Next Article in Journal
Synthesis, Computational Study, and In Vitro α-Glucosidase Inhibitory Action of 1,3,4-Thiadiazole Derivatives of 3-Aminopyridin-2(1H)-ones
Previous Article in Journal
Small-Angle X-ray Scattering (SAXS) Used for the Identification of Nicomorphine Polymorphic Changes at the Early Stage to Avoid Varied Stability and Possible Side Effects
Previous Article in Special Issue
In Silico and In Vitro Evaluation of the Mechanism of Action of Three VX809-Based Hybrid Derivatives as Correctors of the F508del CFTR Protein
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Pyrazolyl Thioureas Active against the Staphylococcus Genus

1
Department of Surgical Sciences and Integrated Diagnostics (DISC), University of Genoa, Viale Benedetto XV, 6, 16132 Genoa, Italy
2
Department of Pharmacy (DIFAR), Section of Medicinal Chemistry, University of Genoa, Viale Benedetto XV, 3, 16132 Genoa, Italy
3
Department of Experimental Medicine (DIMES), University of Genoa, Via L.B. Alberti, 2, 16132 Genoa, Italy
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2024, 17(3), 376; https://doi.org/10.3390/ph17030376
Submission received: 27 November 2023 / Revised: 23 February 2024 / Accepted: 13 March 2024 / Published: 15 March 2024

Abstract

:
To meet the urgent need for new antibacterial molecules, a small library of pyrazolyl thioureas (PTUs) was designed, synthesized and tested against difficult-to-treat human pathogens. The prepared derivatives are characterized by a carboxyethyl functionality on C4 and different hydroxyalkyl chains on N1. Compounds 1ao were first evaluated against a large panel of Gram-positive and Gram-negative pathogens. In particular, the majority of PTUs proved to be active against different species of the Staphylococcus genus, with MIC values ranging from 32 to 128 µg/mL on methicillin-resistant Staphylococcus strains, often responsible for severe pulmonary disease in cystic fibrosis patients. Time-killing experiments were also performed for the most active compounds, evidencing a bacteriostatic mechanism of action. For most active derivatives, cytotoxicity was evaluated in Vero cells, and at the tested concentrations and at the experimental exposure time of 24 h, none of the compounds analysed showed significant toxicity. In addition, favourable drug-like, pharmacokinetic and toxicity properties were predicted for all new synthesized derivatives. Overall, the collected data confirmed the PTU scaffold as a promising chemotype for the development of novel antibacterial agents active against Gram-positive multi-resistant strains frequently isolated from cystic fibrosis patients.

1. Introduction

The World Health Organization (WHO) estimates that there are 700,000 casualties per year worldwide due to drug-resistant infections with a projection of 10 million deaths by 2050 and a general cost for the global economy up to USD 100 trillion. Unfortunately, the WHO also noted that in 2020, none of the 43 antibiotics in clinical use had fully solved the problem of drug resistance [1]. In this regard, the bacteria included in the ESKAPE group (Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumonia, Acinetobacter baumannii, Pseudomonas aeruginosa, Enterobacter species) have been recognized by the Infectious Diseases Society of America (IDSA) as the most dangerous pathogens, due to their remarkable resistance to the most common conventional antibiotics. Additional concerns derived from the isolation of strains resistant to vancomycin and linezolid, considered the last line of defence against Gram-positive bacterial infections [2,3,4]. The clinical management of infections caused by these exceptional pathogens is often complex and problematic, especially in patients hospitalized or suffering from concomitant diseases.
In this regard, cystic fibrosis (CF) is a genetic disorder associated with the production of sticky and thick mucus that accumulates in different organs, including the lungs. This condition facilitates the adhesion and growth of pathogens in the respiratory epithelium and exposes CF patients to respiratory infections, mainly caused by methicillin-resistant Staphylococcus aureus (MRSA). Noteworthy, chronic lung infections with this Gram-positive multi-resistant pathogen have been associated with more severe pulmonary disease and increase the decline in lung function in CF patients [5,6]. Clinical studies on CF patients have shown that the incidence of methicillin-sensitive S. aureus (MSSA) is significantly lower than that of MRSA [7,8,9].
The pyrazole scaffold represents a privileged substructure in medicinal chemistry research and a number of pyrazole derivatives have been evaluated as effective compounds in different therapeutic areas [10,11,12,13,14,15,16], including infectious diseases [17,18] and also against MRSA [19]. In previous studies, pyrazole derivatives I (Figure 1) showed significative antibacterial activity against Gram-positive antibiotic-resistant strains [20]. The prepared molecules were characterized by a N1-hydroxy-2-phenylethyl chain, a carboxyethyl or tert-butyl substituents on C3 or C4 and a (thio)ureido moiety at position 5 of the pyrazole ring. The structure activity relationships (SARs) evidenced that a carboxyethyl group on C4 and a substituted thioureido function on C5 are key structural determinants for the antimicrobial activity of this series (derivatives II, Figure 1). Pyrazolyl thioureas (PTUs) II resulted inactive against all tested Gram-negative species but showed a good antibacterial potency against strains of the Staphylococcus genus resistant to methicillin and linezolid (MIC values between 32 and 64 µg/mL) and also against vancomycin-resistant Enterococcus strains (MIC values between 32 and 64 µg/mL).
The subject of this work is the design and synthesis of a novel small library of PTU molecules (compounds 1ao) to confirm if this scaffold is a pharmaceutically relevant chemotype for the development of novel antibacterial agents, particularly active against Gram-positive species.

2. Results and Discussion

The novel pyrazole small library of PTUs (compounds 1ao, Table 1) are characterized by: (1) a carboxyethyl function on C4 as previous derivatives II, (2) an unsubstituted or fluoro-substituted benzoyl thiourea segment on C5 and (3) various N1 hydroxyalkyl chains with different length, replacing the 2-hydroxy-2-phenylethyl chain of I and II. In detail, hydroxypropyl (compounds 1ad), hydroxybutyl (compounds 1eh), hydroxypentyl (compounds 1il) and finally hydroxyhexyl (compounds 1mo) chains have been inserted on the N1 position of the pyrazole scaffold.
Novel PTU library 1 was evaluated for their antibacterial activity against several Gram-positive and Gram-negative species, using oxacillin as the reference compound. Time-killing experiments on MRSA strains and cytotoxicity evaluation on Vero cells were also performed for the most active compounds. Additionally, in silico prediction of pharmacokinetic properties, drug-likeness and toxicity (ADMET) of all novel pyrazole compounds was performed.

2.1. Chemistry

Novel pyrazole library was obtained following a consolidate procedure [20], following Scheme 1. Briefly, the condensation of the commercially available oxiranes 2ad with hydrazine monohydrate led to the corresponding hydrazino-ethanols 3ad, that were reacted with ethyl ethoxymethylene cyanoacetate in anhydrous toluene (3a) or absolute ethanol (3bd) at 70–80 °C to give the suitable pyrazole intermediates 4ad as yellow solids [21]. Finally, the thiourea moiety was introduced via a one-pot reaction in anhydrous THF for 12 h between the 5-amino-pyrazoles 4ad and the proper benzoyl isothiocyanate 5ad, commercially available or prepared according to the literature method [22]. Derivatives 1 were obtained in yields ranging from 35% to 94% (Table 1) as yellow oils or crystalline white solids.

