Next Article in Journal
Preliminary Studies of Antimicrobial Activity of New Synthesized Hybrids of 2-Thiohydantoin and 2-Quinolone Derivatives Activated with Blue Light
Next Article in Special Issue
Glycosides of Nadifloxacin—Synthesis and Antibacterial Activities against Methicillin-Resistant Staphylococcus aureus
Previous Article in Journal
Nanosorbents as Materials for Extraction Processes of Environmental Contaminants and Others
Previous Article in Special Issue
Discovery of Novel Pleuromutilin Derivatives as Potent Antibacterial Agents for the Treatment of MRSA Infection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Biological Evaluation and Molecular Docking Studies of 5-Indolylmethylen-4-oxo-2-thioxothiazolidine Derivatives

1
Department of Chemistry, Danylo Halytsky Lviv National Medical University, Pekarska 69, 79010 Lviv, Ukraine
2
Department of Pharmaceutical Chemistry, School of Pharmacy, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
3
InterBioScreen, Chernogolovka, 142432 Moscow Region, Russia
4
Department of Chemistry, Ivan Franko National University of Lviv, Kyryla i Mefodia 6, 79005 Lviv, Ukraine
5
Institute of Biomedical Chemistry, Pogodinskaya Street 10 Bldg.8, 119121 Moscow, Russia
6
Department of Pharmacology, School of Pharmacy, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
7
Department of Life and Health Sciences, University of Nicosia, Nicosia CY-1700, Cyprus
8
Mycological Laboratory, Department of Plant Physiology, Institute for Biological Research, Siniša, Stankovic-National Institute of Republic of Serbia, University of Belgrade, Bulevar Despota Stefana 142, 11000 Belgrade, Serbia
*
Author to whom correspondence should be addressed.
Submission received: 23 December 2021 / Revised: 31 January 2022 / Accepted: 31 January 2022 / Published: 5 February 2022
(This article belongs to the Special Issue Antibacterial Agents 2021)

Abstract

:
Background: Infectious diseases represent a significant global strain on public health security and impact on socio-economic stability all over the world. The increasing resistance to the current antimicrobial treatment has resulted in the crucial need for the discovery and development of novel entities for the infectious treatment with different modes of action that could target both sensitive and resistant strains. Methods: Compounds were synthesized using the classical organic chemistry methods. Prediction of biological activity spectra was carried out using PASS and PASS-based web applications. Pharmacophore modeling in LigandScout software was used for quantitative modeling of the antibacterial activity. Antimicrobial activity was evaluated using the microdilution method. AutoDock 4.2® software was used to elucidate probable bacterial and fungal molecular targets of the studied compounds. Results: All compounds exhibited better antibacterial potency than ampicillin against all bacteria tested. Three compounds were tested against resistant strains MRSA, P. aeruginosa and E. coli and were found to be more potent than MRSA than reference drugs. All compounds demonstrated a higher degree of antifungal activity than the reference drugs bifonazole (6–17-fold) and ketoconazole (13–52-fold). Three of the most active compounds could be considered for further development of the new, more potent antimicrobial agents. Conclusion: Compounds 5b (Z)-3-(3-hydroxyphenyl)-5-((1-methyl-1H-indol-3-yl)methylene)-2-thioxothiazolidin-4-one and 5g (Z)-3-[5-(1H-Indol-3-ylmethylene)-4-oxo-2-thioxo-thiazolidin-3-yl]-benzoic acid as well as 5h (Z)-3-(5-((5-methoxy-1H-indol-3-yl)methylene)-4-oxo-2-thioxothiazolidin-3-yl)benzoic acid can be considered as lead compounds for further development of more potent and safe antibacterial and antifungal agents.

1. Introduction

During the last century, several dozen infections have grown and affected the health of millions of people all over the world [1]. In addition to emerging infections, antimicrobial resistance accounts for at least 50,000 deaths each year in Europe and the United States and it is expected that drug resistant infections will be responsible for even larger losses worldwide in the near future [2,3]. Resistant pathogens threaten patients in medical facilities and may emerge in the general population due to the irrational use of antimicrobial agents [4]. Moreover, microbes can transit to the biofilm-growing form to mitigate the harsh environmental conditions or tolerate the presence of a drug. Existing antimicrobial treatment often fails to prevent or eliminate such biofilms [5,6].
Unfortunately, only few novel classes of antibacterial agents (i.e., oxazolidinones, pleuromutilins, tiacumicins, diarylquinolines lipopeptides and streptogramins) have been marketed in the recent decades to solve these problems. Most of them are for the management of Gram-positive bacterial infections [7,8]. However, drug discovery and development is still one of the major ways to ease the burden of microbial infections. Thus, novel molecules with antimicrobial activity are needed and knowledge on their activity profile including both bacterial and molecular targets and off-targets is needed too to provide the possibility to fix such problems as emerging novel infections, drug resistance and tolerance in a rational way.
Among the natural compounds, there are some indole alkaloids, such as echinulin (1), cristatumin A (2), cristatumin D (3) and tardioxopiperazine A (4) from Eurotium cristatum EN-220 that were able to inhibit the growth of Escherichia coli and S. aureus bacteria [9] (Figure 1).
It is known that rhodanin is one of the most notorious key materials for the development of effective antibiotics. The favorable antimicrobial activity of rhodanines is due to the similarity of their structure with the chemical structure of penicillin, which has been proved by several researchers [10,11,12,13]. Rhodanine-3-alkanecarboxylic acid derivatives with p-N,N-benzylidenedialkyl (phenyl)amine moieties on benzene ring 5 were found to be active against staphylococcus, micrococcus and streptococcus strains [14]. Rhodanines bearing N-arylsulfonylindole fragment derivatives 6 exhibited inhibitory activity against S. aureus including methicillin resistant strains (MRSA) [15], while rhodanine 7 was found to be potent against methicillin-resistant Staphylococcus aureus, Staphylococcus epidermidis, Staphylococcus aureus, Enterococcus sp. and Mycobacteria (Figure 2).
It is also known that synthetic thiohydantoin (rhodanine) analogues 8, 9, in addition to antitumor and anti-HIV activity, exhibit pronounced antibacterial properties [16], along with some hetarylidene thiazolidines containing pyridine 9 [17] and furan 10 [12,18] fragments in the side chain being beta-lactamase inhibitors (Figure 3).
Furthermore, 5-arylidene derivatives of rhodanines were found to possess various types of activity, in particular antitumor [19], antiviral [20,21], anti-inflammatory, antidiabetic [22,23,24], antioxidant [25], LOX and cholinesterase inhibitory activities [26,27], as well as aldose reductase inhibitor activity [28]. There are many references in the literature regarding antimicrobial activity of rhodanine derivatives [14,29,30,31,32].
Prompted by everything mentioned above, as well as based on our previous results [33,34], we designed and synthesized new derivatives incorporating two pharmacophores in the frame of one molecule, indole and thiazolidinone, using a hybridization approach. The aim of this approach is mainly to improve the activity profile and reduce undesired side effects.
As is known, rhodanine derivatives are synthesized by several methods, in particular by dithiocarbamate, bis(carboxymethyl)trithiocarbonate (the Holmberg method), and thiocyanate [35]. The starting 3-arylrhodanines were synthesized using the Holmberg method, making it possible to obtain compounds based on arylamines in high yields and sufficient purity, avoiding the formation of thiourea impurities. To obtain the target products, the interaction of 3-arylrhodanines with aldehydes under the conditions of the Knoevenagel reaction was used [36].

2. Results and Discussion

2.1. Chemistry

The starting materials for the synthesis of the described products were 3-aryl-2-thioxothiazolidin-4-ones 3ae. They were obtained by the reaction of aromatic amines 1ae with bis (carboxymethyl) trithiocarbonate 2 (Scheme 1). In the second step, the 3-aryl-2-thioxothiazolidin-4-ones further undergo Knovenagel condensation with 1H-indole-carbaldehyde to give the title compounds 5al. We used the optimized procedure to do this; the details of the procedure are given in Table 1. Unfortunately, we were not able to obtain 2-hydroxy-4-(4-oxo-2-thioxo-thiazolidin-3-yl)-benzoic acid 3f using 4-amino-2-hydroxybenzoic acid as the starting material 1f, since spontaneous decarboxylation with the formation of 3a occurred under the reaction conditions.
We studied the interaction of 3-aryl-2-thioxothiazolidin-4-ones 3ae with 1H-indole-3-carbaldehydes 4ad. It was found that, when these reagents are boiled in acetic acid in the presence of ammonium acetate, 3-aryl-5-(1H-indol-3-ylmethylene) -2-thioxothiazolidin-4-ones 5al are formed in good yields (Scheme 1).
An analysis of the results of obtaining the initial three and target five substances makes it possible to detect a correlation between the substituents’ nature and the products’ yields. Thus, para-substituents in the aromatic ring in position 3 of thiazolidine have a positive effect on the yields compared to the corresponding meta-substituents. In this case, the nature of the substituent also matters. Hydroxyl substituents have a more favorable effect on yields than carboxyl substituents.
The structures of all synthesized compounds were confirmed by 1H and 13C NMR spectroscopy. In the 1H NMR spectra, signals of all protons were present in regions that correspond to the structure of the obtained compounds. In particular, the signals of the methylene group of 3-aryl-2-thioxothiazolidin-4-ones 3ae are in the range 4.33–4.40 ppm. Signals of the methylidene proton of 3-aryl-5-(1H-indol-3-ylmethylene)-2-thioxothiazolidin-4-ones 5al resonate at 7.99–8.17 ppm, which indicates the Z configuration of the double bond at the 5th position of the thiazolidine ring [37,38]. The signals of hydroxy groups were observed at 9.49–9.87 ppm, while those of the NH group were observed at 11.97–12.37 ppm. The indol ring protons appeared in the aromatic area in the range of 6.73–8.15 ppm. The detailed explanation is given in the experimental part.

2.2. PASS-Based In Silico Assessment of Compounds’ Activity

Activities of twelve chemical structures had been assessed using PASS [39] and PASS-based web applications to predict antifungal, antibacterial, and kinase inhibitory activity [37,38].
PASS predicted some activities associated with antibacterial and antifungal effects. The antibacterial activity itself has been predicted for each of the 12 structures. Pa-Pi values were in the range from 0.003 to 0.577. According to the PASS assessments, the most probable mechanism of antibacterial effect of studied chemical compounds are:
(1)
Inhibition of Enoyl-[acyl-carrier-protein] reductase (predicted for 12 compounds with Pa-Pi values in range from 0.136 to 0.577).
(2)
Inhibition of (R)-Pantolactone dehydrogenase (predicted for five compounds with Pa-Pi values in the range from 0.02 to 0.247).
(3)
Either with lower Pa-Pi values or for the smaller number of compounds, the several other mechanisms were predicted: inhibition of D-Ala-D-Ala ligase and histidine kinase, antagonism with para-aminobenzoic acid, and antagonism with human tumor necrosis factor-alpha.
AntiBac-Pred assessed the probable activity for 12 structures against multiple bacterial strains and species with Pa-Pi values ranging from 0.0001 to 0.387. Overall, some bacteria were predicted as targets for each of the 12 molecules. However, Pa-Pi values were low, only for seven compounds out of 12 did they exceeded 0.3. These assessments were obtained for the activity of compounds against the Bacillus subtilis subsp. subtilis str. 168.
PASS predicted general antifungal activity as well as some putative mechanisms of antifungal action (Heat shock protein 90 antagonist, Kinase inhibitor) for all 12 compounds with Pa-Pi = 0.005 ÷ 0.612. Application of AntiFun-Pred, however, did not allow identification of the specificity of antifungal action against the particular fungi strains.
Among the PASS-predicted mechanisms of antibacterial and antifungal activities, inhibition of some kinases was identified. To estimate the action on the studied compounds on 20 kinases in more detail, we applied the KinScreen web application. As a result, for all twelve compounds, inhibitory activity was predicted against one or more kinases with Pa-Pi > 0.5. For six compounds, inhibition of three kinases (serine/threonine-protein kinase haspin, serine/threonine-protein kinase Nek11, and serine/threonine-protein kinase SRPK1) was estimated with Pa-Pi ≥ 0.7. Therefore, the studied compounds’ antibacterial action may be also due to kinases’ inhibition, including those belonging to the host (human) organism.
Taken together, the results of PASS-based activity evaluation suggest that the studied compounds may exhibit antibacterial and antifungal activities. Relatively low, but not negative, Pa-Pi, values obtained with PASS Online, AntiBac-Pred and AntiFun-Pred indicate that either the studied compounds (1) have a significant structural novelty compared to the compounds from the available training sets or (2) structurally similar compounds may be found among both active and inactive examples in the training set. Experimental studies in antibacterial assays could clarify the selectivity of the compounds’ action on particular bacteria. In turn, docking of the studied chemical structures to the targets could clarify molecular mechanism of their action.

