Next Article in Journal
New N-acyl Thiourea Derivatives: Synthesis, Standardized Quantification Method and In Vitro Evaluation of Potential Biological Activities
Next Article in Special Issue
Antimicrobial Resistance: Two-Component Regulatory Systems and Multidrug Efflux Pumps
Previous Article in Journal
The African Wastewater Resistome: Identifying Knowledge Gaps to Inform Future Research Directions
Previous Article in Special Issue
Efflux, Signaling and Warfare in a Polymicrobial World
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Inhibition of Staphylococcus pseudintermedius Efflux Pumps by Using Staphylococcus aureus NorA Efflux Pump Inhibitors

by
Elisa Rampacci
1,*,
Tommaso Felicetti
2,
Giada Cernicchi
2,
Valentina Stefanetti
1,
Stefano Sabatini
2 and
Fabrizio Passamonti
1
1
Department of Veterinary Medicine, University of Perugia, Via San Costanzo 4, 06126 Perugia, Italy
2
Department of Pharmaceutical Sciences, Via Del Liceo 1, 06123 Perugia, Italy
*
Author to whom correspondence should be addressed.
Antibiotics 2023, 12(5), 806; https://doi.org/10.3390/antibiotics12050806
Submission received: 24 March 2023 / Revised: 13 April 2023 / Accepted: 20 April 2023 / Published: 24 April 2023
(This article belongs to the Special Issue Efflux Pumps in Bacteria: What They Do and How We Can Stop Them)

Abstract

:
One promising approach in treating antibiotic-resistant bacteria is to “break” resistances connected with antibacterial efflux by co-administering efflux pump inhibitors (EPIs) with antibiotics. Here, ten compounds, previously optimized to restore the susceptibility to ciprofloxacin (CIP) of norA-overexpressing Staphylococcus aureus, were evaluated for their ability to inhibit norA-mediated efflux in Staphylococcus pseudintermedius and synergize with CIP, ethidium bromide (EtBr), gentamycin (GEN), and chlorhexidine digluconate (CHX). We focused efforts on S. pseudintermedius as a pathogenic bacterium of concern within veterinary and human medicine. By combining data from checkerboard assays and EtBr efflux inhibition experiments, the hits 2-arylquinoline 1, dihydropyridine 6, and 2-phenyl-4-carboxy-quinoline 8 were considered the best EPIs for S. pseudintermedius. Overall, most of the compounds, except for 2-arylquinoline compound 2, were able to fully restore the susceptibility of S. pseudintermedius to CIP and synergize with GEN as well, while the synergistic effect with CHX was less significant and often did not show a dose-dependent effect. These are valuable data for medicinal chemistry optimization of EPIs for S. pseudintermedius and lay the foundation for further studies on successful EPIs to treat staphylococcal infections.

1. Introduction

The rise of bacterial pathogens resistant to antibacterial agents poses a major threat to human and animal health worldwide. The ever-increasing demand for novel antimicrobial strategies, in contrast to the current level of investments, has led to the development of several early phase-hit compounds [1]. Among these, efflux pump inhibitors (EPIs) are antimicrobial resistance breakers targeting bacterial membrane transporters, known as efflux pumps (EPs), which are involved in the extrusion of noxious compounds, including antimicrobial agents [2]. EPIs have the potential to reverse resistance to several antibacterials [3,4], reduce biofilm formation associated with overexpressed EPs [4], and suppress the evolution of resistance [5]. Particularly, EPIs can be used to optimize the process of prevention of antibacterial efflux, thus boosting the efficacy of drugs acting on intracellular bacterial targets.
The need to address antimicrobial resistance mediated by NorA, one of the most studied EPs in Staphylococcus aureus [5,6,7], led to the identification of potent EPIs by exploitation of in silico drug repositioning approaches and medicinal chemistry campaigns [8,9,10]. We have been working for years on the design and synthesis of novel S. aureus NorA EPIs by comparing their biological activity on strains overexpressing norA or norA-deleted, resulting in the identification of several molecules able to fully restore the susceptibility to fluoroquinolone ciprofloxacin (CIP) of resistant S. aureus strains overexpressing NorA EP [8].
Knowledge of the efficacy of these compounds against bacterial species other than S. aureus or in combination with antimicrobial classes different from fluoroquinolones is limited to recent works reporting some derivatives also acting as nontuberculous mycobacteria EPIs [11]. However, as far as we know, no studies have considered using EPIs for pathogens of concern in the veterinary sector. Among these, Staphylococcus pseudintermedius has been identified by the EFSA Panel on Animal Health and Welfare among the most relevant antimicrobial-resistant bacteria due to its frequent implication in clinical diseases in dogs and cats, zoonotic risk, and the high levels of resistance to clinically relevant antibiotics that worryingly limit treatment options [12]. S. pseudintermedius constitutes about 90% of staphylococci isolated from healthy canine carriers and is the most prevalent cause of canine bacterial infections [13]. It is responsible for a large variety of opportunistic infections in dogs, mainly skin infections and otitis externa but also infections of the cornea, urinary, respiratory and reproductive tract [12,14,15]. These are often prolonged inflammatory disorders difficult to treat due to a high rate of biofilm production [16]. S. pseudintermedius is also frequently isolated from feline pyoderma [13], and it has the potential to be virulent in human hosts [17,18].
It was recently demonstrated that the overexpression of EPs in S. pseudintermedius, mainly NorA, is implicated in antimicrobial resistance development, particularly to CIP, which is the major metabolite of the veterinary fluoroquinolone enrofloxacin in dogs [19], gentamycin (GEN), and chlorhexidine gluconate (CHX), and it contributes to a remarkable increase of biofilm production [4]. These findings make using combinations of EPI/antibiotic or EPI/biocide an attractive strategy for preserving the efficacy of veterinary first-line drugs and antibiofilm activity. However, data concerning the relationship between the chemical structure of potential EPI compounds and the biological activity of S. pseudintermedius are lacking, as well as structural information about the NorA protein and its molecular interaction with EPIs.
Overall, NorA function is supposed to be conserved across the staphylococcal species [20]. However, we know that the nucleotide identity of most S. pseudintermedius EPs with S. aureus is approximately 63–73% [4]. Focusing on NorA, the S. pseudintermedius norA gene has a 70% nucleotide identity to S. aureus norA [4], corresponding to a 70% structural similarity at the protein level. These results confirm that genetic variability exists, which may impact the inhibitor design [20].
In this work, we investigated the ability of some compounds previously reported by us as S. aureus NorA EPIs to inhibit S. pseudintermedius EPs to lay the foundation of medicinal chemistry campaigns aimed at identifying novel NorA S. pseudintermedius EPIs.