2.2. Antibacterial Activity

The antibacterial potency of PTU 1ao was evaluated against a panel of fifteen bacterial isolates (Table 2), representative of clinically relevant Gram-positive (eleven strains) and Gram-negative (four strains) species including four S. aureus strains (MRSA), three S. epidermidis isolates (two MRSE and one resistant to methicillin and linezolid), two E. faecalis strains (one vancomycin-sensitive and one vancomycin resistant, VRE), two E. faecium isolates (one vancomycin-sensitive and one VRE), two E. coli isolates resistant to carbapenem (one was a New Delhi metallo-β-lactamase (NDM)-producing isolate) and two P. aeruginosa (multidrug-resistant isolates, MDR) strains.
All tested compounds resulted inactive (MIC > 128 µg/mL) against the Gram-negative and Enterococcus genus. Conversely, the majority of derivatives (10 out of 15) showed a widespread activity against the most clinically relevant Staphylococcus species (i.e., S. aureus MRSA and S. epidermidis MRSE), with MIC values lower in some cases than those of oxacillin, used as a reference compound (32–128 mg/mL against 128–512 mg/mL). Interestingly, PTUs 1i and 1m were identified as the most active derivatives of the series because they showed MIC values in a close range (32–64 µg/mL) against seven of the considered Gram-positive isolates. Also compounds 1d, 1e, 1k and 1n resulted active against six Gram-positive strains.
To further define the antibacterial properties of the series, PTUs 1a,d,e,h,i,m were selected as representative examples of differently N1 substituted pyrazoles (1a and 1d, R = Me; 1e and 1h, R = Et; 1i, R = nPr; 1m, R = nBu) and tested against additional Staphylococcus species for a total of 14 isolates (Table 3). All analysed compounds proved to be ineffective against S. saprophyticus, S. warneri and S. simulans, but showed relevant antibacterial activity against S. lugdunensis and S. auricularis species (MIC value range = 16–32 µg/mL). Moreover, derivative 1d specifically inhibited the proliferation of methicillin-resistant S. capitis 71 strain (MIC = 64 µg/mL), whereas compound 1h affected the growth of S. hominis 124, without influencing the S. hominis 125. Finally, derivatives 1d,e,h,i showed similar activity against S. haemoliticus 115 isolate, resulting ineffective against the other two considered S. haemoliticus strains. Regarding S. auricularis, compounds 1e,i,m resulted the most potent, displaying MIC values of 16 µg/mL.
To investigate whether PTUs act as bacteriostatic or bactericidal, time-killing experiments were carried out on MRSA isolates, very relevant in the daily clinical practice of CF patients. Compounds 1c,d,n,h were selected as representative examples of the chemical diversity of the series and tested against four different MRSA strains (i.e., S. aureus 17, 18, 187 and 195). The experiments were carried out at concentrations four times the MIC values. As exemplified in Figure 2, the four tested compounds proved to act as a bacteriostatic agent, because they were all able to maintain virtually unchanged (105 CFU/mL) the concentration of the initial bacterial inocula for all 24 h of this study. Similar trends were obtained for all the analysed MRSA strains.

2.3. Cytotoxicity Evaluation

To verify if PTUs here reported are characterized by a cytotoxicity activity, selected compounds 1c (R = methyl) and 1n (R = butyl), chosen among the most active ones and as representative examples of the chemical diversity of the series, were tested on Vero cells at the most representative MIC values obtained (32 µg/mL and 64 µg/mL for both compounds, Figure 3).

2.4. Pharmacokinetic Properties, Drug-Likeness and Toxicity (ADMET) Prediction

To further characterize the pharmaceutical potentials of PTUs 1, the drug-likeness and pharmacokinetic properties of the series were calculated using the SwissADME program [23]. Derivatives IIa,b (Figure 1) were used as reference molecules (Table 4). Collectively, this in silico evaluation predicted for PTUs 1 favourable physiochemical and DMPK properties that, in some case, would result better than those of previously described compounds II.
In detail, the replacement of the II phenyl group with aliphatic, linear chains (namely, methyl, ethyl, propyl or butyl R substituents, Table 1) led to an increase in the Csp3 fraction with an improvement of the predicted bioavailability. Respect to previous II, in most cases the number of rotatable bonds of H bond acceptors and H bond donors are the same of previous II. Moreover, 1 and II would display the similar polarity, as indicated by the topological polar surface area (TPSA) descriptor (137.57 Å2). This descriptor has proved to be indispensable for predicting the permeability of a molecule towards biological membranes. It has in fact been demonstrated that when the TPSA value is greater than 140 Å2, the molecules have difficulty permeating the barriers; on the contrary, when it is less than 140 Å2, the passage through the lipophilic barriers is easier.
The different N1 substituents of the pyrazole ring would also affect the lipophilicity of the compounds, being the logPvalues within the desired range (logP between −0.7 and +5.0) for all analysed PTUs. Except for 1mo (N1 hydroxyhexyl derivatives), all compounds were predicted to be water soluble rather than moderately soluble as compounds II.
Compounds 1ai,1m would be highly absorbed in the gastrointestinal tract, whereas derivatives 1jl and 1no, characterized by a fluoro-substituted phenyl ring on thioureido function and a more embedded chain on N1 (hydroxypentyl and hydroxyexyl) would be poorly absorbed. As derivatives II, the novel compounds would not be able to pass the blood–brain barrier (BBB) and enter in the central nervous system. Furthermore, unlike the previous II, derivatives 1 would be substrates of the P-gp efflux pump.
The CYP inhibition properties of compounds 1 would be different from that predicted for derivatives II. Thus, 1A2 and 2D6 isoforms would not be affected by PTUs 1, while 2C19, 2C9 and 3A4 enzymes would be inhibited by derivatives 1eo.
As derivatives II, no violations of the Lipinski rules have been identified for PTUs 1 that would not display any pan-assay interference compounds (PAINS) alerts. According to the Brenk filters, the thiourea thiocarbonyl group on the C5 position was spotted as a problematic fragment [24].
Table 4. Predicted pharmacokinetics and drug-like properties of compounds 1ao in comparison with previously synthesized derivatives IIa,b.
Table 4. Predicted pharmacokinetics and drug-like properties of compounds 1ao in comparison with previously synthesized derivatives IIa,b.
1a1b–d1e1f–h1i1j–l1m1n–oIIa,b
Physicochemical Property
MW (g/mol)376.43394.42390.46408.45404.48422.47418.51436.50456.49
Fraction Csp30.290.290.330.330.370.370.400.400.18
Rotatable bonds101011111212131311
H-bond acceptors565656566
H-bond donors333333333
TPSA a2)137.57137.57137.57137.57137.57137.57137.57137.57137.57
Lipophilicity
LogP b2.062.162.592.692.953.053.493.593.26
Water solubility
Solubility (mg/mL) c0.2820.2030.1330.0950.0800.0570.0370.0260.018
Solubility class dSSSSSSMSMSMS
Pharmacokinetics
GI absorptionhighhighhighhighhighlowhighlowlow
BBB permeantnonononononononono
P-gp substrateyesyesyesyesyesyesyesyesno
CYP1A2 inhibitornonononononononono
CYP2C19 inhibitornonoyesyesyesyesyesyesyes
CYP2C9 inhibitornonoyesyesyesyesyesyesyes
CYP2D6 inhibitornonononononononoyes
CYP3A4 inhibitornonoyesyesyesyesyesyesyes
Drug-likeness
Lipinski violations000000000
Medicinal chemistry
PAINS alerts000000000
Brenk alerts111111111
a Topological polar surface area. b Predicted according to the XLOGP3 program. c Values predicted by the ESOL method [25]. d S = soluble; MS = moderately soluble.
Collectively, this in silico evaluation predicted for this new PTU library good physicochemical, lipophilicity and water solubility properties; in some cases, better than previous II.
In addition, also predicted acute toxicity (lethal dose, LD50) for rats after oral administration (Table 5) was calculated using the Advanced Chemistry Development (ACD) Percepta platform (ACD/Percepta Platform. Advanced Chemistry Development, Inc.; Toronto, ON, Canada: 2015).
Reliability index values are shown as R.I. (values higher than 0.30 are ranked as reliable by the software). The software prediction is performed based on the software implemented training libraries, which include experimentally determined pharmacokinetic and safety properties for different series of compounds.
Notably, the newly developed compounds 1ao were predicted to have LD50 values in the 1600–3300 mg/kg range.

3. Materials and Methods

3.1. Chemistry

3.1.1. General Information

All solvents and reagents were purchased from Chiminord s.r.l. (Milan, Italy) and Merk (Aldrich Chemical, Milan, Italy). Solvents were reagent grade. All commercial reagents were used without further purification. For thin-layer chromatography (TLC), aluminium-backed silica gel plates (Merck DC-Alufolien Kieselgel 60 F254, Darmstad, Germany) were used. For chromatography, Merck silica gel, 230–400 mesh, was used. Flash chromatography was performed using the Isolera One instrument (Biotage, Uppsala, Sweden) using a silica gel column. Melting points were not “corrected” and were obtained with a Buchi M-560 instrument (Buchi instruments, Flawil, Switzerland). NMR spectra were recorded on a JEOL JNM ECZ-400S/L1 (400 MHz, Tokyo, Japan).
Elemental analysis was determined with an elemental analyser EA 1110 (Fison-Instruments, Milan, Italy); compounds have been considered pure when the difference between calculated and found values is ± 0.4 (Table S1 in Supporting Material). Hydrazines 3 and pyrazole intermediates 4ac were prepared according to the already published procedures [21,26]. Benzoyl isothiocyanates 5bd were prepared according to the literature method [20,22].