2.3. Pharmacophore Modelling Study

Taking into account the well obtained antibacterial results against S. aureus of the synthesized compounds of our previews work [34], we selected these molecules for the pharmacophore modelling study in order to search for more active compounds. The structures of the training set (19) and test set compounds (1014) are displayed in Table 2. In order to evaluate the common features of these compounds, crucial for the antibacterial activity, the LigandScout program was used.
Conformation generation within 20 kcal/mol energy range were generated and submitted to the alignment procedure. Pharmacophore run resulted in the generation of 10 hypothesis models categorized by their rank score and mapping into all training set molecules (Table 3). From all models, model-1 was selected as the best hypothesis for further analysis based on its highest rank score and mapping (Figure 4). Model-1 contains twelve features, two hydrophobes (H), four hydrogen bond acceptors (HBA), four aromatics (AR) one positive ionizable features (PI) and one hydrogen bond donor (HBD).
All compounds (training and test sets) were mapped onto model-1 with scoring of the orientation of a mapped compound within the hypothesis features using a “fit value” score. As a primary validation of model-1, mapping of all compounds revealed a good correlation between the biological activity and the fit value score (Table 4). The compounds with the higher antibacterial activity showed a range of fit values of 123.59–108.35, while the rest showed a range of fit values of 97.23–85.10. This primary correlation encouraged us to generate a linear model based on “fit value” in order to predict the antibacterial activity of the understudy compounds. This model (Equation (1), Figure 5) showed good statistics with R2 = 0.807. Thus, based on these findings, we use this linear model in order to calculate the activity of the tested compounds (5a5l) (Table 5).
−logMIC(μM) = 22.105 fitvalue–840.56, R = 0.898, R2 = 0.807
n = 14, st error: slope = 3.120, Y-intercept = 327.5, 1/slope = 0.04524
where n is the number of compounds and R is the multiple correlation coefficient.
The best aligned poses of the most active compounds 1 and 2 and the less active 14, superposed with model-1, are presented in Figure 6. As is obvious, some structural features play a key role for the activity. The 1H-indole moiety is thought to be critical for activity. Additionally, the absence of hydrophobic groups can partially explain their lack of activity. The other features that are common for all compounds are three HBA features and the positive ionizable interaction of carboxyl group.
As we mention above, we use this model in order to predict the antibacterial activity of synthesized compounds 5a5l. Results are presented in Table 4 and revealed that compounds 5ac have the highest fit value score and therefore probably the highest activity. Indeed, the experimental data revealed that the predicted by linear model and the experimental MIC of each compound were in the same range (Table 5).

2.4. Biological Evaluation

2.4.1. Antibacterial Action

Compounds 5a5l were tested for antimicrobial activity against the panel of Gram-positive, Gram-negative bacteria as well as eight fungal species.
The determination of minimal inhibitory and minimal bactericidal/fungicidal activity was performed using the microdilution method. The evaluation revealed that all compounds showed antibacterial activity with MIC at 3.78–12.77 µmol/mL × 10−2 and MBC at 4.09–17.03 µmol/mL × 10−2 (Table 6).
The antibacterial potency of tested compounds can be presented as follows: 5b > 5h > 5g > 5c > 5l > 5a > 5j > 5k > 5i > 5d > 5f > 5e. The best activity was observed for compound 5b with MIC and MBC at 3.00–12.28 µmol/mL × 10−2 and 4.09–16.31 µmol/mL × 10−2, respectively, while compound 5e displayed the lowest activity with MIC in the range of 3.12–12.77 µmol/mL × 10−2 and MBC at 4.26–17.03 µmol/mL × 10−2. It should be noticed that bacteria, in general, showed different sensitivity towards the compound tested.
Thus, the sensitivity of S. aureus, the most sensitive bacterium among Gram-positive bacteria, can be presented as: 5b > 5h > 5a > 5c > 5g > 5l > 5j > 5d > 5k > 5f > 5e > 5i. Some similarities in sensitivity with S. aureus was observed for En. cloacae, the most sensitive Gram-negative bacterium: 5b > 5e > 5h > 5a > 5c > 5g > 5l > 5f > 5d > 5j > 5k > 5i. Sensitivity towards compounds of B. cereus, the most resistant Gram-positive bacterium appeared to be completely different: 5g > 5d = 5l > 5h = 5j > 5a = 5k > 5f > 5c > 5i > 5b > 5e, while for the most resistant Gram-negative bacterium, E. coli the sensitivity can be presented as: 5k > 5a > 5d = 5l > 5h = 5j > 5e > 5c = 5f > 5g = 5i > 5b. In this case too some similarities were observed. Thus, both bacteria B. cereus and E. coli exhibited the same sensitivity to compounds 5h and 5j, as well as 5d and 5l.
Compounds 5a5d and 5g exhibited good activity against En. cloacae with MIC values in the range of 3.12–3.94 µmol/mL × 10−2 and MBC at 4.09–7.89 µmol/mL × 10−2. Some of these compounds (5a, 5b, 5g) also showed good activity against P. aeruginosa with MIC and MBC at 3.0–3.94 µmol/mL × 10−2 and 4.26–7.89 µmol/mL × 10−2, respectively, while compound 5b and 5g displayed good activity against S. aureus as well (MIC/MBC at 3–3.94/4.09–7.89 µmol/mL × 10−2).
S. aureus was found to be the most sensitive bacteria among all bacteria involved in the study, whereas E. coli was the most resistant one.
Precisely, for Gram-positive bacteria, the range of MIC and MBC was 3.00–12.77 µmol/mL × 10−2 and 7.57–17.03 µmol/mL × 10−2, respectively, while for Gram negative bacteria the MIC and MBC were in range from 3.78 to 14.07 µmol/mL × 10−2 and 7.57 to 28.14 µmol/mL × 10−2, respectively, indicating that tested compounds are more potent against Gram-positive bacteria than against Gram-negative bacteria.
All compounds were more potent than ampicillin (MIC at 24.8–74.4 µmol/mL × 10−2 and MBC at 37.2–124.0 µmol/mL × 10−2) and almost all compounds were more potent than streptomycin. This is true for almost all bacterial species, except for B. cereus and S. typhimurium.
Four the most active compounds were tested against three resistant strains: methicillin-resistant S. aureus, MRSA, P. aeruginosa and E. coli (Table 7). Two of these strains MRSA and P. aeruginosa belong to ESKAPE pathogens (Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, and Enterobacter species). Tackling the problem of MRSA is a top priority for public health systems worldwide. As far as P. aeruginosa is concerned, the incidences of diseases caused by it are fairly low in the general population, but are higher in hospital inpatients, especially those which are immunocompromised.
All compounds appeared to be more potent against MRSA than ampicillin, displaying better bactericidal activity also against resistant strains P. aeruginosa and E. coli than ampicillin, which did not show any bactericidal effect. These compounds were tested also for their effect on biofilm formation. The antibiofilm effect of selected compounds was less promising compared to their potential to inhibit growth of planktonic bacterial cells. Compounds 5b and 5g had a slight effect on biofilm inhibition, 11.59 and 18.19% inhibition, respectively (Table 7).
The structure–activity relationship revealed that the presence of the methyl group in the indole ring as well as the hydroxy group in position 3 of benzene ring (5b) of (Z)-3-(3-hydroxyphenyl)-5-((1-methyl-1H-indol-3-yl)methylene)-2-thioxothiazolidin-4-one is beneficial for antibacterial activity. Displacement of the methyl group by hydrogen in the indole ring and at the same time introduction of the 5-MeO group to the indole ring from one hand and replacement of the 3-OH group in the benzene ring by 3-COOH gave compound 5h, second in the order of activity. On the other hand, removal of methyl as a substituent on the nitrogen and the 5-OMe group of the indole ring resulted in compound 5g, with decreased activity being the third one in the order of activity of compounds. In general, among 5-OMe- indole derivatives more favorable for activity is the presence of 3-COOH (5h) than the 3-OH group (5c) in benzene ring. On the other hand, shifting of the 3-OH group to position 4 of benzene ring was negative, leading to one of the less potent compounds 5f, whereas removal of the 5-MeO group of compounds 5f appeared to be detrimental, resulting in the less active compound 5e. In case of 6-OMe derivatives, the presence of 4-OH, 3-COOH substitution is (5l) beneficial, while the opposite 3-OH, 4-COOH (5d) was negative. On the other hand, substitution in position 4 of the benzene ring with the hydroxy group resulted in compound (5j), with increased activity compared to 5d. From all these compounds mentioned above, it is clear that the antibacterial activity of designed and synthesized compounds depends not only on substituents and their position in the indole ring but also on substituents and their position in the benzene ring.
All the compounds, except 5g, were additive with streptomycin (FICI 1.5, Table 8), implying that efficient combination with this antibiotic might be further developed. Compound 5g was indifferent in combination with streptomycin (FICI 2).

2.4.2. Efficient P. aeruginosa Bactericidal Effect after 1 h

Application of the selected compounds has significantly reduced the number of viable P. aeruginosa colonies (Figure 7). After 2 h of application, the number of viable P. aeruginosa colonies was reduced for more than 84% (5b and 5h) and more than 74% (5g). All of the examined compounds, after 6h of treatment, reduced the number of P. aeruginosa CFU by more than 98%.
It should be mentioned that even the less active compound 5b appeared to be able to reduce the number of P. aeruginosa CFUs.

2.4.3. Antifungal Action

The ability of compounds to inhibit fungi growth is presented in Table 9 and follows the order 5k > 5l > 5f > 5g > 5a > 5h > 5i > 5c > 5j > 5b > 5d > 5e. The minimum inhibitory concentration of compounds was in the range of 0.97–34.05 mmol/mL × 10−2, while the minimal fungicidal concentration (MFC) varied from 1.95 to 68.1 µmol/mL × 10−2. The best antifungal activity was achieved by compound 5k with MIC and MFC in range of 1.88–3.52 µmol/mL × 10−2 and 3.52–7.03 µmol/mL × 10−2, respectively. The lowest activity was observed for compound 5e with an MIC at 3.12–34.05µmol/mL × 10−2 and MFC at 4.26–68.10µmol/mL × 10−2.
Ketoconazole demonstrated antifungal potency with an MIC in the range of 38–475 µmol/mL × 10−2 and MFC at 95–570 µmol/mL × 10−2, whereas MIC and MFC of bifonazole were at 48–64 mmol/mL × 10−2 and 64–80 µmol/mL × 10−2, respectively.
The sensitivity of fungi towards tested compounds was different. The sensitivity of the most vulnerable fungi T. viride to the compounds can be presented as follows: 5h > 5d > 5a- = 5k > 5f > 5g = 5i > 5b > 5e > 5c > 5l > 5j, while the sensitivity of the most resistant one, A. fumigatus, differs significantly: 5l > 5a = 5k > 5g = 5i > 5b > 5j > 5h > 5c = 5f > 5e. Species belonging to the Aspergillus genus are among the most frequent causes of human fungal infections, associated with significant mortality. A member of this genus, A. fumigatus, is estimated as a cause of 90% of aspergillus infections [40], especially in individuals with immunodeficiency. Compound 5h displayed very good activity against all Aspergilus species, except for A. fumigatus, and Penicillium species, except for P.v.c, with MIC at 1.95 µmol/mL × 10−2 and MFC at 3.55 µmol/mL × 10−2, as well as against T. viride (MIC/MFC at 0.97/1.95 µmol/mL × 10−2). Compound 5f showed good activity against all fungal species except for A. fumigatus with MIC/MFC at 2.09/3.92 µmol/mL × 10−2, while compound 5d showed very good activity against A. ochraceus and T. viride with MIC at 1.88 µmol/mL × 10−2 and MFC at 3.52 µmol/mL × 10−2.
In general, most compounds showed good activity against A. ochraceus and T. viride. Thus, compounds 5a5d, 5f5i and 5k exhibited excellent activity against T. viride with MIC in range of 0.97–2.12 µmol/mL × 10−2, while some of these compounds (5b5d, 5f5i, 5k) showed good activity also against A. ochraceus with MIC in range of 1.88–2.10 µmol/mL × 10−2. Four compounds 5g, 5k, 5l (MIC at 1.88–3.78 µmol/mL × 10−2) as well as 5i (MIC 3.94 µmol/mL × 10−2) showed good activity against the most resistant to compounds tested fungal A. fumigatus.
Our results indicate that the described compounds are 13–52 times more potent than ketoconazole and 6- to 17-fold more potent than bifonazole.
According to the study of structure–activity relationships, it seems that the presence of 3-COOH, 3-OH groups on the benzene ring of indole derivatives (5k) is beneficial for antifungal activity, whereas attachment of the 6-methoxy group to the indole ring of compound 5k results in the less potent derivative 5l.
The order of activity of indole derivatives can be presented as 5k > 5g > 5a > 5i > 5e. Thus, the combination of 3-COOH, 4-OH groups attached to benzene ring (5k) leads to the increase in activity of indole derivatives, while the opposite (5a) is less favorable. Additionally, the 3-COOH group on the benzene ring produced a positive effect on\ antifungal activity of indole derivatives. In the case of 5-methoxy indole derivatives, the 4-OH substitution (5f) was the most favorable, while 3-COOH (5h) and 3-OH substitution (5c) were much less favorable. Concerning the 6-methoxy indole derivatives: the combination of 3-COOH with 4-OH substituents in the benzene ring is favorable, whereas the combination of 3-OH and 4-COOH produced a negative effect on antifungal activity. Thus, the antifungal activity as, in case of antibacterial compounds, depends not only on the nature of substituents, but also on their relative positions.
Our results indicate that antibacterial and antifungal activities have different relationships with the chemical structure: the compound that showed the best antibacterial activity (5b) appeared to be one of the least active as antifungals. On the other hand, compounds 5g, 5d and 5e expressed almost the same behavior against bacteria and fungi.

2.5. Docking Studies

2.5.1. Docking to Antibacterial Targets

In order to elucidate the probable mechanism of action of designed compound docking studies to several antibacterial targets, such as E. coli DNA Gyrase, S. aureus Thymidylate kinase, E. coli Primase and E. coli MurA and MurB, were performed. The docking studies showed that the assessments of binding free energy to E. coli DNA Gyrase, Thymidylate kinase, E. coli Primase and E. coli MurA were higher than that to E. coli MurB; therefore, it may be resolved that the inhibition of E. coli MurB enzyme is the most probable, among the considered mechanism of action of the compounds, where binding scores were consistent with biological activity (Table 10).
The most active compound 5b in E. coli MurB enzyme showed three favorable hydrogen bond interactions. Two hydrogen bonds were observed between the hydroxyl substituent of the benzene ring of the compound and the residues Tyr157 and Lys261 (distance 1.72 Å and 2.58 Å, respectively), and another hydrogen bond interaction between the O atom of the carbonic group of the thiazolidine ring of the compound and Ser228 (distance 2.65 Å). The benzene ring interacts hydrophobically with the residues Tyr124, Gly122, Asn232 and Arg158, while the thizolidine ring interacts hydrophobically with the residues Tyr189 and Leu289 (Figure 8 and Figure 9). These interactions stabilize the complex compound enzyme and play a crucial role in the increased inhibitory action of the compound 5b. Moreover, the hydrogen bond formation with the residue Ser228 is crucial for the inhibitory action of this compound, because this residue takes part in the proton transfer at the second stage of peptidoglycan synthesis [41]. Hydrogen bond interactions with the residue Ser228 were also observed for the rest of the studied compounds.