2. Results

2.1. Synergistic Activity of EPI/Antimicrobial Combinations

Ten ‘in-house’ molecules (compounds 110 depicted in Figure 1) were selected based on: (i) their own activity as S. aureus NorA EPIs (compound 2 was included as inactive EPI), and (ii) their structural difference in terms of central scaffold (compound 6–nicardipine was also included since it was identified as S. aureus NorA EPI by a drug repurposing approach). Accordingly, seven different scaffolds can be identified within the ten selected molecules: (i) the mostly explored 2-arylquinoline present in compounds 1 [21], 2 [22], 3 [23], and 4 [24]; (ii) the pyrazolobenzothiazine in compound 5 [25]; (iii) the dihydropyridine present in nicardipine (compound 6) [26]; (iv) the 3-phenylquinolone in compound 7 [11]; (v) the 2-phenyl-4-carboxy-quinoline in compound 8 [27]; (vi) the 2-phenyl-quinazoline in compound 9 [28]; and (vii) the benzimidazole in compound 10 [27].
To establish the concentrations of each EPI to be used in synergistic studies, we determined the Minimum Inhibitory Concentration (MIC) of ten selected compounds for S. pseudintermedius ATCC 49444 (wild-type) and its efflux mutant S. pseudintermedius_P8 (overexpressing norA and harboring an 11 bp deletion in the promoter region of the norA gene). The MICs are reported in Table 1. Overall, all compounds, when used alone, had weak antimicrobial activity against S. pseudintermedius, except for the pyrazobenothiazine analog 5, which showed MIC values of 6.25 mg/L against both strains. MICs of all tested compounds exhibited the same values against both strains, suggesting that these derivatives are not S. pseudintermedius NorA substrates. On the other hand, being the susceptibility of S. pseudintermedius_P8 to these molecules affected by the increased efflux (MIC values shown in Table 1), EtBr, GEN, CIP, and CHX are considered to be substrates of S. pseudintermedius efflux system.
Checkerboard assays were then performed by combining EtBr, GEN, CIP, and CHX at scalar concentrations with serial dilutions of each compound used from MIC/4 to MIC/128 to prevent any synergistic effect due to an intrinsic antibacterial activity of the putative tested EPIs 110.
When tested against norA overexpressing strain S. pseudintermedius_P8, all compounds exhibited modest to excellent synergistic activities depending on the combined antimicrobial. Compounds 1, 3, 58, and 10 were synergistic with EtBr and all the antimicrobials tested (see Table 2 for modulation factor “MF” values). Of note, since all compounds were tested at serial dilutions starting from MIC/4 to MIC/128, it is essential to consider their effective concentration when comparing data (concentrations used are shown in Table 2). Overall, 2-arylquinoline derivatives 1, 3, and 4 exhibited a significant synergistic effect with CIP and EtBr and a weaker activity when combined with GEN and CHX. This discrepancy in synergistic activity can be imputed to a reduced efflux of these two antimicrobials. Indeed, MIC values of GEN and CHX are significantly less affected (4-fold difference between the two strains) than those of CIP and EtBr (64-fold for both). Of note, at MIC/128 (1.57 mg/L), compound 1 reduced the CIP MIC by 8-fold, thus exhibiting a very potent synergistic effect. On the contrary, the 2-arylquinoline analog 2 showed a weaker (or absent) synergistic effect in accordance with data previously obtained against S. aureus strains (the synergism with CIP was considered ineffective since it was not dose-dependent). Results of the pyrazolobenzothiazine derivative 5 were affected by its low MIC values, forcing us to use it at very low concentrations in checkerboard assays. However, up to MIC/8 (0.78 mg/L), compound 5 displayed a good synergistic effect (MF = 4) with GEN, CIP, and EtBr. The nicardipine drug (compound 6) exhibited a very good profile by synergizing with GEN, CIP, and EtBr up to very low concentrations. In particular, at 1.57 mg/L (MIC/128), compound 6 still retained a MF of 4 in combination with CIP, similar to the 3-phenylquinolone analog 7 and the 2-phenyl-4-carboxy-quinoline derivative 8. Promising results were also obtained with the quinazoline derivative 9 and the benzimidazole 10 that showed a significant synergistic effect with GEN, CIP, and EtBr up to low concentrations (6.25 and 12.5 mg/L, respectively). In addition, both of them reduced the CIP MIC 4-fold at 1.57 mg/L, thus highlighting a potent synergistic activity.
As a confirmation of the synergistic effect, the minimum fractional inhibitory concentration index (FICI) obtained for each combination EPI/antimicrobial is reported in Table 3. Data supported the potent synergism observed for most of the compounds, confirming a greater effect of the EPIs in combination with the fluoroquinolone CIP and EtBr. As expected, by also analyzing the effect of compounds in combinations with antimicrobials and EtBr against the wild-type strain (ATCC 49444), it is evident a poor or absent synergistic effect with antimicrobials (Table 3). On the other hand, a modest synergism is present when most of the compounds were combined with EtBr, which is a known nonspecific substrate of bacterial EPs, thus suggesting that most of the compounds likely possess the ability to reduce the efflux deriving by the basal expression of EPs.

2.2. Comparative Inhibition of EtBr Efflux

Before testing the inhibiting activity of selected compounds on S. pseudintermedius ATCC 49444 and S. pseudintermedius_P8 EPs, the capability of the bacterial strains to accumulate EtBr was tested. Predictably, EtBr accumulation was lower inside norA-overexpressing S. pseudintermedius_P8 than the parent strain (Figure 2a). Overall, the assay performed in the presence of CCCP showed a detectable accumulation for both strains.
On the strength of the data from checkerboard assays, compounds 1, 3, and 68 were advanced toward fluorometry-based EtBr efflux inhibition test. The EPIs were tested at MIC/8 (compounds 1 and 68, 25 mg/L; compound 3, 6.25 mg/L) and MIC/64 (compounds 1 and 68, 3.13 mg/L; compound 3, 0.8 mg/L). The results are presented in Figure 2. The use of compounds 1, 3, and 68 slightly influenced the EtBr efflux activity of S. pseudintermedius ATCC 49444 (Figure 2b). As expected, S. pseudintermedius_P8 showed high EtBr efflux in the absence of EPI. The residual fluorescence of EtBr in S. pseudintermedius_P8 was indeed 58% lower than S. pseudintermedius ATCC 49444, confirming the presence of highly overexpressed EPs. Compounds 3 and 68 used at MIC/64 did not alter the capability of S. pseudintermedius_P8 to extrude EtBr, while compound 1 at MIC/64 showed a weak inhibitory effect (Figure 2c). At the end of the test, all compounds used at MIC/8 increased EtBr fluorescence as follows: 7 22.5% (p = 0.078); 3 29.9% (p = 0.012); 1 32.6% (p = 0.006); 6 35.1% (p = 0.003); 8 46.1% (p = 0.001) (Figure 2c). These findings confirm that the synergistic effect observed by checkerboard assays is due to the inhibition of drug efflux.