3.1.2. Synthesis of Ethyl 5-amino-1-(2-hydroxyhexyl)-1H-pyrazole-4-carboxylate 4d

A mixture of ethyl ethoxymethylene acetate (3.38 g, 20 mmol) and 2-hydrazinohexane-2-ol 3d (2.64 g, 20 mmol) [26] in absolute ethanol (40 mL) was heated at 70–80 °C for 8 h. After cooling at room temperature, the mixture was evaporated under vacuum; 6N HCl (30 mL) was added and the acid solution was washed with diethyl ether (20 mL), then alkalized with 4M NaOH to obtain a yellow solid that was filtered and recrystallized from a mixture of diethyl ether/ligroin (b.p. 40–60 °C) (1:1).
Mp: 103–105 °C. Yield: 57%. 1H-NMR (400 MHz, DMSO-d6): δ 0.80 (t, J = 7.0, 3H, CH3), 1.07–1.34 (m, 9H, CH3 + 3CH2), 3.73–3.82 (m, 3H, CHOH + NCH2), 4.11 (q, J = 7.0, 2H, CH2O), 4.91 (d, 1H, OH, exchangeable with D2O), 6.05 (br s, 2H, NH2 exchangeable with D2O), 7.41 (s, 1H, H-3). 13C-NMR (101 MHz, DMSO-d6): δ 160.71, 149.11, 143.44, 98.61, 69.38, 55.67, 53.06, 31.80, 27.42, 23.44, 16.27. Anal. (C12H21N3O3) calcd for C, H, N.

