Next Article in Journal
Al-Salahi, R. Synthesis and Reactivity of [1,2,4]Triazolo-annelated Quinazolines. Molecules 2010, 15, 7016-7034
Previous Article in Journal
Acute Oral Toxicity and Brine Shrimp Lethality of Elaeis guineensis Jacq., (Oil Palm Leaf) Methanol Extract
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigating the Spectrum of Biological Activity of Ring-Substituted Salicylanilides and Carbamoylphenylcarbamates

by
Jan Otevrel
1,
Zuzana Mandelova
1,2,
Matus Pesko
3,
Jiahui Guo
4,
Katarina Kralova
5,
Frantisek Sersen
5,
Marcela Vejsova
6,
Danuta S. Kalinowski
7,
Zaklina Kovacevic
7,
Aidan Coffey
4,
Jozef Csollei
1,
Des R. Richardson
7 and
Josef Jampilek
1,2,*
1
Department of Chemical Drugs, Faculty of Pharmacy, University of Veterinary and Pharmaceutical Sciences, Palackeho 1/3, 61242 Brno, Czech Republic
2
Zentiva k.s., U kabelovny 130, 102 37 Prague, Czech Republic
3
Department of Ecosozology and Physiotactics, Faculty of Natural Sciences, Comenius University, Mlynska dolina Ch-2, 842 15 Bratislava, Slovakia
4
Department of Biological Sciences, Cork Institute of Technology, Bishopstown, Cork, Ireland
5
Institute of Chemistry, Faculty of Natural Sciences, Comenius University, Mlynska dolina Ch-2, 842 15 Bratislava, Slovakia
6
Department of Biological and Medical Sciences, Faculty of Pharmacy in Hradec Kralove, Charles University in Prague, Heyrovskeho 1203, 500 05 Hradec Kralove, Czech Republic
7
Department of Pathology and Bosch Institute, University of Sydney, Sydney, New South Wales 2006, Australia
*
Author to whom correspondence should be addressed.
Molecules 2010, 15(11), 8122-8142; https://doi.org/10.3390/molecules15118122
Submission received: 13 August 2010 / Revised: 5 November 2010 / Accepted: 9 November 2010 / Published: 10 November 2010

Abstract

:
In this study, a series of twelve ring-substituted salicylanilides and carbamoylphenylcarbamates were prepared and characterized. The compounds were analyzed using RP-HPLC to determine lipophilicity. They were tested for their activity related to the inhibition of photosynthetic electron transport (PET) in spinach (Spinacia oleracea L.) chloroplasts. Moreover, their site of action in the photosynthetic apparatus was determined. Primary in vitro screening of the synthesized compounds was also performed against mycobacterial, bacterial and fungal strains. Several compounds showed biological activity comparable with or higher than the standards 3-(3,4-dichlorophenyl)-1,1-dimethylurea, isoniazid, penicillin G, ciprofloxacin or fluconazole. The most active compounds showed minimal anti-proliferative activity against human cells in culture, indicating they would have low cytotoxicity. For all compounds, the relationships between lipophilicity and the chemical structure are discussed.

1. Introduction

Salicylanilides are an important class of aromatic compounds with a wide range of pharmacological activities. A number of them show anti-bacterial [1,2,3,4], anti-mycobacterial [5,6], anti-fungal [5,7] and anti-protozoal/molluscicidal [8] as well as anti-inflammatory [9,10] or anti-neoplastic activities [11,12,13]. Recently, a number of organic carbamates have been found to be potential anti-bacterial, anti-mycobacterial and anti-viral agents [14,15]. The carbamate residue present in these new molecules contributes as a core component [16] or, incorporated into a known molecule, contributes to the improvement of its pharmacodynamic and pharmacokinetic properties [17]. In particular, the carbamate group was successfully used to protect phenolic drugs [18].
The presence of an amide or thioamide (-NHCO- or -NHCS-) group is characteristic of a number of herbicides acting as photosynthesis inhibitors (acylanilides, thioacylanilides, phenylcarbamates, ureas, etc.), e.g., [19,20,21,22]. The wide spectrum of biological effects of salicylanilides includes herbicidal activity [23,24] and they were also found to be uncouplers of photosynthetic phosphorylation [25].
The inhibitory activity of the R’ substituted salicylanilides related to the inhibition of oxygen evolution in spinach chloroplasts was correlated with the hydrophobic parameter π- and the Hammett constant (σ) of the R’ substituents [19,20,21,22,23,24]. These compounds were additionally found to interact with the intermediate D (TyrD) in the photosynthetic apparatus of spinach chloroplasts [23]. Intermediate D is situated at the 161st position in the D2 protein occurring at the donor side of photosystem (PS) 2. In the presence of R’ substituted salicylanilides, the decreased intensity of the fluorescence emission band at 686 nm (belonging to the chlorophyll-protein complexes mainly in PS 2 [26]) suggested PS 2 as the site of action of the studied inhibitors.
Upon addition of 1,5-diphenylcarbazide (DPC, an artificial electron donor of PS 2 with a known site of action in the intermediate Z/D on the donor side of PS 2 [27]) to chloroplasts treated with R’ substituted salicylanilides, the inhibition of the oxygen evolution rate was completely restored. This indicated that the core of PS 2 (P 680) and a part of the electron transport chain, at least up to plastoquinone, remained intact. These results are in accordance with those obtained for 3-bromo- and 3,5-dibromosalicylanilides [28] and 3-methylsalicylanilides [24]. However, it should be stressed that 3,5-dibromosalicylanilides with R’ = 3-F and R = 3-Cl interacted not only with the intermediate D (TyrD), but also with the Z intermediate (TyrZ), which is situated in the 161st position of the D1 protein occurring at the donor side of PS 2 [28].
The R’-substituted salicylanilides and their 5-nitro, 5-fluoro and 5-bromo analogues also reduced chlorophyll production in suspensions of Chlorella vulgaris. Indeed, the anti-algal activity of the studied compounds depended on their lipophilicity and the electronic parameter σ of the R’ substituent [29]. Reduction of the chlorophyll content in Chlorella vulgaris was also observed in the presence of 3-methylsalicylanilides [30].
Following on from our previous investigations, [31,32,33,34,35,36,37,38], we examined the syntheses and in particular the herbicidal activity of various aromatic ring-substituted like-carboxamide derivatives. The compounds were tested for their photosynthesis-inhibiting activity (the inhibition of photosynthetic electron transport) in spinach chloroplasts (Spinacia oleracea L.) and their site of action in the photosynthetic apparatus was determined. Relationships between the structure of the new compounds and their in vitro anti-microbial activities or/and inhibitory activity of photosynthetic electron transport (PET) in spinach chloroplasts are discussed. The compounds were also assessed for activity against various bacterial, mycobacterial and fungal strains. Salicylic derivatives are known as strong chelators of the essential nutrient iron that is needed for cellular proliferation [39,40]. Therefore, the compounds were evaluated as potential iron chelating agents with possible anti-proliferative activity against neoplastic cells.

2. Results and Discussion

2.1. Chemistry

The synthetic routes used are shown in Scheme 1. Condensation of commercially available 2-(chlorocarbonyl)phenyl acetate with ring-substituted anilines [32] and subsequent hydrolysis [41] yielded a series of salicylic acid anilides 1-5.
4-Nitrobenzaldehyde was oxidized using KMnO4 [42] to yield 4-nitrobenzoic acid (6) that was chlorinated with SOCl2 and 4-nitrobenzoyl chloride (7) was thus obtained [32]. The chloride 7 was condensed with ring-substituted amines, as described previously [32], to afford amides 8-10. Nitrobenzamides 9 and 10 were hydrogenated over activated Raney Ni (Ra-Ni) [43,44] to give aminobenzamides 11 and 12. Methylphenylcarbamates 13 and 14 were generated by means of the reaction of aminobenzamides 11 and 12 with methyl chloroformiate [45].

2.2. Lipophilicity

Many low molecular weight drugs cross biological membranes through passive transport, which strongly depends on their lipophilicity. Lipophilicity is a property that has a major effect on absorption, distribution, metabolism, excretion, and toxicity (ADME/Tox) properties as well as pharmacological activity. Lipophilicity has been studied and applied as an important drug property for decades [46].
Hydrophobicities (log P/Clog P) of the compounds 1-5 and 8-14 were calculated using two commercially available programs (ChemDraw Ultra 10.0 and ACD/LogP) and also measured by means of the RP-HPLC determination of capacity factors k with subsequent calculation of log k. The procedure was performed under isocratic conditions with methanol as an organic modifier in the mobile phase using an end-capped non-polar C18 stationary RP column. The results are shown in Table 1 and illustrated in Figure 1.
The results obtained with all the compounds show that the experimentally-determined lipophilicities (log k) of compounds 1-5 and 8-11, 13, 14 are lower than those indicated by the calculated log P/Clog P, as shown in Figure 1. The results indicate that experimentally-determined log k values correlate relatively poorly with the calculated values (see Table 1 or Figure 1), which is probably caused by intramolecular interactions of these highly functionalized molecules.
As expected, aminobenzamide 11 showed the lowest lipophilicity (log k), while compounds 5 and 10 exhibited the highest log k values. Generally, nitro compounds 8-10 showed relatively high lipophilicity in comparison with amino derivatives, e.g. 11 < 13 < 9. Within the series of ring-substituted salicylanilides 1-5, the determined log k values corresponded to the expected lipophilicity increase 1 (2-CH3) < 3 (2-OCH3) < 2 (2,6-CH3) < 4 (2-OC2H5) < 5 (2-OC3H7). This dependence is approximately linear and thus, it can be assumed that log k values specify lipophilicity within the individual series of the studied compounds more precisely than calculated log P/Clog P data.