3. Discussion

The overuse and misuse of antibiotics in human and veterinary healthcare and agriculture have accelerated the emergence and spread of antimicrobial-resistant bacteria worldwide, posing a threat to the effective treatment of infections. [29,30,31]. One proposed approach to treat antibiotic-resistant bacteria is to “break” resistances by co-administering appropriate non-antibiotic drugs with failing antibiotics to re-sensitizing resistant bacteria [32]. Among the antibiotic resistance breakers, EPIs can reduce resistance by blocking bacterial EPs and thus increasing intracellular antibiotic concentration.
A wide array of compounds with EPI activity have been reported to date [32,33]. Most of them were optimized to target ESKAPE pathogens [34,35] that are considered critically important owing to their role in many human infections and the frequency of antibiotic resistance worryingly limiting treatment options [36]. Particularly, research efforts were addressed to target S. aureus NorA using EPIs from natural products, de novo synthesis, and repurposing of previously-approved drugs [8,9,10]. Pursuing the development of S. aureus NorA inhibitors, we have previously identified via drug repurposing approaches and medicinal chemistry campaigns a wide array of NorA EPIs characterized by different chemical scaffolds [21,22,23,25,26,28]. Most of these compounds exhibited NorA inhibition activity by restoring at very low concentrations CIP MIC against resistant S. aureus strains and overexpressing norA while not showing any effect against S. aureus strains not expressing norA.
Here, we validated the hypothesis that EPIs specifically optimized to reverse S. aureus NorA-mediated fluoroquinolone resistance can successfully reverse resistance to CIP, GEN, and CHX associated with the overexpression of norA in S. pseudintermedius. Our study clearly shows that most of the compounds could fully restore the susceptibility of S. pseudintermedius P8 to CIP, except for compound 2 (included as inactive S. aureus NorA EPI). This is important because enrofloxacin is largely metabolized to CIP in dogs [19]. Moreover, even if to a lesser extent, all the compounds synergized with GEN as well. On the other hand, the synergistic effect with CHX was less significant and often not showing a dose-dependent effect.
Although these results suggest that S. aureus NorA EPIs can also be used as S. pseudintermedius NorA EPIs, structure-activity relationship (SAR) information acquired over the years about S. aureus NorA inhibition cannot be completely transferred for designing novel S. pseudintermedius NorA EPIs. Indeed, the introduction of a methoxy group at the C-6 position of the quinoline core (compound 3), as well as the replacement of the 2-phenyl ring with a chloro-thiophene moiety (compound 4), led to an increase of the S. aureus NorA EPI activity with respect to “naked” 2-phenylquinoline derivatives such as compound 1. Herein, instead, we observed that the activity of compound 1 was comparable to or slightly better than compounds 3 and 4, thus suggesting that SAR information should be revised. Similarly, replacing the quinoline core with a quinazoline scaffold (compound 9) led to an improvement in terms of S. aureus NorA inhibition activity that was not confirmed for S. pseudintermedius NorA EP. On the other hand, 3-phenylquinolone derivative 7, which exhibited a weaker S. aureus NorA EPI activity than 2-phenylquinoline derivatives (such as compounds 1, 3, and 4), retained a promising S. pseudintermedius NorA inhibition. Interestingly, compound 7 is also endowed with good EPI activity against nontuberculous mycobacteria resistant to clarithromycin and overexpressing EPs MAV_1406 and MAV_1695 [11]. Nicardipine (compound 6), the carboxy-quinoline derivative 8, and the benzimidazole analog 10 showed a comparable inhibition of both NorA EPs. The pyrazolobenzothiazine analog 5 was instead affected by low MIC values that compromised its use as EPI.
The results from the fluorometric tests provided further evidence for the capability of compounds 1, 3, and 68 to inhibit S. pseudintermedius EPs, validating their inhibition as the main mechanism involved in the synergistic effect with antimicrobials.
By combining data from checkerboard assays and EtBr efflux inhibition experiments, compounds 1, 6, and 8 could be considered the best NorA EPIs able to strongly synergize with CIP and EtBr against S. pseudintermedius P8, also showing a modest synergism with GEN and CHX. In addition, considering that their EPI activity was significantly dependent on the overexpression of norA EP (poor synergistic effect and no EtBr efflux inhibition were observed against the wild-type S. pseudintermedius strain), we are confident that these three derivatives act by inhibiting NorA efflux mechanisms. Based on the previously reported cytotoxicity evaluation of these compounds [26,27], it is interesting to note that all of them showed synergistic activity with the fluoroquinolone CIP and with the aminoglycoside GEN at concentrations significantly lower than their CC50 values towards human cells. Special attention should be given to nicardipine (compound 6), which is a non-antibiotic approved drug for the treatment of high blood pressure and angina. At 3.13 mg/L, nicardipine was able to reduce CIP MIC by 8-fold (from 8 to 1 mg/L) against S. pseudintermedius P8 while showing a CC50 of 188.75 and 68.73 mg/L towards HepG2 and A549 cell lines, respectively [26]. Although further studies should be performed, nicardipine (compound 6) could represent a promising candidate for pre-clinical studies in in vivo models. In parallel, results obtained for the 2-phenylquinoline derivative 1 and the 4-carboxyquinoline analog 8 give us great hope for future medicinal chemistry efforts aimed at improving S. pseudintermedius NorA EPI activity, safety, and pharmacokinetic profile.
Concerns have been raised questioning the therapeutic usefulness of EPIs at the community level against bacteria harboring mutations in antibiotic targets, such as gyrase-coding genes for fluoroquinolone resistance [37,38]. However, it was recently demonstrated that overexpressed EPs (specifically NorA in S. aureus) could create a high-resistance-evolvability bacterial niche by promoting the accumulation of antibiotic-resistance mutations or increasing the fitness benefit provided by resistance mutations [5,39,40]. Moreover, high efflux appears to be linked to the downregulation of DNA repair and mutagenesis [39]. From this perspective, EPIs might be used to prevent resistance evolution and preserve the efficacy of existing antibiotics. To further our research, future work should concentrate on using these EPI derivatives to inhibit the efflux-mediated evolvability of staphylococcal species.