3.1.3. General Synthesis of 5-Thioureido Pyrazoles 1ao

The proper 5-amino-pyrazoles 4ad (1 mmol) and the suitable benzoyl isothiocyanate 5ad (1 mmol), commercially available or previously prepared modifying the literature method [22], in anhydrous THF (10 mL) was refluxed for 12 h. After cooling to room temperature, the solution was concentrated under reduced pressure and the crude was dissolved in ethyl acetate (20 mL); the organic phase was washed with 6N HCl (10 mL), then with NaHCO3 saturated solution (10 mL) and water (10 mL), dried (MgSO4) and concentrated under reduced pressure. The crude was purified by flash chromatography using a mixture of diethyl ether/ligroin (b.p. 40–60 °C) 3/1 as the eluent.
Ethyl 5-(3-benzoylthioureido)-1-(2-hydroxypropyl)-1H-pyrazole-4-carboxylate 1a. White solid. M.p.: 106–108 °C. Yield: 64%. 1H-NMR (400 MHz, DMSO-d6): δ 1.07 (d, J = 6.0, 3H, CH3), 1.18 (t, J = 7.0, 3H, CH3), 3.91–4.05 (m, 3H, CHOH + CH2N), 4.13 (q, J = 7.0, 2H, CH2O), 5.05 (d, J = 4.8, 1H, OH exchangeable with D2O), 7.37–8.03 (m, 6H, 5 Ar + H-3), 12.10 (br s, 1H, NH, exchangeable with D2O), 12.20 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 180.47, 165.88, 161.44, 140.22, 139.93, 134.94, 132.03, 128.50, 128.09, 103.44, 66.87, 60.12, 57.55, 20.94, 14.28. Anal. (C17H20N4O4S) calcd for C, H, N, S.
Ethyl 5-(3-(2-fluorobenzoyl)thioureido)-1-(2-hydroxypropyl)-1H-pyrazole-4-carboxylate 1b. Yellow oil. Yield: 45%. 1H-NMR (400 MHz, DMSO-d6): δ 1.05 (d, J = 6.0, 3H, CH3), 1.18 (t, J = 7.0, 3H, CH3), 3.90–4.09 (m, 3H, CHOH + CH2N), 4.12 (q, J = 7.0, 2H, CH2O), 5.05 (br s, 1H, OH exchangeable with D2O), 7.29–7.87 (m, 4H, Ar), 7.89 (s, 1H, H-3), 11.89 (br s, 1H, NH, exchangeable with D2O), 12.16 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 176.97, 164.60, 161.87, 158.29, 140.22, 131.76, 129.99, 122.75, 122.22, 115.82, 104.44, 67.63, 59.28, 55.77, 23.81, 16.39. Anal. (C17H19N4O4SF) calcd for C, H, N, S.
Ethyl 5-(3-(3-fluorobenzoyl)thioureido)-1-(2-hydroxypropyl)-1H-pyrazole-4-carboxylate 1c. Yellow oil. Yield: 35%. 1H-NMR (400 MHz, DMSO-d6): δ 1.07 (d, J = 6.0, 3H, CH3), 1.22 (t, J = 7.0, 3H, CH3), 3.90–4.04 (m, 3H, CHOH + CH2N), 4.13 (q, J = 7.0, 2H, CH2O), 5.05 (d, J = 4.8, 1H, OH exchangeable with D2O), 7.52–7.87 (m, 4H, Ar), 7.92 (s, 1H, H-3), 12.04 (br s, 1H, NH, exchangeable with D2O), 12.16 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 178.28, 165.68, 162.48, 160.01, 158.56, 140.22, 140.09, 139.93, 130.66, 122.97, 115.03, 114.47, 104.20, 68.20, 60.08, 57.55, 21.16, 13.60. Anal. (C17H19N4O4SF) calcd for C, H, N, S.
Ethyl 5-(3-(4-fluorobenzoyl)thioureido)-1-(2-hydroxypropyl)-1H-pyrazole-4-carboxylate 1d. White solid. M.p.: 131–133 °C. Yield: 39%. 1H-NMR (400 MHz, DMSO-d6): δ 1.16 (t, J = 6.0, 3H, CH3), 1.42 (t, J = 7.0, 3H, CH3), 4.09–4.30 (m, 3H, CHOH + CH2N), 4.32 (q, J = 7.0, 2H, CH2O), 5.05 (br s, 1H, OH exchangeable with D2O), 7.34–7.71 (m, 4H, Ar), 7.84 (s, 1H, H-3), 11.86 (br s, 1H, NH, exchangeable with D2O), 11.98 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 175.35, 166.01, 163.54, 161.44, 139.81, 130.17, 129.67, 115.99, 101.92, 68.20, 60.12, 57.55, 16.27, 9.97. Anal. (C17H19N4O4SF) calcd for C, H, N, S.
Ethyl 5-(3-benzoylthioureido)-1-(2-hydroxybutyl)-1H-pyrazole-4-carboxylate 1e. Yellow oil. Yield: 55%. 1H-NMR (400 MHz, DMSO-d6): δ 0.88 (t, J = 6.0, 3H, CH3), 1.18 (t, J = 7.0, 3H, CH3), 1.22–1.47 (m, 2H, CH2), 3.72–4.04 (m, 3H, CHOH + CH2N), 4.14 (q, J = 7.0, 2H, CH2O), 5.02 (d, J = 4.8, 1H, OH exchangeable with D2O), 7.45–8.03 (m, 6H, 5Ar + H-3), 12.10 (br s, 1H, NH, exchangeable with D2O), 12.20 (br s, 1H, NH, exchangeable with D2O). 13C -NMR (101 MHz, DMSO-d6): δ 174.97, 167.07, 158.96, 141.86, 140.22, 136.90, 132.41, 128.50, 128.09, 101.68, 74.10, 60.95, 53.62, 30.22, 16.59, 7.13. Anal. (C18H22N4O4S) calcd for C, H, N, S.
Ethyl 5-(3-(2-fluorobenzoyl)thioureido)-1-(2-hydroxybutyl)-1H-pyrazole-4-carboxylate 1f. Yellow oil. Yield: 41%. 1H-NMR (400 MHz, DMSO-d6): δ 0.88 (t, J = 6.0, 3H, CH3), 1.03–1.46 (m, 5H, CH3 + CH2), 3.74–4.04 (m, 3H, CHOH + CH2N), 4.14 (q, J = 7.0, 2H, CH2O), 5.05 (d, J = 4.8, 1H, OH exchangeable with D2O), 7.26–7.90 (m, 4H, Ar), 7.92 (s, 1H, H-3), 11.92 (br s, 1H, NH, exchangeable with D2O), 12.15 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 176.97, 164.60, 161.47, 158.99, 140.22, 132.22, 129.99, 125.06, 123.62, 116.34, 103.44, 71.50, 60.12, 54.73, 27.50, 14.28, 9.79. Anal. (C18H21N4O4SF) calcd for C, H, N, S.
Ethyl 5-(3-(3-fluorobenzoyl)thioureido)-1-(2-hydroxybutyl)-1H-pyrazole-4-carboxylate 1g. Yellow oil. Yield: 40%. 1H-NMR (400 MHz, DMSO-d6): δ 0.88 (t, J = 6.0, 3H, CH3), 1.16–1.43 (m, 5H, CH3 + CH2), 3.74–4.04 (m, 3H, CHOH + CH2N), 4.14 (q, J = 7.0, 2H, CH2O), 5.02 (d, J = 4.8, 1H, OH exchangeable with D2O), 7.38–7.89 (m, 4H, Ar), 7.92 (s, 1H, H-3), 12.07 (br s, 1H, NH, exchangeable with D2O), 12.16 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 177.08, 166.55, 162.79, 161.44, 160.33, 140.22, 137.37, 130.66, 123.42, 118.99, 115.19, 105.38, 74.10, 63.07, 54.64, 30.54, 15.56, 11.31. Anal. (C18H21N4O4SF) calcd for C, H, N, S.
Ethyl 5-(3-(4-fluorobenzoyl)thioureido)-1-(2-hydroxybutyl)-1H-pyrazole-4-carboxylate 1h. White solid. M.p.: 120–122 °C. Yield: 50%. 1H-NMR (400 MHz, DMSO-d6): δ 1.21 (t, J = 6.0, 3H, CH3), 1.60 (t, J = 7.0, 3H, CH3), 1.63–1.79 (m, 2H, CH2CH3), 4.13–4.28 (m, 3H, CHOH + CH2N), 4.52 (q, J = 7.0, 2H, CH2O), 5.05 (d, J = 4.8, 1H, OH exchangeable with D2O), 7.33–7.38 and 7.65–7.72 (2m, 4H, Ar), 7.88 (s, 1H, H-3), 11.72 (br s, 1H, NH, exchangeable with D2O), 11.82 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 177.08, 166.01, 165.26, 163.54, 159.75, 139.50, 130.17, 118.10, 101.44, 71.50, 54.17, 47.55, 24.23, 10.29, 5.80. Anal. (C18H21N4O4SF) calcd for C, H, N, S.
Ethyl 5-(3-benzoylthioureido)-1-(2-hydroxypentyl)-1H-pyrazole-4-carboxylate 1i. White solid. M.p.: 126–128 C. Yield: 55%. 1H-NMR (400 MHz, DMSO-d6): δ 0.83 (t, J = 6.0, 3H, CH3), 1.18 (t, J = 7.0, 3H, CH3), 1.25–1.43 (m, 4H, 2CH2), 3.81–4.03 (m, 3H, CHOH + CH2N), 4.14 (q, J = 7.0, 2H, CH2O), 5.00 (d, J = 4.8, 1H, OH exchangeable with D2O), 7.55–8.02 (m, 6H, 5Ar + H-3), 12.10 (br s, 1H, NH, exchangeable with D2O), 12.19 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 177.06, 165.63, 161.55, 140.22, 140.08, 135.30, 132.14, 128.66, 128.50, 128.09, 128.01, 103.44, 70.17, 60.12, 55.34, 37.09, 18.44, 14.28, 14.12. Anal. (C19H24N4O4S) calcd for C, H, N, S.
Ethyl 5-(3-(2-fluorobenzoyl)thioureido)-1-(2-hydroxypentyl)-1H-pyrazole-4-carboxylate 1j. Yellow oil. Yield: 36%. 1H-NMR (400 MHz, DMSO-d6): δ 0.83 (t, J = 6.0, 3H, CH3), 1.15–1.42 (m, 7H, CH3 +2CH2), 3.82–4.05 (m, 3H, CHOH + CH2N), 4.15 (q, J = 7.0, 2H, CH2O), 5.02 (d, J = 4.8, 1H, OH exchangeable with D2O), 7.34–7.90 (m, 4H, Ar), 7.92 (s, 1H, H-3), 11.92 (br s, 1H, NH, exchangeable with D2O), 12.10 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 175.74, 164.60, 161.47, 158.99, 140.22, 132.22, 132.14, 125.06, 123.62, 116.34, 106.17, 73.08, 60.12, 55.34, 37.09, 20.30, 11.86. Anal. (C19H23N4O4SF) calcd for C, H, N, S.
Ethyl 5-(3-(3-fluorobenzoyl)thioureido)-1-(2-hydroxypentyl)-1H-pyrazole-4-carboxylate 1k. Yellow oil. Yield: 51%. 1H-NMR (400 MHz, DMSO-d6): δ 0.84 (t, J = 6.0, 3H, CH3), 1.20–1.42 (m, 7H, CH3 +2CH2), 3.82–4.02 (m, 3H, CHOH + CH2N), 4.14 (q, J = 7.0, 2H, CH2O), 4.99 (br s, 1H, OH exchangeable with D2O), 7.52–7.87 (m, 4H, Ar), 7.91 (s, 1H, H-3), 12.11 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 178.12, 167.65, 162.79, 161.44, 160.33, 140.22, 135.37, 130.66, 123.42, 119.36, 103.44, 69.14, 59.92, 54.41, 35.50, 18.44, 14.12. Anal. (C19H23N4O4SF) calcd for C, H, N, S.
Ethyl 5-(3-(4-fluorobenzoyl)thioureido)-1-(2-hydroxypentyl)-1H-pyrazole-4-carboxylate 1l. White solid. M.p.: 134–136 °C. Yield: 61%. 1H-NMR (400 MHz, DMSO-d6): δ 0.84 (t, J = 6.0, 3H, CH3), 1.10–1.40 (m, 7H, CH3 +2CH2), 3.82–3.99 (m, 3H, CHOH + CH2N), 4.13 (q, J = 7.0, 2H, CH2O), 5.00 (br s, 1H, OH exchangeable with D2O), 7.38–7.44-(m, 2H, Ar), 7.92 (s, 1H, H-3), 8.05–8.12 (m, 2H, Ar), 12.14 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 174.73, 166.01, 163.54, 161.44, 140.22, 130.26, 115.99, 105.38, 70.72, 60.71, 55.43, 37.09, 18.44, 13.99, 13.20. Anal. (C19H23N4O4SF) calcd for C, H, N, S.
Ethyl 5-(3-benzoylthioureido)-1-(2-hydroxyhexyl)-1H-pyrazole-4-carboxylate 1m. White solid. M.p.: 108–110 °C. Yield: 43%. 1H-NMR (400 MHz, DMSO-d6): δ 0.80 (t, J = 6.0, 3H, CH3), 1.15 (t, J = 7.0, 3H, CH3), 1.20–1.38 (m, 6H, 3CH2), 3.76–4.07 (m, 3H, CHOH + CH2N), 4.12 (q, J = 7.0, 2H, CH2O), 4.96 (br s, 1H, OH exchangeable with D2O), 7.41–7.98 (m, 6H, 5Ar + H-3), 12.06 (br s, 1H, NH, exchangeable with D2O), 12.18 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 174.73, 166.52, 159.98, 140.22, 140.08, 137.13, 133.20, 128.50, 128.09, 127.68, 102.23, 71.74, 63.63, 57.32, 36.52, 25.26, 21.32, 14.46, 13.20. Anal. (C20H26N4O4S) calcd for C, H, N, S.
Ethyl 5-(3-(3-fluorobenzoyl)thioureido)-1-(2-hydroxyhexyl)-1H-pyrazole-4-carboxylate 1n. Yellow oil. Yield: 47%. 1H-NMR (400 MHz, DMSO-d6): δ 0.85–0.96 (m, 6H, 2CH3), 1.27–1.31 (m, 6H, 3CH2), 3.95–4.25 (m, 3H, CHOH + CH2N), 4.30 (q, J = 7.0, 2H, CH2O), 7.35–7.92 (m, 4H, Ar), 7.98 (s, 1H, H-3), 9.49 (br s, 1H, NH, exchangeable with D2O), 12.09 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 174.95, 168.91, 162.24, 161.44, 159.77, 140.29, 136.15, 130.66, 123.42, 118.99, 114.95, 103.44, 70.23, 60.12, 55.30, 35.74, 27.42, 22.70, 14.28, 12.41. Anal. (C20H25N4O4SF) calcd for C, H, N, S.
Ethyl 5-(3-(4-fluorobenzoyl)thioureido)-1-(2-hydroxyhexyl)-1H-pyrazole-4-carboxylate 1o. White solid. M.p.: 114–116 °C. Yield: 94%. 1H-NMR (400 MHz, DMSO-d6): δ 0.84 (t, J = 6.0, 3H, CH3), 1.07–1.39 (m, 9H, CH3 +3CH2), 3.80–4.03 (m, 3H, CHOH + CH2N), 4.13 (q, J = 7.0, 2H, CH2O), 5.00 (d, J = 4.8, 1H, OH exchangeable with D2O), 7.38–7.44 (m, 2H, Ar), 7.92 (s, 1H, H-3), 8.07–8.12 (m, 2H, Ar), 12.14 (br s, 1H, NH, exchangeable with D2O). 13C-NMR (101 MHz, DMSO-d6): δ 175.68, 167.91, 167.00, 165.45, 163.13, 139.81, 138.95, 132.11, 130.17, 117.86, 115.76, 102.23, 67.88, 59.53, 53.30, 33.61, 26.67, 22.26, 11.23, 10.92. Anal. (C20H25N4O4SF) calcd for C, H, N, S.