2.3. Biological activities

The compounds under investigation could be divided into three groups based on their chemical structure: Group 1 included salicylanilides 1-5; Group 2 contained nitro derivatives 8-10; and Group 3 was composed of amino derivatives or carbamates 11-14. The compounds showed a wide range of biological activities and some interesting structure-activity relationships were observed. All the results are shown in Table 2 (compound 12 did not show any biological activity in any of the assays, therefore it is not included in Table 2).

2.3.1. Inhibition of photosynthetic electron transport (PET) in spinach chloroplasts

The majority of the studied compounds inhibited PET in spinach chloroplasts, as shown in Table 2. Four compounds showed high inhibitory IC50 values: 1.0 µmol/L (14), 1.6 µmol/L (13), 1.6 µmol/L (3) and 2.7 µmol/L (1), which was comparable with the standard DCMU (IC50 = 1.9 µmol/L). The activity of the rest of the studied compounds was moderate or low relative to the standard. PET inhibition by 8 or 9 could not be determined due to precipitation of the compounds during the experiments. With respect to these small but closed specifically substituted groups of compounds some structure-activity relationships (SAR) can be proposed.
Within Group 1 (salicylanilides; compounds 1-5), the highest PET-inhibiting activity was shown by compounds 3 (2-OMe) and 1 (2-Me). Within the series of 2-alkoxy substituted compounds (3-5), their activity decreased with increasing lipophilicity (log k) and higher substituent bulkiness (substituent bulkiness (MR) and Hammett’s parameter (σ) are described in Table 1 [47]). It is probable that σ (especially the electron-donating effect) is also important for explaining biological activity and searching for structure-activity relationships within a series of compounds [47]. The inhibitory activity of compound 2 was two orders lower (IC50 = 331.4 μmol/L) than the activity of compounds 1 and 3, with comparable log k values (0.6661 and 0.6770, respectively). Therefore, when compounds 1 and 2 are considered, similar statements can be made: higher lipophilicity and substituent bulkiness decreases the inhibition of PET.
Generally, Group 2 (4-nitrobenzencarboxamides) showed only slight PET inhibition caused by the low solubility of compounds 8 and 9 in the testing medium. Compound 10 alone showed medium PET inhibiting activity.
Group 3 (4-aminobenzencarboxamide derivatives) showed very interesting PET activity. Compound 11 with a primary amino moiety possessed very low activity, while substitution of hydrogen in the 4-NH2 group of (11) by methyl acetate generating compounds 13 and 14 (carbamoylphenylcarbamates), resulted in a strong increase in the inhibitory activity of PET. This finding underlined the importance of the amide group, which can interact with amino acid residues or peptide bonds of proteins situated in photosystems, which can result in PET inhibition. It can be stated that an increase in lipophilicity positively influences PET-inhibiting activity, contrary to Group 1 compounds. Thus, it can be concluded that PET-inhibiting activity is increased by the electron-withdrawing effect and bulkiness of substituents in the anilide part of the molecule.
Experiments were then performed to determine the site of action of the studied compounds in the photosynthetic apparatus. In Figure 2, EPR spectra of untreated chloroplasts and those treated with the very effective PET inhibitor, compound 3 (Table 2), are presented. The EPR spectrum of untreated chloroplasts recorded in the dark consists of the typical signal IIslow (g = 2.0046, ΔBpp= 2 mT, see Figure 2A - full line). An increase of EPR signal intensity in the light represents signal IIvery fast (g = 2.0046, ΔBpp= 2 mT; in Figure 2A it is the difference between dotted and full lines). These EPR signals belong to the intermediates D and Z respectively, which are tyrosine radicals situated in the 161st positions in D2 (for D) and D1 (for Z) proteins at the donor side of PS 2 [48,49]. The treatment of chloroplasts with compound 3 evoked a marked increase in the intensity of the EPR signal in the irradiated chloroplasts (Figure 2B, dotted line). On the other hand, in the absence of light, no changes between EPR signal of chloroplasts treated with compound 3 and that of untreated chloroplasts were observed (Figure 2B, full line). The above-mentioned EPR signal (g = 2.0028, ΔBpp= 0.8 mT, Figure 2B, dotted line) is denoted as signal I and belongs to the oxidized chlorophyll a dimer in the core of PS 1 [50]. The pronounced increase of signal I intensity is caused by an interruption of PET between PS 2 and PS 1. In untreated chloroplasts, after the excitation of chlorophyll a dimer (P700) in the core of PS 1 and its subsequent oxidation by the first electron acceptor of PS 1, P700 is reduced by electrons supplied from PS 2, so the intensity of EPR signal I is very small. On the other hand, the disruption of PET between PS 2 and PS 1 results in a great increase of signal I intensity.
To further determine the site of action of the studied compounds, experiments with 1,5-diphenylcarbazide (DPC) were performed. DPC is an artificial electron donor for PS 2 with a known site of action in the Z/D intermediate [27]. The site of action of the artificial electron acceptor 2,6-dichlorophenol-indophenol (DCPIP) used for monitoring the Hill reaction is plastoquinone on the acceptor side of PS 2 [51]. Thus, if the part of the photosynthetic electron transport chain between manganese cluster and Z/D is inhibited, then DPC will restore PET through PS 2. It was found that after the addition of DPC (2 mmol/L) to chloroplasts treated with the most active compounds 3 and 13, PET through PS 2 was not restored. On the other hand, after the addition of 40 μmol/L of DCPIPH2, which is an artificial electron donor of PS 1 and which has a site of action in plastocyanin on the donor side of PS 1 [51], completely restored PET through PS 1. Based on the above-mentioned experimental results, we suggest that the site of action of the studied compounds could be in QB, which is the second quinone acceptor situated on the oxidizing (acceptor) site of PS 2.

2.3.2. In vitro anti-fungal and anti-bacterial susceptibility testing

All discussed above, compounds were tested for their in vitro anti-fungal and anti-bacterial activity, although only the salicylanilides 1-5 (Group 1), showed some broad spectrum activity. The rest of the compounds did not show any significant activity, potentially due to low solubility in the testing medium and their precipitation during the incubation period. The results are shown in Table 2. Among a number of fungi strains, Absidia corymbifera, as a common human pathogen causing pulmonary, rhinocerebral, disseminated, CNS or cutaneous infections and Trichophyton mentagrophytes as an example of a dermatophyte carrying a variety of cutaneous infections were selected.
All the compounds in Group 1 showed similar structure-activity relationships with respect to anti-fungal and anti-bacterial activity. Compounds 2 (2,6-CH3), 3 (2-OCH3) and 1 (2-CH3) exhibited low biological activity. Compounds with the longer alkoxy chain 4 (2-OC2H5) and 5 (2-OC3H7) showed higher anti-fungal and anti-bacterial activity. With respect to these specific alkyl/alkoxy substituted compounds it can be assumed that a lipophilicity increase and a bulkier substituent seem to be important factors for increasing both of these previously described activities. 2-Hydroxy-N-(2-propoxyphenyl)benzamide (5) showed similar chemotherapeutic activity against all tested fungal and bacterial strains. Notably, compound 5 demonstrated much higher activity against Absidia corymbifera than the standard fluconazole and also higher activity against methicillin resistant Staphylococcus aureus, and Staphylococcus epidermidis than the common clinically used antibiotics penicillin G and ciprofloxacin.

2.3.3. In vitro antimycobacterial evaluation

Although all the compounds discussed above were evaluated for their in vitro anti-mycobacterial activity against atypical mycobacterial strains, only compound 1 showed moderate activity, therefore no thorough structure-activity relationships could be established. According to the results presented in Table 2, it can be concluded that compound 1 exhibited activity against M. avium paratuberculosis and that its efficacy was comparable with that of the standard, isoniazid.

2.3.4. In vitro anti-proliferative activity

Carbonyl-containing moieties are potential iron chelating agents, with many demonstrating anti-proliferative activity due to their ability to bind cellular iron, which is required for proliferation [52]. Therefore, the anti-proliferative activity of compounds 1-5 was examined against the human SK-N-MC neuroepithelioma cell line to determine if there were any undesired cytotoxic side-effects. The SK-N-MC cell line was chosen as the effect of iron chelators on the proliferation of these cells has been extensively examined [53,54]. The anti-proliferative activity of the evaluated compounds was assessed in comparison to the well-known clinically used iron chelator, desferrioxamine (DFO), and the highly cytotoxic chelator, di-2-pyridylketone 4,4-dimethyl-3-thiosemicarbazone (Dp44mT) [52].
All tested compounds 1-5 demonstrated poor anti-proliferative effects against the SK-N-MC cell line, with IC50 values >6.25 µmol/L. As expected from our previous studies [53,54], the iron chelator, DFO, demonstrated poor anti-cancer activity, with an IC50 value of 17.07 ± 3.77 µmol/L, while the cytotoxic chelator, Dp44mT (IC50 = 0.01 ± 0.01 µmol/L), showed potent anti-proliferative effects. These results suggest that the poor anti-proliferative activity of the anilides 1-5 will lead to limited cytotoxicity if ever used as anti-fungal, anti-bacterial or anti-mycobacterial agents.

