributions

ME and YZ contributed to the study conception and design, and performed material preparation, data collection and analysis. The first draft of the manuscript was written by ME and both authors commented on previous versions of the manuscript. ME and YZ confirm the authenticity of all the raw data. Both authors have read and approved the final manuscript.

Ethics approval and consent to participate

The present study was performed in line with the principles of The Declaration of Helsinki. Approval was granted by the Gaziantep University Clinical Research Ethics Committee (date, January 27, 2021; approval no. 2021/11; Gaziantep, Turkey).

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Drzewiecka D: Significance and roles of proteus spp. Bacteria in natural environments. Microb Ecol. 72:741–758. 2016.PubMed/NCBI View Article : Google Scholar

2 

Mobley HLT: Proteus mirabilis overview. Methods Mol Biol. 2021:1–4. 2019.PubMed/NCBI View Article : Google Scholar

3 

Armbruster CE, Mobley HLT and Pearson MM: Pathogenesis of Proteus mirabilis infection. EcoSal Plus. 8(10)2018.PubMed/NCBI View Article : Google Scholar

4 

Wasfi R, Hamed SM, Amer MA and Fahmy LI: Proteus mirabilis Biofilm: Development and therapeutic strategies. Front Cell Infect Microbiol. 10(414)2020.PubMed/NCBI View Article : Google Scholar

5 

Yuan F, Huang Z, Yang T, Wang G, Li P, Yang B and Li J: Pathogenesis of Proteus mirabilis in catheter-associated urinary tract infections. Urol Int. 105:354–361. 2021.PubMed/NCBI View Article : Google Scholar

6 

El-Sayed Ahmed MAE, Zhong LL, Shen C, Yang Y, Doi Y and Tian GB: . Colistin and its role in the Era of antibiotic resistance: An extended review (2000-2019). Emerg Microbes Infect. 9:868–885. 2020.PubMed/NCBI View Article : Google Scholar

7 

Gogry FA, Siddiqui MT, Sultan I and Haq QMR: Current update on intrinsic and acquired colistin resistance mechanisms in bacteria. Front Med (Lausanne). 8(677720)2021.PubMed/NCBI View Article : Google Scholar

8 

Vinuesa V and McConnell MJ: Recent advances in iron chelation and gallium-based therapies for antibiotic resistant bacterial infections. Int J Mol Sci. 22(2876)2021.PubMed/NCBI View Article : Google Scholar

9 

Cassini A, Högberg LD, Plachouras D, Quattrocchi A, Hoxha A, Simonsen GS, Colomb-Cotinat M, Kretzschmar ME, Devleesschauwer B, Cecchini M, et al: Attributable deaths and disability-adjusted life-years caused by infections with antibiotic-resistant bacteria in the EU and the European economic area in 2015: A population-level modelling analysis. Lancet Infect Dis. 19:56–66. 2019.PubMed/NCBI View Article : Google Scholar

10 

Andrews SC, Robinson AK and Rodríguez-Quiñones F: Bacterial iron homeostasis. FEMS Microbiol Rev. 27:215–237. 2003.PubMed/NCBI View Article : Google Scholar

11 

Sousa Gerós A, Simmons A, Drakesmith H, Aulicino A and Frost JN: The battle for iron in enteric infections. Immunology. 161:186–199. 2020.PubMed/NCBI View Article : Google Scholar

12 

Seyoum Y, Baye K and Humblot C: Iron homeostasis in host and gut bacteria-a complex interrelationship. Gut Microbes. 13:1–19. 2021.PubMed/NCBI View Article : Google Scholar

13 

Palmer LD and Skaar EP: Transition metals and virulence in bacteria. Annu Rev Genet. 50:67–91. 2016.PubMed/NCBI View Article : Google Scholar

14 

Sassone-Corsi M, Chairatana P, Zheng T, Perez-Lopez A, Edwards RA, George MD, Nolan EM and Raffatellu M: Siderophore-based immunization strategy to inhibit growth of enteric pathogens. Proc Natl Acad Sci USA. 113:13462–13467. 2016.PubMed/NCBI View Article : Google Scholar

15 

Zhou T, Winkelmann G, Dai ZY and Hider RC: Design of clinically useful macromolecular iron chelators. J Pharm Pharmacol. 63:893–903. 2011.PubMed/NCBI View Article : Google Scholar

16 

Liu ZD and Hider RC: Design of clinically useful iron(III)-selective chelators. Med Res Rev. 22:26–64. 2022.PubMed/NCBI View Article : Google Scholar

17 

Byrne SL, Krishnamurthy D and Wessling-Resnick M: Pharmacology of iron transport. Annu Rev Pharmacol Toxicol. 53:17–36. 2013.PubMed/NCBI View Article : Google Scholar