3.2. Microbiological Evaluation

3.2.1. Bacterial Species Considered in This Study

A total of 29 clinical strains were used in this study, all belonging to a collection obtained from the School of Medicine and Pharmacy of University of Genoa (Italy), identified by VITEK® 2 (Biomerieux, Firenze, Italy) or the matrix-assisted laser desorption/ionization time-of-flight (MALDI-TOF) mass spectrometric technique (Biomerieux, Firenze, Italy).
Particularly, fifteen strains were used to preliminary test the 15 new PTU derivatives, which were eleven Gram-positive strains (four MRSA Staphylococcus aureus strains), threeStaphylococcus epidermidis isolates (two MRSE and one resistant to methicillin and linezolid), two Enterococcus faecalis strains (one vancomycin-sensitive and one vancomycin-resistant (VRE)), two Enterococcus faecium isolates (one vancomycin-sensitive and one (VRE)) and four Gram-negative isolates (two Escherichia coli isolates, one was a New Delhi metallo-β-lactamase (NDM)-producing isolate and both carbapenem resistant, and two Pseudomonas aeruginosa MDR isolates).
In addition, fourteen isolates of the Staphylococcus genus were also evaluated, including one S. saprophyticus, two S. capitis (one was resistant to methicillin), one S. warneri (resistant to methicillin), two S. simulans (both resistant to methicillin), two S. lugdunensis (one was resistant to methicillin), three S. haemoliticus (two were resistant to methicillin), two S. hominis (both resistant to methicillin) and one S. auricularis (resistant to methicillin).

3.2.2. Determination of the Minimal Inhibitory Concentrations (MICs)

MIC values were determined following the microdilution procedures detailed by the European Committee on Antimicrobial Susceptibility Testing EUCAST [27], as also reported in our previous studies [20,28]. Briefly, serial two-fold dilutions in Mueller–Hinton (MH) broth (Merck, Darmstadt, Germany) of all the fifteen samples (dissolved in DMSO), ranging from 128 to 1 µg/mL, were used. DMSO was also tested as a control to verify the absence of antibacterial activity of the solvent used for the experiments. Cultures of all the selected bacteria, after overnight incubation, were diluted to yield a standardized inoculum of 1.5 × 108 CFU/mL. Appropriate aliquots of each suspension were added to 96-well microplates containing dilutions of compounds to be tested to yield a final concentration of about 5 × 105 cells/mL. After 24 h of incubation at 37 °C, the lowest concentration of sample that prevented visible growth was recorded as the MIC. All MICs were obtained at least in triplicate and results were expressed reporting the modal value; that is, the value that has been observed most frequently. In case of equivocal or not clear results, more than three determinations of MICs were carried out.

3.2.3. Killing Curves

Killing curve assays for most interesting compounds (1a, 1e, 1i and 1m) were performed on the four MRSA isolates selected for this study as previously reported [20,28].
A mid logarithmic phase bacterial culture was diluted in MH broth (10 mL) containing 4× MIC of all the compounds to give a final inoculum of 3.0 × 105 CFU/mL. The same inoculum was added to MH broth, as a growth control. Tubes were incubated at 37 °C with constant shaking for 24 h. Samples of 0.20 mL from each tube were removed at 0, 2, 4, 6 and 24 h, diluted appropriately with a 0.9% sodium chloride solution to avoid carryover of compounds being tested, plated onto MH plates and incubated for 24 h at 37 °C. Growth controls were run in parallel. The percentage of surviving bacterial cells was determined for each sampling time by comparing colony counts with those of standard dilutions of the growth control. Results have been expressed as log10 of viable cell numbers (CFU/mL) of of surviving bacterial cells over a 24 h period. All time-kill curve experiments were performed in triplicate.

3.3. Maintenance of Cell Cultures

Vero cell line, isolated from kidney epithelia cells extracted from Cercopithecus aethiops, was certified by STR DNA profile analysis by Biological Bank, a Core Facility of the IRCCS San Martino University Hospital-IST National Institute for Cancer Research (Genoa, Italy). Vero cells were routinely cultured al 37 °C under 5% CO2 in a DMEM medium (Euroclone, Milan, Italy) plus 10% heat inactivated FBS serum (Euroclone, Milan, Italy). No antibiotic or anti-fungine solutions were added to the standard or experimental medium in order to avoid any potential interference of these drugs with the experimental conditions. The medium was changed every 2 to 3 days and cells were sub-cultured by TripLETM Express (Invitrogen, Life Technologies, Carlsbad, CA, USA) treatment when the original flask was approximately 75% confluent. All cell cultures were found to be mycoplasma-free during regular checks with the Reagent Set Mycoplasma Euroclone (Euroclone, Milan, Italy).

3.3.1. Cell Viability Index

MTT Test

At the end of each experimental treatment, the cell viability was assessed by thiazolyl blue tetrazolium dye reduction assay (MTT) (Euroclone, Milan, Italy) [29]. The optical densities (OD) of the dissolved formazan crystals (for the MTT test) were determined spectrophotometrically at 570 nm.
A chemical compound was considered toxic if the cell viability was reduced by 15% compared to untreated cultures, according to the ECVAM’s guidelines testing any cytotoxic effects of the compounds and in parallel for measuring cell proliferation, according to manufacturer’s instructions.