3. Experimental

3.1. General

All reagents were purchased from Aldrich. Kieselgel 60, 0.040-0.063 mm (Merck, Darmstadt, Germany) was used for column chromatography. TLC experiments were performed on alumina-backed silica gel 40 F254 plates (Merck, Darmstadt, Germany). The plates were illuminated under UV (254 nm) and evaluated in iodine vapour. The melting points were determined on Boetius PHMK 05 (VEB Kombinat Nagema, Radebeul, Germany) and are uncorrected. The purity of the final compounds was checked by a HPLC separation module (Waters Alliance 2695 XE, Waters Corp., Milford, MA, USA). The detection wavelength of 210 nm was chosen. The peaks in the chromatogram of the solvent (blank) were deducted from the peaks in the chromatogram of the sample solution. The purity of individual compounds was determined from the area peaks in the chromatogram of the sample solution. UV spectra (λ, nm) were determined on a Waters Photodiode Array Detector 2996 (Waters Corp., Milford, MA, USA) in ca. 6 × 10-4 M methanolic solution and log ε (the logarithm of molar absorption coefficient ε) was calculated for the absolute maximum λmax of individual target compounds. Infrared (IR) spectra were recorded on a Smart MIRacle™ ATR ZnSe for Nicolet™ Impact 410 FT-IR Spectrometer (Thermo Scientific, USA). The spectra were obtained by accumulation of 256 scans with 2 cm-1 resolution in the region of 4000-600 cm-1. All 1H and 13C NMR spectra were recorded on a Bruker Avance-500 FT-NMR spectrometer (500 MHz for 1H and 125 MHz for 13C, Bruker Comp., Karlsruhe, Germany). Chemicals shifts are reported in ppm (δ) using internal Si(CH3)4 as the reference, with diffuse, easily exchangeable signals being omitted. Mass spectra were measured using a LTQ Orbitrap Hybrid Mass Spectrometer (Thermo Electron Corporation, USA) with direct injection into an APCI source (400 °C) in the positive mode.

3.2. Synthesis

3.2.1. General procedure for synthesis of carboxamide derivatives 1-5

2-(Chlorocarbonyl)phenyl acetate (0.014 mol) dissolved in dry acetone (50 mL) was added drop-wise to a stirred solution of the corresponding substituted aniline (0.02 mmol) in dry pyridine (50 mL) kept at room temperature. After the addition was completed, stirring was continued for another 1 h. The reaction mixture was then poured into cold water (100 mL). The mixture was extracted with Et2O, dried over anhydrous MgSO4, filtered and the solvent removed in vacuo. The crude acetylsalicylanilide was dissolved in a 1 M solution of KOH stirred at 50 °C until clarification. The mixture was then neutralized using 1 M HCl. The crude amide was collected and recrystallized from aqueous ethanol and dried.
2-Hydroxy-N-(2-methylphenyl)benzamide (1). Yield 35%; m.p. 153-155 °C (m.p. 145-147°C [55]); HPLC purity 99.41%; UV (nm), λmax/log ε: 271.1/3.42; IR (Zn/Se ATR, cm-1): 3324 (N-H), 1611 (C=O), 3741 (O-H), 1542 (C=C); 1H-NMR (DMSO-d6), δ: 12.07 (bs, 1H, OH), 10.31 (bs, 1H, NH), 8.04 (dd, 1H, J = 7.9 Hz, J = 1.8 Hz, Ar), 7.80 (d, 1H, J = 7.9 Hz, Ar), 7.50-7.40 (m, 1H, Ar), 7.28 (d, 1H, J = 7.4 Hz, Ar), 7.23 (ddd, 1H, J = 7.6 Hz, J = 7.6 Hz, J = 1.5 Hz, Ar), 7.13 (ddd, 1H, J = 7.6 Hz, J = 7.6 Hz, J = 1.5 Hz, Ar), 7.06-6.94 (m, 2H, Ar), 2.28 (s, 3H, CH3); 13C-NMR (DMSO-d6), δ: 165.62, 158.08, 136.16, 133.65, 130.66, 130.30, 129.51, 126.25, 125.14, 123.99, 119.26, 117.20, 117.15, 17.71; HR-MS: for C14H14O2N [M+H]+ calculated 228.098 m/z, found 228.0892 m/z.
N-(2,6-Dimethylphenyl)-2-hydroxybenzamide (2). Yield 37%; m.p. 137-140 °C; HPLC purity 99.73%; UV (nm), λmax/log ε: 237.9/3.45; IR (Zn/Se ATR, cm-1): 3278 (N-H), 1602 (C=O), 1538 (C=C); 1H-NMR (DMSO-d6), δ: 12.34 (bs, 1H, OH), 10.07 (s, 1H, NH), 8.05 (dd, 1H, J = 8.2 Hz, J = 1.8 Hz, Ar), 7.51-7.43 (m, 1H, Ar), 7.22-7.10 (m, 3H, Ar), 7.02-6.92 (m, 2H, Ar), 2.19 (s, 6H, CH3); 13C-NMR (DMSO-d6), δ: 167.68, 160.06, 135.48, 134.28, 133.95, 128.14, 127.79, 127.00, 118.81, 117.45, 115.38, 17.97; HR-MS: for C15H16O2N [M+H]+ calculated 242.118 m/z, found 242.1084 m/z.
2-Hydroxy-N-(2-methoxyphenyl)benzamide (3). Yield 39%; m.p. 112-115 °C (m.p. 115-117 °C [55]); HPLC purity 99.92%; UV (nm), λmax/log ε: 298.4/3.49; IR (Zn/Se ATR, cm-1): 3288 (N-H), 1600 (C=O), 1228 (C-O), 1547 (C=C); 1H-NMR (DMSO-d6), δ: 11.70 (bs, 1H, OH), 10.81 (bs, 1H, NH), 8.43-8.36 (m, 1H, Ar), 8.03 (dd, 1H, J = 7.9 Hz, J = 1.8 Hz, Ar), 7.45-8.38 (m, 1H, Ar), 7.12-7.07 (m, 2H, Ar), 7.02 (dd, 1H, J = 8.2 Hz, J = 0.9 Hz, Ar), 7.01-6.93 (m, 2H, Ar), 3.89 (s, 1H, CH3); 13C-NMR (DMSO-d6), δ: 163.43, 156.28, 148.69, 133.32, 130.67, 127.84, 123.89, 120.57, 120.30, 119.59, 118.74, 116.90, 110.96, 55.99; HR-MS: for C14H14O3N [M+H]+ calculated 244.098 m/z, found 244.0980 m/z.
N-(2-Ethoxyphenyl)-2-hydroxybenzamide (4). Yield 35%; m.p. 105-107 °C; HPLC purity 99.94%; UV (nm), λmax/log ε: 299.6/3.48; IR (Zn/Se ATR, cm-1): 3418 (N-H), 1604 (C=O), 3736 (O-H), 1216 (C-O); 1H-NMR (DMSO-d6), δ: 11.54 (bs, 1H, OH), 10.93 (bs, 1H, NH), 8.50-8.42 (m, 1H, Ar), 8.04 (dd, 1H, J = 7.9 Hz, J = 1.8 Hz, Ar), 7.45-8.39 (m, 1H, Ar), 7.10-7.02 (m, 3H, Ar), 7.01-6.97 (m, 1H, Ar), 6.97-6.92 (m, 1H, Ar), 4.12 (q, 2H, J = 6.9 Hz, CH2), 1.44 (t, 3H, J = 6.9 Hz, CH3); 13C-NMR (DMSO-d6), δ: 163.01, 156.02, 147.65, 133.25, 130.87, 128.32, 123.57, 120.54, 119.70, 119.65, 118.98, 116.85, 111.79, 64.08, 14.75; HR-MS: for C15H16O3N [M+H]+ calculated 258.118 m/z, found 258.0889 m/z.
2-Hydroxy-N-(2-propoxyphenyl)benzamide (5). Yield 31%; m.p. 92-95 °C; HPLC purity 99.98%; UV (nm), λmax/log ε: 299.6/3.39; IR (Zn/Se ATR, cm-1): 3281 (N-H), 1598 (C=O), 1544 (C=C), 1241 (C-O); 1H-NMR (DMSO-d6), δ: 11.49 (bs, 1H, OH), 10.86 (bs, 1H, NH), 8.49-8.44 (m, 1H, Ar), 8.04 (dd, 1H, J = 7.9 Hz, J = 1.8 Hz, Ar), 7.44-7.37 (m, 1H, Ar), 7.11-7.02 (m, 3H, Ar), 7.02-6.96 (m, 1H, Ar), 6.96-6.92 (m, 1H, Ar), 4.02 (t, 2H, J = 6.5 Hz, CH2CH2), 1.84 (q, 2H, J = 7.0 Hz CH2CH2), 1.04 (t, 3H, J = 7.3 Hz, CH3); 13C-NMR (DMSO-d6), δ: 163.07, 156.06, 147.79, 133.25, 130.88, 128.17, 123.64, 120.41, 119.87, 119.64, 118.95, 116.86, 111.60, 69.82, 22.16, 10.51; HR-MS: for C16H18O3N [M+H]+ calculated 272.128 m/z, found 272.1280 m/z.
4-Nitrobenzoic acid (6). 4-Nitrobenzaldehyde (0.15 mol) in acetone (600 mL) and 10% aqueous solution of KMnO4 (400 mL) was refluxed for 90 min. The mixture was then filtered and the resulting colourless solution was concentrated in vacuo to 300 mL. The solution was acidified by 35% HCl. The crude product was filtered, washed by water, recrystallized from aqueous ethanol and dried. Yield: 75%; m.p. 237-240 °C.
4-Nitrobenzoyl chloride (7). Acid 6 (0.05 mol) and thionyl chloride (0.15 mmol) was refluxed for about 1 h. The excess of thionyl chloride was removed by repeated evaporation with dry toluene in vacuo. The crude product was recrystallized form anhydrous petroleum ether. Yield: 100%; m.p. 75 °C.