18 

Kontoghiorghes GJ, Kolnagou A, Skiada A and Petrikkos G: The role of iron and chelators on infections in iron overload and non iron loaded conditions: Prospects for the design of new antimicrobial therapies. Hemoglobin. 34:227–239. 2010.PubMed/NCBI View Article : Google Scholar

19 

Holbein BE and Mira de Orduña R: Effect of trace iron levels and iron withdrawal by chelation on the growth of Candida albicans and Candida vini. FEMS Microbiol Lett. 307:19–24. 2010.PubMed/NCBI View Article : Google Scholar

20 

Thompson MG, Corey BW, Si Y, Craft DW and Zurawski DV: Antibacterial activities of iron chelators against common nosocomial pathogens. Antimicrob Agents Chemother. 56:5419–5421. 2012.PubMed/NCBI View Article : Google Scholar

21 

Gokarn K and Pal RB: Activity of siderophores against drug-resistant Gram-positive and Gram-negative bacteria. Infect Drug Resist. 11:61–75. 2018.PubMed/NCBI View Article : Google Scholar

22 

International Organization for Standardization (ISO): ISO 20776-1:2019(en) Susceptibility testing of infectious agents and evaluation of performance of antimicrobial susceptibility test devices. Part 1: Broth micro-dilution reference method for testing the in vitro activity of antimicrobial agents against rapidly growing aerobic bacteria involved in infectious diseases. ISO, Geneva, Switzerland, 2019. https://www.iso.org/obp/ui/en/#iso:std:iso:20776:-1:ed-2:v2:en. Accessed March 24, 2022.

23 

Xu X, Xu L, Yuan G, Wang Y, Qu Y and Zhou M: Synergistic combination of two antimicrobial agents closing each other's mutant selection windows to prevent antimicrobial resistance. Sci Rep. 8(7237)2018.PubMed/NCBI View Article : Google Scholar

24 

Stein C, Makarewicz O, Bohnert JA, Pfeifer Y, Kesselmeier M, Hagel S and Pletz MW: Three dimensional checkerboard synergy analysis of colistin, meropenem, tigecycline against multidrug-resistant clinical Klebsiella pneumonia isolates. PLoS One. 10(e0126479)2015.PubMed/NCBI View Article : Google Scholar

25 

Thwaites M, Hall D, Stoneburner A, Shinabarger D, Serio AW, Krause KM, Marra A and Pillar C: Activity of plazomicin in combination with other antibiotics against multidrug-resistant Enterobacteriaceae. Diagn Microbiol Infect Dis. 92:338–345. 2018.PubMed/NCBI View Article : Google Scholar

26 

Clinical and Laboratory Standards Institute (CLSI): M26-A Methods for determining bactericidal activity of antimicrobial agents; Approved guideline. Volume 19. CLSI, Wayne, PA, USA, 1999.

27 

Gómez-Junyent J, Benavent E, Sierra Y, El Haj C, Soldevila L, Torrejón B, Rigo-Bonnin R, Tubau F, Ariza J and Murillo O: Efficacy of ceftolozane/tazobactam, alone and in combination with colistin, against multidrug-resistant Pseudomonas aeruginosa in an in vitro biofilm pharmacodynamic model. Int J Antimicrob Agents. 53:612–619. 2019.PubMed/NCBI View Article : Google Scholar

28 

The European Committee on Antimicrobial Susceptibility Testing (EUCAST): Breakpoint tables for interpretation of MICs and zone diameters Version 12.0, 2022. https://www.eucast.org/fileadmin/src/media/PDFs/EUCAST_files/Breakpoint_tables/v_12.0_Breakpoint_Tables.pdf. Accessed July 15, 2022.

29 

Scott C, Arora G, Dickson K and Lehmann C: Iron chelation in local infection. Molecules. 26(189)2021.PubMed/NCBI View Article : Google Scholar

30 

Soofi S, Cousens S, Iqbal SP, Akhund T, Khan J, Ahmed I, Zaidi AK and Bhutta ZA: Effect of provision of daily zinc and iron with several micronutrients on growth and morbidity among young children in Pakistan: A cluster-randomised trial. Lancet. 382:29–40. 2013.PubMed/NCBI View Article : Google Scholar

31 

Isanaka S, Aboud S, Mugusi F, Bosch RJ, Willett WC, Spiegelman D, Duggan C and Fawzi WW: Iron status predicts treatment failure and mortality in tuberculosis patients: A prospective cohort study from Dar es Salaam, Tanzania. PLoS One. 7(e37350)2012.PubMed/NCBI View Article : Google Scholar