4. Materials and Methods

4.1. Bacterial Strains and Growth Conditions

The strains used in this study were S. pseudintermedius ATCC 49444 and its derivative S. pseudintermedius_P8. This was obtained by adapting S. pseudintermedius ATCC 49444 to increasing concentrations of EtBr, a known substrate of bacterial efflux system [3,41]. More details on the growth conditions of S. pseudintermedius_P8 are given in our previous paper [4]. S. pseudintermedius_P8 overexpresses norA and harbors an 11 bp deletion in the norA promoter region. S. pseudintermedius_P8 shows higher MICs of EtBr, CIP, GEN, and CHX than the parent strain, as reported in Table 1.

4.2. Efflux Pump Inhibitors

Compounds tested as EPIs have been selected within an ‘in-house’ library of previously published S. aureus NorA EPIs. Seven different scaffolds identified within ten selected molecules were investigated: 2-arylquinoline present in compounds 1, 2, 3, and 4, the pyrazolobenzothiazine in compound 5, the dihydropyridine present in compound 6, the 3-phenylquinolone in compound 7, the 2-phenyl-4-carboxy-quinoline in compound 8, the 2-phenyl-quinazoline in compound 9, and the benzimidazole in compound 10 [11,21,22,23,24,25,26,27,28].

4.3. MIC Determination

The lowest concentration of compounds 110 that inhibits visible bacterial growth (MIC) was determined in triplicate by broth microdilution according to CLSI recommendations [42]. A total of 96-well plates were inoculated with 100 µL of 2-fold serial dilutions of each compound in cation-adjusted Mueller-Hinton broth (CAMHB) to test a dose range of 100–3.125 mg/L. S. pseudintermedius ATCC 49444 and P8 colonies were resuspended in a sterile medium, and the suspension turbidity was measured spectrophotometrically at OD600. The bacterial suspensions were then adjusted in CAMHB to 5 × 105 CFU/mL. Plates were inoculated with 100 µL of bacterial suspensions and incubated at 37 °C for 20 h.

4.4. Sinergy Studies

The synergistic activity of compounds 110 combined with EtBr, CIP, GEN, and CHX was evaluated on S. pseudintermedius ATCC 49444 and S. pseudintermedius P8 using two-dimensional checkerboard assays using 96-well microtiter plates according to the standards [43]. Each EPI was tested in a concentration range between MIC/4 and MIC/128. For S. pseudintermedius P8, the antibiotic dilutions tested ranged from the MIC to -1 doubling dilution (1 log2) below the MIC value for the original strain. For S. pseudintermedius ATCC 49444, antibiotics were tested from the MIC to MIC/8. Inoculum preparation was performed in CAMHB by colony suspension from 24 h cultures on Cation-adjusted Mueller Hinton agar. One hundred microliters of the bacterial suspension were inoculated in each well at a final concentration of 5 × 104 CFUs. The plates were incubated aerobically at 37 °C for 20 h. After reading well optical turbidity, the FICI was calculated for synergy interpretation as follows:
M I C d r u g     c o m b i n a t i o n M I C d r u g   a l o n e + M I C E P I   c o m b i n a t i o n M I C E P I   a l o n e
The combination was considered synergistic when the FICI was ≤0.5. MF, Modulation Factor, represents the n-fold reduction of the MIC of the corresponding antimicrobial when combined with the EPI.

4.5. Fluorometric Tests

Our experimental setup is based on the one proposed by Kaatz [44]. S. pseudintermedius ATCC 49444 and P8 were grown overnight at 37 °C in static conditions in 10 mL of TSB without EtBr or supplemented with EtBr at 16 mg/L, respectively.
Preliminarily, we evaluated the capability of the strains to accumulate EtBr. Bacteria were pelleted by centrifugation at 8000 rpm for 10 min and washed twice with sterile PBS. The turbidity of the suspensions was adjusted to 0.6 OD600 in PBS. Bacterial cells were loaded with EtBr at 10 mg/L using 20 mg/L of CCCP. The fluorescence of DNA-bound EtBr was measured at 37 °C at 60 s intervals for 30 min using a TECAN Infinite 200 PRO reader at excitation/emission 530/600 nm.
To assay the inhibitory effect of selected compounds on EtBr efflux, S. pseudintermedius ATCC 49444 and P8 were loaded with EtBr at 10 mg/L using 20 mg/L of CCCP, as described. After 20 min of incubation at room temperature in the dark, the bacterial suspensions were pelleted, and cells were resuspended in sterile PBS at 0.6 OD600. One hundred microliters of each suspension were added to wells in a white microtiter plate containing, in duplicate: (i) 100 μL of PBS with glucose (final concentration 0.4%) and without EPIs (conditions of efflux); (ii) 100 μL of compound 1 at a final concentration of 25 mg/L and (iii) 3.13 mg/L in the presence of 0.4% glucose; (iv) 100 μL of compound 6 at a final concentration of 25 mg/L and (v) 3.13 mg/L in the presence of 0.4% glucose; (vi) 100 μL of compound 8 at a final concentration of 25 mg/L and (vii) 3.13 mg/L in the presence of 0.4% glucose; (viii) 100 μL of compound 3 at a final concentration of 6.25 mg/L and (ix) 0.8 mg/L in the presence of 0.4% glucose; (x) 100 μL of compound 7 at a final concentration of 25 mg/L and (xi) 3.13 mg/L in the presence of 0.4% glucose. The plate was immediately read by a TECAN Infinite 200 PRO reader at excitation/emission 530/600 nm. The fluorescence of DNA-bound EtBr was measured at 37 °C at 60 s intervals for 30 min. The assay was repeated two times. Relative fluorescence remaining at each time point was normalized against the value measured at time 0 (T0), as follows:
100 × F I t F I t 0
where FIt is the fluorescence intensity at different time points, and FIt0 is the fluorescence intensity at T0.

4.6. Statistical Analyses

Statistical analyses were performed using SPSS (version 17.0). The Mann–Whitney U test was applied to compare the biological replicates between treatments in fluorescence assays. A p-value < 0.05 was assumed as significant.

5. Conclusions

To summarize, this work provided valuable data for the medicinal chemistry optimization of EPIs for S. pseudintermedius. It has led us to conclude that (1) a small set of ‘in-house’ compounds optimized to reverse CIP resistance of S. aureus overexpressing norA have EPI activity against S. pseudintermedius; (2) all compounds synergized with CIP and GEN; (3) the synergistic effect resulted dependent on overexpressed norA as observed by fluorometric assays with EtBr; (4) compounds 1, 6, and 8 emerged as able to strongly inhibit the activity of S. pseudintermedius NorA EP and completely restore the activity of CIP against the resistant strain S. pseudintermedius_P8.
These data pave the way for further studies on antimicrobial-resistant staphylococcal species other than S. aureus, which have highly virulent features for both humans and companion animals.