3.2.2. General procedure for synthesis of carboxamide derivatives 8-10

The acyl chloride 7 (0.014 mol) dissolved in dry acetone (50 mL) was added drop-wise to a stirred solution of the corresponding substituted amine (0.02 mmol) in 50 mL of dry pyridine kept at the room temperature. After the addition was completed, stirring continued for another 1 h. The reaction mixture was then poured into 100 mL of cold water and the crude amide was collected and recrystallized from aqueous ethanol and dried.
N-(2-Chlorophenyl)-4-nitrobenzamide (8). Yield 75%; m.p. 165-167 °C (m.p. 157-158 °C [56]); HPLC purity 99.73%; UV (nm), λmax/log ε: 264.0/3.44; IR (Zn/Se ATR, cm-1): 3299 (N-H), 1653 (C=O), 680 (C-Cl), 1514, 1342 (NO2), 1584 (C=C); 1H-NMR (DMSO-d6), δ: 10.45 (bs, 1H, NH), 8.39 (d, 2H, J = 8.9 Hz, Ar), 8.22 (d, 2H, J = 8.9 Hz, Ar), 7.63-7.55 (m, 2H, Ar), 7.42 (ddd, 1H, J = 7.7 Hz, J = 7.7 Hz, J = 1.5 Hz, Ar), 7.34 (ddd, 1H, J = 7.7 Hz, J = 7.7 Hz, J = 1.7 Hz, Ar); 13C-NMR (DMSO-d6), δ: 163.91, 149.33, 139.57, 134.53, 129.69, 129.62, 129.20, 128.62, 127.94, 127.56, 123.66; HR-MS: for C13H10O3N2Cl [M+H]+ calculated 277.038 m/z, found 277.0354 m/z.
N-(2-Chlorobenzyl)-4-nitrobenzamide (9). Yield 78%; m.p. 178 °C; HPLC purity 99.81%; UV (nm), λmax/log ε: 265.2/3.41; 1H-NMR (DMSO-d6), δ: 9.39 (bt, 1H, J = 5.6 Hz, NH), 8.34 (d, 2H, J = 8.9 Hz, Ar), 8.14 (d, 2H, J = 8.9 Hz, Ar), 7.50-7.43 (m, 1H, Ar), 7.43-7.37 (m, 1H, Ar), 7.37-7.28 (m, 2H, Ar), 4.58 (d, 2H, J = 5.6 Hz, CH2); 13C-NMR (DMSO-d6), δ: 164.82, 149.10, 139.67, 135.82, 132.06, 129.16, 128.87, 128.83, 128.72, 127.20, 123.55, 40.81; HR-MS: for C14H12O3N2Cl [M+H]+ calculated 291.058 m/z, found 291.0530 m/z.
N-[2-Chloro-5-(trifluoromethyl)phenyl]-4-nitrobenzamide (10). Yield 79%; m.p. 182 °C; HPLC purity 99.95%; UV (nm), λmax/log ε: 268.7/3.42; 1H-NMR (DMSO-d6), δ: 10.66 (bs, 1H, NH), 8.40 (d, 2H, J = 8.9 Hz, Ar), 8.22 (d, 2H, J = 8.9 Hz, Ar), 8.05 (d, 1H, J = 2.0 Hz, Ar), 7.85 (d, 1H, J = 8.4 Hz, Ar), 7.70 (dd, 1H, J = 8.4 Hz, J = 2.0 Hz, Ar); 13C-NMR (DMSO-d6), δ: 164.19, 149.47, 139.18, 135.56, 133.76, 130.93, 129.33, 128.17 (q, 1C, 2JFC = 32.8 Hz), 124.97 (q, 1C, 3-1JFC = 3.8 Hz), 124.33 (q, 1C, 3-2JFC = 3.8 Hz), 123.67, 123.51 (q, 1C, 1JFC = 272.5 Hz); HR-MS: for C14H9O3N2ClF3 [M+H]+ calculated 345.028 m/z, found 345.0238 m/z.
4-Amino-N-(2-chlorobenzyl)benzamide (11). Nitrobenzamide 9 (0.01 mol) was dissolved in MeOH (50 mL) and suspension of activated Ra-Ni (3 g) [43] was added. The suspension was mixed and vigorously stirred for 24 h under hydrogen atmosphere from balloon. Then the reaction mixture was filtered and evaporated in vacuo. The crude product was recrystallized from aqueous ethanol and dried. Yield 85%; m.p. 137 °C; HPLC purity 99.51%; UV (nm), λmax/log ε: 279.4/3.38; 1H-NMR (DMSO-d6), δ: 8.56 (bt, 1H, J = 5.8 Hz, NH), 7.64 (d, 2H, J = 8.6 Hz, Ar), 7.47-7.39 (m, 1H, Ar), 7.35-7.22 (m, 3H, Ar), 6.56 (d, 2H, J = 8.6 Hz, -CH=C(NH2)-CH=), 5.64 (bs, 2H, NH2), 4.48 (d, 2H, J = 5.8 Hz, CH2); 13C-NMR (DMSO-d6), δ: 166.39, 151.79, 136.98, 131.74, 128.96, 128.82, 128.35, 128.26, 127.03, 120.67, 112.53, 40.22; HR-MS: for C14H14ON2Cl [M+H]+ calculated 261.078 m/z, found 261.0789 m/z.
4-Amino-N-[2-chloro-5-(trifluoromethyl)phenyl]benzamide (12). Nitrobenzamide 10 (0.01 mol) was dissolved in mixture MeOH:THF 1:1 (50 mL) and suspension of activated Ra-Ni (3 g) [44] was added. Hydrogenation was realized in an autoclave under 65-75 psi for 24 h. Then the reaction mixture was filtered and evaporated in vacuo. The crude product was recrystallized from aqueous ethanol and dried. Yield 80%; m.p. 151 °C; 1H-NMR (DMSO-d6), δ: 10.15 (bs, 1H, NH), 8.09 (d, 1H, J = 2.0 Hz, Ar), 8.07 (d, 1H, J = 8.4 Hz, Ar), 8.05 (dd, 1H, J = 8.4 Hz, J=2.0 Hz, Ar), 7.79 (d, 2H, J = 8.6 Hz, Ar), 6.64 (d, 2H, J=8.6 Hz, -CH=C(NH2)-CH=), 5.86 (bs, 2H, NH2); 13C-NMR (DMSO-d6), δ: 165.66, 152.50, 136.67, 132.25, 130.59, 129.56, 129.35 (q, 1C, 2JFC = 32.04 Hz), 125.34 (q, 1C, 1JFC = 272.53 Hz), 124.76 (d, 1C, 3-1JFC = 3.42 Hz), 123.47 (d, 1C, 3-2JFC = 3.62 Hz), 120.38, 112.58.
Methyl 4-(2-chlorobenzylcarbamoyl)phenylcarbamate (13). Aminobenzamide 11 (0.01 mol) was dissolved in acetone (30 mL) and K2CO3 (0.015 mol) was added. Then methyl chloroformiate (0.015 mol) was added drop-wise. The mixture was refluxed for 24 h. After cooling the solid was filtered and solution was evaporated in vacuo. The crude product was washed by water and crystallized from EtOH. Yield 90%; m.p. 205 °C; HPLC purity 97.24%; UV (nm), λmax/log ε: 264.6/3.39; IR (Zn/Se ATR, cm-1): 3302 (N-H), 3345 (OCON-H), 1704 (OC=O), 1630 (C=O), 1527 (C=C), 1229 (C-O), 733 (C-Cl); 1H-NMR (DMSO-d6), δ: 9.93 (bs, 1H, OCONH), 8.89 (bt, 1H, NH, J = 5.8 Hz), 7.86 (d, 2H, J = 8.7 Hz, Ar), 7.55 (d, 2H, J = 8.7 Hz, Ar), 7.45 (dd, 1H, J = 7.4 Hz, J = 1.6 Hz, Ar), 7.38-7.24 (m, 3H, Ar), 4.53 (d, 2H, J = 5.8 Hz, CH2), 3.69 (s, 3H, CH3); 13C-NMR (DMSO-d6), δ: 165.90, 153.84, 142.02, 136.50, 131.85, 129.04, 128.50, 128.44, 128.23, 127.75, 127.11, 117.22, 51.77, 40.42; HR-MS: for C16H16O3N2Cl [M+H]+ calculated 319.088 m/z, found 319.0844 m/z.
Methyl 4-[2-chloro-5-(trifluoromethyl)phenylcarbamoyl]phenylcarbamate (14). Aminobenzamide 12 (0.01 mol) was dissolved in acetone (30 mL), and K2CO3 (0.015 mol) was added. Then methyl chloroformiate (0.015 mol) was added dropwise. The mixture was refluxed for 24 h. After cooling the solid was filtered and solution was evaporated in vacuo. The crude product was washed by water and crystallized from hexane. Yield 76%; Mp. 173°C; HPLC purity 93.33%; UV (nm), λmax/log ε: 278.2/3.43; IR (Zn/Se ATR, cm-1): 3296 (N-H), 1714 (OC=O), 1653 (C=O), 1532 (C=C), 1257 (C-O), 1171, 1117 (C-F), 662 (C-Cl); 1H-NMR (DMSO-d6), δ: 10.08 (s, 1H, NH), 10.03 (s, 1H, OCONH), 8.08-8.01 (m, 1H, Ar), 7.96 (d, 2H, J = 8.4 Hz, Ar), 7.81 (d, 1H, J = 8.4 Hz, Ar), 7.67-7.62 (m, 1H, Ar), 7.62 (d, 2H, J = 8.8 Hz, Ar), 3.71 (s, 3H, CH3); 13C-NMR (DMSO-d6), δ: 164.93, 153.84, 142.78, 136.18, 133.14, 130.76, 128.87, 128.04 (q, 1C, 2JFC = 32.6 Hz), 126.99, 124.27 (q, 1C, 3-1JFC = 3.8 Hz), 123.63 (q, 1C, 1JFC = 272.48 Hz), 123.49 (q, 1C, 3-2JFC = 3.8 Hz), 117.30, 51.85; HR-MS: for C16H13O3N2ClF3 [M+H]+ calculated 373.738 m/z, found 373.0562 m/z.