2.5.2. Docking to Antifungal Targets

All the synthesized compounds and the reference drug ketoconazole were docked to lanosterol 14a-demethylase of C. albicans and DNA topoisomerase IV (Table 11).
Docking results showed that the most active compound 5k take place inside the active site of the enzyme interacting with the heme group of CYP51Ca throughout its –COOH substituent of the benzene ring forming negative ionizable interactions with the heme group. Furthermore, the oxygen of the –OH substituent interacts with the Fe iron of the heme group and with the N atom of Heme, forming a hydrogen bond. Another hydrogen bond is formed between the N atom of indole moiety and Ser378. Hydrophobic interactions were also detected with the residues Thr311, Ley376, Phe233, Phe380 and Met308. Interaction with the heme group was also observed with the benzene ring of ketoconazole, which forms hydrophobic and aromatic interactions (Figure 10 and Figure 11). However, compound 5k forms more and stronger interactions than ketoconazole and a more stable complex of the ligand with the enzyme. This is probably the reason why compound 5k has better antifungal activity than ketoconazole.

2.6. Cytotoxicity Assessment

The assessment of cellular cytotoxicity of the compounds in normal human MRC-5 cells was evaluated at two concentrations in culture, i.e., 1 × 10−5 M (Figure 12A,B) and 1 × 10−6 M (Figure 12B,C). No substantial effect on cell proliferation after 48 h exposure has been observed in cultures, since the growth was ≥80% for all the tested agents compared to control untreated cultures (Figure 12A,C). Moreover, the percentage of dead cells accumulated in cultures was very low, since the maximum number did not exceed that of 2–2.5% (Figure 12B,D).

3. Materials and Methods

3.1. Chemistry

3.1.1. General Procedure for the Synthesis 3-Aryl-2-thioxothiazolidin-4-ones 3ae

The mixture of bis (carboxymethyl) trithiocarbonate, (0.2 mol), of aromatic amine (0.2 mol) and 100 mL of the solvent was boiled for 5–8 h, cooled, the precipitate is filtered off, washed successively with diluted alcohol 1: 2, water, dried and recrystallized.
3-(3-Hydroxyphenyl)-2-thioxothiazolidin-4-one (3a).Yield 70%; m.p. 194–195 °C (AcOH). IR (cm−1): 3362.73 (OH), 1723.31 (C=O), 1596.98 (C=S). 1H-NMR (400 MHz, DMSO-d6, ppm) δ 9.79 (s, 1H, OH), 7.29 (t, J = 8.0 Hz, 1H, C6H4), 6.86 (ddd, J = 8.2, 2.3, 1.0 Hz, 1H, C6H4), 6.67–6.62 (m, 2H, C6H4), 4.36 (s, 2H, CH2). 13C NMR (101 MHz, dmso) δ 203.50, 173.94, 157.94, 136.39, 129.87, 119.04, 116.20, 115.59, 37.01. Anal. Calcd. for C9H7NO2S2 (%): C 47.98; H, 3.13; N, 6.22; S, 28.46. Found (%):C 48.06; H, 3.04; N, 6.31; S, 28.54.
3-(4-Hydroxyphenyl)-2-thioxothiazolidin-4-one (3b). Yield 85%. m.p. 250–252 °C (AcOH). IR (cm−1): 3442.77 (OH), 1742.6 (C=O), 1596.02 (C=S). 1H-NMR (400 MHz, DMSO-d6, ppm) δ 9.82 (s, 1H, OH), 7.05–6.99 (m, 2H, C6H4), 6.88–6.83 (m, 2H, C6H4), 4.34 (s, 2H, CH2). 13C NMR (101 MHz, dmso) δ 203.93, 174.16, 157.90, 129.67, 126.39, 115.69, 36.69. Anal. Calcd. for C9H7NO2S2 (%): C 47.98; H, 3.13; N, 6.22; S, 28.46. Found (%):C 47.89; H, 3.06; N, 6.15; S, 28.41.
3-(4-Oxo-2-thioxothiazolidin-3-yl)benzoic acid (3c). Yield 83%. m.p. 252–254 °C (AcOH-DMFA). IR (cm−1): 1740.67 (C=O), 1688.6 (C=O), 1586.38. (C=S). 1H-NMR (400 MHz, DMSO-d6, ppm) δ 13.24 (s, 1H, COOH), 8.06–8.01 (m, 1H, C6H4), 7.87 (t, J = 1.7 Hz, 1H, C6H4), 7.66 (t, J = 7.8 Hz, 1H, C6H4), 7.54 (ddd, J = 7.9, 2.1, 1.2 Hz, 1H, C6H4), 4.37 (s, 2H, CH2). 13C NMR (101 MHz, dmso) δ 203.72, 174.00, 166.37, 135.84, 133.29, 131.96, 129.97, 129.76, 129.60, 37.28. Anal. Calcd. for C10H7NO3S2 (%): C 47.42; H, 2.79; N, 5.53; S, 25.32. Found (%): C 47.54; H, 2.68; N, 5.41; S, 25.41.
4-(4-Oxo-2-thioxothiazolidin-3-yl)benzoic acid (3d). Yield 89%. m.p. >270 °C (AcOH-DMFA). IR (cm−1): 1722.35 (C=O), 1697.28 (C=O), 1503.44. (C=S). 1H-NMR (400 MHz, DMSO-d6, ppm) δ 13.12 (s, 1H, COOH), 8.10–8.05 (m, 2H, C6H4), 7.45–7.39 (m, 2H, C6H4), 4.40 (s, 1H, CH2). 13C NMR (101 MHz, dmso) δ 203.35, 173.81, 166.55, 139.30, 131.53, 130.16, 129.11, 37.29. 13C NMR (101 MHz, dmso) δ 203.95, 174.06, 171.02, 161.26, 135.86 (s), 130.69, 126.52, 117.93, 113.50, 36.99.Anal. Calcd. for C10H7NO3S2 (%): C 47.42; H, 2.79; N, 5.53; S, 25.32. Found (%): C 47.36; H, 2.71; N, 5.62; S, 25.22.
2-Hydroxy-5-(4-oxo-2-thioxo-thiazolidin-3-yl)-benzoic acid (3e). Yield 61%. m.p. >270 °C (AcOH-DMFA). IR (cm−1): 1748.39 (C=O), 1672.2 (C=O), 1486.08. (C=S). 1H-NMR (400 MHz, DMSO-d6, ppm) δ 7.71 (d, J = 2.6 Hz, 1H, C6H3), 7.39 (dd, J = 8.8, 2.6 Hz, 1H, C6H3), 7.09 (d, J = 8.8 Hz, 1H, C6H3), 4.33 (s, 2H, CH2). Anal. Calcd. for C10H7NO4S2 (%): C 44.60; H, 2.62; N, 5.20; S, 23.81. Found (%): C C 44.73; H, 2.55; N, 5.13; S, 23.90.