4. Conclusions

The obtained results pointed at the PTU scaffold as a pharmaceutically relevant chemotype for the development of novel antibacterial agents active against Gram-positive species. In fact, novel PTUs here reported proved to be equally or more active that previous derivatives I. Compounds 1ao evidenced interesting activity towards Gram-positive resistant pathogen, often associated with severe pulmonary disease in CF patients [30,31,32,33,34]; in fact, several novel PTUs proved to be effective against different species of the Staphylococcus genus, with MIC values ranging from 32 to 128 µg/mL on MRSA and MRSE strains. In addition, time-killing experiments confirm the bacteriostatic actions of this class of compounds.
For all the synthesized compounds, favourable pharmacokinetic properties were calculated, evidencing for a major part of PTUs’ good ADMET properties. Finally, considering the preliminary cytotoxicity results obtained on Vero cells, it is reasonable to assume that the PTU library here reported have low toxicity.
Collected data allow us to draw the following SAR considerations regarding N1 and thiourea moiety substitution (Figure 4): (1) the replacement of hydroxy-2-phenylethyl chain with hydroxyalkyl substituent at N1 improves the antimicrobial activity of the compounds, but make PTUs here reported P-gp efflux pump substrates; (2) the presence of the N1 hydroxyexyl chain at N1 reduced solubility (compounds 1mo); (3) the ortho-fluoro substitution of the thiourea benzoyl ring is detrimental for activity (as in derivatives 1b,f,j), whereas para and meta fluoro-substituted compounds evidenced a potency comparable to their non-substituted analogues (1a,e,i,m); (4) the simultaneous presence of the fluorine atom and longer hydroxyalkyl chain (hydroxypentyl 1jl or hydroxyexyl 1n,1o) increases the Csp3 fraction with an improvement of the predicted bioavailability; (5) the carboxyethyl function at C4 position resulted relevant for biological activity, as previously evidenced for derivatives IIa,b.
Additional chemical modifications on the C4 and C3 pyrazole nucleus (as in previous derivatives I [20]) will be performed to extend the SAR consideration about the PTU chemotype.
As in our previous studies on pyrazole derivatives I and II [20], these collected results supported that, upon nano-formulation with proper polymer matrices, the new synthesized compounds could provide novel pyrazole-based drug delivery systems with enhanced and enlarged-spectrum of antibacterial activity, particularly against Gram-positive MRSA and MRSE species.
In conclusion, PTUs here reported could represent a new starting point for the development of new antibacterial pyrazole-based agents [35,36,37,38] able to counteract strains resistant to common antibiotics and consequently a new therapeutic approach for CF patients [39,40,41]. Additional studies will be necessary to identify the mechanism of action of this class of molecules.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ph17030376/s1, Elemental analysis, 1H NMR (400 MHz) and 13C NMR (101 MHz) of compounds 4d, 1ao are reported.