3.3. Lipophilicity determination by HPLC (capacity factor k/calculated log k)

A Waters Alliance 2695 XE HPLC separation module and a Waters Photodiode Array Detector 2996 (Waters Corp., Milford, MA, USA) were used. A Symmetry® C18 5 μm, 4.6 × 250 mm, Part No. WAT054275 (Waters Corp., Milford, MA, USA) chromatographic column was used. The HPLC separation process was monitored by Empower™ 2 Chromatography Data Software, Waters 2009 (Waters Corp., Milford, MA, USA). A mixture of MeOH p.a. (70%) and H2O-HPLC – Mili-Q Grade (30%) was used as a mobile phase. The total flow of the column was 1.0 mL/min, injection volume 30 μL, column temperature 30 °C and sample temperature 10°C. The detection wavelength of 210 nm was chosen. The KI methanolic solution was used for the dead time (tD) determination. Retention times (tR) were measured in minutes. The capacity factors k were calculated using the Empower™ 2 Chromatography Data Software according to formula k = (tR - tD)/tD, where tR is the retention time of the solute, whereas tD denotes the dead time obtained using an unretained analyte. Log k, calculated from the capacity factor k, is used as the lipophilicity index converted to log P scale. The log k values of the individual compounds are shown in Table 1.

3.4. Lipophilicity calculations

Log P, i.e. the logarithm of the partition coefficient for n-octanol/water, was calculated using the programs CS ChemOffice Ultra ver. 10.0 (CambridgeSoft, Cambridge, MA, USA) and ACD/LogP ver. 1.0 (Advanced Chemistry Development Inc., Toronto, Canada). Clog P values (the logarithm of n-octanol/water partition coefficient based on established chemical interactions) were generated by means of CS ChemOffice Ultra ver. 10.0 (CambridgeSoft, Cambridge, MA, USA) software. The results are shown in Table 1.

3.5. Study of inhibition photosynthetic electron transport (PET) in spinach chloroplasts

Chloroplasts were prepared from spinach (Spinacia oleracea L.) according to Masarovicova and Kralova [57]. The inhibition of photosynthetic electron transport (PET) in spinach chloroplasts was determined spectrophotometrically (Genesys 6, Thermo Scientific, USA), using an artificial electron acceptor 2,6-dichlorophenol-indophenol (DCIPP) according to Kralova et al. [58], and the rate of photosynthetic electron transport was monitored as a photoreduction of DCPIP. The measurements were carried out in phosphate buffer (0.02 mol/L, pH 7.2) containing sucrose (0.4 mol/L), MgCl2 (0.005 mol/L) and NaCl (0.015 mol/L). The chlorophyll content was 30 mg/L in these experiments and the samples were irradiated (~100 W/m2) from 10 cm distance with a halogen lamp (250 W) using a 4 cm water filter to prevent warming of the samples (suspension temperature 22 °C). The studied compounds were dissolved in DMSO due to their limited water solubility. The applied DMSO concentration (up to 4%) did not affect the photochemical activity in spinach chloroplasts. The inhibitory efficiency of the studied compounds was expressed by IC50 values, i.e. by molar concentration of the compounds causing 50% decrease in the oxygen evolution rate relative to the untreated control. The comparable IC50 value for a selective herbicide 3-(3,4-dichlorophenyl)-1,1-dimethylurea, DCMU (Diurone®) was about 1.9 μmol/L [59]. The results are summarized in Table 2.
EPR spectra were registered by the equipment ERS 230 (ZWG, Acad. Sci., Berlin, Germany), which operates in X-band (~9.3 GHz), with modulation amplitude 0.5 mT and microwave power 5 mW at room temperature. The samples containing chlorophyll (3.2 g/L) were measured in a flat quartz cell and their irradiation (~400 μE/m2s PAR) was carried out directly in the resonance cavity with a 250 W halogen lamp from 0.5 m distance through 5 cm water filter.

3.6. In vitro anti-fungal susceptibility testing

The broth microdilution test [60] was used for the assessment of in vitro anti-fungal activity of the synthesized compounds against Absidia corymbifera 272 (AC) and Trichophyton mentagrophytes 445 (TM). Fluconazole (FLU) was used as the standard since it is a clinically used anti-mycotic drug. The procedure was performed with a two-fold dilution of the compounds in RPMI 1640 (Sevapharma a.s., Prague, Czech Republic) buffered to pH 7.0 with 0.165 mol of 3-morpholino-propane-1-sulfonic acid (MOPS, Sigma, Germany). The final concentrations of the compounds ranged from 500 to 0.975 μmol. Drug–free controls were included. The minimum inhibitory concentration (MIC) determination was performed according to the Clinical and Laboratory Standards Institute (M38-A) for moulds and is an IC50 value. IC50 values were defined as a 50% reduction of growth in comparison with the control. The values of the minimum inhibitory concentration (MICs) were determined after 24 and 48 h of static incubation at 35 °C. For T. mentagrophytes, the final MICs were determined after 72 and 120 h of incubation. The results are summarized in Table 2.

3.7. In vitro anti-bacterial susceptibility testing

The synthesized compounds were evaluated for in vitro anti-bacterial activity against methicilin resistant Staphylococcus aureus H 5996/08 (MRSA), and Staphylococcus epidermidis H 6966/08 (SE). Penicillin G (PEN), and ciprofloxacin (CPF) were used as standards since they are clinically used anti-bacterial drugs. All strains were sub-cultured on nutrient agar (HiMedia) and maintained on the same medium at 4°C. Prior to testing, each strain was passaged onto nutrient agar and bacterial inocula were prepared by suspending a small portion of bacterial colony in sterile 0.85% saline. The cell density was adjusted to 0.5 McFarland units using a densitometer (Densi-La-Meter, PLIVA Lachema Diagnostika). The final inoculum was made by 1:20 dilution of the suspension with the test medium (Mueller-Hinton broth). The compounds were dissolved in DMSO and the anti-bacterial activity was determined using Mueller-Hinton broth (MH broth, HiMedia, pH 7.0 ± 0.2). Controls consisted of MH broth and DMSO alone. The final concentration of DMSO in the MH broth did not exceed 1% (v/v) of the total solution composition. The MIC were defined as 90% inhibition of bacterial growth compared to control and were determined after 24 and 48 h of static incubation at 37 °C. The results are shown in Table 2.

3.8. In vitro anti-mycobacterial evaluation

Clinical isolates of Mycobacterium avium complex CIT19/06 (MAC), M. avium paratuberculosis ATCC19698 (MAP), and M. kansasii CIT11/06 (MK) were grown in Middlebrook broth (MB), supplemented with Oleic, Albumin, Dextrose, Catalase supplement (OADC, Becton Dickinson, UK). Identification of these isolates was performed using biochemical and molecular protocols. At log phase growth, culture (10 mL) was centrifuged at 15,000 rpm/20 min using a bench top centrifuge (Model CR 4-12 Jouan Inc., UK). Following removal of the supernatant, the pellet was washed in fresh Middlebrook 7H9GC broth and re-suspended in fresh supplemented MB (10 mL). The turbidity was adjusted to match McFarland standard No. 1 (3 × 108 cfu) with MB broth. A further 1:20 dilution of the culture was then performed in MB broth.
The anti-microbial susceptibility of all three mycobacterial species was investigated in a 96-well plate format. In these experiments, sterile deionised water (300 µL) was added to all outer-perimeter wells of the plates to minimize evaporation of the medium in the test wells during incubation. Each evaluated compound (100 µL) was incubated with each of the mycobacterial species (100 µL). Dilutions of each compound were prepared in duplicate. For all synthesized compounds, final concentrations ranged from 1,000 µg/mL to 8 µg/mL. All compounds were prepared in DMSO and subsequent dilutions were made in supplemented MB. The plates were sealed with parafilm and incubated at 37°C, for 5 days in the case of M. kansasii and M. avium complex and 7 days in the case of M. avium paratuberculosis. Following incubation, a 10% addition of alamarBlue (AbD Serotec) was mixed into each well and readings at 570 nm and 600 nm were taken, initially for background subtraction and subsequently after 24 h re-incubation. The background subtraction is necessary for strongly coloured compounds, where the colour may interfere with the interpretation of any colour change [61]. For non-interfering compounds, a blue colour in the well was interpreted as an absence of growth and a pink colour was scored as growth. The MIC was initially defined as the lowest concentration which prevented a visual colour change from blue to pink. Isoniazid (INH) was used as the standard as it is a clinically used anti-mycobacterial drug. The results are shown in Table 2. The MIC for mycobacteria was defined as a 90% or greater (IC90) reduction of growth in comparison with the control. The MIC/IC90 value is routinely and widely used in bacterial assays and is a standard detection limit according to the Clinical and Laboratory Standards Institute (CLSI, www.clsi.org/).