Author Contributions

Conceptualization, E.R., T.F., S.S. and F.P.; methodology, E.R. and T.F.; validation, S.S. and F.P.; formal analysis, E.R.; investigation, E.R., T.F., G.C. and V.S.; resources, F.P. and S.S.; data curation, E.R. and T.F.; writing—original draft preparation, E.R. and T.F.; writing—review and editing, V.S., G.C., S.S. and F.P.; visualization, E.R.; supervision, F.P. and S.S.; project administration, F.P.; funding acquisition, F.P. and S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been funded by the European Union-NextGenerationEU under the Italian Ministry of University and Research (MUR) National Innovation Ecosystem grant ECS00000041-VITALITY. We acknowledge Università degli Studi di Perugia and MUR for support within the project Vitality. This research was funded in part by the University of Perugia, Fondi per la ricerca di base 2019 grant number 6RICBASEPREM19.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Miethke, M.; Pieroni, M.; Weber, T.; Brönstrup, M.; Hammann, P.; Halby, L.; Arimondo, P.B.; Glaser, P.; Aigle, B.; Bode, H.B.; et al. Towards the sustainable discovery and development of new antibiotics. Nat. Rev. Chem. 2021, 5, 726–749. [Google Scholar] [CrossRef]
  2. Costa, S.S.; Viveiros, M.; Amaral, L.; Couto, I. Multidrug efflux pumps in Staphylococcus aureus: An update. Open Microbiol. J. 2013, 7, 59–71. [Google Scholar] [CrossRef] [PubMed]
  3. Rampacci, E.; Marenzoni, M.L.; Cannalire, R.; Pietrella, D.; Sabatini, S.; Giovagnoli, S.; Felicetti, T.; Pepe, M.; Passamonti, F. Ethidium bromide exposure unmasks an antibiotic efflux system in Rhodococcus equi. J. Antimicrob. Chemother. 2021, 76, 2040–2048. [Google Scholar] [CrossRef] [PubMed]
  4. Rampacci, E.; Felicetti, T.; Pietrella, D.; Sabatini, S.; Passamonti, F. Drug efflux transporters in Staphylococcus pseudintermedius: In silico prediction and characterization of resistance. J. Antimicrob. Chemother. 2022, 77, 3283–3290. [Google Scholar] [CrossRef]
  5. Papkou, A.; Hedge, J.; Kapel, N.; Young, B.; MacLean, R.C. Efflux pump activity potentiates the evolution of antibiotic resistance across S. aureus isolates. Nat. Commun. 2020, 11, 3970. [Google Scholar] [CrossRef] [PubMed]
  6. Yu, X.H.; Hao, Z.H.; Liu, P.L.; Liu, M.M.; Zhao, L.L.; Zhao, X. Increased expression of efflux pump NorA drives the rapid evolutionary trajectory from tolerance to resistance against ciprofloxacin in Staphylococcus aureus. Antimicrob. Agents Chemother. 2022, 66, e0059422. [Google Scholar] [CrossRef] [PubMed]
  7. Brawley, D.N.; Sauer, D.B.; Li, J.; Zheng, X.; Koide, A.; Jedhe, G.S.; Suwatthee, T.; Song, J.; Liu, Z.; Arora, P.S.; et al. Structural basis for inhibition of the drug efflux pump NorA from Staphylococcus aureus. Nat. Chem. Biol. 2022, 18, 706–712. [Google Scholar] [CrossRef]
  8. Cernicchi, G.; Felicetti, T.; Sabatini, S. Microbial efflux pump inhibitors: A journey around quinoline and indole derivatives. Molecules 2021, 26, 6996. [Google Scholar] [CrossRef]
  9. Lamut, A.; Peterlin Mašič, L.; Kikelj, D.; Tomašič, T. Efflux pump inhibitors of clinically relevant multidrug resistant bacteria. Med. Res. Rev. 2019, 39, 2460–2504. [Google Scholar] [CrossRef]
  10. Monteiro, K.L.C.; de Aquino, T.M.; Mendonça Junior, F.J.B. An update on Staphylococcus aureus Nora efflux pump inhibitors. Curr. Top. Med. Chem. 2020, 20, 2168–2185. [Google Scholar] [CrossRef]
  11. Felicetti, T.; Machado, D.; Cannalire, R.; Astolfi, A.; Massari, S.; Tabarrini, O.; Manfroni, G.; Barreca, M.L.; Cecchetti, V.; Viveiros, M.; et al. Modifications on c6 and c7 positions of 3-phenylquinolone efflux pump inhibitors led to potent and safe antimycobacterial treatment adjuvants. ACS Infect. Dis. 2019, 5, 982–1000. [Google Scholar] [CrossRef]
  12. EFSA Panel on Animal Health and Welfare (AHAW); Nielsen, S.S.; Bicout, D.J.; Calistri, P.; Canali, E.; Drewe, J.A.; Garin-Bastuji, B.; Gonzales Rojas, J.L.; Gortazar Schmidt, C.; Herskin, M.; et al. Assessment of animal diseases caused by bacteria resistant to antimicrobials: Dogs and cats. EFSA J. 2021, 19, e06680. [Google Scholar] [PubMed]
  13. Bannoehr, J.; Guardabassi, L. Staphylococcus pseudintermedius in the dog: Taxonomy, diagnostics, ecology, epidemiology and pathogenicity. Vet. Dermatol. 2012, 23, 253-e52. [Google Scholar] [CrossRef]
  14. Lynch, S.A.; Helbig, K.J. The Complex Diseases of Staphylococcus pseudintermedius in Canines: Where to Next? Vet. Sci. 2021, 8, 11. [Google Scholar] [CrossRef] [PubMed]
  15. LoPinto, A.J.; Mohammed, H.O.; Ledbetter, E.C. Prevalence and risk factors for isolation of methicillin-resistant Staphylococcus in dogs with keratitis. Vet. Ophthalmol. 2015, 18, 297–303. [Google Scholar] [CrossRef]