32 

Subashchandrabose S and Mobley HL: Back to the metal age: battle for metals at the host-pathogen interface during urinary tract infection. Metallomics. 7:935–942. 2015.PubMed/NCBI View Article : Google Scholar

33 

Bauckman KA and Mysorekar IU: Ferritinophagy drives uropathogenic Escherichia coli persistence in bladder epithelial cells. Autophagy. 12:850–863. 2016.PubMed/NCBI View Article : Google Scholar

34 

Mava Y, Ambe JP, Bello M, Watila I and Nottidge VA: Urinary tract infection in febrile children with sickle cell anaemia. West Afr J Med. 30:268–272. 2011.PubMed/NCBI

35 

Collard KJ: Iron homeostasis in the neonate. Pediatrics. 123:1208–1216. 2009.PubMed/NCBI View Article : Google Scholar

36 

Arshad M and Seed PC: Urinary tract infections in the infant. Clin Perinatol. 42:17–28. 2015.PubMed/NCBI View Article : Google Scholar

37 

Matsumoto T: Urinary tract infections in the elderly. Curr Urol Rep. 2:330–333. 2001.PubMed/NCBI View Article : Google Scholar

38 

Pfrimer K, Micheletto RF, Marchini JS, Padovan GJ, Moriguti JC and Ferriolli E: Impact of aging on urinary excretion of iron and zinc. Nutr Metab Insights. 7:47–50. 2014.PubMed/NCBI View Article : Google Scholar

39 

Detweiler K, Mayers D and Fletcher SG: Bacteruria and urinary tract infections in the elderly. Urol Clin North Am. 42:561–568. 2015.PubMed/NCBI View Article : Google Scholar

40 

Carver PL: The battle for iron between humans and microbes. Curr Med Chem. 25:85–96. 2018.PubMed/NCBI View Article : Google Scholar

41 

Kovacevic Z, Kalinowski DS, Lovejoy DB, Quach P, Wong J and Richardson DR: Iron chelators: Development of novel compounds with high and selective anti-tumour activity. Curr Drug Deliv. 7:194–207. 2010.PubMed/NCBI View Article : Google Scholar

42 

Iribarnegaray V, González MJ, Caetano AL, Platero R, Zunino P and Scavone P: Relevance of iron metabolic genes in biofilm and infection in uropathogenic Proteus mirabilis. Curr Res Microb Sci. 2(100060)2021.PubMed/NCBI View Article : Google Scholar

43 

Silva B and Faustino P: An overview of molecular basis of iron metabolism regulation and the associated pathologies. Biochim Biophys Acta. 1852:1347–1359. 2015.PubMed/NCBI View Article : Google Scholar

44 

Bauckman KA, Matsuda R, Higgins CB, DeBosch BJ, Wang C and Mysorekar IU: Dietary restriction of iron availability attenuates UPEC pathogenesis in a mouse model of urinary tract infection. Am J Physiol Renal Physiol. 316:F814–F822. 2019.PubMed/NCBI View Article : Google Scholar

45 

VanEpps JS and Younger JG: Implantable device-related infection. Shock. 46:597–608. 2016.PubMed/NCBI View Article : Google Scholar

46 

Vestby LK, Grønseth T, Simm R and Nesse LL: Bacterial Biofilm and its role in the pathogenesis of disease. Antibiotics (Basel). 9(59)2020.PubMed/NCBI View Article : Google Scholar

47 

Singh PK, Tack BF, McCray PB Jr and Welsh MJ: Synergistic and additive killing by antimicrobial factors found in human airway surface liquid. Am J Physiol Lung Cell Mol Physiol. 279:L799–L805. 2000.PubMed/NCBI View Article : Google Scholar

48 

Rogan MP, Taggart CC, Greene CM, Murphy PG, O'Neill SJ and McElvaney NG: Loss of microbicidal activity and increased formation of biofilm due to decreased lactoferrin activity in patients with cystic fibrosis. J Infect Dis. 190:1245–1253. 2004.PubMed/NCBI View Article : Google Scholar

49 

Sharma P, Dube D, Sinha M, Yadav S, Kaur P, Sharma S and Singh TP: Structural insights into the dual strategy of recognition by peptidoglycan recognition protein, PGRP-S: Structure of the ternary complex of PGRP-S with lipopolysaccharide and stearic acid. PLoS One. 8(e53756)2013.PubMed/NCBI View Article : Google Scholar

50 

Moreau-Marquis S, O'Toole GA and Stanton BA: Tobramycin and FDA-approved iron chelators eliminate Pseudomonas aeruginosa biofilms on cystic fibrosis cells. Am J Respir Cell Mol Biol. 41:305–313. 2009.PubMed/NCBI View Article : Google Scholar