3.9. Cell culture and in vitro antiproliferative activity

The human SK-N-MC neuroepithelioma cell type was obtained from the American Type Culture Collection (Manassas, VA, USA) and was cultured in minimum essential medium (MEM; Gibco, Melbourne, Australia) containing 10% (v/v) FBS, 1.0 mM sodium pyruvate (Gibco), 1% (v/v) non-essential amino acids (Gibco), 2 mM L-glutamine (Gibco), 100 U/mL penicillin (Gibco), streptomycin (Gibco) and 0.28 μg/mL fungizone (Squibb Pharmaceuticals, Montreal, Canada). Cells were cultured under standard conditions at 37°C, in a humidified atmosphere at 5% CO2.
The effect of the compounds on cellular proliferation was determined by the MTT [1-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium] assay using standard techniques [53,54]. Previous studies have demonstrated that MTT reduction is proportional to viable cell counts using SK-N-MC cells [62]. The SK-N-MC cells were seeded in 96-well microtiter plates at 1.5 × 104 cells/well in medium containing human diferric transferrin (Tf) at 1.25 μmol/L ([Fe] = 2.5 μmol/L) and compounds at a range of concentrations (0-25 μmol/L). Control samples contained medium with Fe-transferrin (1.25 μmol/L) without the compounds. The chelators, DFO and Dp44mT, were also included as internal controls, as their effects are well characterized in this cell line [53,54]. The cells were incubated at 37°C in a humidified atmosphere containing 5% CO2 and 95% air for 72 h. After 72 h, 10 µL of MTT solution (stock solution: 5 mg/mL) was added to each well and incubated for 2 h at 37°C. After solubilization of the cells with 100 μL of 10% SDS-50% isobutanol in 0.01 M HCl, the plates were read at 570 nm using a scanning multi-well spectrophotometer. The inhibitory concentration (IC50) was defined as the compound concentration necessary to reduce the absorbance to 50% of the untreated control.

4. Conclusions

A series of twelve ring-substituted salicylanilides and carbamoylphenylcarbamates were prepared and characterized. Their lipophilicity was determined using a well established RP-HPLC method. The prepared compounds were tested for their ability to inhibit photosynthetic electron transport (PET) in spinach chloroplasts (Spinacia oleracea L.) and for their anti-fungal, anti-bacterial, anti-mycobacterial and anti-tumor activity. Four compounds, methyl 4-[2-chloro-5-(trifluoromethyl) phenylcarbamoyl]-phenylcarbamate (14), methyl 4-(2-chlorobenzylcarbamoyl)phenylcarbamate (13), 2-hydroxy-N-(2-methoxyphenyl)benzamide (3), and 2-hydroxy-N-(2-methylphenyl)benzamide (1) showed PET inhibition comparable with or higher than the standard DCMU. The site responsible for the inhibitory action of the studied compounds in the photosynthetic apparatus could be QB, the second quinone acceptor situated on the oxidizing (acceptor) site of PS 2.
Interestingly, 2-hydroxy-N-(2-propoxyphenyl)benzamide (5) demonstrated higher anti-fungal activity against Absidia corymbifera than the standard fluconazole and also higher activity against methicillin-resistant Staphylococcus aureus and Staphylococcus epidermidis than penicillin G and ciprofloxacin. Compound 1 exhibited activity against Mycobacterium avium paratuberculosis that was comparable with the standard isoniazid. None of the agents showed effective anti-tumor activity in vitro.

Acknowledgements

This study was supported by the Ministry of Education of the Czech Republic MSM 6215712403, by the Irish Department of Education and Science TSR Strand1-06/CR08 and by Sanofi-Aventis Pharma Slovakia. D.R.R. was supported by a NHMRC Senior Principal Research Fellowship and Project grants and an ARC Discovery Grant. D.S.K. was supported by a Cancer Institute New South Wales Early Career Development Fellowship and Innovation Grant.