  16. Singh, A.; Walker, M.; Rousseau, J.; Weese, J.S. Characterization of the biofilm forming ability of Staphylococcus pseudintermedius from dogs. BMC Vet. Res. 2013, 9, 93. [Google Scholar] [CrossRef] [PubMed]
  17. Yarbrough, M.L.; Lainhart, W.; Burnham, C.A. Epidemiology, clinical characteristics, and antimicrobial susceptibility profiles of human clinical isolates of Staphylococcus intermedius group. J. Clin. Microbiol. 2018, 56, e01788-17. [Google Scholar] [CrossRef] [PubMed]
  18. Bhooshan, S.; Negi, V.; Khatri, P.K. Staphylococcus pseudintermedius: An undocumented, emerging pathogen in humans. GMS Hyg. Infect. Control 2020, 15, Doc32. [Google Scholar]
  19. Cester, C.C.; Toutain, P.L. A comprehensive model for enrofloxacin to ciprofloxacin transformation and disposition in dog. J. Pharm. Sci. 1997, 86, 1148–1155. [Google Scholar] [CrossRef]
  20. Ferreira, C.; Abrantes, P.; Costa, S.S.; Viveiros, M.; Couto, I. Occurrence and variability of the efflux pump gene norA across the Staphylococcus genus. Int. J. Mol. Sci. 2022, 23, 15306. [Google Scholar] [CrossRef]
  21. Sabatini, S.; Gosetto, F.; Manfroni, G.; Tabarrini, O.; Kaatz, G.W.; Patel, D.; Cecchetti, V. Evolution from a natural flavones nucleus to obtain 2-(4-propoxyphenyl)quinoline derivatives as potent inhibitors of the S. aureus NorA efflux pump. J. Med. Chem. 2011, 54, 5722–5736. [Google Scholar] [CrossRef] [PubMed]
  22. Sabatini, S.; Gosetto, F.; Iraci, N.; Barreca, M.L.; Massari, S.; Sancineto, L.; Manfroni, G.; Tabarrini, O.; Dimovska, M.; Kaatz, G.W.; et al. Re-evolution of the 2-phenylquinolines: Ligand-based design, synthesis, and biological evaluation of a potent new class of Staphylococcus aureus NorA efflux pump inhibitors to combat antimicrobial resistance. J. Med. Chem. 2013, 56, 4975–4989. [Google Scholar] [CrossRef] [PubMed]
  23. Felicetti, T.; Cannalire, R.; Pietrella, D.; Latacz, G.; Lubelska, A.; Manfroni, G.; Barreca, M.L.; Massari, S.; Tabarrini, O.; Kieć-Kononowicz, K.; et al. 2-phenylquinoline S. aureus NorA efflux pump inhibitors: Evaluation of the importance of methoxy group introduction. J. Med. Chem. 2018, 61, 7827–7848. [Google Scholar] [CrossRef] [PubMed]
  24. Felicetti, T.; Mangiaterra, G.; Cannalire, R.; Cedraro, N.; Pietrella, D.; Astolfi, A.; Massari, S.; Tabarrini, O.; Manfroni, G.; Barreca, M.L.; et al. C-2 phenyl replacements to obtain potent quinoline-based Staphylococcus aureus NorA inhibitors. J. Enzym. Inhib. Med. Chem. 2020, 35, 584–597. [Google Scholar] [CrossRef]
  25. Sabatini, S.; Gosetto, F.; Serritella, S.; Manfroni, G.; Tabarrini, O.; Iraci, N.; Brincat, J.P.; Carosati, E.; Villarini, M.; Kaatz, G.W.; et al. Pyrazolo[4,3-c][1,2]Benzothiazines 5,5-Dioxide: A promising new class of Staphylococcus aureus NorA efflux pump inhibitors. J. Med. Chem. 2012, 55, 3568–3572. [Google Scholar] [CrossRef]
  26. Astolfi, A.; Felicetti, T.; Iraci, N.; Manfroni, G.; Massari, S.; Pietrella, D.; Tabarrini, O.; Kaatz, G.W.; Barreca, M.L.; Sabatini, S.; et al. Pharmacophore-based repositioning of approved drugs as novel Staphylococcus aureus NorA efflux pump inhibitors. J. Med. Chem. 2017, 60, 1598–1604. [Google Scholar] [CrossRef]
  27. Cannalire, R.; Mangiaterra, G.; Felicetti, T.; Astolfi, A.; Cedraro, N.; Massari, S.; Manfroni, G.; Tabarrini, O.; Vaiasicca, S.; Barreca, M.L.; et al. Structural modifications of the quinolin-4-yloxy core to obtain new Staphylococcus aureus NorA inhibitors. Int. J. Mol. Sci. 2020, 21, 7037. [Google Scholar] [CrossRef]
  28. Cedraro, N.; Cannalire, R.; Astolfi, A.; Mangiaterra, G.; Felicetti, T.; Vaiasicca, S.; Cernicchi, G.; Massari, S.; Manfroni, G.; Tabarrini, O.; et al. From quinoline to quinazoline-based S. aureus NorA efflux pump inhibitors by coupling a focused scaffold hopping approach and a pharmacophore search. ChemMedChem. 2021, 16, 3044–3059. [Google Scholar] [CrossRef]
  29. Darby, E.M.; Trampari, E.; Siasat, P.; Gaya, M.S.; Alav, I.; Webber, M.A.; Blair, J.M.A. Molecular mechanisms of antibiotic resistance revisited. Nat. Rev. Microbiol. 2023, 21, 280–295. [Google Scholar] [CrossRef]
  30. Antimicrobial Resistance Collaborators. Global burden of bacterial antimicrobial resistance in 2019: A systematic analysis. Lancet 2022, 399, 629–655. [Google Scholar] [CrossRef]
  31. Urban-Chmiel, R.; Marek, A.; Stępień-Pyśniak, D.; Wieczorek, K.; Dec, M.; Nowaczek, A.; Osek, J. Antibiotic resistance in bacteria—A review. Antibiotics 2022, 11, 1079. [Google Scholar] [CrossRef] [PubMed]
  32. Laws, M.; Shaaban, A.; Rahman, K.M. Antibiotic resistance breakers: Current approaches and future directions. FEMS Microbiol. Rev. 2019, 43, 490–516. [Google Scholar] [CrossRef] [PubMed]
  33. AlMatar, M.; Albarri, O.; Makky, E.A.; Köksal, F. Efflux pump inhibitors: New updates. Pharmacol. Rep. 2021, 73, 1–16. [Google Scholar] [CrossRef] [PubMed]