51 

Zughaier SM and Cornelis P: Editorial: Role of iron in bacterial pathogenesis. Front Cell Infect Microbiol. 8(344)2018.PubMed/NCBI View Article : Google Scholar

52 

Kramer J, Özkaya Ö and Kümmerli R: Bacterial siderophores in community and host interactions. Nat Rev Microbiol. 18:152–163. 2020.PubMed/NCBI View Article : Google Scholar

53 

Ito A, Sato T, Ota M, Takemura M, Nishikawa T, Toba S, Kohira N, Miyagawa S, Ishibashi N, Matsumoto S, et al: In Vitro Antibacterial Properties of Cefiderocol, a Novel Siderophore Cephalosporin, against Gram-Negative Bacteria. Antimicrob Agents Chemother. 62:e01454–e01417. 2017.PubMed/NCBI View Article : Google Scholar

54 

Sheldon JR and Heinrichs DE: Recent developments in understanding the iron acquisition strategies of gram positive pathogens. FEMS Microbiol Rev. 39:592–630. 2015.PubMed/NCBI View Article : Google Scholar

55 

Sabnis A, Hagart KL, Klöckner A, Becce M, Evans LE, Furniss RCD, Mavridou DA, Murphy R, Stevens MM, Davies JC, et al: Colistin kills bacteria by targeting lipopolysaccharide in the cytoplasmic membrane. Elife. 10(e65836)2021.PubMed/NCBI View Article : Google Scholar

56 

Plachouras D, Karvanen M, Friberg LE, Papadomichelakis E, Antoniadou A, Tsangaris I, Karaiskos I, Poulakou G, Kontopidou F, Armaganidis A, et al: Population pharmacokinetic analysis of colistin methanesulfonate and colistin after intravenous administration in critically ill patients with infections caused by gram-negative bacteria. Antimicrob Agents Chemother. 53:3430–3436. 2009.PubMed/NCBI View Article : Google Scholar

57 

Qiao J, Liu Z, Cui S, Nagy T and Xiong MP: Synthesis and evaluation of an amphiphilic deferoxamine: Gallium-conjugated cationic random copolymer against a murine wound healing infection model of Pseudomonas aeruginosa. Acta Biomater. 126:384–393. 2021.PubMed/NCBI View Article : Google Scholar

58 

Banin E, Brady KM and Greenberg EP: Chelator-induced dispersal and killing of Pseudomonas aeruginosa cells in a biofilm. Appl Environ Microbiol. 72:2064–2069. 2006.PubMed/NCBI View Article : Google Scholar

59 

Hancock V, Dahl M and Klemm P: Abolition of biofilm formation in urinary tract Escherichia coli and Klebsiella isolates by metal interference through competition for fur. Appl Environ Microbiol. 76:3836–3841. 2010.PubMed/NCBI View Article : Google Scholar

60 

Zhang Y, Lin Y, Zhang X, Chen L, Xu C, Liu S, Cao J, Zheng X, Jia H, Chen L, et al: Combining Colistin with furanone C-30 rescues Colistin resistance of gram-negative bacteria in vitro and in vivo. Microbiol Spectr. 9(e0123121)2021.PubMed/NCBI View Article : Google Scholar

61 

Mojica MF, Bonomo RA and Fast W: B1-Metallo-β-Lactamases: Where do we stand? Curr Drug Targets. 17:1029–1050. 2016.PubMed/NCBI View Article : Google Scholar

Related Articles

Journal Cover

September-2023
Volume 26 Issue 3

Print ISSN: 1792-0981
Online ISSN:1792-1015

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Erinmez M and Erinmez M: Effects of deferoxamine on intrinsic colistin resistance of <em>Proteus mirabilis</em>. Exp Ther Med 26: 459, 2023
APA
Erinmez, M., & Erinmez, M. (2023). Effects of deferoxamine on intrinsic colistin resistance of <em>Proteus mirabilis</em>. Experimental and Therapeutic Medicine, 26, 459. https://doi.org/10.3892/etm.2023.12158
MLA
Erinmez, M., Zer, Y."Effects of deferoxamine on intrinsic colistin resistance of <em>Proteus mirabilis</em>". Experimental and Therapeutic Medicine 26.3 (2023): 459.
Chicago
Erinmez, M., Zer, Y."Effects of deferoxamine on intrinsic colistin resistance of <em>Proteus mirabilis</em>". Experimental and Therapeutic Medicine 26, no. 3 (2023): 459. https://doi.org/10.3892/etm.2023.12158