References

  1. Vinsova, J.; Imramovsky, A. Salicylanilides: Still a topical potential antibacterially active group. Ces. Slov. Farm. 2004, 53, 294–299. [Google Scholar]
  2. De la Fuente, R.; Sonawane, N.D.; Arumainayagam, D.; Verkman, A.S. Small molecules with antimicrobial activity against E. coli and P. aeruginosa identified by high-throughput screening. Br. J. Pharmacol. 2006, 149, 551–559. [Google Scholar] [CrossRef] [PubMed]
  3. Dahlgren, M.K.; Kauppi, A.M.; Olsson, I.M.; Linusson, A.; Elofsson, M. Design, synthesis, and multivariate quantitative structure–activity relationship of salicylanilidess–potent inhibitors of type III secretion in Yersinia. J. Med. Chem. 2007, 50, 6177–6188. [Google Scholar] [CrossRef] [PubMed]
  4. Stephenson, K.; Yamaguchi, Y.; Hoch, J.A. The mechanism of action of inhibitors of bacterial two-component signal transduction systems. J. Biol. Chem. 2000, 275, 38900–38904. [Google Scholar] [CrossRef] [PubMed]
  5. Vinsova, J.; Imramovsky, A.; Buchta, V.; Ceckova, M.; Dolezal, M.; Staud, F.; Jampilek, J.; Kaustova, J. Salicylanilide acetates: Synthesis and antibacterial evaluation. Molecules 2007, 12, 1–12. [Google Scholar] [CrossRef] [PubMed]
  6. Imramovsky, A.; Vinsova, J.; Ferriz, J.M.; Dolezal, R.; Jampilek, J.; Kaustova, J.; Kunc, F. New antituberculotics originated from salicylanilides with promising in vitro activity against atypical mycobacterial strains. Bioorg. Med. Chem. 2009, 17, 3572–3579. [Google Scholar] [CrossRef] [PubMed]
  7. Imramovsky, A.; Vinsova, J.; Ferriz, J.M.; Buchta, V.; Jampilek, J. Salicylanilide esters of N-protected amino acids as novel antimicrobial agents. Bioorg. Med. Chem. Lett. 2009, 19, 348–351. [Google Scholar] [CrossRef] [PubMed]
  8. Hassan, G.S.; Hegazy, G.H.; Safwat, H.M. Synthesis of Furo-salicylanilides and their heterocyclic derivatives with anticipated molluscicidal activity. Arch. Pharm. Chem. Life Sci. 2006, 339, 448–455. [Google Scholar] [CrossRef] [PubMed]
  9. Daidone, G.; Raffa, D.; Plescia, S.; Matera, M.; Caruso, A.; Leone, V.; Amico-Roxas, M. Synthesis and evaluation of the analgesic and antiinflammatory activities of N-substituted salicylamides. Farmaco 1989, 44, 465–473. [Google Scholar] [PubMed]
  10. Brown, M.E.; Fitzner, J.N.; Stevens, T.; Chin, W.; Wright, C.D.; Boyce, J.P. Salicylanilides: Selective inhibitors of interleukin-12p40 production. Bioorg. Med. Chem. 2008, 16, 8760–8764. [Google Scholar] [CrossRef] [PubMed]
  11. Liechti, C.; Sequin, U.; Bold, G.; Furet, P.; Meyer, T.; Traxler, P. Salicylanilides as inhibitors of the protein tyrosine kinase epidermal growth factor receptor. Eur. J. Med. Chem. 2004, 39, 11–26. [Google Scholar] [CrossRef] [PubMed]
  12. Deng, W.; Guo, Z.; Guo, Y.; Feng, Z.; Jiang, Y.; Chu, F. Acryolylamino-salicylanilides as EGFR PTK inhibitors. Bioorg. Med. Chem. Lett. 2006, 16, 469–472. [Google Scholar] [CrossRef] [PubMed]
  13. Kamath, S.; Buolamwini, J.K. Targeting EGFR and HER-2 receptor tyrosine kinases for cancer drug discovery and development. Med. Res. Rev. 2006, 26, 569–594. [Google Scholar] [CrossRef] [PubMed]
  14. Ray, S.; Pathak, S.R.; Chaturvedi, D. Organic carbamates in drug development. Part II: antimicrobial agents - Recent reports. Drugs Future 2005, 30, 161–180. [Google Scholar] [CrossRef]
  15. Ferriz, J.M.; Vavrova, K.; Kunc, F.; Imramovsky, A.; Stolarikova, J.; Vavrikova, E.; Vinsova, J. Salicylanilide carbamates: Antitubercular agents active against multidrug-resistant Mycobacterium tuberculosis strains. Bioorg. Med. Chem. 2010, 18, 1054–1061. [Google Scholar] [CrossRef] [PubMed]
  16. Agouridas, C.; Denis, A.; Auger, J.M.; Benedetti, Y.; Bonnefoy, A.; Bretin, F.; Chantot, J.F.; Dussarat, A.; Fromentin, C.; D‘Ambrieres, S.G.; Lachaud, S.; Laurin, P.; Le Martret, O.; Loyau, V.; Tessot, N. Synthesis and antibacterial activity of ketolides (6-O-methyl-3-oxoerythromycin derivatives): A new class of antibacterials highly potent against macrolide-resistant and -susceptible respiratory pathogens. J. Med. Chem. 1998, 41, 4080–4100. [Google Scholar] [CrossRef] [PubMed]
  17. Meng, Q.; Luo, H.; Liu, Y.; Li, W.; Zhang, W.; Yao, Q. Synthesis and evaluation of carbamate prodrugs of SQ109 as antituberculosis agents. Bioorg. Med. Chem. 2009, 19, 2808–2810. [Google Scholar] [CrossRef] [PubMed]
  18. Thorberg, S.; Berg, S.; Lundstrom, L.; Pettersson, B.; Wijkstrom, A.; Sanchez, D.; Lindberg, P.; Nilsson, J.G. Carbamate ester derivatives as potential prodrugs of the presynaptic dopamine autoreceptor agonist (-)-3-(3-hydroxyphenyl)-N-propylpiperidine. J. Med. Chem. 1987, 30, 2008–2012. [Google Scholar] [CrossRef] [PubMed]
  19. Good, N.E. Inhibitors of the Hill reaction. Plant Physiol. 1961, 36, 788–803. [Google Scholar] [CrossRef] [PubMed]
  20. Kralova, K.; Sersen, F.; Cizmarik, J. Inhibitory effect of piperidinoethylesters of alkoxyphenylcarbamic acids on photosynthesis. Gen. Physiol. Biophys. 1992, 11, 261–267. [Google Scholar]
  21. Kralova, K.; Sersen, F.; Kubicova, L.; Waisser, K. Inhibitory effects of substituted benzanilides on photosynthetic electron transport in spinach chloroplasts. Chem. Pap. 1999, 53, 328–331. [Google Scholar] [CrossRef]
  22. Kralova, K.; Sersen, F.; Kubicova, L.; Waisser, K. Inhibition of photosynthetic electron transport in spinach chloroplasts by 3- and 4-halogeno substituted benzanilides and thiobenzanilides. J. Trace Microprobe Technol. 2000, 18, 251–256. [Google Scholar]
  23. Kubicova, L.; Kralova, K.; Sersen, F.; Gregor, J.; Waisser, K. Effects of substituted salicylanilides on the photosynthetic apparatus of spinach chloroplasts. Folia Pharm. Univ. Carol. 2000, 25, 89–96. [Google Scholar]
  24. Pravda, M.; Hrnciarova, D.; Kralova, K. 3-Methylthiosalicylanilides – inhibitors of Hill reaction. Chem. Listy 2003, 97, 1122–1123. [Google Scholar]
  25. Black, C.C. Photosynthetic phosphorylation and associated reactions in the presence of a new group of uncouplers: Salicylanilides. Biochim. Biophys. Acta 1968, 162, 294–296. [Google Scholar] [CrossRef]
  26. Govindjee, S. Sixty-three years since Kautsky: Chlorophyll a fluorescence. Aust. J. Plant Physiol. 1995, 22, 131–160. [Google Scholar] [CrossRef]
  27. Jegerschold, C.; Styring, S. Fast oxygen-independent degradation of D1 reaction center protein in photosystem II. FEBS Lett. 1991, 280, 87–90. [Google Scholar] [CrossRef]
  28. Kralova, K.; Kubicova, L.; Sersen, F.; Waisser, K. Inhibition of Hill reaction in spinach chloroplasts by 5-bromo- and 3,5-dibromosalicylanilides. Proceedings of 51st Congress of Chemical Societies, Nitra, Slovakia, 6-9 September 1999. [Google Scholar]
  29. Kubicova, L.; Kissova, K.; Waisser, K. Inhibition of chlorophyll production in Chlorella vulgaris by substituted salicylanilides. Folia Pharm. Univ. Carol. 2000, 25, 67–72. [Google Scholar]
  30. Pravda, M.; Sustr, M.; Hrnciarova, D.; Kubicova, L.; Kralova, K. Effects of 3-methylthiosalicylanilides on chlorophyll content in freshwater alga Chlorella vulgaris. In Proceedings of ECOpole’03, Opole, Poland, 16–18 October 2003; Society of Ecological Chemistry and Engineering: Opole, Poland, 2003; pp. 105–108. [Google Scholar]
  31. Jampilek, J.; Dolezal, M.; Kunes, J.; Buchta, V.; Kralova, K. Quinaldine derivatives: Preparation and biological activity. Med. Chem. 2005, 1, 591–599. [Google Scholar] [CrossRef] [PubMed]
  32. Dolezal, M.; Palek, L.; Vinsova, J.; Buchta, V.; Jampilek, J.; Kralova, K. Substituted pyrazinecarboxamides: Synthesis and biological evaluation. Molecules 2006, 11, 242–256. [Google Scholar] [CrossRef] [PubMed]
  33. Musiol, R.; Jampilek, J.; Kralova, K.; Richardson, D.R.; Kalinowski, D.; Podeszwa, B.; Finster, J.; Niedbala, H.; Palka, A.; Polanski, J. Investigating biological activity spectrum for novel quinoline analogues. Bioorg. Med. Chem. 2007, 15, 1280–1288. [Google Scholar] [CrossRef] [PubMed]
  34. Musiol, R.; Tabak, D.; Niedbala, H.; Podeszwa, B.; Jampilek, J.; Kralova, K.; Dohnal, J.; Finster, J.; Mencel, A.; Polanski, J. Investigating biological activity spectrum for novel quinoline analogues 2: Hydroxyquinolinecarboxamides with photosynthesis inhibiting activity. Bioorg. Med. Chem. 2008, 16, 4490–4499. [Google Scholar] [CrossRef] [PubMed]
  35. Dolezal, M.; Cmedlova, P.; Palek, L.; Vinsova, J.; Kunes, J.; Buchta, V.; Jampilek, J.; Kralova, K. Synthesis and antimycobacterial evaluation of substituted pyrazinecarboxamides. Eur. J. Med. Chem. 2008, 43, 1105–1113. [Google Scholar] [CrossRef] [PubMed]
  36. Jampilek, J.; Musiol, R.; Pesko, M.; Kralova, K.; Vejsova, M.; Carroll, J.; Coffey, A.; Finster, J.; Tabak, D.; Niedbala, H.; Kozik, V.; Polanski, J.; Csollei, J.; Dohnal, J. Ring-substituted 4-hydroxy-1H-quinolin-2-ones: Preparation and biological activity. Molecules 2009, 14, 1145–1159. [Google Scholar] [CrossRef] [PubMed]
  37. Jampilek, J.; Musiol, R.; Finster, J.; Pesko, M.; Carroll, J.; Kralova, K.; Vejsova, M.; O’Mahony, J.; Coffey, A.; Dohnal, J.; Polanski, J. Investigating biological activity spectrum for novel styrylquinazoline analogues. Molecules 2009, 14, 4246–4265. [Google Scholar] [CrossRef] [PubMed]
  38. Musiol, R.; Jampilek, J.; Nycz, J.E.; Pesko, M.; Carroll, J.; Kralova, K.; Vejsova, M.; O’Mahony, J.; Coffey, A.; Mrozek, A.; Polanski, J. Investigating the activity spectrum for ring-substituted 8-hydroxyquinolines. Molecules 2010, 15, 288–304. [Google Scholar] [CrossRef] [PubMed]
  39. Josuu, R.M.; Patel, M.M. Chelation ion-exchange properties of salicylic acid-urea-formaldehyde copolymers. J. Chem. Sci. 1982, 91, 351–358. [Google Scholar]
  40. Mahmoud, M.E.; Soliman, E.M. Study of the selective extraction of iron (III) by silica-immobilized 5-Formyl-3-Arylazo-salicylic acid derivatives. Talanta 1997, 44, 1063–1071. [Google Scholar] [CrossRef]
  41. Rho, H.S.; Baek, H.S.; You, J.W.; Kim, S.J.; Kim, M.K.; Kim, D.H.; Chang, I.S. Biological activities of 3,5-dihydroxy-N-(4-hydroxyphenyl)benzamide: A mimic compound of trans-resveratrol. Bull. Korean Chem. Soc. 2007, 28, 837–839. [Google Scholar]
  42. Althuis, T.H.; Hess, H.J. Synthesis and identification of the major metabolites of prazosin formed in dog and rat. J. Med. Chem. 1977, 20, 146–149. [Google Scholar] [CrossRef] [PubMed]
  43. Brown, R.K.; Nelson, N.A. 6-Aminoindole. J. Am. Chem. Soc. 1954, 76, 5149–5150. [Google Scholar] [CrossRef]
  44. Fellows, I.M.; Kaelin, D.E.; Martin, S.F. Application of ring-closing metathesis to the formal total synthesis of (+)−FR900482. J. Am. Chem. Soc. 2000, 122, 10781–10787. [Google Scholar] [CrossRef]
  45. Hiraj, K.; Yano, T.; Matsukawa, T.; Ugai, S.; Nagato, S.; Hori, M. Synthesis and herbicidal activity of new oxazolidinedione derivates. J. Pestic. Sci. 1999, 24(2), 156–169. [Google Scholar]
  46. Kerns, E.H.; Li, D. Drug-like Properties: Concept, Structure Design and Methods; Elsevier: San Diego, CA, USA, 2008. [Google Scholar]
  47. Norrington, F.E.; Hyde, R.M.; Williams, S.G.; Wotton, R. Physicochemical-activity relations in practice. 1. Rational and self-consistent data bank. J. Med. Chem. 1975, 18, 604–607. [Google Scholar] [CrossRef] [PubMed]
  48. Svensson, B.; Vass, I.; Styring, S. Sequence analysis of D1 and D2 reaction center proteins of photosystem II. Z. Naturforsch C. 1991, 46c, 765–776. [Google Scholar] [CrossRef]
  49. Noren, G.H.; Barry, B.A. The YF161D1 mutant of synechocystis 6803 exhibits an EPR signal from a light-induced photosystem II radical. Biochemistry 1992, 31, 3335–3342. [Google Scholar] [CrossRef] [PubMed]
  50. Hoff, A.J. Application of ESR in photosynthesis. Phys. Rep. 1979, 54, 75–200. [Google Scholar] [CrossRef]
  51. Izawa, S. Acceptors and Donors for Chloroplast Electron Transport; San Pietro, A., Ed.; Academic Press: London, UK, 1980; Volume 69, pp. 413–434. [Google Scholar]
  52. Kalinowski, D.S.; Richardson, D.R. The evolution of iron chelators for the treatment of iron overload disease and cancer. Pharmacol. Rev. 2005, 57, 547–583. [Google Scholar] [CrossRef] [PubMed]
  53. Richardson, D.R.; Sharpe, P.C.; Lovejoy, D.B.; Senaratne, D.; Kalinowski, D.S.; Islam, M.; Bernhardt, P.V. Dipyridyl thiosemicarbazone chelators with potent and selective antitumor activity form iron complexes with redox activity. J. Med. Chem. 2006, 49, 6510–6521. [Google Scholar] [CrossRef] [PubMed]
  54. Kalinowski, D.S.; Yu, Y.; Sharpe, P.C.; Islam, M.; Liao, Y.T.; Lovejoy, D.B.; Kumar, N.; Bernhardt, P.V.; Richardson, D.R. Design, synthesis, and characterization of novel iron chelators: Structure−activity relationships of the 2-benzoylpyridine thiosemicarbazone series and their 3-nitrobenzoyl analogues as potent antitumor agents. J. Med. Chem. 2007, 50, 3716–3729. [Google Scholar] [CrossRef] [PubMed]
  55. Wagner, G.; Singer, D.; Weuffen, W. Untersuchungen uber 2-hydroxythiobenzamide und 2-hydroxythiobenzanilide. Pharmazie 1966, 21, 161–166. [Google Scholar] [PubMed]
  56. Bahrami, K.; Khodaei, M.M.; Farrokhi, A. H2O2/SOCl2: A useful reagent system for the conversion of thiocarbonyls to carbonyl compounds. Tetrahedron 2009, 65, 7658–7661. [Google Scholar] [CrossRef]
  57. Masarovicova, E.; Kralova, K. Approaches to measuring plant photosynthesis activity. In Handbook of Photosynthesis, 2nd ed.; Pessarakli, M., Ed.; Taylor & Francis Group: Boca Raton, FL, USA, 2005; pp. 617–656. [Google Scholar]
  58. Kralova, K.; Sersen, F.; Sidoova, E. Photosynthesis inhibition produced by 2-alkylthio-6-R-benzothiazoles. Chem. Pap. 1992, 46, 348–350. [Google Scholar]
  59. Fedke, C. Biochemistry and Physiology of Herbicide Action; Springer Verlag: New York, NY, USA, 1982. [Google Scholar]
  60. National Committee for Clinical Laboratory Standards. Method for Antifungal Disk Diffusion Susceptibility Testing of Yeasts: Approved Guideline M44-A; National Committee for Clinical Laboratory Standards: Wayne, PA, USA, 2004. [Google Scholar]
  61. Carroll, J.; Douarre, P.; Coffey, A.; Buckley, J.; Cashman, B.; O’Farrell, K.; O’Mahony, J. Optimization of a rapid viability assay for Mycobacterium avium paratuberculosis by using alamarBlue. Appl. Environ. Microbiol. 2009, 75, 7870–7872. [Google Scholar] [CrossRef] [PubMed]
  62. Richardson, D.R.; Tran, E.H.; Ponka, P. The potential of iron chelators of the pyridoxal isonicotinoyl hydrazone class as effective antiproliferative agents. Blood 1995, 86, 4295–4306. [Google Scholar] [PubMed]
Sample Availability: Samples of the compounds are available from the authors.
Scheme 1. Synthesis of the studied compounds.
Scheme 1. Synthesis of the studied compounds.
Molecules 15 08122 sch001
Reagents and conditions: a) acetone, pyridine; b) KOH, HCl; c) KMNO4, acetone, H2O; d), SOCl2, toluene; e) H2, Ra-Ni, MeOH, THF; f) ClCO2Me, acetone, K2CO3.
Figure 1. Comparison of the log P data calculated using the two programs with the experimentally found log k values. The compounds are arranged in the ascending manner according to the experimental log k values.
Figure 1. Comparison of the log P data calculated using the two programs with the experimentally found log k values. The compounds are arranged in the ascending manner according to the experimental log k values.
Molecules 15 08122 g001
Figure 2. EPR spectra of control chloroplasts (A) and chloroplasts treated with 0.05 M of compound 3 (B). Full lines were recorded in darkness. Dotted lines were recorded under irradiation (~400 μE/m2s PAR) directly in the resonance cavity with a 250 W halogen lamp from 0.5 m distance through a 5 cm water filter.
Figure 2. EPR spectra of control chloroplasts (A) and chloroplasts treated with 0.05 M of compound 3 (B). Full lines were recorded in darkness. Dotted lines were recorded under irradiation (~400 μE/m2s PAR) directly in the resonance cavity with a 250 W halogen lamp from 0.5 m distance through a 5 cm water filter.
Molecules 15 08122 g002
Table 1. Comparison of the calculated lipophilicities (log P/Clog P) with the determined log k values, Hammett’s parameter (σ) and bulk parameter (MR, reflecting bulkiness).
Table 1. Comparison of the calculated lipophilicities (log P/Clog P) with the determined log k values, Hammett’s parameter (σ) and bulk parameter (MR, reflecting bulkiness).
Molecules 15 08122 i001
Comp.R1R2log klog P/Clog P
ChemOffice
log P
ACD/LogP
σ [47]MR [47]
12-OH Molecules 15 08122 i0020.66612.94 / 3.12123.73 ± 0.370.104.7
22-OH Molecules 15 08122 i0030.68073.42 / 2.97024.19 ± 0.380.209.4
32-OH Molecules 15 08122 i0040.67702.32 / 2.75763.17 ± 0.390.006.5
42-OH Molecules 15 08122 i0050.75152.66 / 3.28663.70 ± 0.390.0211.3
52-OH Molecules 15 08122 i0060.84013.15 / 3.81564.23 ± 0.39NF15.9
84-NO2 Molecules 15 08122 i0070.67823.15 / 2.77823.19 ± 0.340.674.7
94-NO2 Molecules 15 08122 i0080.69662.82 / 3.71703.19 ± 0.350.674.7
104-NO2 Molecules 15 08122 i0090.82694.13 / 3.77264.94 ± 0.431.108.7
114-NH2 Molecules 15 08122 i0100.58862.66 / 2.63502.47 ± 0.390.674.7
124-NH2 Molecules 15 08122 i011ND3.52 / 3.02994.12 ± 0.421.108.7
134-CH3OCONH Molecules 15 08122 i0120.63452.96 / 3.25103.21 ± 0.430.674.7
144-CH3OCONH Molecules 15 08122 i0130.77823.81 / 3.64594.86 ± 0.501.108.7
NF = not found in literature; ND = not determined/analyzed.
Table 2. IC50 values of compounds 1-11, 13 and 14 related to photosynthetic electron transport (PET) inhibition in spinach chloroplasts in comparison with the 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU) standard; in vitro anti-fungal and anti-bacterial activity (IC50/IC90) of compounds 1-5 compared with the fluconazole (FLU), penicillin G (PEN), ciprofloxacin (CPF) standards and the anti-mycobacterial activity (IC90) of compound 1 in comparison with the standard isoniazid (INH).
Table 2. IC50 values of compounds 1-11, 13 and 14 related to photosynthetic electron transport (PET) inhibition in spinach chloroplasts in comparison with the 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU) standard; in vitro anti-fungal and anti-bacterial activity (IC50/IC90) of compounds 1-5 compared with the fluconazole (FLU), penicillin G (PEN), ciprofloxacin (CPF) standards and the anti-mycobacterial activity (IC90) of compound 1 in comparison with the standard isoniazid (INH).
Comp.PET inhibition
IC50 [μmol/L]
MIC [µmol/L]MIC/IC90 [µg/mL]
ACaTMaMRSAbSEbMACMAPMK
24 h
48 h
72 h
120 h
24 h
48 h
24 h
48 h
12.762.50
250
62.50
125
125
250
125
250
250250125
2331.4>250
>250
250
500
>500
>500
>500
>500
NANANA
31.6125
125
125
125
500
>500
>500
>500
NANANA
415.062.50
125
62.50
62.50
125
125
125
125
NANANA
544.831.25
31.25
31.25
31.25
31.25
31.25
31.25
31.25
NANANA
1050.1NANANANANANANA
11398.4NANANANANANANA
131.6NANANANANANANA
141.0NANANANANANANA
DCMU1.9
FLU>125
>125
1.95
3.91
PEN125
125
31.62
125
CPF500
500
250
250
INH<10>250<10
MIC determination (MIC: minimum inhibitory concentration): amoulds-fungi (IC50 value), bbacteria (IC90 value), NA = no significant activity, AC: Absidia corymbifera, TM: Trichophyton mentagrophytes, MRSA: methicillin-resistant Staphylococcus aureus, SE: Staphylococcus epidermidis, MAC: Mycobacterium avium complex, MAP: M. avium paratuberculosis, MK: M. kansasii).