3.1.2. General Procedure for the Synthesis 3-Aryl-5-(1H-indol-3-ylmethylene)-2-thioxothiazolidin-4-ones 5al

A mixture of 3-aryl-2-thioxothiazolidin-4-one 3ae (5 mmol), the corresponding 1H-indole-3-carbaldehyde 4ad (6 mmol) and ammonium acetate (5 mmol, 0.39g) in acetic acid (10 mL) was heated to boiling for 2 h, cooled, the precipitate is filtered off, washed successively with acetic acid, alcohol and water, dried and recrystallized with AcOH-DMFA mixture.
(Z)-3-(3-Hydroxyphenyl)-5-(1H-indol-3-ylmethylene)-2-thioxothiazolidin-4-one (5a). Yield 78%; m.p. 271–273 °C. 1H-NMR (400 MHz, DMSO-d6, ppm) δ 12.38 (s, 1H, NH), 9.83 (s, 1H, OH), 8.09 (s, 1H, CH = ), 7.97 (dd, J = 9.3, 5.2 Hz, 2H, H4 +H7, indole), 7.53 (d, J = 7.8 Hz, 1H, H2 indole), 7.37–7.19 (m, 3H, C6H4 + H5 +H6, indole), 6.90 (ddd, J = 8.3, 2.4, 0.9 Hz, 1H, C6H4), 6.82–6.73 (m, 2H, C6H4). 13C-NMR (101 MHz, DMSO-d6, ppm) δ 192.73, 166.58, 157.99, 136.45, 136.39, 130.46, 129.90, 126.75, 125.86, 123.34, 121.48, 119.13, 118.54, 116.27, 115.68, 115.28, 112.55, 111.11. MS (ESI): m/z = 353.0 [M + H]+. Anal. Calcd. for C18H12N2O2S2 (%): C, 61.34; H, 3.43; N, 7.95; S, 18.20. Found (%): C, 61.45; H, 3.36; N, 7.86; S, 18.29.
(Z)-3-(3-Hydroxyphenyl)-5-(1-methyl-1H-indol-3-ylmethylene)-2-thioxothiazolidin-4-one (5b). Yield 92%; m.p. >275 °C. IR (cm−1): 3356.94 (OH), 1707.88 (C=O), 1670.27 (C=O), 1595.05 (C=C), 1572.87 (C=S). 1H-NMR (300 MHz, DMSO-d6, ppm) δ 9.49 (s, 1H, OH), 8.02 (s, 1H, CH = ), 7.89 (d, J = 7.6 Hz, 1H, H4 indole), 7.85 (s, 1H, H7 indole), 7.50 (d, J = 7.9 Hz, 1H, H2 indole), 7.36–7.20 (m, 3H, C6H4 + H5 +H6, indole), 6.94–6.88 (m, 1H, C6H4), 6.72–6.66 (m, 2H, C6H4), 3.99 (s, 3H, CH3N). 13C-NMR (101 MHz, DMSO-d6, ppm) δ 192.66, 166.54, 157.98, 136.92, 136.42, 133.76, 129.92, 127.24, 125.17, 123.39, 121.80, 119.12, 118.62, 116.26, 115.65, 114.98, 110.90, 110.09, 33.38. MS (ESI): m/z = 367.2 [M + H]+. Anal. Calcd. for C19H14N2O2S2 (%): C, 62.27; H, 3.85; N, 7.64; S, 17.50. Found (%):C, 62.34; H, 3.77; N, 7.59; S, 17.41.
(Z)-3-(3-Hydroxyphenyl)-5-(5-methoxy-1H-indol-3-ylmethylene)-2-thioxothiazolidin-4-one (5c). Yield 82%; m.p. 268–270°C. IR (cm−1): 3247.01 (OH), 3263.4 (NH) 1676.06 (C=O), 1591.2 (C=C), 1572.87 (C=S) 1H-NMR (300 MHz, DMSO-d6, ppm) δ 12.08 (s, 1H, NH), 9.55 (s, 1H, OH), 8.07 (s, 1H, CH = ), 7.68 (d, J = 2.4 Hz, 1H, H4 indole), 7.40–7.25 (m, 3H, C6H4 + H2 +H7, indole), 6.89 (d, J = 7.9 Hz, 1H, C6H4), 6.82 (d, J = 9.5 Hz, 1H, H6 indole), 6.70–6.61 (m, 2H, C6H4), 3.86 (s, 3H, CH3O). 13C-NMR (101 MHz, DMSO-d6, ppm) δ 192.68, 166.53, 157.99, 155.27, 136.51, 131.16, 130.50, 129.91, 127.76, 126.45, 119.13, 116.24, 115.68, 114.40, 113.55, 113.33, 111.22, 100.37, 55.48. MS (ESI): m/z = 383.2 [M + H]+. Anal. Calcd. for C19H14N2O3S2 (%): C, 59.67; H, 3.69; N, 7.32; S, 16.77. Found (%): C, 59.51; H, 3.74; N, 7.26; S, 16.85.
(Z)-3-(3-Hydroxyphenyl)-5-(6-methoxy-1H-indol-3-ylmethylene)-2-thioxothiazolidin-4-one (5d). Yield 87%; m.p. 272–274 °C. IR (cm−1): 3343.44 (OH), 3263.4 (NH) 1683.78 (C=O), 1592.16 (C=C), 1570.95 (C=S) 1H-NMR (300 MHz, DMSO-d6, 1H-NMR (300 MHz, DMSO-d6, ppm) δ 11.97 (s, 1H, NH), 9.55 (s, 1H, OH), 7.99 (s, 1H, CH = ), 7.72 (d, J = 8.6 Hz, 1H, H4 indole), 7.61 (s, 1H, H2 indole), 7.29 (t, J = 8.1 Hz, 1H, C6H4), 6.95 (s, 1H, H7 indole), 6.89 (d, J = 8.0 Hz, 1H, C6H4), 6.81 (d, J = 8.2 Hz, 1H, H5 indole), 6.66 (s, 2H, C6H4), 3.84 (s, 3H, CH3O). 13C-NMR (101 MHz, DMSO-d6, ppm) δ 192.66, 166.55, 157.98, 156.84, 137.31, 136.46, 129.90, 129.55, 126.11, 120.68, 119.33, 119.12, 116.26, 115.66, 115.06, 111.44, 111.28, 95.36, 55.27. MS (ESI): m/z = 383.2 [M + H]+. Anal. Calcd. for C19H14N2O3S2 (%): C, 59.67; H, 3.69; N, 7.32; S, 16.77. Found (%): C, 59.75; H, 3.76; N, 7.27; S, 16.64.
(Z)-3-(4-Hydroxyphenyl)-5-(1H-indol-3-ylmethylene)-2-thioxothiazolidin-4-one (5e). Yield 94%; m.p. >275 °C. IR (cm−1): 3382.98 (OH, NH) 1679.92 (C=O), 1592.16 (C=C), 1574.8 (C=S) 1H-NMR (400 MHz, DMSO-d6, ppm) δ 12.37 (s, 1H, NH), 9.86 (s, 1H, OH), 8.09 (s, 1H, CH = ), 8.00–7.92 (m, 2H, H4 +H7, indole), 7.53 (d, J = 8.0 Hz, 1H, H2 indole), 7.32–7.19 (m, 2H, H5 +H6, indole), 7.16 (d, J = 8.8 Hz, 2H, C6H4), 6.89 (d, J = 8.8 Hz, 2H, C6H4). 13C-NMR (101 MHz, DMSO-d6, ppm) δ 193.15, 166.78, 157.98, 136.38, 130.39, 129.76, 126.75, 126.43, 125.77, 123.32, 121.46, 118.51, 115.70, 115.19, 112.54, 111.10. MS (ESI): m/z = 353.2 [M + H]+. Anal. Calcd. for C18H12N2O2S2 (%): C, 61.34; H, 3.43; N, 7.95; S, 18.20. Found (%): C, 61.30; H, 3.29; N, 7.99; S, 18.12.
(Z)-3-(4-Hydroxyphenyl)-5-(5-methoxy-1H-indol-3-ylmethylene)-2-thioxothiazolidin-4-one (5f). Yield 88%; m.p. 264–266 °C. IR (cm−1): 3250.86 (OH, NH) 1669.31(C=O), 1570.95 (C=C), 1510.19 (C=S). 1H-NMR (300 MHz, DMSO-d6, ppm) δ 12.25 (s, 1H, NH), 9.87 (s, 1H, OH), 8.14 (s, 1H, CH = ), 7.87 (d, J = 2.9 Hz, 1H, H4 indole), 7.52 (s, 1H, H2 indole), 7.40 (d, J = 8.8 Hz, 1H, H7 indole), 7.15 (d, J = 8.6 Hz, 2H, C6H4), 6.92–6.86 (m, 3H, C6H4 + H6 indole), 3.83 (s, 3H, CH3O). 13C-NMR (101 MHz, DMSO-d6, ppm) δ 193.08, 166.75, 157.94, 155.25, 131.12, 130.39, 129.74, 127.73, 126.48, 126.33, 115.72, 114.30, 113.51, 113.33, 111.20, 100.29, 55.45. MS (ESI): m/z = 383.2 [M + H]+. Anal. Calcd. for C19H14N2O3S2 (%): C, 59.67; H, 3.69; N, 7.32; S, 16.77. Found (%): C, 59.71; H, 3.61; N, 7.24; S, 16.83.
(Z)-3-[5-(1H-Indol-3-ylmethylene)-4-oxo-2-thioxo-thiazolidin-3-yl]-benzoic acid (5g). Yield 82%; m.p. >275 °C. IR (cm−1): 3230.61 (OH, NH), 1722.35 (C=O), 1693.42 (C=O), 1596.02 (C=C), 1577.7 (C=S).1H-NMR (300 MHz, DMSO-d6, ppm) δ 12.73 (s, COOH), 12.25 (s, NH), 8.15–8.06 (m, 2H, C6H4 + CH = ), 7.96–7.84 (m, 2H, C6H4 + H4 indole), 7.78 (d, J = 2.9 Hz, 1H, H7 indole), 7.66 (t, J = 7.8 Hz, 1H, C6H4), 7.56–7.46 (m, 2H, C6H4 + H2 indole), 7.28–7.14 (m, 2H, H5 +H6, indole). 13C NMR (101 MHz, dmso) δ 192.94, 166.54, 136.41, 135.85, 133.28, 132.18, 130.56, 130.02, 129.83, 129.58, 126.75, 126.03, 123.36, 121.51, 118.52, 115.32, 112.57, 111.09. Anal. Calcd. for C19H12N2O3S2 (%): C, 59.99; H, 3.18; N, 7.36; S, 16.86. Found (%):C, 60.08; H, 3.09; N, 7.45; S, 16.92.
(Z)-3-[5-(5-Methoxy-1H-indol-3-ylmethylene)-4-oxo-2-thioxothiazolidin-3-yl]-benzoic acid (5h). Yield 72%; m.p. >275 °C. IR (cm−1): 3248.93 (OH, NH), 1706.92 (C=O), 1587.34 (C=C), 1575.77 (C=S).1H-NMR (400 MHz, DMSO-d6, ppm) δ δ 13.22 (s, 1H, COOH), 12.29 (s, 1H, NH), 8.17 (s, 1H, CH = ), 8.07 (dd, J = 6.3, 2.5 Hz, 1H, C6H4), 7.98 (s, 1H, H4 indole), 7.90 (d, J = 2.8 Hz, 1H, C6H4), 7.73–7.66 (m, 2H, C6H4 + H7 indole), 7.54 (d, J = 2.0 Hz, 1H, C6H4), 7.40 (t, J = 7.4 Hz, 1H, H2 indole), 6.89 (dd, J = 8.7, 2.3 Hz, 1H, H6 indole), 3.83 (s, 3H, CH3O). 13C-NMR (101 MHz, DMSO-d6, ppm) δ 192.89, 166.49, 155.30, 135.92, 133.36, 132.00, 131.16, 130.57, 130.01, 129.84, 129.62, 127.76, 126.62, 114.44, 113.54, 113.34, 111.21, 100.38, 55.48. MS (ESI): m/z = 411.2 [M + H]+. Anal. Calcd. C20H14N2O4S2 (%): C, 58.52; H 3.44; N, 6.82; S, 15.62. Found (%):C, 58.65; H 3.33; N, 6.76; S, 15.71.
(Z)-4-[5-(1H-Indol-3-ylmethylene)-4-oxo-2-thioxothiazolidin-3-yl]-benzoic acid (5i). Yield 82%; m.p. >275 °C. IR (cm−1) 3433.13 (OH), 3228.68 (NH), 1712.71 (C=O), 1689.56 (C=O), 1596.98 (C=C), 1574.8 (C=S).1H-NMR (300 MHz, DMSO-d6, ppm) δ 12.90 (s, COOH), 12.25 (s, NH), 8.15 (d, J = 8.4 Hz, 2H, C6H4), 8.09 (s, 1H, CH = ), 7.88 (d, J = 7.2 Hz, 1H, H4 indole), 7.78 (d, J = 3.0 Hz, 1H, H7 indole), 7.49 (d, J = 7.3 Hz, 1H, H2 indole), 7.42 (d, J = 8.4 Hz, 2H, C6H4), 7.28–7.13 (m, 2H, H5 +H6, indole). 13C-NMR (101 MHz, DMSO-d6, ppm) δ 192.61, 166.53, 139.38, 136.41, 131.58, 130.65, 130.15, 129.23, 126.76, 126.21, 123.38, 121.54, 118.54, 115.13, 112.58, 111.09. MS (ESI): m/z = 381.2 [M + H]+. Anal. Calcd. C19H12N2O3S2 (%): C, 59.99; H, 3.18; N, 7.36; S, 16.86. Found (%):C, 60.07; H, 3.09; N, 7.28; S, 16.95.
(Z)-4-[5-(6-Methoxy-1H-indol-3-ylmethylene)-4-oxo-2-thioxothiazolidin-3-yl]-benzoic acid (5j). Yield 84%; m.p. >275 °C. IR (cm−1) 3254.72 (OH), 1704.99 (C=O), 1691.49 (C=O), 1596.98 (C=C), 1576.73 (C=S).1H-NMR (300 MHz, DMSO-d6, ppm) δ 12.59 (s, 1H, COOH), 12.04 (s, 1H, NH), 8.14 (d, J = 5.6 Hz, 2H, C6H4), 8.02 (s, 1H, CH = ), 7.73 (d, J = 7.7 Hz, 1H, H4 indole), 7.65 (s, 1H, H2 indole), 7.42 (d, J = 6.6 Hz, 2H, C6H4), 6.95 (s, 1H, H7 indole), 6.82 (d, J = 6.1 Hz, 1H, H5 indole), 3.83 (s, 3H, CH3O). 13C-NMR (101 MHz, DMSO-d6, ppm) δ 192.54 (s), 166.51, 156.87, 139.39, 137.34, 131.54, 130.13, 129.75, 129.22, 126.46, 120.68, 119.35, 114.91, 111.49, 111.27, 95.38, 55.26. MS (ESI): m/z = 411.2 [M + H]+. Anal. Calcd. C20H14N2O4S2 (%): C, 58.52; H 3.44; N, 6.82; S, 15.62. Found (%):C, 58.47; H 3.49; N, 6.79; S, 15.54.
(Z)-2-Hydroxy-5-[5-(1H-indol-3-ylmethylene)-4-oxo-2-thioxothiazolidin-3-yl]-benzoic acid (5k). Yield 93%; m.p. >275 °C. IR (cm−1) 3226.75 (OH, NH), 1718.49 (C=O), 1678.95 (C=O), 1597.95 (C=C), 1577.7 (C=S).1H-NMR (300 MHz, DMSO-d6, ppm) δ 12.24 (s, 1H, NH), 11.54 (s, 1H, COOH), 8.07 (s, 1H, CH = ), 7.87 (d, J = 7.4 Hz, 1H, H4 indole), 7.76 (s, 2H, C6H3 + H7 indole), 7.49 (d, J = 7.5 Hz, 1H, H2 indole), 7.39 (dd, J = 8.8, 2.4 Hz, 1H, C6H3), 7.25–7.16 (m, 2H, H5 +H6, indole), 7.06 (d, J = 8.8 Hz, 1H, C6H3). 13C-NMR (101 MHz, DMSO-d6, ppm) δ 193.19, 171.09, 166.68, 161.33, 136.37, 135.92, 130.76, 130.44, 126.73, 126.51, 125.89, 123.36, 121.51, 118.47, 117.94, 115.28, 113.54, 112.56, 111.07. MS (ESI): m/z = 397.0 [M + H]+. Anal. Calcd. C19H12N2O4S2 (%): C, 57.56; H 3.05; N, 7.07; S, 16.18. Found (%):C, 57.64; H 3.12; N, 7.01; S, 16.25.
(Z)-2-Hydroxy-5-[5-(6-methoxy-1H-indol-3-ylmethylene)-4-oxo-2-thioxothiazolidin-3-yl]-benzoic acid (5l). Yield 72%; m.p. >275 °C. IR (cm−1) 3247.97 (OH, NH), 1717.53 (C=O), 1677.03 (C=O), 1599.88 (C=C), 1576.73 (C=S).1H-NMR (300 MHz, DMSO-d6, ppm) δ 12.00 (s, NH), 11.52 (s, 1H, COOH), 8.00 (s, CH = ), 7.79–7.69 (m, 2H, C6H3, H4 indole), 7.62 (d, J = 2.1 Hz, 1H, H2 indole), 7.37 (dd, J = 8.8, 2.2 Hz, 1H, C6H3), 7.05 (d, J = 8.8 Hz, 1H, C6H3), 6.94 (s, 1H, H7 indole), 6.81 (d, J = 8.7 Hz, 1H, H5 indole), 3.84 (s, 3H, CH3O). 13C-NMR (101 MHz, DMSO-d6, ppm) δ 193.15, 171.08, 166.65, 161.34, 156.84, 137.30, 135.88, 130.74, 129.55, 126.49, 126.12, 120.67, 119.29, 117.91, 115.08, 113.61, 111.46, 111.24, 95.36, 55.26. MS (ESI): m/z = 427.0 [M + H]+. Anal. Calcd. C20H14N2O5S2 (%): C, 56.33; H 3.31; N, 6.57; S, 15.04. Found (%):C, 56.26; H 3.28; N, 6.66; S, 14.95.

3.2. PASS and PASS-Based Web Applications

PASS predictions are based on the structure–activity relationships derived from the data on over eight thousand biological activities of more than one million molecules included in the training set [37,42,43,44]. Structure–activity relationships are examined using MNA (Multilevel Neighborhoods of Atoms) structural descriptors [45] and modified the Naïve Bayes approach [46]. Structural formulae presented as MDL MOL or SDF files [47] are used as input information. PASS output is the list of predicted activities with two assessments: Pa is the estimate of the probability of belonging to the class of active compounds, and Pi is the estimate of the probability of belonging to the class of inactive ones [46]. The higher the probability difference Pa-Pi, the higher the chance to confirm the prediction in the following experiment.
Since 1999 [48], the PASS Online web application has been freely available via the Internet. It provides an opportunity to predict several thousand biological activities with an average probability of about 95% [39]. Comparing PASS Online with some other freely available web services predicting biological activity profiles demonstrated its superiority in performance [49]. Using the special training sets created based on ChEMBL data [50] we developed several specialized PASS-based web applications: AntiBac-Pred [38,51], AntiFun-Pred [52], KinScreen [53,54], which predict the detailed antibacterial, antifungal and kinase inhibitory activity profiles, respectively. These web applications differ from the standard PASS version only by the training sets focused on the particular pharmacotherapeutic fields. Therefore, the interpretation of the prediction results is the same as described above.

3.3. Ligand-Based Pharmacophore Modeling

The LigandScout program (Advanced version 4.4.7) [55] with default settings was used to perform the pharmacophore modeling studies. Prior to the generation of pharmacophore hypotheses, all tested dataset compounds were built using ChemDraw Ultra (CambridgeSoft, version 12.0) and converted into the 3D format. The lowest energy conformations were generated using MMFF94 (Merck molecular force field) and the BEST conformation model generation method was used during conformer generation with a maximum number of 250 conformations, an energy threshold value of 20 kcal/mol above the global energy minimum and an RMS threshold of 0.8.
Compounds 1–9 were selected for training. The training set molecules play a key role in determining the quality of the pharmacophore models generated, while the test set compounds serve to validate the resultant pharmacophore solutions.
Based on the feature mapping results, five matching features were selected, including hydrophobic features (H, yellow), aromatic rings (AR, blue), hydrogen bond donors (HBD, green), hydrogen bond acceptors (HBA, red) and negative ionizable features (red star). The quality of generated hypotheses was ranked based on the pharmacophore fit score, which indicates the modality of the mapping between a molecule and a model. A value of 1 reflects the best prediction [56]. The highest rank score hypotheses for antibacterial activity were considered statistically the best hypotheses and selected for the further analysis.

3.3.1. Pharmacophore Validation

The generated pharmacophore hypothesis was validated using a test set and leave-one-out methods.