Author Contributions

Conceptualization, synthesis and characterization C.B. and B.T.; writing, pharmacokinetic properties, drug-likeness and toxicity prediction A.S. and E.C.; writing—original draft preparation, C.B. and A.S.; antimicrobial activity, D.C. and A.M.S.; cytotoxicity test, S.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors thank F. Rapetti, M. Anzaldi and R. Raggio for spectral recording and elemental analysis.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Breijyeh, Z.; Karaman, R. Design and Synthesis of Novel Antimicrobial Agents. Antibiotics 2023, 12, 628. [Google Scholar] [CrossRef]
  2. McGuinness, W.A.; Malachowa, N.; DeLeo, F.R. Vancomycin Resistance in Staphylococcus aureus. YJBM 2017, 90, 269–281. [Google Scholar]
  3. Barros, E.M.; Martin, M.J.; Selleck, E.M.; Lebreton, F.; Sampaio, J.L.M.; Gilmore, M.S. Daptomycin Resistance and Tolerance Due to Loss of Function in Staphylococcus aureus dsp1 and asp23. Antimicrob. Agents Chemother. 2018, 63, e01542-18. [Google Scholar] [CrossRef] [PubMed]
  4. Egan, S.A.; Shore, A.C.; O’Connell, B.; Brennan, G.I.; Coleman, D.C. Linezolid resistance in Enterococcus faecium and Enterococcus faecalis from hospitalized patients in Ireland: High prevalence of the MDR genes optrA and poxtA in isolates with diverse genetic backgrounds. J. Antimicrob. Chemother. 2020, 75, 1704–1711. [Google Scholar] [CrossRef] [PubMed]
  5. Dasenbrook, E.C.; Checkley, W.; Merlo, C.A.; Konstan, M.W.; Lechtzin, N.; Boyle, M.P. Association between respiratory tract methicillin-resistant Staphylococcus aureus and survival in cystic fibrosis. JAMA 2010, 303, 2386–2392. [Google Scholar] [CrossRef] [PubMed]
  6. Ren, C.L.; Morgan, W.J.; Konstan, M.W.; Schechter, M.S.; Wagener, J.S.; Fisher, K.A.; Regelmann, W.E. Presence of methicillin resistant Staphylococcus Aureus in respiratory cultures from cystic fibrosis patients is associated with lower lung function. Pediatr. Pulmonol. 2007, 42, 513–518. [Google Scholar] [CrossRef]
  7. Xu, X.; Zhang, X.; Zhang, G.; Tadi, D.A. Prevalence of antibiotic resistance of Staphylococcus aureus in cystic fibrosis infection: A systematic review and meta-analysis. JGAR 2013, in press. [Google Scholar] [CrossRef]
  8. Akil, N.; Muhlebach, M.S. Biology and management of methicillin resistant Staphylococcus aureus in cystic fibrosis. Pediatr. Pulmonol. 2018, 53, S64–S74. [Google Scholar] [CrossRef] [PubMed]
  9. Molina, A.; Del Campo, R.; Máiz, L.; Morosini, M.I.; Lamas, A.; Baquero, F.; Cantón, R. High prevalence in cystic fibrosis patients of multiresistant hospital-acquired methicillin-resistant Staphylococcus aureus ST228-SCCmecI capable of biofilm formation. J. Antimicrob. Chemother. 2008, 62, 961–967. [Google Scholar] [CrossRef] [PubMed]
  10. Lusardi, M.; Spallarossa, A.; Brullo, C. Amino-Pyrazoles in Medicinal Chemistry: A Review. Int. J. Mol. Sci. 2023, 24, 7834. [Google Scholar] [CrossRef]
  11. Verma, R.; Verma, S.K.; Rakesh, K.P.; Girish, Y.R.; Ashrafizadeh, M.; Sharath Kumar, K.S.; Rangappa, K.S. Pyrazole-based analogs as potential antibacterial agents against methicillin-resistance staphylococcus aureus (MRSA) and its SAR elucidation. Eur. J. Med. Chem. 2021, 212, 113134. [Google Scholar] [CrossRef]
  12. Abu-Hashem, A.; El-Shazly, M. Synthesis of quinoxaline, pyrimidine, and pyrazole furochromone derivatives as cytotoxic agents. Monatsh. Chem. 2017, 148, 1853–1863. [Google Scholar] [CrossRef]
  13. Abu-Hashem, A.A.; Aly, A.S. Synthesis of New Pyrazole, Triazole, Thiazolidine, -Pyrimido[4,5-b] quinoline derivatives with Potential Antitumor Activity. Arch. Pharm. Res. 2012, 35, 437–445. [Google Scholar] [CrossRef] [PubMed]
  14. Abu-Hashem, A.; Gouda, M.A. Synthesis and Antimicrobial Activity of Some Novel Quinoline, Chromene, Pyrazole Derivatives Bearing Triazolopyrimidine Moiety. J. Heter. Chem. 2017, 54, 850–858. [Google Scholar] [CrossRef]
  15. Abu-Hashem, A. Synthesis, and biological activity of pyrimidines, quinolines, thiazines and pyrazoles bearing a common thieno moiety. Acta Pol. Pharm.-Drug Res. 2018, 75, 59–70. [Google Scholar]
  16. Abu-Hashem, A. Synthesis of new pyrazoles, oxadiazoles, triazoles, pyrrolotriazines and pyrrolotriazepines as potential cytotoxic agents. J. Heter. Chem. 2021, 58, 805–821. [Google Scholar] [CrossRef]
  17. Bennani, F.E.; Doudach, L.; El Rhayam, Y.; Karrouchi, K.; Cherrah, Y.; Tarib, A.; Ansar, M.; Faouzi, M.E.A. Identification of the new progress on Pyrazole Derivatives Molecules as Antimicrobial and Antifungal Agents. West Afr. J. Med. 2022, 39, 1217–1244. [Google Scholar]
  18. Alam, M.A. Pyrazole: An emerging privileged scaffold in drug discovery. Future Med. Chem. 2023, 15, 2011–2023. [Google Scholar] [CrossRef]
  19. Ambade, S.S.; Gupta, V.K.; Bhole, R.P.; Khedekar, P.B.; Chikhale, R.V. A Review on Five and Six-Membered Heterocyclic Compounds Targeting the Penicillin-Binding Protein 2 (PBP2A) of Methicillin-Resistant Staphylococcus aureus (MRSA). Molecules 2023, 28, 7008. [Google Scholar] [CrossRef]
  20. Brullo, C.; Caviglia, D.; Spallarossa, A.; Alfei, S.; Franzblau, S.G.; Tasso, B.; Schito, A.M. Microbiological Screening of 5-Functionalized Pyrazoles for the Future Development of Optimized Pyrazole-Based Delivery Systems. Pharmaceutics 2022, 14, 1770. [Google Scholar] [CrossRef]
  21. Bruno, O.; Brullo, C.; Bondavalli, F.; Schenone, S.; Spisani, S.; Falzarano, M.S.; Varani, K.; Barocelli, E.; Ballabeni, V.; Giorgio, C.; et al. 1-Methyl and 1-(2-hydroxyalkyl)-5-(3-alkyl/cycloalkyl/phenyl/naphthylureido)-1H-pyrazole-4-carboxylic acid ethyl esters as potent human neutrophil chemotaxis inhibitors. Bioorg. Med. Chem. 2009, 17, 3379–3387. [Google Scholar] [CrossRef]
  22. Lewis, R.T.; Bode, C.M.; Choquette, D.M.; Potashman, M.; Romero, K.; Stellwagen, J.C.; Teffera, Y.; Moore, E.; Whittington, D.A.; Chen, H.; et al. The discovery and optimization of a novel class of potent, selective, and orally bioavailable anaplastic lymphoma kinase (ALK) inhibitors with potential utility for the treatment of cancer. J. Med. Chem. 2012, 55, 6523–6540. [Google Scholar] [CrossRef]
  23. Daina, A.; Michielin, O.; Zoete, V. SwissADME: A free web tool to evaluate pharmacokinetics, drug-likeness and medicinal chemistry friendliness of small molecules. Sci. Rep. 2017, 7, 42717. [Google Scholar] [CrossRef]
  24. Brenk, R.; Schipani, A.; James, D.; Krasowski, A.; Gilbert, I.H.; Frearson, J.; Wyatt, P.G. Lessons learnt from assembling screening libraries for drug discovery for neglected diseases. ChemMedChem 2008, 3, 435–444. [Google Scholar] [CrossRef] [PubMed]
  25. Delaney, J.S. ESOL: Estimating Aqueous Solubility Directly from Molecular Structure. J. Chem. Inf. Model. 2004, 44, 1000–1005. [Google Scholar] [CrossRef] [PubMed]
  26. Lusardi, M.; Wehrle-Haller, B.; Sidibe, A.; Ponassi, M.; Iervasi, E.; Rosano, C.; Brullo, C.; Spallarossa, A. Novel 5-aminopyrazoles endowed with anti-angiogenetic properties: Design, synthesis and biological evaluation. Eur. J. Med. Chem. 2023, 260, 115727. [Google Scholar] [CrossRef] [PubMed]
  27. EUCAST. European Committee on Antimicrobial Susceptibility Testing. Available online: https://www.eucast.org/ast_of_bacteria (accessed on 23 June 2023).
  28. Schito, A.M.; Piatti, G.; Caviglia, D.; Zuccari, G.; Zorzoli, A.; Marimpietri, D.; Alfei, S. Bactericidal Activity of Non-Cytotoxic Cationic Nanoparticles against Clinically and Environmentally Relevant Pseudomonas spp. Isolates. Pharmaceutics 2021, 13, 1411. [Google Scholar] [CrossRef]
  29. Mosmann, T. Rapid colorimetric assay for cellular growth and survival: Application to proliferation and cytotoxicity assays. J. Immunol. Meth. 1983, 65, 55–63. [Google Scholar] [CrossRef] [PubMed]
  30. Conway, S.P.; Brownlee, K.G.; Denton, M.; Peckham, D.G. Antibiotic treatment of multidrug-resistant organisms in cystic fibrosis. Am. J. Respir. Med. 2003, 2, 321–332. [Google Scholar] [CrossRef] [PubMed]
  31. Chen, H.; Zhang, J.; He, Y.; Lv, Z.; Liang, Z.; Chen, J.; Li, P.; Liu, J.; Yang, H.; Tao, A.; et al. Exploring the Role of Staphylococcus aureus in Inflammatory Diseases. Toxins 2022, 14, 464. [Google Scholar] [CrossRef]
  32. Biswas, L.; Götz, F. Molecular Mechanisms of Staphylococcus and Pseudomonas Interactions in Cystic Fibrosis. Front. Cell. Infect. Microbiol. 2022, 11, 824042. [Google Scholar] [CrossRef]
  33. Parkins, M.D.; Elborn, J.S. Newer antibacterial agents and their potential role in cystic fibrosis pulmonary exacerbation management. J. Antimicrob. Chemother. 2010, 65, 1853–1861. [Google Scholar] [CrossRef] [PubMed]
  34. Rusnati, M.; D’Ursi, P.; Pedemonte, N.; Urbinati, C.; Ford, R.C.; Cichero, E.; Uggeri, M.; Orro, A.; Fossa, P. Recent Strategic Advances in CFTR Drug Discovery: An Overview. Int. J. Mol. Sci. 2020, 21, 2407. [Google Scholar] [CrossRef] [PubMed]
  35. Raslan, M.A.; Raslan, S.A.; Shehata, E.M.; Mahmoud, A.S.; Sabri, N.A. Advances in the Applications of Bioinformatics and Chemoinformatics. Pharmaceuticals 2023, 16, 1050. [Google Scholar] [CrossRef] [PubMed]
  36. Balbi, A.; Anzaldi, M.; Macciò, C.; Aiello, C.; Mazzei, M.; Gangemi, R.; Castagnola, P.; Miele, M.; Rosano, C.; Viale, M. Synthesis and biological evaluation of novel pyrazole derivatives with anticancer activity. Eur. J. Med. Chem. 2011, 46, 5293–5309. [Google Scholar] [CrossRef] [PubMed]