Share and Cite

MDPI and ACS Style

Otevrel, J.; Mandelova, Z.; Pesko, M.; Guo, J.; Kralova, K.; Sersen, F.; Vejsova, M.; Kalinowski, D.S.; Kovacevic, Z.; Coffey, A.; et al. Investigating the Spectrum of Biological Activity of Ring-Substituted Salicylanilides and Carbamoylphenylcarbamates. Molecules 2010, 15, 8122-8142. https://doi.org/10.3390/molecules15118122

AMA Style

Otevrel J, Mandelova Z, Pesko M, Guo J, Kralova K, Sersen F, Vejsova M, Kalinowski DS, Kovacevic Z, Coffey A, et al. Investigating the Spectrum of Biological Activity of Ring-Substituted Salicylanilides and Carbamoylphenylcarbamates. Molecules. 2010; 15(11):8122-8142. https://doi.org/10.3390/molecules15118122

Chicago/Turabian Style

Otevrel, Jan, Zuzana Mandelova, Matus Pesko, Jiahui Guo, Katarina Kralova, Frantisek Sersen, Marcela Vejsova, Danuta S. Kalinowski, Zaklina Kovacevic, Aidan Coffey, and et al. 2010. "Investigating the Spectrum of Biological Activity of Ring-Substituted Salicylanilides and Carbamoylphenylcarbamates" Molecules 15, no. 11: 8122-8142. https://doi.org/10.3390/molecules15118122

Article Metrics

Back to TopTop