Pharmacophore Validation Using Test Set

The test set method is used to clarify whether the generated pharmacophore model is capable to predict the antibacterial activity of compounds other than the training set compounds and categorize them properly in their activity scale. For the test set, compounds 1014 were selected. For the test set compounds, the conformation generation was performed using default values and BEST conformation analysis algorithms [57,58].

Pharmacophore Validation Using Leave-One-Out

The pharmacophore model was cross-validated by the leave-one-out method. In this method, pharmacophore models are recomputing again by leaving one compound at a time from the training set compounds, until each compound was left out once, and its affinity is predicted using that new model [59]. This validation is performed to verify that the correlation of the original pharmacophore model does not depend only on one particular compound [57,60]. By leaving each one of the 9 training set compounds, 9 new models were generated. Thus, we did not obtain any meaningful differences between Model-1 and each model generated from this method, validating our pharmacophore model.

3.4. Biological Evaluation

3.4.1. Antibacterial Action

The following Gram-negative bacteria: Escherichia coli (ATCC 35210), Enterobacter cloacae (clinical isolate), Salmonella Typhimurium (ATCC 13311), Pseudomonas aeruginosa (ATCC 27853) as well as Gram-positive bacteria: Listeria monocytogenes (NCTC 7973), Bacillus cereus (clinical isolate), Micrococcus luteus (ATCC 10240) and Staphylococcus aureus (ATCC 6538) were used. The organisms were obtained from the Mycological Laboratory, Department of Plant Physiology, Institute for Biological Research “Siniša Stankovic”-National Institute of Republic of Serbia, Belgrade, Serbia.
The minimum inhibitory (MIC) and minimum bactericidal (MBC) concentrations were determined by the modified microdilution method, as previously reported [34,60].
Resistant strains used in microdilution assay were isolates of S. aureus (strain isolated from cow), E. coli (strain isolated form pig) and P. aeruginosa (strain isolated from cat) obtained as described in Kartsev et al. [61].

3.4.2. Inhibition of Biofilm Formation

The method was performed as described [62] with some modifications. Briefly, the P. aeruginosa resistant strain was incubated with MIC and subMIC of tested compounds in Triptic soy broth enriched with 2% glucose at 37 °C for 24 h. After 24 h, each well was washed twice with sterile PBS (phosphate buffered saline, pH 7.4) and fixed with methanol for 10 min. Methanol was then removed and the plate was air dried. Biofilm was stained with 0.1% crystal violet (Bio-Merieux, Lyon, France) for 30 min. Wells were washed with water, air dried and 100 μL of 96% ethanol (Zorka, Serbia) was added. The absorbance was read at 620 nm on a Multiskan™ FC Microplate Photometer, Thermo Scientific™. The percentage of inhibition of biofilm formation was calculated by the formula:
[(A620 control − A620 sample)/A620 control] × 100.

3.4.3. Checkboard Assay

It was carried out with 96-well microplates containing TSB medium for the resistant P. aeruginosa strain, supplemented with examined compounds in concentrations ranging from 1/16 to 4 × MIC as described previously [60] in the checkboard manner. The fractional inhibitory concentration index (FICI) was calculated by the following equation as described in our previous paper [63]:
FICI = FIC10/MIC10 + FIC20/MIC20
FIC10 and FIC20 are the MICs of a combination of tested compounds and antibiotics, and MIC10 and MIC20, represent the MIC values of individual agents. The following cut-offs: FIC ≤ 0.5 synergistic, >0.5 <2 additive, ≥2 <4 indifferent, and FIC > 4 antagonistic effects were used for the discussion of obtained results.

3.4.4. Antifungal Activity

For the antifungal bioassays, six fungi were used: Aspergillus niger (ATCC 6275), Aspergillus fumigatus (ATCC 1022), Aspergillus versicolor (ATCC 11730), Aspergillus ochraceus (ATCC 12066), Penicillium funiculosum (ATCC 36839), Trichoderma viride (IAM 5061), Penicillium verrucosum var. cyclopium (food isolate), Penicillium ochrochloron (ATCC 9112). The organisms were obtained from the Mycological Laboratory, Department of Plant Physiology, Institute for Biological Research “Siniša Stankovic”, Belgrade, Serbia [64,65].

3.5. Docking Studies

Τhe AutoDock 4.2® software was used for the docking stimulation. The free energy of binding (ΔG) of E. coli DNA GyrB, Thymidylate kinase, E. coli MurA, E. coli primase, E. coli MurB, DNA topoIV and CYP51 of C. albicans in complex with the inhibitors were generated using this molecular docking program. Regarding the X-ray crystal structures, data of all the enzymes used were obtained from the Protein Data Bank (PDB ID: 1KZN, AQGG, 1DDE, JV4T, 2Q85, 1S16 and 5V5Z, respectively). All procedures were performed according to our previous paper [66].

3.6. Assessment of Cytotoxicity

The normal human lung fibroblast MRC-5 cell line is stored and used in our laboratory in a routine manner (passage < 40). MRC-5 cells were grown in culture (37 °C, humidified atmosphere containing 5% v/v CO2) in DMEM medium supplemented with 10% v/v FBS, 1% PS penicillin–streptomycin). The compounds tested were dissolved in DMSO and stored in 4 °C. For the assessment of cytotoxicity, the cells were seeded in a 96-well plate at an initial concentration of 5 × 104 cells/mL and allowed to attach for at least 3h before the addition of the compounds at two different concentrations: 1 × 10−5 M (10 μΜ) and 1 × 10−6 M (1 μΜ). Note that the concentration of DMSO in culture was ≤0.2% v/v, in which no detectable effect on cell proliferation is observed [67]. To assess the cytotoxicity of each compound, the cells were allowed to grow for additional 48 h before their number is estimated in culture using the Neubauer counting chamber under an optical microscope. Cell growth in each treated culture is expressed as the percentage compared to that seen for the untreated control cells. Moreover, the number of dead cells was also measured using the Trypan-blue method, as previously described [65,66,67]. Statistical t-test analysis was performed via the use of GraphPad Prism 6.0 program.

4. Conclusions

Twelve 3-aryl-5-(1H-indol-3-ylmethylene)-2-thioxothiazolidin-4-ones 5al were designed, synthesized and evaluated in silico and experimentally for their antimicrobial actions against the panel of Gram positive and Gram negative bacteria and fungi.
It should be mentioned that all compounds appeared to be more potent than ampicillin against all bacteria tested and streptomycin against all bacteria except B. cereus, and En. Cloacae. The most sensitive bacteria were found to be S. aureus, while L. monocytogenes was the most resistant one. Compounds also appeared to be active against three resistant strains MRSA, E. coli and P. aeruginosa, showing better activity against MRSA than both reference drugs, while showing better activity against the other two resistant strains than ampicillin.
Concerning antifungal action, the tested compounds exhibited very good activity against all the fungal species tested, being more active than ketoconazole and bifonazole. The most sensitive fungal strain appeared to be T. viride, while the most resistant filamentous A. fumigatus.
It can be observed that the growth of both Gram-negative and Gram-positive bacteria and fungi responded differently to the tested compounds, which indicates that different substituents may lead to different modes of action or that the metabolism of some bacteria/fungi was better able to overcome the effect of the compounds or adapt to it.
Docking analysis to DNA Gyrase, Thymidylate kinase and E. coli MurB indicated a probable involvement of MurB inhibition in the antibacterial mechanism of compounds tested while docking analysis to 14α-lanosterol demethylase (CYP51) and tetrahydrofolate reductase of Candida albicans indicated a probable implication of CYP51 reductase at the antifungal activity of the compounds and secondary involvement of dihydrofolate reductase inhibition at the mechanism of action of the most active compounds.
Finally, compounds 5b (Z)-3-(3-hydroxyphenyl)-5-((1-methyl-1H-indol-3-yl)methylene)-2-thioxothiazolidin-4-one Z)-3-[5-(1H-Indol-3-ylmethylene)-4-oxo-2-thioxo-thiazolidin-3-yl]-benzoic acid as well as 5h (Z)-3-(5-((5-methoxy-1H-indol-3-yl)methylene)-4-oxo-2-thioxothiazolidin-3-yl)benzoic acid can be considered as lead compounds for further development of more potent and safe antibacterial and antifungal agents.

Author Contributions

Conceptualization, A.G. and V.K.; methodology, V.H.; software, A.P. and P.P.; formal analysis, V.M.; investigation, M.I., M.K., M.S., T.A.P. and I.S.V.; data curation, A.G., V.P. and M.S., original draft preparation, A.G. and P.P.; review and editing, A.G. and V.P.; supervision, A.G. and V.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by the Serbian Ministry of Education, Science and Technological Development (Contract No. 451-03-9/2021-14/200007).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