  34. Santajit, S.; Indrawattana, N. Mechanisms of antimicrobial resistance in ESKAPE pathogens. Biomed. Res. Int. 2016, 2016, 2475067. [Google Scholar] [CrossRef]
  35. Boucher, H.W.; Talbot, G.H.; Bradley, J.S.; Edwards, J.E.; Gilbert, D.; Rice, L.B.; Scheld, M.; Spellberg, B.; Bartlett, J. Bad bugs, no drugs: No ESKAPE! an update from the infectious diseases society of America. Clin. Infect. Dis. 2009, 48, 1–12. [Google Scholar] [CrossRef] [PubMed]
  36. Brown, D. Antibiotic resistance breakers: Can repurposed drugs fill the antibiotic discovery void? Nat. Rev. Drug Discov. 2015, 14, 821–832. [Google Scholar] [CrossRef]
  37. Sharma, A.; Gupta, V.K.; Pathania, R. Efflux pump inhibitors for bacterial pathogens: From bench to bedside. Indian J. Med. Res. 2019, 149, 129–145. [Google Scholar]
  38. Costa, S.S.; Falcão, C.; Viveiros, M.; Machado, D.; Martins, M.; Melo-Cristino, J.; Amaral, L.; Couto, I. Exploring the contribution of efflux on the resistance to fluoroquinolones in clinical isolates of Staphylococcus aureus. BMC Microbiol. 2011, 11, 241. [Google Scholar] [CrossRef]
  39. Bhattacharyya, S.; Bhattacharyya, M.; Pfannenstiel, D.M.; Nandi, A.K.; Hwang, Y.; Ho, K.; Harshey, R.M. Efflux-linked accelerated evolution of antibiotic resistance at a population edge. Mol. Cell 2022, 82, 4368–4385.e6. [Google Scholar] [CrossRef]
  40. Langevin, A.M.; El Meouche, I.; Dunlop, M.J. Mapping the role of AcrAB-TolC efflux pumps in the evolution of antibiotic resistance reveals near-MIC treatments facilitate resistance acquisition. mSphere 2020, 5, e01056-20. [Google Scholar] [CrossRef]
  41. Couto, I.; Costa, S.S.; Viveiros, M.; Martins, M.; Amaral, L. Efflux-mediated response of Staphylococcus aureus exposed to ethidium bromide. J. Antimicrob. Chemother. 2008, 62, 504–513. [Google Scholar] [CrossRef] [PubMed]
  42. The Clinical and Laboratory Standards Institute (CLSI). Methods for Dilution Antimicrobial Susceptibility Tests for Bacteria That Grow Aerobically, 11th ed.; CLSI Standard M07; CLSI: Wayne, PA, USA, 2018. [Google Scholar]
  43. Pillai, S.K.; Moellering, R.C.; Eliopoulos, G.M. Antimicrobial combinations. In Antibiotics in Laboratory Medicine, 5th ed.; Lorian, V., Ed.; Lippincott Williams & Wilkins: Philadelphia, PA, USA, 2005; pp. 365–440. [Google Scholar]
  44. Kaatz, G.W.; Seo, S.M.; O’Brien, L.; Wahiduzzaman, M.; Foster, T.J. Evidence for the existence of a multidrug efflux transporter distinct from NorA in Staphylococcus aureus. Antimicrob. Agents Chemother. 2000, 44, 1404–1406. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structures of the selected compounds previously reported as S. aureus NorA EPIs.
Figure 1. Chemical structures of the selected compounds previously reported as S. aureus NorA EPIs.
Antibiotics 12 00806 g001
Figure 2. (a) Assessment of ethidium bromide accumulation in Staphylococcus pseudintermedius ATCC 49444 and Staphylococcus pseudintermedius_P8 in the presence of CCCP. (b,c) Relative fluorescence of ethidium bromide bound to Staphylococcus pseudintermedius ATCC 49444 (b) or P8 (c) DNA remaining at 30 min in the presence/absence of the compounds 1, 3, 6, and 78 used at MIC/8 and MIC/64. * Statistical significance (p < 0.05) between no EPI treatment and treatments with compounds 1, 3, 6, and 78.
Figure 2. (a) Assessment of ethidium bromide accumulation in Staphylococcus pseudintermedius ATCC 49444 and Staphylococcus pseudintermedius_P8 in the presence of CCCP. (b,c) Relative fluorescence of ethidium bromide bound to Staphylococcus pseudintermedius ATCC 49444 (b) or P8 (c) DNA remaining at 30 min in the presence/absence of the compounds 1, 3, 6, and 78 used at MIC/8 and MIC/64. * Statistical significance (p < 0.05) between no EPI treatment and treatments with compounds 1, 3, 6, and 78.
Antibiotics 12 00806 g002
Table 1. MIC of ethidium bromide (EtBr), ciprofloxacin (CIP), gentamycin (GEN), chlorhexidine digluconate (CHX) [4], and selected EPIs for S. pseudintermedius ATCC 49444 and its laboratory derivative S. pseudintermedius_P8.
Table 1. MIC of ethidium bromide (EtBr), ciprofloxacin (CIP), gentamycin (GEN), chlorhexidine digluconate (CHX) [4], and selected EPIs for S. pseudintermedius ATCC 49444 and its laboratory derivative S. pseudintermedius_P8.
MIC (mg/L)
CompoundSP ATCC 49444SP_P8
1>100>100
2>100>100
35050
42525
56.256.25
6>100>100
7>100>100
8>100>100
9>100>100
10>100>100
EtBr0.532
CIP≤0.1258
GEN0.251
CHX14
Table 2. MICs of gentamycin (GEN), chlorhexidine digluconate (CHX), ciprofloxacin (CIP), and ethidium bromide (EtBr) in the presence of compounds 110 used at scalar concentrations (ranging from 1/4 to 1/128 MIC) against norA-overexpressing S. pseudintermedius_P8.
Table 2. MICs of gentamycin (GEN), chlorhexidine digluconate (CHX), ciprofloxacin (CIP), and ethidium bromide (EtBr) in the presence of compounds 110 used at scalar concentrations (ranging from 1/4 to 1/128 MIC) against norA-overexpressing S. pseudintermedius_P8.