  37. Delpe-Acharige, A.; Zhang, M.; Eschliman, K.; Dalecki, A.; Covarrubias-Zambrano, O.; Minjarez-Almeida, A.; Shrestha, T.; Lewis, T.; Al-Ibrahim, F.; Leonard, S.; et al. Pyrazolyl Thioureas and Carbothioamides with an NNSN Motif against MSSA and MRSA. ACS Omega 2021, 6, 6088–6099. [Google Scholar] [CrossRef] [PubMed]
  38. Ommi, O.; Naiyaz Ahmad, M.; Gajula, S.N.R.; Wanjari, P.; Sau, S.; Agnivesh, P.K.; Sahoo, S.K.; Kalia, N.P.; Sonti, R.; Nanduri, S.; et al. Synthesis and pharmacological evaluation of 1,3-diaryl substituted pyrazole based (thio)urea derivatives as potent antimicrobial agents against multi-drug resistant Staphylococcus aureus and Mycobacterium tuberculosis. RSC Med. Chem. 2023, 14, 1296–1308. [Google Scholar] [CrossRef]
  39. Allen, L.; Allen, L.; Carr, S.B.; Davies, G.; Downey, D.; Egan, M.; Forton, J.T.; Gray, R.; Haworth, C.; Horsley, A.; et al. Future therapies for cystic fibrosis. Nat. Commun. 2023, 14, 693. [Google Scholar] [CrossRef]
  40. Esposito, C.; Kamper, M.; Trentacoste, J.; Galvin, S.; Pfister, H.; Wang, J. Advances in the Cystic Fibrosis Drug Development Pipeline. Life 2023, 13, 1835. [Google Scholar] [CrossRef]
  41. Singh, J.; Yeoh, E.; Fitzgerald, D.A.; Selvadurai, H. A systematic review on the use of bacteriophage in treating Staphylococcus aureus and Pseudomonas aeruginosa infections in cystic fibrosis. Paediatr. Respir. Rev. 2023, 48, 3–9. [Google Scholar] [CrossRef]
Figure 1. Structure of previous pyrazoles I and II and workflow of the applied strategies leading to the design of a novel small PTU library (compounds 1ao).
Figure 1. Structure of previous pyrazoles I and II and workflow of the applied strategies leading to the design of a novel small PTU library (compounds 1ao).
Pharmaceuticals 17 00376 g001
Scheme 1. Synthesis of compounds 1ao. Reagents and conditions: (a) hydrazine monohydrate, 90–95 °C, 15 min–1 h 79–82%; (b) ethyl ethoxymethylene cyanoacetate, toluene, 70–80 °C, 8 h (4a) or abs. EtOH, 70–80 °C, 8 h, 61–72% (4bd); (c) suitable benzoyl isothiocyanates 5ad, an. THF, reflux, 12 h, 35–94%.
Scheme 1. Synthesis of compounds 1ao. Reagents and conditions: (a) hydrazine monohydrate, 90–95 °C, 15 min–1 h 79–82%; (b) ethyl ethoxymethylene cyanoacetate, toluene, 70–80 °C, 8 h (4a) or abs. EtOH, 70–80 °C, 8 h, 61–72% (4bd); (c) suitable benzoyl isothiocyanates 5ad, an. THF, reflux, 12 h, 35–94%.
Pharmaceuticals 17 00376 sch001
Figure 2. Time-killing curves performed with compounds 1c, 1d, 1n, 1h at a concentration of 4× MIC on S. aureus 18 (MRSA). Error bars represent standard deviations (n = 3) of the mean values. Similar results were obtained for S. aureus strains 17, 187 and 195.
Figure 2. Time-killing curves performed with compounds 1c, 1d, 1n, 1h at a concentration of 4× MIC on S. aureus 18 (MRSA). Error bars represent standard deviations (n = 3) of the mean values. Similar results were obtained for S. aureus strains 17, 187 and 195.
Pharmaceuticals 17 00376 g002
Figure 3. Cytotoxicity of compounds 1c (in light blue) and 1n (in blue) performed at concentrations 32 and 64 μg/mL on Vero cells. Data, expressed as percentage of viability versus untreated cultures and extrapolated by MTT assay, are the means ± SD of three separate experiments performed in triplicate. * = p < 0.01 versus untreated cultures (ANOVA and Dunnett test). At a concentration of 32 µg/mL, tested compounds showed the highest viability rates. According to ECVAM guidelines, compounds 1c and 1n can therefore be considered low toxic (at 32 µg/mL) and moderately toxic (at 64 µg/mL).
Figure 3. Cytotoxicity of compounds 1c (in light blue) and 1n (in blue) performed at concentrations 32 and 64 μg/mL on Vero cells. Data, expressed as percentage of viability versus untreated cultures and extrapolated by MTT assay, are the means ± SD of three separate experiments performed in triplicate. * = p < 0.01 versus untreated cultures (ANOVA and Dunnett test). At a concentration of 32 µg/mL, tested compounds showed the highest viability rates. According to ECVAM guidelines, compounds 1c and 1n can therefore be considered low toxic (at 32 µg/mL) and moderately toxic (at 64 µg/mL).
Pharmaceuticals 17 00376 g003
Figure 4. SAR considerations about PTUs here reported.
Figure 4. SAR considerations about PTUs here reported.
Pharmaceuticals 17 00376 g004
Table 1. Chemical structure, melting point and reaction yield of novel PTU derivatives 1ao.
Table 1. Chemical structure, melting point and reaction yield of novel PTU derivatives 1ao.
Pharmaceuticals 17 00376 i001
CompoundRXMelting Point (°C)Yield (%)
1aCH3H106–108 °C64%
1bCH3o-FYellow oil45%
1cCH3m-FYellow oil35%
1dCH3p-F131–133 °C39%
1eC2H5HYellow oil55%
1fC2H5o-FYellow oil41%
1gC2H5m-FYellow oil40%
1hC2H5p-F120–122 °C50%
1iC3H7H126–128 °C55%
1jC3H7o-FYellow oil36%
1kC3H7m-FYellow oil51%
1lC3H7p-F134–136 °C61%
1mC4H9H108–110 °C43%
1nC4H9m-FYellow oil47%
1oC4H9p-F114–116 °C94%
Table 2. MIC values (expressed as µg/mL) of PTU 1 and reference compound oxacillin (oxa) against bacteria of the Gram-positive and Gram-negative species, obtained from experiments carried out at least in triplicate. * denotes resistant to methicillin; ** denotes resistant to methicillin and linezolid; # denotes resistance to vancomycin; NDM: New Delhi metallo-beta-lactamase producer; MDR: multidrug-resistant strain; CR carbapenem resistant. In bold are evidenced the most active compounds.
Table 2. MIC values (expressed as µg/mL) of PTU 1 and reference compound oxacillin (oxa) against bacteria of the Gram-positive and Gram-negative species, obtained from experiments carried out at least in triplicate. * denotes resistant to methicillin; ** denotes resistant to methicillin and linezolid; # denotes resistance to vancomycin; NDM: New Delhi metallo-beta-lactamase producer; MDR: multidrug-resistant strain; CR carbapenem resistant. In bold are evidenced the most active compounds.
MIC (µg/mL)
Strain1a1b1c1d1e1f1g1h1i1j1k1l1m1n1ooxa
Gram-positive
S. aureus 17 *128>12864128128>1286412864>12864>1286464>128256
S. aureus 18 *128>128646464>12812812864>12864>1286464>128512
S. aureus 187 *64>128323264>128646464>12832>1283232>128512
S. aureus 195 *64>128646464>128646464>12864>1283264>128256
S. epidermidis 22 *32>128643232>128>1286432>12864>1283264>128256
S. epidermidis 171 *64>1281283264>1281283264>12864>1286464>128128
S. epidermidis 181 **64>128643264>1281283264>128128>12864128>128256
E. faecalis 365 #>128>128>128>128>128>128>128>128>128>128>128>128>128>128>128-
E. faecalis 28>128>128>128>128>128>128>128>128>128>128>128>128>128>128>128-
E. faecium 152 #>128>128>128>128>128>128>128>128>128>128>128>128>128>128>128-
E. faecium 158>128>128>128>128>128>128>128>128>128>128>128>128>128>128>128-
Gram-negative
E. coli 462 (NDM, CR)>128>128>128>128>128>128>128>128>128>128>128>128>128>128>128-
E. coli 475 (CR)>128>128>128>128>128>128>128>128>128>128>128>128>128>128>128-
P. aeruginosa 1V MDR>128>128>128>128>128>128>128>128>128>128>128>128>128>128>128-
P. aeruginosa 6G MDR>128>128>128>128>128>128>128>128>128>128>128>128>128>128>128-
Table 3. MIC values of some PTUs 1 and reference compound oxacillin (oxa) against other Staphylococcus species, obtained from experiments carried out at least in triplicate, expressed as µg/mL. * denotes resistant to methicillin. In bold are evidenced the most active compounds.
Table 3. MIC values of some PTUs 1 and reference compound oxacillin (oxa) against other Staphylococcus species, obtained from experiments carried out at least in triplicate, expressed as µg/mL. * denotes resistant to methicillin. In bold are evidenced the most active compounds.
MIC (µg/mL)
Strains1a1d1e1h1i1moxa
S. saprophyticus 41>128>128>128>128>128>1280.5
S. capitis 71 *1286412812812812864
S. capitis 121>128128>128128>128>1280.25
S. warneri 74 *>128>128>128>128>128>12864
S. simulans 94 *>128>128>128>128>128>12816
S. simulans 163 *>128>128>128>128>128>12816
S. lugdunensis 963232323232321
S. lugdunensis 137 *32321632163216
S. haemoliticus 115 *1286464646412864
S. haemoliticus 193 *>128128128128>128>12816
S. haemoliticus 174>128>128128>128>128>1280.25
S. hominis 124 *12812812864128>12816
S. hominis 125 *12812812812812812824
S. auricularis 136 *32321632161616
Table 5. Predicted toxicity properties of compounds 1ao.
Table 5. Predicted toxicity properties of compounds 1ao.
CpdLD50 (mg/kg)Reliability (R.I.)
1a26000.45
1b–d16000.39
1e28000.45
1fh17000.39
1i30000.45
1jl19000.40
1m33000.45
1no20000.40
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

Schito, A.M.; Caviglia, D.; Penco, S.; Spallarossa, A.; Cichero, E.; Tasso, B.; Brullo, C. New Pyrazolyl Thioureas Active against the Staphylococcus Genus. Pharmaceuticals 2024, 17, 376. https://doi.org/10.3390/ph17030376

AMA Style

Schito AM, Caviglia D, Penco S, Spallarossa A, Cichero E, Tasso B, Brullo C. New Pyrazolyl Thioureas Active against the Staphylococcus Genus. Pharmaceuticals. 2024; 17(3):376. https://doi.org/10.3390/ph17030376

Chicago/Turabian Style

Schito, Anna Maria, Debora Caviglia, Susanna Penco, Andrea Spallarossa, Elena Cichero, Bruno Tasso, and Chiara Brullo. 2024. "New Pyrazolyl Thioureas Active against the Staphylococcus Genus" Pharmaceuticals 17, no. 3: 376. https://doi.org/10.3390/ph17030376

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