PASS predictions (P.P. and V.P.) were performed in the framework of the Russian Federation Fundamental Research Program for the long-term period for 2021–2030.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mukherjee, S. Emerging Infectious Diseases: Epidemiological Perspective. Indian J. Dermatol. 2017, 62, 459–467. [Google Scholar]
  2. O’Neill, J. Antimicrobial Resistance: Tackling a Crisis for the Health and Wealth of Nations; The Review on Antimicrobial Resistance: London, UK, 2014; pp. 1–16. Available online: https://wellcomecollection.org/works/rdpck35v/items (accessed on 22 December 2021).
  3. Singer, A.C.; Shaw, H.; Rhodes, V.; Hart, A. Review of antimicrobial resistance in the environment and its relevance to environmental regulators. Front. Microbiol. 2016, 7, 1728–1750. [Google Scholar] [CrossRef] [Green Version]
  4. Michael, C.A.; Dominey-Howes, D.; Labbate, M. The antimicrobial resistance crisis: Causes, consequences, and management. Public Health Front. 2014, 16, 145. [Google Scholar] [CrossRef] [PubMed]
  5. Simões, M.; Simoes, L.C.; Vieira, M.J. A review of current and emergent biofilm control strategies. LWT Food Sci. Technol. 2010, 43, 573–583. [Google Scholar] [CrossRef] [Green Version]
  6. Anderson, G.G.; Palermo, J.J.; Schilling, J.D.; Roth, R.; Heuser, J.; Hultgren, S.J. Intracellular biofilm-like pods in urinary tract infections. Science 2003, 301, 105–107. [Google Scholar] [CrossRef] [Green Version]
  7. Brown, E.D.; Wright, G.D. Antibacterial drug discovery in the resistance era. Nature 2016, 529, 336. [Google Scholar] [CrossRef]
  8. Coates, A.R.; Halls, G.; Hu, Y. Novel classes of antibacterial agents or more of the same? Br. J. Pharm. 2011, 163, 184–194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Du, F.-Y.; Li, X.; Li, X.-M.; Zhu, L.-W.; Wang, B.-G. Indolediketopiperazine alkaloids from Eurotium cristatum EN-220, an endophytic fungus isolated from the marine alga Sargassum thunbergii. Mar. Drugs 2017, 15, 24. [Google Scholar] [CrossRef] [Green Version]
  10. Jin, X.; Zheng, C.-J.; Song, M.-X. Synthesis and antimicrobial evaluation of L-phenylalanine-derived C5-substituted rhodamine and chalcone derivatives containing thiobarbituric acid or 2-thi-oxo-4-thiazolidinone. Eur. J. Med. Chem. 2012, 56, 203–209. [Google Scholar] [CrossRef]
  11. Lie, X.; Wakim, B.; Sem, D.S. Chemical proteomics-based drug design: Target and antitarget fishing with a catecholrhodanine privileged scaffold for NAD (P)(H) binding proteins. J. Med. Chem. 2008, 51, 4571–4580. [Google Scholar]
  12. Xiang, Y.; Chen, C.; Wang, W.M.; Xu, L.W.; Yang, K.W.; Oelschlaeger, P.; He, Y. Rhodanine as a Potent Scaffold for the Development of Broad-Spectrum Metallo-β-lactamase Inhibitors. ACS Med. Chem. Lett. 2018, 9, 359–364. [Google Scholar] [CrossRef] [PubMed]
  13. Xu, L.L.; Zheng, C.J.; Sun, L.P.; Miao, J.; Piao, H.R. Synthesis of novel 1,3-diaryl pyrazole derivatives bearing rhodanine-3-fatty acid moieties as potential antibacterial agents. Eur. J. Med. Chem. 2012, 48, 174–178. [Google Scholar] [CrossRef] [PubMed]
  14. Tejchman, W.; Korona-Glowniak, I.; Malm, A. Antibacterial properties of 5-substituted derivatives of rhodanine-3-carbox-yalkyl acids. Med. Chem. Res. 2017, 26, 1316–1324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Song, M.X.; Li, S.H.; Peng, J.Y.; Guo, T.T.; Xu, W.H.; Xiong, S.F.; Deng, X.Q. Synthesis and Bioactivity Evaluation of N-Arylsulfonylindole Analogs Bearing a Rhodanine Moiety as Antibacterial Agents. Molecules 2017, 22, 970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Suzen, S.; Buyukbingol, E. Studies on the synthesis and structure–activity relationships of 5-(3′-indolyl)-2-thiohydantoin derivatives as aldose reductase enzyme inhibitors. Farmaco 1994, 49, 443–447. [Google Scholar]
  17. Trotsko, N.; Kosikowska, U.; Paneth, A.; Plech, T.; Malm, A.; Wujec, M. Synthesis and antibacterial activity of new thiazolidine-2,4-dione-based chlorophenylthiosemicarbazone hybrids. Molecules 2018, 23, 1023. [Google Scholar] [CrossRef] [Green Version]
  18. Grant, E.B.; Guiadeen, D.; Baum, E.Z.; Foleno, B.D.; Jin, H.; Montenegro, D.A.; Nelson, E.A.; Bush, K.; Hlasta, D.J. The synthesis and SAR of rhodanines as novel class C beta-lactamase inhibitors. Bioorg. Med. Chem. Lett. 2000, 19, 2179–2182. [Google Scholar] [CrossRef]
  19. He, X.Y.; Lu, L.; Qiu, J.; Zou, P.; Yu, F.; Jiang, X.K.; Li, L.; Jiang, S.; Liu, S.; Xie, L. Small molecule fusion inhibitors: Design, synthesis and biological evaluation of (Z)-3-(5-(3-benzyl-4-oxo-2-thioxothiazolidinylidene) methyl)-N-(3-carboxy-4-hydroxy) phe-nyl-2, 5-dimethylpyrroles and related derivatives targeting HIV-1gp41. Bioorg. Med. Chem. 2013, 21, 7539–7548. [Google Scholar] [CrossRef]
  20. Yingchoncharoen, P.; Kalinowski, D.S.; Richardson, D.R. Lipid-baseddrug delivery systems in cancer therapy: What is available andwhat is yet to come. Pharmacol. Rev. 2016, 68, 701–787. [Google Scholar] [CrossRef] [Green Version]
  21. Nitsche, C.; Schreier, V.N.; Behnam, M.A.M. Thiazolidinone–peptide hybrids as dengue virus protease inhibi-tors with antiviral activity in cell culture. J. Med. Chem. 2013, 56, 8389–8403. [Google Scholar] [CrossRef]
  22. Bari, S.B.; Firake, S.D. Exploring anti-inflammatory Potential of thiazolidinone derivatives of benzenesulfonamide via synthesis, molecular docking and biological evaluation. Antiinflamm Antiallergy. Agents Med. Chem. 2016, 15, 44–53. [Google Scholar]
  23. Khaled, R.A.; Mohamed, A.A.; Heba, A.H.; Shahinda, S.R. Design, synthesis and biological screening of new 4-thiazolidinone derivatives with promising COX-2 selectivity, anti-inflammatory activity and gastric safety profile. Bioorg. Chem. 2016, 64, 1–12. [Google Scholar]
  24. Yasmin, S.; Capone, F.; Laghezza, A.; Dal Piaz, F.; Loiodice, F.; Vijayan, V.; Devadasan, V.; Mondal, S.; Atlı, O.; Baysal, M.; et al. Novel benzylidene thiazolidinedione derivatives as partial PPARγ agonists and their antidiabetic effects on type 2 diabetes. Sci. Rep. 2017, 7, 14453. [Google Scholar] [CrossRef]
  25. Djukic, M.; Fesatidou, M.; Xenikakis, I.; Geronikaki, A.; Angelova, V.; Savic, V.; Pasic, M.; Krilovic, B.; Djukic, D.; Gobeljic, B.; et al. In vitro antioxidant activity of thiazolidinone derivatives of 1,3-thiazole and 1,3,4-thiadiazole. Chem.-Biol. Interact. 2018, 286, 119–131. [Google Scholar] [CrossRef] [PubMed]
  26. Shafii, N.; Khoobi, M.; Amini, M. Synthesis and biologicalevaluation of 5-benzylidenerhodanine-3-acetic acid derivatives asAChE and 15-LOX inhibitors. J. Enzyme Inhib. Med. Chem. 2015, 30, 389–395. [Google Scholar] [CrossRef] [Green Version]
  27. Kratky, M.; Stepankova, S.; Vorcakova, K.; Vinšová, J. Synthesis and invitro evaluation of novel rhodanine derivatives as potential cho-linesterase inhibitors. Bioorg. Chem. 2016, 68, 23–29. [Google Scholar] [CrossRef] [PubMed]
  28. El-Sayed, S.; Metwally, K.; El-Shanawani, A.A.; Abdel-Aziz, L.M.; El-Rashedy, A.A.; Soliman, M.E.S.; Quattrini, L.; Coviello, V.; la Motta, C. Quinazolinone-based rhodanine-3-acetic acids as potent aldose reductase inhibitors: Synthesis, functional evaluation and molecular modeling study. Bioorg. Med. Chem. Lett. 2017, 27, 4760–4764. [Google Scholar] [CrossRef] [PubMed]
  29. Song, M.X.; Zheng, C.J.; Deng, X.Q.; Wei, Z.-Y.; Piao, H.-R. The synthesis and anti-bacterial activities of N-carboxymethyl rhodanines. Med. Chem. 2014, 4, 441–448. [Google Scholar]
  30. Krátký, M.; Vinšová, J.; Stolaříková, J. Antimicrobial activity of rhodanine-3-acetic acid derivatives. Bioorg. Med. Chem. 2017, 25, 1839–1845. [Google Scholar] [CrossRef]
  31. Zheng, C.J.; Song, M.X.; Sun, L.P.; Wu, Y.; Hong, L.; Piao, H.R. Synthesis and biological evaluation of 5-aryloxypyrazole derivatives bearing a rhodanine-3-aromatic acid as potential antimicrobial agents. Bioorg. Med. Chem. Lett. 2012, 22, 7024–7028. [Google Scholar] [CrossRef]
  32. Zakaria, A.; Osman, M.; Dabboussi, F.; Rafei, R.; Mallat, H.; Papon, N.; Bouchara, J.P.; Hamze, M. Recent trends in the epidemiology, diagnosis, treatment, and mechanisms of resistance in clinical Aspergillus species: A general review with a special focus on the Middle Eastern and North African region. J. Infect. Public Health 2020, 13, 1–10. [Google Scholar] [CrossRef]
  33. Horishny, V.; Kartsev, V.G.; Matiychuk, V.S.; Geronikaki, A.; Petrou, A.; Pogodin, D.; Poroikov, V.; Ivanov, M.; Kostic, M.; Sokovic, M. 3-Amino-5- (indol-3-yl) methylene-4-oxo-2-thioxothiazolidine derivatives as antimicrobial agents: Synthesis, computational and biological evaluation. Pharmaceuticals 2020, 13, 229. [Google Scholar] [CrossRef] [PubMed]
  34. Horishny, V.; Kartsev, V.G.; Matiychuk, V.S.; Geronikaki, A.; Petrou, A.; Glamoclija, J.; Ciric, A.; Sokovic, M. 5-(1H-Indol-3-ylmethylene)-4-oxo-2-thioxothiazolidin-3-yl)alkancarboxylic acids as antimicrobial agents. Synthesis, biological evaluation and molecular docking studies. Molecules 2020, 25, 1964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Buzun, K.; Kryshchyshyn-Dylevych, A.; Senkiv, J.; Roman, O.; Gzella, A.; Bielawski, K.; Bielawska, A.; Lesyk, R. Synthesis and Anticancer Activity Evaluation of 5-[2-Chloro-3- (4-nitrophenyl)-2-propenylidene]-4-thiazolidinones. Molecules 2021, 26, 3057. [Google Scholar] [CrossRef] [PubMed]
  36. Kaminskyy, D.; Kryshchyshyn, A.; Lesyk, R. 5-Ene-4-thiazolidinones–An ecient tool in medicinal chemistry. Eur. J. Med. Chem. 2017, 140, 542–594. [Google Scholar] [CrossRef]
  37. Pogodin, P.V.; Lagunin, A.A.; Rudik, A.V.; Druzhilovskiy, D.S.; Filimonov, D.A.; Poroikov, V.V. AntiBac-Pred: A Web Application for Predicting Antibacterial Activity of Chemical Compounds. J. Chem. Inf. Model. 2019, 59, 4513–4518. [Google Scholar] [CrossRef]
  38. Druzhilovskiy, D.S.; Rudik, A.V.; Filimonov, D.A.; Gloriozova, T.A.; Lagunin, A.A.; Dmitriev, A.V.; Pogodin, P.V.; Dubovskaya, V.I.; Ivanov, S.M.; Tarasova, O.A.; et al. Computational platform Way2Drug: From the prediction of biological activity to drug repurposing. Russ. Chem. Bull. 2017, 66, 1832–1841. [Google Scholar] [CrossRef]
  39. Poroikov, V.V.; Filimonov, D.A.; Gloriozova, T.A.; Lagunin, A.A.; Druzhilovskiy, D.S.; Rudik, A.V.; Stolbov, L.A.; Dmitriev, A.V.; Tarasova, O.A.; Ivanov, S.M.; et al. Computer-aided prediction of biological activity spectra for organic compounds: The possibilities and limitations. Russ. Chem. Bull. 2019, 68, 2143–2154. [Google Scholar] [CrossRef]
  40. Benson, T.E.; Walsh, C.T.; Massey, V. Kinetic characterization of wild-type and S229A mutant MurB: Evidence for the role of Ser 229 as a general acid. Biochemistry 1997, 36, 796–805. [Google Scholar] [CrossRef]
  41. Geronikaki, A.; Fesatidou, M.; Kartsev, V.; Macaev, F. Synthesis and Biological Evaluation of Potent Antifungal Agents. Curr. Top. Med. Chem. 2013, 13, 2684–2733. [Google Scholar] [CrossRef]
  42. Bender, A.; Cortés-Ciriano, I. Artificial intelligence in drug discovery: What is realistic, what are illusions? Part 1: Ways to make an impact, and why we are not there yet. Drug Discov. Today 2021, 26, 511–524. [Google Scholar] [CrossRef] [PubMed]
  43. Filimonov, D.A.; Druzhilovskiy, D.S.; Lagunin, A.A.; Gloriozova, T.A.; Rudik, A.V.; Dmitriev, A.V.; Pogodin, P.V.; Poroikov, V.V. Computer-aided prediction of biological activity spectra for chemical compounds: Opportunities and limitations. Biomed. Chem. Res. Methods 2018, 1, e00004. [Google Scholar] [CrossRef] [Green Version]
  44. Poroikov, V.V. Computer-aided drug design: From discovery of novel pharmaceutical agents to systems pharmacology. Biochem. Suppl. Ser. B Biomed. Chem. 2020, 14, 216–227. [Google Scholar]
  45. Filimonov, D.; Poroikov, V.; Borodina, Y.; Gloriozova, T. Chemical similarity assessment through multilevel neighborhoods of atoms: Definition and comparison with the other descriptors. J. Chem. Inf. Comput. Sci. 1999, 39, 666–670. [Google Scholar] [CrossRef]
  46. Filimonov, D.A.; Lagunin, A.A.; Gloriozova, T.A.; Rudik, A.V.; Druzhilovskii, D.S.; Pogodin, P.V.; Poroikov, V.V. Prediction of the biological activity spectra of organic compounds using the PASS online web resource. Chem. Heterocycl. Compd. 2014, 50, 444–457. [Google Scholar] [CrossRef]
  47. Dalby, A.; Nourse, J.G.; Hounshell, W.D.; Gushurst, A.K.I.; Grier, D.L.; Leland, B.A.; Laufer, J. Description of several chemical structure file formats used by computer programs developed at Molecular Design Limited. J. Chem. Inf. Model. 1992, 32, 244–255. [Google Scholar] [CrossRef]
  48. Lagunin, A.; Stepanchikova, A.; Filimonov, D.; Poroikov, V. PASS: Prediction of activity spectra for biologically active substances. Bioinformatics 2000, 16, 747–748. [Google Scholar] [CrossRef]
  49. Murtazalieva, K.A.; Druzhilovskiy, D.S.; Goel, R.K.; Sastry, G.N.; Poroikov, V.V. How good are publicly available web services that predict bioactivity profiles for drug repurposing? SAR QSAR Environ. Res. 2017, 28, 843–862. [Google Scholar] [CrossRef]
  50. Mendez, D.; Gaulton, A.; Bento, A.P.; Chambers, J.; De Veij, M.; Félix, E.; Magariños, M.P.; Mosquera, J.F.; Mutowo, P.; Nowotka, M.; et al. ChEMBL: Towards direct deposition of bioassay data. Nucleic Acids Res. 2019, 47, D930–D940. [Google Scholar] [CrossRef]
  51. AntiBac-Pred. Available online: http://www.way2drug.com/antibac (accessed on 22 December 2021).
  52. AntiFun-Pred. Available online: http://www.way2drug.com/micF/ (accessed on 22 December 2021).
  53. Pogodin, P.V.; Lagunin, A.A.; Rudik, A.V.; Filimonov, D.A.; Druzhilovskiy, D.S.; Nicklaus, M.C.; Poroikov, V.V. How to achieve better results using PASS-based virtual screening: Case study for kinase inhibitors. Front. Chem. 2018, 6, 133. [Google Scholar] [CrossRef] [Green Version]
  54. KinScreen. Available online: www.way2drug.com/KinScreen (accessed on 22 December 2021).
  55. Wolber, G.; Langer, T. LigandScout: 3-D pharmacophores derived from protein-bound ligands and their use as virtual screening filters. J. Chem. Inf. Model. 2005, 45, 160–169. [Google Scholar] [CrossRef] [PubMed]
  56. Jade, D.D.; Pandey, R.; Kumar, R.; Gupta, D. Ligand-based pharmacophore modeling of TNF-α to design novel inhibitors using virtual screening and molecular dynamics. J. Biomol. Struct. Dyn. 2020, 1, 1–17. [Google Scholar] [CrossRef] [PubMed]
  57. John, S.; Thangapandian, S.; Sakkiah, S. Potent bace-1 inhibitor design using pharmacophore modeling, in silico screening and molecular docking studies. BMC Bioinform. 2011, 12, S28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Friederike, S.; Sven, L.; Thomas, H.; Karsten, S.; Hans-Dieter, H. Pharmacophore definition and three-dimensional quantitative structure–activity relationship study on structurally diverse prostacyclin receptor agonists. Mol. Pharmacol. 2002, 62, 1103–1111. [Google Scholar]
  59. Kostić, M.; Smiljković, M.; Petrović, J.; Glamočilija, J.; Barros, L.; Ferreira, I.C.F.R.; Ćirić, A.; Soković, M. Chemical, nutritive composition and a wide range of bioactive properties of honey mushroom Armillaria mellea (Vahl: Fr.) Kummer. Food Funct. 2017, 8, 3239–3249. [Google Scholar] [CrossRef] [Green Version]
  60. Kartsev, V.; Lichitsky, B.; Geronikaki, A.; Petrou, A.; Smiljkovic, M.; Kostic, M.; Radanovic, O.; Soković, M. Design, synthesis and antimicrobial activity of usnic acid derivatives. MedChemComm 2018, 9, 870–882. [Google Scholar] [CrossRef] [Green Version]
  61. Cady, N.C.; McKean, K.A.; Behnke, J.; Kubec, R.; Mosier, A.P.; Kasper, S.H.; Burz, D.S.; Musah, R.A. Inhibition of biofilm formation, quorum sensing and infection in Pseudomonas aeruginosa by natural products-inspired organosulfur compounds. PLoS ONE 2012, 7, e38492. [Google Scholar] [CrossRef] [Green Version]
  62. Simakov, S.; Kartsev, V.; Petrou, A.; Nicolaou, I.; Geronikaki, A.; Ivanov, M.; Kostic, M.; Glamočlija, J.; Soković, M.; Talea, D.; et al. Vizirianakis 4-(Indol-3-yl)thiazole-2-amines and 4-ιndol-3-yl)thiazoleAcylamines as Novel Antimicrobial Agents: Synthesis, In Silicoand In Vitro Evaluation. Pharmaceuticals 2021, 14, 1096. [Google Scholar] [CrossRef]
  63. Kritsi, E.; Matsoukas, M.T.; Potamitis, C.; Detsi, A.; Ivanov, M.; Sokovic, M.; Zoumpoulakis, P. Novel Hit Compounds as Putative Antifungals: The Case of Aspergillus fumigatus. Molecules 2019, 24, 3853. [Google Scholar] [CrossRef] [Green Version]
  64. Aleksić, M.; Stanisavljević, D.; Smiljković, M.; Vasiljević, P.; Stevanović, M.; Soković, M.; Stojković, D. Pyrimethanil: Between efficient fungicide against Aspergillus rot on cherry tomato and cytotoxic agent on human cell lines. Ann. Appl. Biol. 2019, 175, 228–235. [Google Scholar] [CrossRef]
  65. Petrou, A.; Eleftheriou, P.; Geronikaki, A.; Akrivou, M.G.; Vizirianakis, I.S. Novel Thiazolidin-4-ones as Potential Non-nucleoside Inhibitors of HIV-1 Reverse Transcriptase. Molecules 2019, 24, 3821. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Akrivou, M.G.; Demertzidou, V.P.; Theodoroula, N.F.; Chatzopoulou, F.M.; Kyritsis, K.A.; Grigoriadis, N.; Zografos, A.L.; Vizirianakis, I.S. Uncovering the pharmacological response of novel sesquiterpene derivatives that differentially alter gene expression and modulate the cell cycle in cancer cells. Int. J. Oncol. 2018, 53, 2167–2179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Tseligka, E.D.; Rova, A.; Amanatiadou, E.P.; Calabrese, G.; Tsibouklis, J.; Fatouros, D.G.; Vizirianakis, I.S. Pharmacological Development of Target-Specific Delocalized Lipophilic Cation-Functionalized Carboranes for Cancer Therapy. Pharm. Res. 2016, 33, 1945–1958. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Structure of some indole alkaloids.
Figure 1. Structure of some indole alkaloids.
Molecules 27 01068 g001
Figure 2. Structure of some rhodanine analogues.
Figure 2. Structure of some rhodanine analogues.
Molecules 27 01068 g002
Figure 3. Structure of some synthetic rhodanic analogues.
Figure 3. Structure of some synthetic rhodanic analogues.
Molecules 27 01068 g003
Scheme 1. Synthesis of titled compounds.
Scheme 1. Synthesis of titled compounds.
Molecules 27 01068 sch001
Figure 4. The proposed pharmacophore model of antibacterial activity (in yellow color the hydrophobic (H), red hydrogen bond acceptor (HBA), green hydrogen bond donor (HBD), red star positive ionizable (PI) and blue aromatic (AR) features).
Figure 4. The proposed pharmacophore model of antibacterial activity (in yellow color the hydrophobic (H), red hydrogen bond acceptor (HBA), green hydrogen bond donor (HBD), red star positive ionizable (PI) and blue aromatic (AR) features).
Molecules 27 01068 g004
Figure 5. Negative logarithm of the MIC in μM (dependent value) against the LigandScout program output fit value (independent value).
Figure 5. Negative logarithm of the MIC in μM (dependent value) against the LigandScout program output fit value (independent value).
Molecules 27 01068 g005
Figure 6. (a) Best aligned pose of compound 1 (MIC 0.56 μM × 10−2) superposed with model-1. (b) Best aligned pose of compound 2 (MIC 1.96 μM × 10−2). (c) Best aligned pose of compound 14 (MIC 8.66 μM × 10−2).
Figure 6. (a) Best aligned pose of compound 1 (MIC 0.56 μM × 10−2) superposed with model-1. (b) Best aligned pose of compound 2 (MIC 1.96 μM × 10−2). (c) Best aligned pose of compound 14 (MIC 8.66 μM × 10−2).
Molecules 27 01068 g006
Figure 7. Number of P. aeruginosa CFUs after different time intervals of antimicrobial treatment with the MBC of tested compounds.
Figure 7. Number of P. aeruginosa CFUs after different time intervals of antimicrobial treatment with the MBC of tested compounds.
Molecules 27 01068 g007
Figure 8. (left) Docked conformation of the most active compound 5b in E. coli MurB. (Right) 2D diagrams of the most active compounds 5b (up) and 5h (down) in E. coli MurB.
Figure 8. (left) Docked conformation of the most active compound 5b in E. coli MurB. (Right) 2D diagrams of the most active compounds 5b (up) and 5h (down) in E. coli MurB.
Molecules 27 01068 g008
Figure 9. Docked conformation of compounds 5h (green), 5b (blue) and FAD (yellow) in E. coli MurB.
Figure 9. Docked conformation of compounds 5h (green), 5b (blue) and FAD (yellow) in E. coli MurB.
Molecules 27 01068 g009
Figure 10. Docked conformation of ketoconazole in lanosterol 14alpha-demethylase of C. albicans (CYP51ca).
Figure 10. Docked conformation of ketoconazole in lanosterol 14alpha-demethylase of C. albicans (CYP51ca).
Molecules 27 01068 g010
Figure 11. Docked conformation of compound 5k in lanosterol 14alpha-demethylase of C. albicans (CYP51ca).
Figure 11. Docked conformation of compound 5k in lanosterol 14alpha-demethylase of C. albicans (CYP51ca).
Molecules 27 01068 g011
Figure 12. Assessment of cell proliferation of MRC-5 cells exposed to different compounds in culture. MRC-5 cells grown in culture were separately incubated with each of the compounds at concentrations 1 × 10−5 M (10μΜ) (panels A,B) and 1 × 10−6 M (1 μM) (panels C,D) for 48 h. The cell number was measured in cultures under the microscope using the Neubauer counting chamber, as indicated in “Methods” and expressed as % of the proliferation of control–untreated cultures (panels A,C). Moreover, the evaluation of cell death was assessed by the trypan-blue exclusion-dye assay, as shown in “Methods” (panels B,D). The results shown above indicate the mean numbers ± SD of two independent biological experiments (n ≥ 3). The diagrams shown above and the t-test statistical analysis were carried out using the GraphPad Prism 6.0 program. Notably, no statistical significance between the control–untreated culture with each one of treated compounds was obtained.
Figure 12. Assessment of cell proliferation of MRC-5 cells exposed to different compounds in culture. MRC-5 cells grown in culture were separately incubated with each of the compounds at concentrations 1 × 10−5 M (10μΜ) (panels A,B) and 1 × 10−6 M (1 μM) (panels C,D) for 48 h. The cell number was measured in cultures under the microscope using the Neubauer counting chamber, as indicated in “Methods” and expressed as % of the proliferation of control–untreated cultures (panels A,C). Moreover, the evaluation of cell death was assessed by the trypan-blue exclusion-dye assay, as shown in “Methods” (panels B,D). The results shown above indicate the mean numbers ± SD of two independent biological experiments (n ≥ 3). The diagrams shown above and the t-test statistical analysis were carried out using the GraphPad Prism 6.0 program. Notably, no statistical significance between the control–untreated culture with each one of treated compounds was obtained.
Molecules 27 01068 g012
Table 1. Formation of 3-aryl-2-thioxothiazolidin-4-ones 3ae.
Table 1. Formation of 3-aryl-2-thioxothiazolidin-4-ones 3ae.
NoStarting AminReaction ProductSolventReaction Time (Hours)Yield, %
11a3ai-PrOH:H2O/1:2872
21b3bi-PrOH:H2O/1:1885
31c3ci-PrOH:H2O/1:2883
41d3dH2O589
51e3eDMFA: i-PrOH:H2O/1:1:1561
61f3ai-PrOH:H2O/1:2670
Table 2. Structure of compounds of the training and test sets.
Table 2. Structure of compounds of the training and test sets.
NoStructureNoStructure
1 Molecules 27 01068 i0018 Molecules 27 01068 i002
2 Molecules 27 01068 i0039 Molecules 27 01068 i004
3 Molecules 27 01068 i00510 Molecules 27 01068 i006
4 Molecules 27 01068 i00711 Molecules 27 01068 i008
5 Molecules 27 01068 i00912 Molecules 27 01068 i010
6 Molecules 27 01068 i01113 Molecules 27 01068 i012
7 Molecules 27 01068 i01314 Molecules 27 01068 i014
Table 3. Generated pharmacophores of the antibacterial activity against S. aureus.
Table 3. Generated pharmacophores of the antibacterial activity against S. aureus.
HypothesisFeaturesRank Score
Model-1HHAAAADP0.9345
Model-2HHAAAADP0.9338
Model-3HHAAADP0.9240
Model-4HHAAADP0.9197
Model-5HHAAAAD0.9191
Model-6HHAAADP0.9177
Model-7HHAAADP0.9125
Model-8HHAAAAP0.9035
Model-9HHAAAAP0.8975
Model-10HHAAAP0.8832
Table 4. Correlation between the biological activity and the fit value score against S. aureus.
Table 4. Correlation between the biological activity and the fit value score against S. aureus.
CompoundsMIC (μM × 102)−log MIC (μM)Fit ValuePred. −log MIC (μM) Residual
10.562.251123.591.994−0.257
21.961.707118.601.857−0.150
31.991.701117.831.836−0.135
42.941.531109.001.594−0.063
53.691.432109.001.594−0.162
63.711.430107.821.562−0.132
73.981.400108.251.573−0.173
84.141.38297.191.2700.112
94.161.38097.231.2710.109
102.261.645107.871.5630.082
113.991.399107.421.551−0.212
127.681.11486.950.9900.124
138.611.06485.780.9580.106
148.661.06285.100.9390.123
Table 5. Experimental and estimated MIC values of tested compounds based on pharmacophore model-1 against S. aureus.
Table 5. Experimental and estimated MIC values of tested compounds based on pharmacophore model-1 against S. aureus.
CompoundsFit ValuePred. −log MIC (μM) MIC (μM × 10−2)−log MIC (μM)Residual
5a106.271.5083.781.4220.158
5b105.121.4833.001.522−0.039
5c104.201.4623.921.4060.056
5d87.661.0977.031.153−0.056
5e78.830.9018.511.070−0.169
5f86.251.0657.641.116−0.051
5g97.891.3453.941.404−0.059
5h97.121.3063.551.449−0.143
5i77.700.8767.891.102−0.226
5j88.231.1095.361.270−0.161
5k88.621.1185.571.254−0.136
5l87.491.0935.161.287−0.194
Table 6. Antibacterial activity of title compounds (MIC/MBC in µmol/mL × 10−2).
Table 6. Antibacterial activity of title compounds (MIC/MBC in µmol/mL × 10−2).
R.br B.cM.fS.aL.mEn.clP.aS.tE. coli
5aMIC7.57 ± 0.0611.35 ± 0.13.78 ± 0.027.57 ± 0.13.78 ± 0.023.78 ± 0.0211.35 ± 0.17.57 ± 0.02
MBC15.14 ± 0.115.14 ± 0.27.57 ± 0.0615.14 ± 0.27.57 ± 0.067.57 ± 0.0615.14 ± 0.215.14 ± 0.1
5bMIC8.19 ± 0.048.19 ± 0.043.00 ± 0.016.00 ± 0.043.00 ± 0.016.00 ± 0.026.00 ± 0.0212.28 ± 0.1
MBC16.31 ± 0.116.31 ± 0.14.09 ± 0.028.19 ± 0.14.09 ± 0.028.19 ± 0.048.19 ± 0.0416.31 ± 0.1
5cMIC7.84 ± 0.047.84 ± 0.043.92 ± 0.027.84 ± 0.053.92 ± 0.023.92 ± 0.025.75 ± 0.0411.77 ± 0.2
MBC15.69 ± 0.0815.69 ± 0.087.84 ± 0.0415.69 ± 0.17.84 ± 0.047.84 ± 0.047.84 ± 0.0815.69 ± 0.1
5dMIC7.03 ± 0.0410.55 ± 0.17.03 ± 0.0410.55 ± 0.17.03 ± 0.047.03 ± 0.047.03 ± 0.0410.55 ± 0.2
MBC14.07 ± 0.214.07 ± 0.214.07 ± 0.214.07 ± 0.214.07 ± 0.114.07 ± 0.214.07 ± 0.214.07 ± 0.2
5eMIC8.51 ± 0.0412.77 ± 0.18.51 ± 0.048.51 ± 0.083.12 ± 0.018.51 ± 0.048.51 ± 0.18.51 ± 0.08
MBC17.03 ± 0.0817.03 ± 0.0817.03 ± 0.0817.03 ± 0.084.26 ± 0.0217.03 ± 0.217.03 ± 0.0817.03 ± 0.1
5fMIC7.64 ± 0.0611.77 ± 0.17.64 ± 0.067.64 ± 0.065.75 ± 0.067.64 ± 0.0611.77 ± 0.111.77 ± 0.1
MBC15.69 ± 0.215.69 ± 0.215.69 ± 0.115.69 ± 0.27.64 ± 0.0815.69 ± 0.115.69 ± 0.115.69 ± 0.1
5gMIC3.94 ± 0.027.89 ± 0.063.94 ± 0.027.89 ± 0.023.94 ± 0.043.94 ± 0.027.89 ± 0.0411.83 ± 0.1
MBC7.89 ± 0.0415.77 ± 0.17.89 ± 0.0815.77 ± 0.27.89 ± 0.087.89 ± 0.0815.77 ± 0.215.77 ± 0.2
5hMIC7.31 ± 0.045.36 ± 0.043.55 ± 0.027.31 ± 0.043.55 ± 0.023.55 ± 0.027.31 ± 0.0410.96 ± 0.1
MBC14.62 ± 0.27.31 ± 0.087.31 ± 0.0814.62 ± 0.27.31 ± 0.047.31 ± 0.0414.62 ± 0.214.62 ± 0.2
5iMIC7.89 ± 0.087.89 ± 0.087.89 ± 0.087.89 ± 0.047.89 ± 0.081.97 ± 0.047.89 ± 0.0811.83 ± 0.1
MBC15.77 ± 0.215.77 ± 0.215.73 ± 0.215.77 ± 0.115.77 ± 0.23.94 ± 0.0415.77 ± 0.115.7 ± 0.2
5jMIC7.31 ± 0.085.36 ± 0.045.36 ± 0.047.31 ± 0.047.31 ± 0.087.31 ± 0.087.31 ± 0.0810.96 ± 0.1
MBC14.62 ± 0.27.31 ± 0.087.31 ± 0.0814.62 ± 0.0814.62 ± 0.114.62 ± 0.114.62 ± 0214.62 ± 0.2
5kMIC7.57 ± 0.085.55 ± 0.045.57 ± 0.087.57 ± 0.087.57 ± 0.047.57 ± 0.047.57 ± 0.083.78 ± 0.04
MBC15.14 ± 0.27.57 ± 0.0815.14 ± 0.215.14 ± 0.115.14 ± 0.115.14 ± 0.215.14 ± 0.27.57 ± 0.08
5lMIC7.03 ± 0.0510.55 ± 0.15.16 ± 0.047.03 ± 0.085.16 ± 0.045.16 ± 0.047.03 ± 0.0810.55 ± 0.1
MBC14.07 ± 0.114.07 ± 0.17.03 ± 0.0814.07 ± 0.27.03 ± 0.087.03 ± 0.0814.07 ± 0.0214.07 ± 0.02
AmMIC 24.80 ± 0.324.80 ± 0.224.80 ± 0.237.20 ± 0.424.8 ± 0.374.4 ± 0.924.80 ± 0.337.2 ± 0.4
MBC37.20 ± 0.437.20 ± 0.437.20 ± 0.274.40 ± 0.837.2 ± 0.3124.0 ± 249.2 ± 0.649.2 ± 0.4
Str MIC 4.30 ± 0.088.60 ± 0.117.20 ± 0.225.80 ± 0.44.3 ± 0.0317.20 ± 0.317.2 ± 0.317.2 ± 0.3
MBC8.60 ± 0.117.20 ± 0.234.40 ± 0.451.60 ± 0.48.6 ± 0.0634.40 ± 0.334.4 ± 0.334.40 ± 0.3
B.c.B. cereus, M.f.M. flavus, S.a.S. aurues, l.m.L. monocytogenes, En.c.En. cloacae, P.a.P. aeruginosa, S.t.S. typhimurium, E.c.E. coli.
Table 7. Antibacterial activity against resistant strains (µmol/mL × 10−2) and the effect on biofilm formation (inhibition percentage), * NE—no effect.
Table 7. Antibacterial activity against resistant strains (µmol/mL × 10−2) and the effect on biofilm formation (inhibition percentage), * NE—no effect.
CompoundsResistant StrainsBiofilm Formation
MRSAP.a.E.c.MIC0.5MIC
5bMIC136 ± 1634 ± 2.7136 ± 14.211.59NE
MBC272 ± 3368 ± 5.2272 ± 31
5hMIC122 ± 16.430.5 ± 2.8122 ± 15.1NENE
MBC244 ± 2861 ± 7244 ± 22.8
5gMIC31.5 ± 2.731.5 ± 3.8126 ± 14.718.19NE
MBC63 ± 8.163 ± 4252 ± 28
StreptomycinMIC17.2 ± 2.18.6 ± 1.217.2 ± 2.171.9455.42
MBC-17.2 ± 1.434.4 ± 4.2
AmpicillineMIC-57.2 ± 6.457.2 ± 7.867.3630.35
MBC///
E.c.E. coli, P.a.P. aeruginosa.
Table 8. FICI indexes of combinations of selected compounds with streptomycin.
Table 8. FICI indexes of combinations of selected compounds with streptomycin.
CompoundFICI
5b1.5
5g2
5h1.5
Table 9. Antifungal activity of title compounds (MIC/MFC in µmol/mL × 10−2).
Table 9. Antifungal activity of title compounds (MIC/MFC in µmol/mL × 10−2).
R.br A.fA.vA.oA.nT.vP.oP.fP.v.c
5aMIC3.78 ± 0.043.78 ± 0.043.78 ± 0.043.78 ± 0.042.02 ± 0.043.78 ± 0.043.78 ± 0.083.78 ± 0.04
MFC7.57 ± 0.087.57 ± 0.087.57 ± 0.087.57 ± 0.083.78 ± 0.047.57 ± 0.17.57 ± 0.17.57 ± 0.08
5bMIC8.19 ± 0.048.19 ± 0.083.00 ± 0.026.00 ± 0.083.00 ± 0.026.00 ± 0.086.00 ± 0.0812.28 ± 0.1
MFC16.31 ± 0.216.31 ± 0.24.09 ± 0.038.19 ± 0.084.09 ± 0.038.19 ± 0.088.19 ± 0.0816.31 ± 0.12
5cMIC15.69 ± 0.23.92 ± 0.062.09 ± 0.025.75 ± 0.042.88 ± 0.043.92 ± 0.083.92 ± 0.043.92 ± 0.08
MFC31.37 ± 0.47.84 ± 0.063.92 ± 0.037.84 ± 0.087.84 ± 0.087.84 ± 0.087.84 ± 0.17.84 ± 0.1
5dMIC29.31 ± 0.43.52 ± 0.041.88 ± 0.023.52 ± 0.041.88 ± 0.023.52 ± 0.023.52 ± 0.043.52 ± 0.04
MFC56.28 ± 1.17.03 ± 0.063.52 ± 0.027.03 ± 0.063.52 ± 0.047.03 ± 0.067.03 ± 0.067.03 ± 0.06
5eMIC34.05 ± 0.64.26 ± 0.083.12 ± 0.044.26 ± 0.053.12 ± 0.044.26 ± 0.086.24 ± 0.088.51 ± 0.08
MFC68.10 ± 0.88.51 ± 0.064.26 ± 0.058.51 ± 0.14.26 ± 0.058.51 ± 0.18.51 ± 0.0817.03 ± 0.2
5fMIC15.69 ± 0.12.09 ± 0.042.09 ± 0.022.09 ± 0.042.09 ± 0.022.09 ± 0.022.09 ± 0.022.09 ± 0.04
MFC31.37 ± 0.43.92 ± 0.063.92 ± 0.043.92 ± 0.043.92 ± 0.063.92 ± 0.043.92 ± 0.043.92 ± 0.06
5gMIC3.94 ± 0.033.94 ± 0.042.10 ± 0.013.94 ± 0.062.10 ± 0.023.94 ± 0.063.94 ± 0.062.63 ± 0.02
MFC7.89 ± 0.087.89 ± 0.063.94 ± 0.067.89 ± 0.063.94 ± 0.067.89 ± 0.087.89 ± 0.067.89 ± 0.1
5hMIC14.62 ± 0.21.95 ± 0.011.95 ± 0.011.95 ± 0.020.97 ± 0.021.95 ± 0.021.95 ± 0.027.31 ± 0.1
MFC29.24 ± 0.43.55 ± 0.023.55 ± 0.043.55 ± 0.041.95 ± 0.043.55 ± 0.043.55 ± 0.0614.62 ± 0.2
5iMIC3.94 ± 0.063.94 ± 0.062.10 ± 0.043.94 ± 0.082.10 ± 0.045.78 ± 0.083.94 ± 0.067.89 ± 0.06
MFC7.89 ± 0.087.89 ± 0.063.94 ± 0.067.89 ± 0.083.94 ± 0.067.89 ± 0.087.89 ± 0.0615.77 ± 0.2
5jMIC14.62 ± 0.23.55 ± 0.063.55 ± 0.063.55 ± 0.045.36 ± 0.063.55 ± 0.053.55 ± 0.067.31 ± 0.06
MFC29.23 ± 0.47.31 ± 0.17.31 ± 0.17.31 ± 0.087.31 ± 0.17.31 ± 0.087.31 ± 0.0814.62 ± 0.09
5kMIC3.78 ± 0.042.02 ± 0.042.02 ± 0.042.02 ± 0.042.02 ± 0.023.78 ± 0.023.78 ± 0.043.78 ± 0.04
MFC7.57 ± 0.083.78 ± 0.043.78 ± 0.043.78 ± 0.043.78 ± 0.087.57 ± 0.087.57 ± 0.17.57 ± 0.08
5lMIC2.58 ± 0.023.52 ± 0.022.58 ± 0.025.16 ± 0.063.52 ± 0.065.16 ± 0.063.52 ± 0.053.52 ± 0.04
MFC3.52 ± 0.047.03 ± 0.053.52 ± 0.057.03 ± 0.063.52 ± 0.047.03 ± 0.087.03 ± 0.087.03 ± 0.08
KetoconazoleMIC38.0 ± 1.2285. ± 6.838.0 ± 1.238.0 ± 0.8475.0 ± 5.8380.0 ± 5.838.0 ± 1.238.00 ± 1.6
MFC95.0 ± 2.3380. ± 8.495.0 ± 1.295.0 ± 0.6570.0 ± 8.6380.0 ± 4.895.0 ± 2.395 ± 2.6
BifonazoleMIC48.0 ± 2.248.0 ± 1.248.0 ± 2.848.0 ± 1.264.0 ± 2.848.0 ± 2.064.0 ± 1.248 ± 2.2
MFC64.0 ± 3.464.0 ± 0.880.0 ± 1.864.0 ± 2.380.0 ± 3.864.0 ± 1.680.0 ± 2.164 ± 3.4
A.f.A. fumigatus, A.v.A. versicolor, A.o.A. ochraceus, A.n.A. niger, T.v.T. viride, P.f.P. funiculosum, P.o.P. ochrochloron, P.v.c.P. cyclpoium var verucosum.
Table 10. Molecular docking binding affinities to antibacterial targets.
Table 10. Molecular docking binding affinities to antibacterial targets.
NoEst. Binding Energy (kcal/mol)I-H
E. coli MurB
Residues
E. coli MurB
E. coli Gyrase
1KZN
S. aureus Thymidylate Kinase
4QGG
E. coli Primase
1DDE
E. coli MurA JV4TE. coli MurB 2Q85
5a−4.52−1.91−6.17−8.592Arg213, Ser228
5b−6.25−2.48−1.26−8.11−12.573Tyr157, Ser228, Lys261
5c−4.27−5.28−9.452Arg213, Ser228
5d−2.74−1.63−2.44−4.12−7.021Ser228
5e−5.20−6.971Ser228
5f−5.23−7.121Ser228
5g−5.21−1.69−1.37−6.76−9.382Ser228, Lys261
5h−6.13−1.47−7.04−10.282Ser228, Asn232
5i−1.32−4.81−7.552Arg213, Ser228
5j−5.33−1.29−6.05−8.142Arg213, Ser228
5k−5.83−7.752Arg213, Ser228
5l−2.55−6.21−8.822Ser228, Asn232
Table 11. Molecular docking binding affinities to antifungal targets.
Table 11. Molecular docking binding affinities to antifungal targets.
N/NEst. Binding Energy (kcal/mol)I-HResidues
CYP51 of C. albicans
Interactions with HEM601
DNA TopoIV
1S16
CYP51 of C. Albicans
5V5Z
5a−1.55−9.961Tyr145Negative Ionizable, Hydrophobic
5b-−8.542Tyr135, Tyr145Hydrophobic
5c−3.34−8.271Tyr145Hydrophobic
5d−1.64−7.83Negative Ionizable, Hydrophobic
5e-−8.112Tyr145, Thr315Hydrophobic
5f-−9.932Ser378, Met508Hydrophobic
5g−4.01−10.801Tyr145Negative Ionizable, Hydrophobic
5h−4.17−9.811Tyr145Negative Ionizable, Hydrophobic
5i−1.31−9.711Th315Negative Ionizable, Hydrophobic
5j−1.09−8.541Tyr145Negative Ionizable, Hydrophobic
5k−2.41−13.013Ser378, HEM601Negative Ionizable, iron binding (Fe), hydrogen bond
5l−4.25−10.851Tyr135Negative Ionizable, Hydrophobic
ketoconazole−8.231Tyr64Hydrophobic, aromatic
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Horishny, V.; Geronikaki, A.; Kartsev, V.; Matiychuk, V.; Petrou, A.; Pogodin, P.; Poroikov, V.; Papadopoulou, T.A.; Vizirianakis, I.S.; Kostic, M.; et al. Synthesis, Biological Evaluation and Molecular Docking Studies of 5-Indolylmethylen-4-oxo-2-thioxothiazolidine Derivatives. Molecules 2022, 27, 1068. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27031068

AMA Style

Horishny V, Geronikaki A, Kartsev V, Matiychuk V, Petrou A, Pogodin P, Poroikov V, Papadopoulou TA, Vizirianakis IS, Kostic M, et al. Synthesis, Biological Evaluation and Molecular Docking Studies of 5-Indolylmethylen-4-oxo-2-thioxothiazolidine Derivatives. Molecules. 2022; 27(3):1068. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27031068

Chicago/Turabian Style

Horishny, Volodymyr, Athina Geronikaki, Victor Kartsev, Vasyl Matiychuk, Anthi Petrou, Pavel Pogodin, Vladimir Poroikov, Theodora A. Papadopoulou, Ioannis S. Vizirianakis, Marina Kostic, and et al. 2022. "Synthesis, Biological Evaluation and Molecular Docking Studies of 5-Indolylmethylen-4-oxo-2-thioxothiazolidine Derivatives" Molecules 27, no. 3: 1068. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27031068

Article Metrics

Back to TopTop