MIC in mg/L (MF) *
EPI (mg/L)GENCHXCIPEtBrEPIGENCHXCIPEtBr
No EPI14832No EPI14832
1 1/4 (50)0.125 (8)0.5 (8)0.125 (64)0.5 (64)2 1/4 (50)0.125 (8)4 (-)2 (4)16 (2)
1 1/8 (25)0.125 (8)0.5 (8)0.125 (64)0.5 (64)2 1/8 (25)0.25 (4)4 (-)2 (4)16 (2)
1 1/16 (12.5)0.25 (4)1 (4)0.25 (32)0.5 (64)2 1/16 (12.5)0.25 (4)4 (-)2 (4)16 (2)
1 1/32 (6.25)0.25 (4)2 (2)1 (8)2 (16)2 1/32 (6.25)0.25 (4)4 (-)2 (4)16 (2)
1 1/64 (3.13)0.5 (2)2 (2)1 (8)4 (8)2 1/64 (3.13)0.5 (2)4 (-)2 (4)16 (2)
1 1/128 (1.57)0.5 (2)2 (2)1 (8)8 (4)2 1/128 (1.57)0.5 (2)4 (-)2 (4)32 (-)
3 1/4 (12.5)0.25 (4)0.5 (8)0.25 (32)0.5 (64)4 1/4 (6.25)0.25 (4)2 (2)0.25 (32)0.5 (64)
3 1/8 (6.25)0.25 (4)1 (4)0.5 (16)1 (32)4 1/8 (3.13)0.25 (4)2 (2)0.5 (16)2 (16)
3 1/16 (3.13)0.5 (2)2 (2)1 (8)4 (8)4 1/16 (1.57)0.5 (2)2 (2)2 (4)8 (4)
3 1/32 (1.57)0.5 (2)2 (2)2 (4)8 (4)4 1/32 (0.78)0.5 (2)2 (2)4 (2)16 (2)
3 1/64 (0.78)0.5 (2)4 (-)2 (4)32 (-)4 1/64 (0.39)0.5 (2)2 (2)4 (2)16 (2)
3 1/128 (0.39)0.5 (2)4 (-)4 (2)32 (-)4 1/128 (0.20)1 (-)4 (-)8 (-)32 (-)
5 1/4 (1.57)0.25 (4)1 (4)0.5 (16)4 (8)6 1/4 (50)0.125 (8)1 (4)0.25 (32)0.5 (64)
5 1/8 (0.78)0.25 (4)2 (2)2 (4)8 (4)6 1/8 (25)0.125 (8)1 (4)0.25 (32)0.5 (64)
5 1/16 (0.39)0.5 (2)2 (2)2 (4)16 (2)6 1/16 (12.5)0.25 (4)2 (2)0.5 (16)1 (32)
5 1/32 (0.20)0.5 (2)2 (2)2 (4)16 (2)6 1/32 (6.25)0.5 (2)2 (2)1 (8)8 (4)
5 1/64 (0.10)0.5 (2)2 (2)2 (4)16 (2)6 1/64 (3.13)0.5 (2)2 (2)1 (8)16 (2)
5 1/128 (0.05)1 (-)4 (-)4 (2)32 (-)6 1/128 (1.57)1 (-)4 (-)2 (4)32 (-)
7 1/4 (50)0.25 (4)1 (4)0.25 (32)0.5 (64)8 1/4 (50)0.25 (4)0.5 (8)0.25 (32)0.5 (64)
7 1/8 (25)0.25 (4)2 (2)0.25 (32)2 (16)8 1/8 (25)0.25 (4)0.5 (8)0.25 (32)0.5 (64)
7 1/16 (12.5)0.25 (4)2 (2)0.5 (16)8 (4)8 1/16 (12.5)0.25 (4)2 (2)0.25 (32)2 (16)
7 1/32 (6.25)0.5 (2)2 (2)1 (8)16 (2)8 1/32 (6.25)0.25 (4)2 (2)0.5 (16)8 (4)
7 1/64 (3.13)0.5 (2)2 (2)1 (8)16 (2)8 1/64 (3.13)0.5 (2)2 (2)1 (8)16 (2)
7 1/128 (1.57)1 (-)4 (-)2 (4)32 (-)8 1/128 (1.57)0.5 (2)4 (-)2 (4)32 (-)
9 1/4 (50)0.25 (4)4 (-)0.5 (16)2 (16)10 1/4 (50)0.25 (4)1 (4)0.5 (16)8 (4)
9 1/8 (25)0.25 (4)4 (-)1 (8)2 (16)10 1/8 (25)0.25 (4)1 (4)0.5 (16)8 (4)
9 1/16 (12.5)0.25 (4)4 (-)1 (8)2 (16)10 1/16 (12.5)0.25 (4)2 (2)0.5 (16)8 (4)
9 1/32 (6.25)0.25 (4)4 (-)1 (8)4 (8)10 1/32 (6.25)0.5 (2)2 (2)1 (8)16 (2)
9 1/64 (3.13)0.5 (2)4 (-)2 (4)8 (4)10 1/64 (3.13)0.5 (2)2 (2)2 (4)16 (2)
9 1/128 (1.57)0.5 (2)4 (-)2 (4)16 (2)10 1/128 (1.57)1 (-)4 (-)2 (4)32 (-)
* MF: Modulation Factor that represents the n-fold reduction of the MIC of the corresponding antimicrobial when combined with the EPI. Hyphen (-) indicates no reduction of the MIC. Bold numbers indicate Modulation Factors (MF) ≥ 4.
Table 3. The lowest Fractional Inhibitory Concentration Index (FICI) obtained by combining compounds 110 with ethidium bromide (EtBr), ciprofloxacin (CIP), gentamycin (GEN), and chlorhexidine digluconate (CHX) for S. pseudintermedius ATCC 49444 and S. pseudintermedius_P8. Bold numbers indicate synergistic combinations.
Table 3. The lowest Fractional Inhibitory Concentration Index (FICI) obtained by combining compounds 110 with ethidium bromide (EtBr), ciprofloxacin (CIP), gentamycin (GEN), and chlorhexidine digluconate (CHX) for S. pseudintermedius ATCC 49444 and S. pseudintermedius_P8. Bold numbers indicate synergistic combinations.
MIC (mg/L)
GENCHXCIPEtBr
EPISP 49444SP P8SP 49444SP P8SP 49444SP P8SP 49444SP P8
10.500.250.560.250.500.130.160.08
22.000.282.001.002.000.262.000.51
32.000.380.750.382.000.160.500.16
40.560.382.000.532.000.190.250.19
50.560.382.000.502.000.282.000.38
60.500.310.560.382.000.130.190.09
70.500.310.560.502.000.130.500.19
82.000.280.500.252.000.090.250.13
92.000.280.561.002.000.160.560.09
100.380.310.560.382.000.130.250.31
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Rampacci, E.; Felicetti, T.; Cernicchi, G.; Stefanetti, V.; Sabatini, S.; Passamonti, F. Inhibition of Staphylococcus pseudintermedius Efflux Pumps by Using Staphylococcus aureus NorA Efflux Pump Inhibitors. Antibiotics 2023, 12, 806. https://doi.org/10.3390/antibiotics12050806

AMA Style

Rampacci E, Felicetti T, Cernicchi G, Stefanetti V, Sabatini S, Passamonti F. Inhibition of Staphylococcus pseudintermedius Efflux Pumps by Using Staphylococcus aureus NorA Efflux Pump Inhibitors. Antibiotics. 2023; 12(5):806. https://doi.org/10.3390/antibiotics12050806

Chicago/Turabian Style

Rampacci, Elisa, Tommaso Felicetti, Giada Cernicchi, Valentina Stefanetti, Stefano Sabatini, and Fabrizio Passamonti. 2023. "Inhibition of Staphylococcus pseudintermedius Efflux Pumps by Using Staphylococcus aureus NorA Efflux Pump Inhibitors" Antibiotics 12, no. 5: 806. https://doi.org/10.3390/antibiotics12050806

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop