Next Article in Journal
Tetrahydrocurcumin-Related Vascular Protection: An Overview of the Findings from Animal Disease Models
Next Article in Special Issue
Synthesis and Biological Activity of N-acyl Anabasine and Cytisine Derivatives with Adamantane, Pyridine and 1,2-Azole Fragments
Previous Article in Journal
Tyrosol Derivatives, Bearing 3,5-Disubstituted Isoxazole and 1,4-Disubstituted Triazole, as Potential Antileukemia Agents by Promoting Apoptosis
Previous Article in Special Issue
Synthesis of Pyrimidine Conjugates with 4-(6-Amino-hexanoyl)-7,8-difluoro-3,4-dihydro-3-methyl-2H-[1,4]benzoxazine and Evaluation of Their Antiviral Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Biological Activity of Unsymmetrical Monoterpenylhetaryl Disulfides

by
Denis V. Sudarikov
1,*,
Yulia V. Gyrdymova
1,
Alexander V. Borisov
2,
Julia M. Lukiyanova
2,
Roman V. Rumyantcev
3,
Oksana G. Shevchenko
4,
Diana R. Baidamshina
5,
Nargiza D. Zakarova
5,
Airat R. Kayumov
5,
Ekaterina O. Sinegubova
6,
Alexandrina S. Volobueva
6,
Vladimir V. Zarubaev
6 and
Svetlana A. Rubtsova
1
1
Institute of Chemistry, FRC “Komi Scientific Centre”, Ural Branch of the Russian Academy of Sciences, Pervomayskaya St. 48, 167000 Syktyvkar, Komi Republic, Russia
2
Department of Chemistry, Nizhny Novgorod State Technical University n.a. R.E. Alekseev, 24 Minin St., 603155 Nizhny Novgorod, Russia
3
Razuvaev Institute of Organometallic Chemistry, Russian Academy of Sciences, Tropinina 49, 603950 Nizhny Novgorod, Russia
4
Center of Collective Usage “Molecular Biology”, Institute of Biology, Komi Science Centre, Ural Branch of Russian Academy of Sciences, 28 Kommunisticheskaya Street, 167982 Syktyvkar, Russia
5
Institute of Fundamental Medicine and Biology, Kazan Federal University, 18 Kremlevskaya St., 420008 Kazan, Russia
6
Saint Petersburg Pasteur Research Institute of Epidemiology and Microbiology, 14 Mira St., 197101 Saint Petersburg, Russia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(16), 5101; https://doi.org/10.3390/molecules27165101
Submission received: 21 July 2022 / Revised: 28 July 2022 / Accepted: 5 August 2022 / Published: 10 August 2022
(This article belongs to the Special Issue Synthesis of Bioactive Compounds)

Abstract

:
New unsymmetrical monoterpenylhetaryl disulfides based on heterocyclic disulfides and monoterpene thiols were synthesized for the first time in 48–88% yields. Hydrolysis of disulfides with fragments of methyl esters of 2-mercaptonicotinic acid was carried out in 73–95% yields. The obtained compounds were evaluated for antioxidant, antibacterial, antifungal activity, cytotoxicity and mutagenicity.

1. Introduction

Disulfide bonds play an important role in living organisms. In the formation of the tertiary structure of proteins, covalent disulfide bonds provide the formation of cross-links that are much stronger than hydrophobic interactions or hydrogen bonds. The biochemical significance of the disulfide bond is also determined by the unique nature of the thiol-disulfide system, in which the S-S bond can be formed under conditions characteristic of biological processes—by disulfide exchange with the participation of glutathione [1]. The reactivity of disulfides is determined by the lability of the S-S bond and its ability to be cleaved by both nucleophiles and electrophiles [1].
In the presence of various thiophilic nucleophiles, disulfides behave as electrophilic reagents, which leads to heterolytic cleavage of the S-S bond. Intracellular nucleophiles such as glutathione and thioredoxins, for example, can rapidly react with di- and trisulfide compounds, disrupting cellular homeostasis [2,3].
Some biologically active lipophilic disulfides of natural origin are known. These include the symmetrical diallyl disulfide (DADS), ajoene, vinyldithiins, and diallyl polysulfides found in Allium sativum. DADS has moderate antioxidant activity at low concentrations (20 µM) [4], however, at higher concentrations (50–100 µM), DADS exhibits pro-oxidant activity and cytotoxicity in vitro through the accumulation of reactive oxygen species (ROS) in cells [5], which is consistent with studies on the antitumor activity of DADS on human neuroblastoma cells SH-SY5Y [6], as well as on human promyelocytic leukemia cell line HL-60 [7], leading to their selective apoptosis. The antimicrobial activity of DADS and diallyl polysulfides and the ability to inhibit the formation of biofilms on surfaces are due to the modification of thiol-containing proteins in Pseudomonas aeruginosa, which leads to a decrease in the level of glutathione, induction of protein aggregation and inactivation of enzymes responsible for the vital activity of microorganisms, as well as inactivation of quorum-sensing genes [4,8]. DADS prevents infection with methicillin-resistant Staphylococcus aureus in diabetic mice [9], has a synergistic effect with gentamicins on Escherichia coli [10,11] and inhibits Helicobacter pylori activity both in vitro and in vivo [12,13]. DADS-loaded 140 ± 30 nm niosomes effectively eliminated the fungal burden in mice infected with Candida albicans and increased the survival rate of infected specimens [14]. DADS has also been reported to inhibit the growth of Aspergillus versicolor and its toxic metabolites [15]. DADS has been found to inhibit the proliferation of HIV-1-infected cells [16]. In addition, DADS had anti-inflammatory and antioxidant effects in the dengue virus study, reducing symptoms and disease severity [17].
The symmetrical Psammaplin A disulfide isolated from the sponge Psammaplysilla and containing bromotyrosine fragments and its unsymmetrical analogs, in particular UVI5008, exhibit in vitro activity against a methicillin-resistant strain of Staphylococcus aureus (MRSA) [18,19]. Isolated from Polycitorella mariae, a new cytorellamine disulfide containing heterocyclic bromoindole fragments has significant antimicrobial activity [20]. Cytorellamine disulfide (originally a misattributed sulfide structure) has demonstrated strong antibacterial activity against Staphylococcus aureus, Bacillus subtilus, and Escherichia coli, as well as cytotoxicity against certain cancer cell lines [21].
Disulfiram, a drug originally used for the treatment of alcohol dependence, and unsymmetrical disulfiram-based disulfides with alkyl and aryl fragments have antibacterial activity against a large selection of gram-positive (such as Staphylococcus aureus—MRSA, VISA, VRSA) and gram-negative pathogenic bacteria (Bacillus anthracisa, Bacillus cereus, Borrelia burgdorferi), and have a synergistic effect together with vancomycin [22,23,24]. Antimicrobial therapy in a limited number of patients has shown the effectiveness of disulfiram in the treatment of borreliosis caused by Borrelia burgdorferi in clinical settings [25]. The authors attribute the mechanism of action of disulfiram and its derivatives to reactions with thiophilic residues in proteins, metabolites, and cofactors, as well as to a decrease in the content of cellular glutathione [22].
Synthetic alkylaryl and dialkyl disulfides show potent antimicrobial activity against Staphylococcus aureus and Bacillus anthracis, including MRSA, through inhibition of the β-ketoacyl acyl carrier protein synthase III, or FabH, a key enzyme in type II fatty acid biosynthesis [26]. The highest antimicrobial activity in vitro was found in unsymmetrical disulfides with nitrophenyl substituents, which may be the result of electronic activation of the leaving arylthio group for a nucleophilic attack on the disulfide bond [26].
Unsymmetrical disulfides with a 1,2,4-triazine substituent exhibit in vitro low micromolar inhibitory properties of MCF-7 human breast cancer cells. Disulfides with methyl and methoxy substituents in the aromatic ring demonstrated activity superior to chlorambucil [27].
Unsymmetrical disulfides are a promising platform for the development of new drugs against SARS. Unsymmetrical aromatic disulfides, incl. with fragments of substituted imidazole, pyrimidine, oxadiazole, etc., proved to be in vitro inhibitors of the protease of the SARS CoV Mpro coronavirus, the main causative agent of atypical pneumonia [28,29].
Semi-synthetic symmetrical and asymmetric disulfides with fragments of mono- and sesquiterpenoids demonstrate antioxidant, membrane-protective, and antiviral properties [30,31,32].
It is interesting to develop the chemistry of disulfide compounds based on natural compounds in order to study their biological properties. Combining in one molecule terpene and heterocyclic fragments connected by an S-S bond can promote the growth of already existing biological activity.

2. Results and Discussion

2.1. Synthesis of Unsymmetrical Monoterpenylhetaryl Disulfides

In our work, we synthesized a series of unsymmetrical disulfides combining monoterpene and heterocyclic fragments, and evaluated the antibacterial, antifungal, antiviral, and antioxidant activities of the synthesized conjugates.
As a heterocyclic component, we used symmetrical dihetaryl disulfides with benzothiazole, pyrimidine, and pyridine fragments, in which the disulfide bond is easily cleft in the reaction with SO2Cl2 to form sulfenyl chlorides [33,34]. The above compounds are highly reactive and often unstable, so sulfenyl chlorides were immediately involved in reactions with terpenthiols without prior isolation to form the corresponding monoterpenylhetaryl disulfides (1–4)a–c in 48–88% yields (Scheme 1), according to the procedure [35,36].
The structure and composition of disulfides (1–4)a–c have been proved by IR, NMR spectroscopy and mass spectrometry. The 1H and 13C NMR spectra contain signals of both terpene and heterocyclic fragments. The structure and configuration of 1c was confirmed by single-crystal X-ray diffraction (Figure 1). The SSHet-fragmet in the molecule 1c is almost flat. The average deviation of non-hydrogen atoms from the plane is 0.044 Å. The cyclohexane ring adopts a chair conformation. The methyl and isopropyl groups are in equatorial positions, while the SSHet-fragment occupy axial positions. The S-S distance in 1c is 2.0238(7) Å and is in good agreement with previously published related disulfides [37].
To assess the effect of the ester and carboxyl groups on the biological activity of disulfides for derivatives of 2-mercaptonicotinic acid, hydrolysis of disulfides 1c–4c having an ester group was carried out. The esters were hydrolyzed in the LiOH–H2O–THF system. As a result, the corresponding acids 1d–4d were obtained in 73–95% yields (Scheme 2).
The structure and composition of disulfides 1d–4d with a fragment of 2-mercaptonicotinic acid were proved by IR, NMR spectroscopy and mass spectrometry. In the 1H and 13C NMR spectra of acids 1d–4d, there are no signals of methyl groups compared to the starting esters 1c–4c. The mass spectra contain peaks of the corresponding molecular ions.
The synthesized disulfide conjugates were evaluated for antioxidant, antibacterial, antifungal and antiviral activities.

2.2. Antioxidant Activity in a Non-Cellular Model System (Brain Homogenates)

In a non-cellular model system containing easily oxidized brain lipids (Figure 2), compounds 3a–c, 4a–c and 3d containing a hydroxyl group in the pinane fragment, in general, significantly more actively inhibit the oxidation of unsaturated fatty acids of brain lipids than similar in structure compounds 1a–d, 2a–d. If we focus on the structure of the heterocyclic substituent, then in all four groups the most active compounds contain a fragment of the methyl ester of 2-mercaptonicotinic acid (1–4c).
The highest antioxidant activity in this model system is demonstrated by the disulfides 3c and 4c, which combine a hydroxyl group in the pinane and an ester group in the heterocyclic fragment (Figure 2).

2.3. Erythrotoxicity, Membrane-Protective and Antioxidant Activity (on Erythrocytes)

A study of hemolytic activity showed that at a concentration of 100 μM, 13 out of 16 disulfides exhibit high erythrotoxicity (Figure 3A). Low hemolytic activity at the indicated concentration was noted only for disulfides 1a, 2a, and 3d. No significant hemolytic activity was detected at a concentration of 1 μM (Figure 3B).
The membrane-protective properties of the compounds were evaluated in the model of oxidative hemolysis of erythrocytes at a concentration of 1 μM in order to avoid the effect of hemolytic activity. It was shown (Figure 4) that even at such a low concentration, individual disulfides have a statistically significant membrane-protective activity comparable to the activity of Trolox, which follows from a decrease in the death of erythrocytes in their presence under the influence of H2O2. The most active are pinane derivatives 2a, 2b, 2c, 3b, 3c, 4a, 4b, 4c and 4d, most of which contain a hydroxyl group in the terpene moiety (3b, 3c, 4a, 4b, 4c and 4d).
Thus, in both test systems used, the antioxidant activity of the obtained derivatives depended mainly on the structure of the terpene fragment, the presence of a hydroxyl group and a heterocyclic fragment in it.

2.4. Antimicrobial Activity

The antimicrobial activity of all compounds was evaluated on S. aureus ATCC 29213 (MSSA), S. aureus MRSA clinical isolate, P. aeruginosa ATCC 27853 and fluconazole-sensitive Candida albicans 703 clinical isolate. As could be seen from Table 1, compounds 1a–c and 2b exhibited antibacterial activity on both S. aureus and P. aeruginosa while the antifungal activity was low. By contrast, compounds 3d and 4d were active against yeast on low antibacterial activity background.
Unsymmetrical disulfides 1a–d with a neomentane fragment showed antimicrobial activity against both S. aureus strains with MICs of 16–32 μg/mL. The antimicrobial properties of disulfides with pipane substituents varied depending on the structure of the heterocyclic fragment. Among the disulfides 2a–d, compound 2b with a pyrimidine substituent (MIC 16 μg/mL) and 2d (MIC 16 μg/mL) showed the highest activity against the MSSA isolate. Compound 2a with a benzothiazole fragment had weak activity (MIC 128 μg/mL), and 2c was inactive against MSSA, while showing activity in the case of MRSA. The introduction of an OH group into the structure of the terpene fragment leads to an almost complete loss of the antibacterial properties of thiotherpenoids 4a–4c. Unsymmetrical pinane disulfides 3a–c, having a hydroxymethyl group, were similar in their biological properties to neomentane thiotherpenoids 1a–c and showed antibacterial activity with an MIC of 32–34 µg/mL. These thiotherpenoids tested showed similar properties against the clinical methylicin-resistant S. aureus isolate.
All disulfides 1a–c with the neomentane structure of the terpene part of the molecule, as well as compounds 2a and 3a with a benzothiazole substituent, and pinane disulfide 2b with a pyrimidine fragment repressed the growth of P. aeruginosa ATCC 27853. All other compounds were inactive against this microorganism.
None of the tested asymmetric disulfides (1–4)a–c had antifungal activity against Candida albicans 703 clinical isolate. However, the presence of a carboxyl group in the 2-mercaptonicotinic acid fragment of disulfides 1–4d led to the appearance of antifungal activity (MIC 16–128 μg/mL).
In general, the synthesized asymmetric monoterpenylhetaryl disulfides (1–4)a–d have a high cytotoxicity (CC50) on the nuclear cells of the embryonic bovine lung (EBL). The least toxicity was shown by pinane disulfides 2a, 2c and 3d. Notably, neomentane disulfides 1a and 1b showed mutagenicity in the Ames test on Salmonella typhimurium TA100 and TA102 strains, respectively, causing point mutations and reversions [38].
Thus, unsymmetrical disulfides containing monoterpene and heterocyclic substituents have antimicrobial properties against clinically significant pathogens. However, these compounds demonstrate relatively high to moderate cytotoxicity, thus reducing the possibility of further use of these compounds as potential antibiotics for systemic application, while can be used as topical agents and can be considered as a start point for the search among unsymmetrical monoterpenylhetaryl disulfides for new substances with low toxicity and selectivity against pathogenic strains of microorganisms.

2.5. Antiviral Activity

The synthesized hybrid compounds were screened for antiviral activity. The cytotoxicity of the compounds (CC50) was evaluated in uninfected MDCK cells (canine kidney cells). The antiviral activity (IC50) of the synthesized substances was studied on the model of influenza infection of MDCK cells caused by the influenza virus strain A/Puerto Rico/8/34 (H1N1), which is resistant to adamantane-type antiviral drugs-amantadine and remantadine. The obtained data were used to calculate the value of the selectivity index (SI) for each of the tested samples of monoterpenylhetaryldisulfides. The results of the experiments performed are shown in the Table 2.
All compounds, with the exception of disulfides 3c, 3d and 4d, showed high virus-inhibiting properties. However, almost all of them showed significant cytotoxicity.
Based on the totality of data and taking into account the highest value of the selectivity index (SI > 10) [39], disulfide 2d with the pinane structure of the terpene and heterocyclic fragment of nicotinoate can be attributed to the active compounds.
Importantly, the virus used in our study is resistant to the reference compound Rimantadine having, similarly to the compound library under investigation, cage fragment in its structure. This is due to amino acid substitution S31N in transmembrane domain of viral proton pump M2 which is target for adamantane-based anti-influenza drugs, Rimantadine and Amantadine. The activity of 2d appeared higher than that of Rimantadine. The results obtained in our study, therefore, suggest that either 2d is able to block M2 channel of rimantadine-resistant virus or, otherwise, it has another target and an alternative mechanism of activity. Further studies are therefore needed to decipher the exact mode of anti-viral action of this class of compounds.

3. Conclusions

Thus, new unsymmetrical monoterpenylhetaryl disulfides based on heterocyclic disulfides and monoterpene thiols have been synthesized for the first time. The described disulfides, in general, have good indicators of antioxidant, antimicrobial and antiviral activity. However, given the high in vitro cytotoxicity in different models, the possibility of further systemic use of these compounds as potential antioxidants, antibiotics, and antiviral agents is limited, their chemotype can be considered as a start point for the among unsymmetrical monoterpenylhetaryl disulfides for new substances with low toxicity and selectivity for pathogenic strains of microorganisms and viruses. In addition, cytotoxicity analysis shows differences in the activity of these compounds in different models (mice erythrocytes, embryonic bovine lung cells, and canine kidney cells), which may indicate a selective effect on different cell types and provides prerequisites for testing them for antitumor activity.

4. Materials and Methods

4.1. General Information

IR spectra were registered on a Shimadzu IR Prestige 21 infrared Fourier spectrometer (Shimadzu, Duisburg, Germany) in a thin layer or in KBr tablets. 1H NMR and 13C NMR spectra were recorded on a Bruker Avance 300 spectrometer (Bruker BioSpin, Ettlingen, Germany) (300.17 and 75.48 MHz) in C6D6, CDCl3 and acetone-d6 using the signal of the indicated solvent as an internal standard (See Supplementary Materials). 13C NMR spectra were registered in the J-modulation mode. The complete assignment of 1H and 13C signals was performed using 2D homo- (1H–1H COSY, 1H–1H NOESY) and heteronuclear experiments (1H–13C HSQC, 1H–13C HMBC). For convenience of interpretation of the NMR spectra on the structures (1–4)a–d, the numbering of carbon atoms is marked, which may not correspond to the recommended IUPAC (See Scheme 1 and details of the NMR spectra in the experimental section). Mass spectra were recorded on a Thermo Finnigan LCQ Fleet instrument (Thermo Fisher Scientific, Waltham, USA) equipped with an MS detector. Masses were scanned in the m/z range 50–2000 (ESI, 20 eV). Melting points were determined on a Sanyo Gallenkamp MPD350BM3.5 instrument (Sanyo Gallenkamp, Southborough, UK) and were not corrected. The angles of optical rotation were measured on an automated digital polarimeter Optical Activity PolAAr 3001 (Optical Activity, Huntingdon, UK). Sorbfil plates were used for thin-layer chromatography; the visualizing agent was a solution of phosphoromolybdic acid in ethanol. Alfa Aesar silica gel (0.06–0.2 mm) was used for column chromatography. The commercially available 2,2′-dithiobisbenzothiazol, 99% (Sigma Aldrich, St. Louis, MO, USA), 4,6-dimethylpyrimidine-2-thiol, 98% (Alfa Chemistry, Heysham, UK), 2-mercaptonicotinic acid, 90% (Alfa Aesar, Heysham, UK) were used without additional purification.
The diffraction data for compound 1c were collected on a Bruker D8 Quest diffractometer (Mo-Kα radiation, ω-scan technique, λ = 0.71073 Å). The intensity data were integrated by the SAINT [40] program. The structure was solved by dual methods [41] and was refined on F hkl 2 using the SHELXTL package [42]. All non-hydrogen atoms were refined anisotropically. All hydrogen atoms were placed in calculated positions and were refined using a riding model (Uiso(H) = 1.5Ueq(C) for CH3 groups and Uiso(H) = 1.2Ueq(C) for other groups). The SADABS program [43] was used to perform absorption corrections. The disorder of C(15) and C(16) methyl groups was modeled with EADP/SADI/ISOR constraints. CCDC 2182740 contains the supplementary crystallographic data. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via https://www.ccdc.cam.ac.uk/structures, accessed on 28 July 2022.

4.2. General Procedure

2,2′-Disulfanediyldinicotinic acid. To a solution of 8 g (0.2 mol) of sodium hydroxide in 200 mL of water 31 g (0.2 mol) of 2-mercaptonicotinic acid was added with stirring. A 7% aqueous solution of iodine with potassium iodide (1:1 eq) was added to the resulting solution with vigorous stirring until the color of iodine ceased to disappear. The precipitate formed was filtered off and thoroughly washed with water and then dried. The dried precipitate was boiled 2 times for 30 min in 100 mL of methylene chloride and filtered off. The yield of the disulfide is 27.4 g (89%) with m.p. 222–224 °C. The spectral data correspond to the literature data [44].
1,2-bis(4,6-dimethylpyrimidin-2-yl)disulfide was prepared similarly, m.p. 167–168 °C, spectral characteristics correspond to the literature [45].
Dimethyl 2,2′-disulfanediyldinicotinate. To a suspension of 6.16 g (20 mmol) of 2,2′-disulfanediyldinicotinic acid in 20 mL of absolute methanol was added 5 g of concentrated sulfuric acid and boiled for 5 h. Then, the methanol was distilled off in a vacuum, the residue was treated with water and filtered off. The filtrate was cooled to 0 °C and carefully neutralized to a slightly alkaline reaction (pH 8–9) with a NaHCO3 solution. The precipitate formed was filtered off and dried. The dried precipitate was dissolved in 10 mL of methylene chloride, the solution was filtered. The filtrate was evaporated in a vacuum, and the residue was boiled 2 times for 30 min in 30 mL of diethyl ether and filtered off. The yield of disulfide is 3.7 g (55%), m.p. 199–200 °C. 1H-NMR (300 MHz, CDCl3): 4.01 (s, 3H, OCH3), 7.10 (m, 1H), 8.23 (d.d, 1H, 3J = 7.6, J = 1.3), 8.48 (d.d., 1H, 3J = 5.6, J = 1.3). IR (KBr), cm−1: 496; 825, 765, 760, 752, 731 (ρArCH); 1287, 1246, 1136, 1067 (νCN,CS), (βArCH); 1576, 1557, 1396 (νArCH); 1713 (ν C=O).
Neomenthanthiol (1) [46], (−)-cis-myrthanethiol (2) [47], 10-hydroxyisopinocampheylthiol (3) [48] and 10-sulfanylisopinocampheol (4) [49] were synthesized according to known methods.

4.2.1. General Procedure for the Synthesis of Unsymmetrical Disulfides

The procedure is based on methods for the synthesis of sulfenyl chlorides [33,34] and their interaction with thiols [35,36].
A total of 0.45 mmol (1 equiv.) of dihetaryl disulfide was dissolved in 10 mL of CH2Cl2 and cooled on ice-bath. A solution of 0.68 mmol (1.5 equiv.) SO2Cl2 in CH2Cl2 (v/v 1:10) was added to the cooled mixture and stirred at 0 °C for 5 min. The solvent and the excess of unreacted SO2Cl2 were removed under reduced pressure. The residue was dissolved in 5 mL of CH2Cl2, and 0.90 mmol (2 equiv.) of terpenthiol in 4 mL of CH2Cl2 was added at 0 °C. After 5–10 min, the ice-bath was removed and left to stir at room temperature. The progress of the reaction was monitored by TLC using the eluents indicated in the descriptions of the compounds. Upon completion of the reaction, the mixture was diluted with 15–20 mL of CH2Cl2 and washed with brine, the organic layer was dried over Na2SO4, and the solvent was removed under vacuum. The products were isolated by silica gel column chromatography using the same eluents as for TLC.
2-(((1S,2S,5R)-2-isopropyl-5-methylcyclohexyl)disulfanyl)benzo[d]thiazole (1a). Yield: 61%. Yellowish oil. Rf = 0.5 (PhH/EtOAc, 20:1). [ α ] D 26 = +96.9 (c = 1.0, CHCl3). IR (KBr): 2951, 2918, 1458, 1427, 1236, 1003, 756, 725, 671. 1H-NMR (300 MHz, C6D6): 0.72–0.83 (m, 1H, H-4a), 0.79 (d, 3H, J = 6.6, Me-7), 0.87 (d, 3H, J = 6.6, Me-10), 0.98–1.12 (m, 2H, H-2, H-6a), 0.99 (d, 3H, J = 6.6, Me-9), 1.10–1.24 (m, 1H, H-3a), 1.60–1.73 (m, 2H, H-3b, H-4b), 1.70 (ddt, 1H, J = 13.0, 10.0, 6.6, H-8), 1.81–2.00 (m, 1H, H-5), 2.30 (dq, 1H, J = 13.9, 2.9, H-6b), 3.52–3.57 (m, 1H, H-1), 6.92 (t, 1H, J = 7.7, H-5′), 7.07 (t, 1H, J = 7.8, H-4′), 7.31 (d, 1H, J = 7.8, H-6′), 7.88 (d, 1H, J = 7.8, H-3′). 13C-NMR (75 MHz, CDCl3): 21.6 (C-10, C-9), 22.5 (C-7), 26.4 (C-3), 27.0 (C-5), 30.7 (C-8), 35.7 (C-4), 40.1 (C-6), 49.4 (C-2), 55.5 (C-1), 121.6 (C-6′), 123.1 (C-3′), 125.0 (C-5′), 126.0 (C-4′), 136.9 (C-7′), 156.2 (C-2′), 173.9 (C-1′). MS (ESI, 20 eV): 338.20 (100) [M+H]+. M = 337.56, C17H23NS3.
2-(((1S,2S,5R)-2-isopropyl-5-methylcyclohexyl)disulfanyl)-4,6-dimethylpyrimidine (1b). Yield: 62%. Light-brown crystal. Rf = 0.39 (PhH/EtOAc, 20:1). [ α ] D 23 = +135.7 (c = 0.8, CHCl3). IR (KBr): 2949, 2916, 1581, 1531, 1446, 1369, 1256, 1028, 880, 849, 546. 1H-NMR (300 MHz, C6D6): 0.86–0.93 (m, 1H, H-4a), 0.99 (d, 3H, J = 6.6, Me-7), 1.03 (d, 3H, J = 6.6, Me-10), 1.09–1.25 (m, 2H, H-2, H-6a, H-2), 1.38 (d, 3H, J = 6.6, Me-9), 1.42–1.56 (m, 1H, H-3a), 1.75–1.88 (m, 2H, H-3b, H-4b), 2.06 (ddt, 1H, J = 13.2, 10.3, 6.6, H-8), 2.18 (s, 6H, Me-5′, Me-6′), 2.32–2.49 (m, 1H, H-5), 2.63 (dq, 1H, J = 13.6, 2.8, H-6b), 3.80–3.84 (m, 1H, H-1), 6.08 (s, 1H, H-3′). 13C-NMR (75 MHz, CDCl3): 21.7 (C-10), 21.9 (C-9), 22.8 (C-7), 23.8 (C-5′, 6′), 26.7 (C-3), 26.9 (C-5), 30.9 (C-8), 36.1 (C-4), 39.9 (C-6), 49.6 (C-2), 53.6 (C-1), 117.0 (C-3′), 167.7 (C-2′, 4′), 172.2 (C-1′). MS (ESI, 20 eV): 311.15 (100) [M+H]+. M = 310.52, C16H26N2S2.
Methyl 2-(((1S,2S,5R)-2-isopropyl-5-methylcyclohexyl)disulfanyl)pyridine-3-carboxylate (1c). Yield: 80%. Colorless crystals. M.p. 96 °C. Rf = 0.26 (PhH). [ α ] D 26 = +119.8 (c = 1.0, CHCl3). IR (KBr): 2951, 2912, 1715 (C=O), 1557, 1445, 1398, 1288, 1240, 1130, 1068, 760. 1H-NMR (300 MHz, C6D6): 0.82–0.96 (m, 1H, H-4a), 0.86 (d, 3H, J = 6.6, Me-7), 0.93 (d, 3H, J = 6.6, Me-10), 0.99–1.15 (m 2H, H-2, H-6a), 1.30 (d, 3H, J = 6.6, Me-9), 1.37–1.51 (m 1H, H-3a), 1.64–1.77 (m 2H, H-3b, H-4b), 2.03 (dq, 1H, J = 9.5, 6.6, H-8), 2.23–2.39 (m, 1H, H-5), 2.48 (dq, 1H, J = 13.8, 2.7, H-6b), 3.37 (s, 3H, Me-7′), 3.73–3.78 (m, 1H, H-1), 6.38 (dd, 1H, J = 7.7, 4.8, H-3′), 7.75 (dd, 1H, J = 7.7, 1.7, H-4′), 8.31 (dd, 1H, J = 4.8, 1.7, H-2′). 13C-NMR (75 MHz, CDCl3): 21.6 (C-9), 21.9 (C-10), 22.8 (C-7), 26.7 (C-3), 27.0 (C-5), 30.9 (C-8), 36.2 (C-4), 40.1 (C-6), 49.9 (C-2), 52.2 (C-6′), 52.9 (C-1), 120.0 (C-3′), 124.6 (C-5′), 138.9 (C-4′), 152.4 (C-2′), 163.6 (C-1′), 165.7 (C-6′). MS (ESI, 20 eV): 340.09 [M+H]+. M = 339.51, C17H25NO2S2. A colorless prismatic crystal of the orthorhombic system had size 0.47 × 0.44 × 0.25 mm, space group P212121, a = 8.9974(4), b = 13.2611(5), c = 15.5898(6) Å, α = β = γ = 90°, V = 1860.10(13) Å3, Z = 4, μ = 0.292 mm−1, dcalc = 1.212 g/cm3, F(000) = 728. A dataset of 27448 reflections was collected at scattering angles 2.016° < θ < 26.007°, of which 3659 were independent (Rint = 0.0215), including 3421 reflections with I > 2σ(I). The final refinement parameters were R1 = 0.0301, wR2 = 0.0687 (all data), R1 = 0.0267, wR2 = 0.0664 [I > 2σ(I)] with GooF = 1.069. Δρe = 0.122/−0.215 e Å−3; Flack parameter = −0.039(11).
2-((((1S,2R,5S)-6,6-dimethylbicyclo [3.1.1]heptan-2-yl)methyl)disulfaneyl)benzo[d]thiazole (2a). Yield: 71%. Yellowish oil. Rf = 0.26 (PE/EtOAc 50:1). [ α ] D 24 = −75.7 (c = 0.3, CHCl3). IR (KBr): 2983, 2911, 1462, 1427, 1308, 1234, 1078, 1003, 756, 725, 667. 1H-NMR (300 MHz, CDCl3): 0.94 (d, 1H, J = 10.2, H-7a), 1.00 (s, 3H, Me-8), 1.20 (s, 3H, Me-9), 1.50–1.65 (m, 1H, H-3a), 1.86–2.15 (m, 5H, H-5, H-4a, H-4b, H-1, H-3b), 2.33–2.52 (m, 2H, H-2, H-7b), 2.99–3.12 (m, 2H, H-10a, H-10b), 7.33 (t, 1H, J = 7.8, H-5′), 7.44 (t, 1H, J = 7.8, H-4′), 7.82 (d, 1H, J = 7.8, H-6′), 7.87 (d, 1H, J = 7.8, H-3′). 13C-NMR (75 MHz, CDCl3): 21.7 (C-3), 23.3 (C-8), 26.0 (C-4), 27.8 (C-9), 33.1 (C-7), 38.7 (C-6), 40.3 (C-2), 41.1 (C-5), 45.3 (C-1), 46.8 (C-10), 121.1 (C-6′), 122.1 (C-3′), 124.4 (C-5′), 126.2 (C-4′), 135.8 (C-7′), 155.2 (C-2′), 173.2 (C-1′). MS (ESI, 20 eV): 336.24 (100) [M+H]+. M = 335.54, C17H21NS3.
2-((((1S,2R,5S)-6,6-dimethylbicyclo[3.1.1]heptan-2-yl)methyl)disulfanyl)-4,6-dimethylpyrimidine (2b). Yield: 57%. Light-brown powder. M.p. 45 °C. Rf = 0.34 (PE/EtOAc 10:1). [ α ] D 24 = −76.4 (c = 0.3, CHCl3). IR (KBr): 2938, 2913, 1582, 1531, 1435, 1340, 1256, 1032, 880, 764. 1H-NMR (300 MHz, CDCl3): 0.88 (d, 1H, J = 9.5, H-7a), 0.99 (s, 3H, Me-8), 1.19 (s, 3H, Me-9), 1.52–1.66 (m, 1H, H-3a), 1.80–2.14 (m, 5H, H-5, H-4a, H-4b, H-1, H-3b), 2.29–2.47 (m, 2H, H-2, H-7b), 2.46 (s, 6H, Me-5′, Me-6′), 2.87–3.03 (m, 2H, H-10a, H-10b), 6.78 (s, 1H, H-3′). 13C-NMR (75 MHz, CDCl3): 21.8 (C-3), 23.2 (C-8), 23.9 (C-5′, C-6′), 26.1 (C-4), 27.9 (C-9), 33.3 (C-7), 38.6 (C-6), 39.9 (C-2), 41.2 (C-5), 45.4 (C-1), 46.0 (C-10), 116.9 (C-3′), 167.6 (C-2′,C-4′), 170.8 (C-1′). MS (ESI, 20 eV): 309.30 (100) [M+H]+. M = 308.50, C16H24N2S2.
Methyl 2-((((1S,2R,5S)-6,6-dimethylbicyclo[3.1.1]heptan-2-yl)methyl)disulfaneyl)pyridine-3-carboxylate (2c). Yield: 74%. Yellowish oil. Rf = 0.24 (PE/EtOAc 10:1). [ α ] D 24 = −63.7 (c = 0.3, CHCl3). IR (KBr): 2983, 2911, 1719 (C=O), 1574, 1553, 1439, 1396, 1285, 1240, 1136, 1063, 762. 1H-NMR (300 MHz, CDCl3): 0.90 (d, 1H, J = 9.5, H-7a), 0.98 (s, 3H, Me-8), 1.18 (s, 3H, Me-9), 1.50-1.63 (m, 1H, H-3a), 1.80–2.13 (m, 5H, H-5, H-4a, H-4b, H-1, H-3b), 2.30–2.46 (m, 2H, H-2, H-7b), 2.83–2.99 (m, 2H, H-10a, H-10b), 3.96 (s, 3H, Me-7′), 7.18 (dd, 1H, J = 7.7, 4.8, H-3′), 8.23 (d, 1H, J = 7.7, H-4′), 8.73 (d, 1H, J = 4.8, H-2′). 13C-NMR (75 MHz, CDCl3): 21.9 (C-3), 23.2 (C-8), 26.1 (C-4), 27.9 (C-9), 33.2 (C-7), 38.6 (C-6), 40.0 (C-2), 41.2 (C-5), 45.5 (C-1), 45.5 (C-10), 52.5 (C-7′), 119.8 (C-3′), 123.7 (C-5′), 139.0 (C-4′), 152.6 (C-2′), 161.8 (C-1′), 165.7 (C-6′). MS (ESI, 20 eV): 338.25 (100) [M+H]+. M = 337.50, C17H23NO2S2.
((1S,2R,3S,5R)-3-(benzo[d]thiazol-2-yldisulfanyl)-6,6-dimethylbicyclo[3.1.1]heptan-2-yl)methanol (3a). Yield: 78%. Yellowish oil. Rf = 0.24 (PhH/EtOAc, 10:1). [ α ] D 26 = +104.7 (c = 1.2, CHCl3). IR (KBr): 3383 (OH), 2924, 1460, 1425, 1238, 1049, 1005, 758, 729. 1H-NMR (300 MHz, CDCl3): 0.94 (s, 3H, Me-8), 1.17 (d, 1H, J = 10.3, H-7a), 1.22 (s, 3H, Me-9), 2.00 (tt, 1H, J = 5.6, 3.0, H-5), 2.11 (td, 1H, J = 5.9, 2.2, H-1), 2.20 (ddd, 1H, J = 13.9, 5.9, 2.9, H-4a), 2.33 (qd, 1H, J = 7.3, 2.2, H-2), 2.39–2.50 (m, 1H, H-7b), 2.57–2.73 (m, 2H, H-4b, OH), 3.46–3.54 (m, 1H, H-3), 3.68 (dd, 1H, J = 10.6, 7.7, H-10a), 3.89 (dd, 1H, J = 10.6, 7.3, H-10b), 7.34 (t, 1H, J = 7.8, H-5′), 7.45 (t, 1H, J = 7.8, H-4′), 7.80 (d, 1H, J = 7.8, H-6′), 7.91 (d, 1H, J = 7.8, H-3′). 13C-NMR (75 MHz, CDCl3): 23.4 (C-8), 27.3 (C-9), 32.5 (C-7), 36.7 (C-4), 38.5 (C-6), 41.8 (C-5), 43.1 (C-1), 44.3 (C-3), 52.1 (C-2), 66.4 (C-10), 121.2 (C-6′), 122.0 (C-3′), 124.7 (C-5′), 126.3 (C-4′), 135.8 (C-7′), 154.3 (C-2′), 171.9 (C-1′). MS (ESI, 20 eV): 352.16 (29) [M+H]+. M = 351.54, C17H21NOS3.
((1S,2R,3S,5R)-3-((4,6-dimethylpyrimidin-2-yl)disulfanyl)-6,6-dimethylbicyclo[3.1.1]heptan-2-yl)methanol (3b). Yield: 57%. Yellowish oil. Rf = 0.26 (PhH/EtOAc, 5:1). [ α ] D 26 = +54.9 (c = 0.7, CHCl3). IR (KBr): 3416 (OH), 2922 1583, 1529, 1437, 1254, 1047, 879, 756, 546. 1H-NMR (300 MHz, CDCl3): 0.94 (s, 3H, Me-8), 1.13 (d, 1H, J = 10.3, H-7a), 1.19 (s, 3H, Me-9), 1.84 (td, 1H, J = 5.9, 1.7, H-1), 1.97 (tt, 1H, J = 5.5, 2.9, H-5), 2.20 (ddd, 1H, J = 14.3, 6.2, 2.9, H-4a), 2.30 (dddd, 1H, J = 10.2, 7.6, 5.0, 2.2, H-2), 2.37–2.46 (m, 1H, H-7b), 2.49 (s, 6H, Me-5′, Me-6′), 2.61–2.72 (m, 1H, H-4b), 2.51–3.59 (m, 1H, H-3), 3.62 (dd, 1H, J = 11.0, 5.1, H-10a), 3.62 (t, 1H, J = 11.0, H-10b), 4.28 (br.s, 1H, OH), 6.83 (s, 1H, H-3′). 13C-NMR (75 MHz, CDCl3): 23.5 (C-8), 23.6 (C-5′, C-6′), 27.5 (C-9), 33.4 (C-7), 38.4 (C-4), 38.5 (C-6), 41.9 (C-5), 44.2 (C-1), 44.7 (C-3), 50.5 (C-2), 68.1 (C-10), 117.5 (C-3′), 168.0 (C-2′, C-4′), 170.9 (C-1′). MS (ESI, 20 eV): 325.21 (100) [M+H]+. M = 324.50, C16H24N2OS2.
Methyl 2-(((1S,2R,3S,5R)-2-(hydroxymethyl)-6,6-dimethylbicyclo[3.1.1]hept-3-yl)disulfanyl)pyridine-3-carboxylate (3c). Yield: 88%. Yellowish oil. Rf = 0.24 (PhH/EtOAc, 10:1). [ α ] D 26 = +192.1 (c = 1.1, CHCl3). IR (KBr): 3343 (OH), 2924, 1719 (C=O), 1572, 1557, 1443, 1396, 1286, 1244, 1138, 1062, 763, 729, 680. 1H-NMR (300 MHz, C6D6): 0.67 (s, 3H, Me-8), 0.97 (s, 3H, Me-9), 1.12 (d, 1H, J = 10.3, H-7a), 1.66–1.75 (m, 2H, H-5, H-1), 2.15–2.25 (m, 1H, H-7b), 2.35 (ddd, 1H, J = 14.1, 5.7, 2.2, H-4a), 2.44 (dddd, 1H, J = 9.6, 7.2, 5.1, 2.2, H-2), 2.52–2.63 (m, 1H, H-4b), 3.37 (s, 3H, Me-7′), 3.53–3.63 (m, 1H, H-3), 3.73 (dd, 1H, J = 10.3, 4.4, H-10a), 4.02 (t, 1H, J = 10.3, H-10b), 4.77 (br.s, 1H, OH), 6.32 (dd, 1H, J = 7.7, 4.8, H-3′), 7.68 (dd, 1H, J = 7.7, 1.7, H-4′), 8.45 (dd, 1H, J = 4.8, 1.7, H-2′). 13C-NMR (75 MHz, CDCl3): 23.9 (C-8), 27.7 (C-9), 33.9 (C-7), 38.7 (C-6), 39.0 (C-4), 42.7 (C-5), 45.0 (C-1), 45.1 (C-3), 52.2 (C-7′), 52.7 (C-2), 68.7 (C-10), 120.4 (C-3′), 124.9 (C-5′), 139.5 (C-4′), 152.8 (C-2′), 162.6 (C-1′), 165.5 (C-6′). MS (ESI, 20 eV): 354.21 (23) [M+H]+. M = 353.50, C17H23NO3S2.
(1S,2S,3S,5R)-2-((benzo[d]thiazol-2-yldisulfanyl)methyl)-6,6-dimethylbicyclo[3.1.1]heptan-3-ol (4a). Yield: 68%. Yellowish oil. Rf = 0.42 (PhH/EtOAc, 5:1). [ α ] D 27 = +29.9 (c = 0.4, CHCl3). IR (KBr): 3377 (OH), 1460, 1427, 1275, 1236, 1038, 1005, 758. 1H-NMR (300 MHz, C6D6): 0.59 (s, 3H, Me-8), 1.04 (s, 3H, Me-9), 1.17 (d, 1H, J = 9.5, H-7a), 1.78 (tt, 1H, J = 5.5, 2.9, H-5), 1.79–1.91 (m, 2H, H-1, H-4a), 2.15–2.23 (m, 1H, H-7b), 2.30–2.39 (m, 2H, H-4b, H-2), 3.50 (m, 1H, H-3), 2.75 (dd, 1H, J = 13.2, 6.6, H-10a), 2.97 (dd, 1H, J = 13.2, 8.8, H-10b), 3.17 (br.s, 1H, OH), 4.01 (dt, 1H, J = 9.2, 4.2, H-3), 6.91 (t, 1H, J = 7.8, H-5′), 7.07 (t, 1H, J = 7.8, H-4′), 7.28 (d, 1H, J = 7.8, H-6′), 7.96 (d, 1H, J = 7.8, H-3′). 13C-NMR (75 MHz, CDCl3): 24.0 (C-8), 27.6 (C-9), 33.4 (C-7), 38.3 (C-6), 39.0 (C-4), 42.0 (C-5), 45.9 (C-1), 46.2 (C-10), 53.1 (C-2), 69.1 (C-3), 121.6 (C-6′), 123.0 (C-3′), 125.2 (C-5′), 126.9 (C-4′), 136.8 (C-7′), 155.5 (C-2′), 171.7 (C-1′). MS (ESI, 20 eV): 352.43 (100) [M+H]+. M = 351.54, C17H21NOS3.
(1S,2S,3S,5R)-2-(((4,6-dimethylpyrimidin-2-yl)disulfanyl)methyl)-6,6-dimethylbicyclo[3.1.1]heptan-3-ol (4b). Yield: 48 %. Yellowish oil. Rf = 0.35 (PE/EtOAc, 2:1). [ α ] D 27 = +153.5 (c = 0.6, CHCl3). IR (KBr): 3343 (OH), 2983, 2918, 1585, 1530, 1433, 1366, 1344, 1294, 1256, 1032, 756. 1H-NMR (300 MHz, CDCl3): 0.86 (s, 3H, Me-8), 1.12 (d, 1H, J = 9.5, H-7a), 1.19 (s, 3H, Me-9), 1.77–1.86 (m, 2H, H-1, H-4a), 1.90–2.00 (m, 1H, H-5), 2.15–2.25 (m, 1H, H-2), 2.27–2.37 (m, 1H, H-7b), 2.47 (s, 6H, Me-5′, Me-6′), 2.46–2.59 (m, 1H, H-4b), 2.83–3.03 (m, 2H, H-10a, H-10b), 4.14 (br.s, 1H, OH), 4.28 (dt, J = 8.6, 4.2, 1H, H-3), 6.80 (s, 1H, H-3′). 13C-NMR (75 MHz, CDCl3): 23.6 (C-5′, C-6′), 23.8 (C-8), 27.0 (C-9), 32.5 (C-7), 37.3 (C-4), 37.8 (C-6), 41.2 (C-5), 46.6 (C-1), 46.7 (C-10), 52.6 (C-2), 68.6 (C-3), 117.4 (C-3′), 168.0 (C-2′,C-4′), 170.3 (C-1′). MS (ESI, 20 eV): 325.21 (100) [M+H]+. M = 324.50, C16H24N2OS2.
Methyl 2-((((1S,2S,3S,5R)-3-hydroxy-6,6-dimethylbicyclo[3.1.1]heptan-2-yl)methyl)disulfanyl)pyridine-3-carboxylate (4c). Yield: 85%. Yellowish gum. Rf = 0.25 (PE/EtOAc 3:1). [ α ] D 27 = +118.7 (c = 1.0, CHCl3). IR (KBr): 3400 (OH), 1715 (C=O), 1572, 1557, 1398, 1288, 1244, 1138, 1065, 762. 1H-NMR (300 MHz, CDCl3): 0.87 (s, 3H, Me-8), 1.10 (d, 1H, J = 9.5, H-7a), 1.19 (s, 3H, Me-9), 1.77–1.86 (m, 2H, H-4a, H-1), 1.90–1.99 (m, 1H, H-5), 2.14–2.22 (m, 1H, H-2), 2.27–2.37 (m, 1H, H-7b), 2.46–2.59 (m, 1H, H-4b), 2.82–3.04 (m, 2H, H-10a, H-10b), 3.96 (s, 3H, Me-7′), 4.32 (dt, 1H, J = 9.0, 4.0, H-3), 4.86 (br.s, 1H, OH), 7.22 (dd, 1H, J = 7.7, 4.8, H-3′), 8.26 (d, 1H, J = 7.7, H-4′), 8.66 (d, 1H, J = 4.8, H-2′). 13C-NMR (75 MHz, CDCl3): 23.8 (C-8), 27.1 (C-9), 32.8 (C-7), 37.3 (C-4), 37.9 (C-6), 41.3 (C-5), 46.3 (C-10), 46.6 (C-1), 52.6 (C-2, C-7′), 69.1 (C-3), 120.3 (C-3′), 123.7 (C-5′), 139.4 (C-4′), 152.3 (C-2′), 161.4 (C-1′), 165.4 (C-6′). MS (ESI, 20 eV): 354.21 (100) [M+H]+. M = 353.50, C17H23NO3S2.

4.2.2. General Method for the Synthesis of Carboxylic Acids

The method is based on the procedure [50].
A solution of 0.90 mmol of LiOH in 3–4 mL of H2O was added with stirring to a solution of 0.45 mmol of substrate 1c–4c in 6 mL of THF and left under stirring for 4–5 h. The progress of the reaction was monitored by TLC. Upon completion of the reaction, the solvent was removed under reduced pressure, the residue was diluted with H2O (15–20 mL), neutralized with 2 M HCl to pH = 4, extracted with CHCl3. The organic layer was dried over Na2SO4, the solvent was removed under reduced pressure. The product was purified by recrystallization from hexane-Et2O system.
2-(((1S,2S,5R)-2-isopropyl-5-methylcyclohexyl)disulfanyl)nicotinic acid (1d). Yield: 80%. White powder. M.p. 160–161 °C with decomposition. [ α ] D 27 = +147.2 (c = 0.180, CHCl3). IR (KBr): 3080, 3028, 2647, 2912, 2870, 2841, 2650, 2549, 1684, 1571, 1552, 1446, 1419, 1386, 1292, 1251, 1149, 1126, 1060, 916, 821, 763, 713, 557. 1H-NMR (300 MHz, CDCl3): 0.85 (d, 3H, J = 6.16, Me-7), 0.90–1.00 (m, 1H, H-4a), 0.95 (d, 3H, J = 6.64 Hz, Me-10), 1.03–1.27 (m, 2H, H-2, H-6), 1.20 (d, 3H, J = 2.64 Hz, Me-9), 1.33 (td, 1H, J = 12.69, 3.08 Hz, H-3a), 1.64–1.99 (m, 3H, H-3b, 4b, H-5), 2.04–2.24 (m, 2H, H-6, H-8), 3.35 (d, 1H, J = 1.76 Hz, H-1), 7.18–7.26 (m, 1H, H-3′), 8.35 (dd, 1H, J = 7.78, 1.61 Hz, H-4′), 8.73 (dd, 1H, J = 4.70, 1.47 Hz, H-2′). 13C-NMR (75 MHz, CDCl3): 20.89 (C-10), 21.35 (C-9), 22.09 (C-7), 25.99 (C-3), 26.43 (C-5), 30.22 (C-8), 35.48 (C-4), 39.42 (C-6), 49.21 (C-2), 53.14 (C-1), 119.91 (C-3′), 122.84 (C-5′), 139.93 (C-4′), 152.99 (C-2′), 163.64 (C-1′), 170.30 (C-6′). MS (ESI, 20 eV): 326.07 (100) [M+H]+. M = 325.47, C16H23NO2S2.
2-((((1S,2R,5S)-6,6-dimethylbicyclo[3.1.1]heptan-2-yl)methyl)disulfanyl)nicotinic acid (2d). Yield: 85%. White powder. M.p. 150 with decomposition. [ α ] D 27 = −63.3 (c = 0.139, CHCl3). IR (KBr): 3041, 2980, 2941, 2910, 2868, 2650, 2553, 1691, 1571, 1552, 1465, 1429, 1388, 1286, 1255, 1230, 1145, 1064, 904, 815, 759, 715, 555. 1H-NMR (300 MHz, CDCl3): 0.84 (d, 1H, J = 9.68, H-7a), 0.92 (s, 3H, Me-8), 1.12 (s, 3H, Me-9), 1.42–1.60 (m, 1H, H-3a), 1.70–2.08 (m, 5H, H-1, H-3b, H-4a, H-4b, H-5), 2.20–2.42 (m, 2H, H-2, H-7b), 2.72–2.95 (m, 2H, H-10a, 10b), 7.14 (dd, 1H, J = 7.63, 4.70 Hz, H-3′), 8.23 (dd, 1H, J = 7.63, 1.47 Hz = H-4′), 8.66 (dd, 1H, J = 4.55, 1.61 Hz, H-2′). 13C-NMR (75 MHz, CDCl3): 21.73 (C-3), 23.03 (C-8), 25.99 (C-4), 27.75 (C-9), 33.07 (C-7), 38.53 (C-6), 39.94 (C-2), 41.12 (C-5), 44.37 (C-1), 45.40 (C-10), 119.81 (C-3′), 124.08 (C-5′), 139.45 (C-4′), 152.36 (C-2′), 161.73 (C-1′), 167.27 (C-6′). MS (ESI, 20 eV): 324.40 (100) [M+H]+. M = 323.47, C16H21NO2S2.
2-(((1S,2R,3S,5R)-2-(hydroxymethyl)-6,6-dimethylbicyclo[3.1.1]heptan-3-yl)disulfanyl)nicotinic acid (3d). Yield: 95%. Yellow powder. M.p. 99 °C. [ α ] D 27 = −225.3 (c = 0.141, CHCl3). IR (KBr): 3238 (OH), 2985, 2926, 1707 (C=O), 1560, 1390, 1271, 1246, 1141, 1062, 1041, 808, 765, 704, 547. 1H-NMR (300 MHz, CDCl3): 0.91 (s, 2H, Me-8), 1.14 (d, 1H, J = 9.98 Hz, H-7a), 1.88 (t, 1H, J = 5.14 Hz, H-1), 1.94–2.05 (m, 1H, H-5), 2.22–2.51 (m, 3H, H-2, H-6a, H-7b), 2.70 (t, 1H, J = 12.03 Hz, H-6b), 3.47 (dt, 1H, J = 9.61, 6.79 Hz, H-3), 3.73 (dd, 1H, J = 10.86, 4.11 Hz, H-10a), 3.93 (t, 1H, J = 10.56 Hz, H-10b), 7.30 (dd, 1H, J = 7.78, 4.84 Hz, H-3′), 9.41 (dd, 1H, J = 7.92, 1.47 Hz, H-4′), 8.77 (dd, 1H, J = 4.70, 1.47 Hz, H-2′). 13C-NMR (75 MHz, CDCl3): 23.66 (C-8), 27.38 (C-9), 33.49 (C-7), 38.31 (C-4), 38.52 (C-6), 41.93 (C-5), 44.35 (C-1), 44.74 (C-3), 51.04 (C-2), 68.17 (C-10), 120.51 (C-3′), 124.17 (C-5′), 140.46 (C-4′), 152.75 (C-2′), 162.30 (C-1′), 168.31 (C-6′). MS (ESI, 20 eV): 340.12 (100) [M+H]+. M = 339.47, C16H21NO3S2.
2-((((1S,2S,3S,5R)-3-hydroxy-6,6-dimethylbicyclo[3.1.1]heptan-2-yl)methyl)disulfanyl)nicotinic acid (4d). Yield: 73%. White powder. m.p. 172–173 °C. [ α ] D 27 = +169.2 (c = 0.130, CHCl3). IR (KBr): 3256 (OH), 2985, 2920, 1703 (C=O), 1562, 1388, 1296, 1259, 1144, 1063, 1032, 810, 764, 704, 542. 1H-NMR (300 MHz, (CD3)O): 0.91 (s, 3H, Me-8), 1.09 (d, 1H, H-7a, J = 9.4 Hz), 1.77 (dt, 1H, H-4, J = 14, 3.3 Hz), 1.87–1.97 (m, 1H, H-5), 2.11–2.24 (m, 2H, H-1, 2), 2.31–2.42 (m, 1H, H-7), 2.42–2.55 (m, 1H, H-4), 2.88 (dd, 1H, H-10a, J = 12.9, 7.9 Hz), 3.14 (dd, 1H, H-140b, J = 12.9, 7.8 Hz), 4.15 (dt, 1H, H-3, J = 9, 4.4 Hz), 7.39 (dd, 1H, H-3′, J = 7.6, 4.7 Hz), 8.35 (d, 1H, H-4′, J = 7.9 Hz), 8.71 (d, 1H, H-2′, J = 4.7 Hz). 13C-NMR (75 MHz, (CD3)O): 24.03 (C-8), 27.84 (C-9), 33.67 (C-7), 38.87 (C-6), 33.19 (C-4), 42.58 (C-5), 45.48 (C-10), 45.98 (C-1), 53.52 (C-2), 69.31 (C-3), 121.51 (C-3′), 133.13 (C-5′), 140.39 (C-4′), 153.39 (C-2′), 168.78 (C-6′). MS (ESI, 20 eV): 340.14 (100) [M+H]+. M = 339.47, C16H21NO3S2.

4.3. Antioxidant Activity (Non-Cellular Model)

The antioxidant activity of monoterpenylhetaryl disulfides was evaluated for their in vitro ability to inhibit lipid peroxidation (LPO) processes in the brain lipids of laboratory mice [51,52]. The brain was removed and homogenized (10%) in physiological saline (pH = 7.4) and centrifuged for 10 min. Then, an aliquot of the supernatant (S1) was taken [51,53], into which the test compounds were added as solutions in ethanol (final concentration of 100 μM and 1 mM). After 30 min, LPO was initiated by adding a freshly prepared solution of FeSO4 and ascorbic acid [54]. Then, the test samples were incubated for one hour at 37 °C with slow stirring in a thermostatic Biosan ES-20 shaker (Biosan, Riga, Latvia). The concentration of LPO secondary products reacting with 2-thiobarbituric acid (thiobarbituric acid reactive substances, TBA-RS) was determined using ThermoSpectromic Genesys 20 spectrophotometer (Thermo Fisher Scientific, Waltham, MA, USA) at λ = 532 nm, the extinction coefficient 1.56 × 105 M−1cm−1 was used for calculations [52,55].

4.4. Erythrotoxicity, Antioxidant Activity, Membrane-Protective Activity (Cellular Model System)

To evaluate the erythrotoxicity, antioxidant and membrane-protective activity of the compounds, we used 0.5% (v/v) suspension of red blood cells from laboratory mice in phosphate-buffered saline (PBS, pH 7.4). The test substances were previously dissolved in ethanol. Erythrotoxicity (hemolytic activity) was evaluated by the degree of hemolysis of red blood cells 1–5 h after the introduction of solutions of the compounds. Membrane-protective and antioxidant activity was determined by the degree of inhibition of H2O2-induced hemolysis. For this purpose, 30 min after the erythrocyte suspension was added to the solutions of the test compounds, hemolysis was initiated with a solution of hydrogen peroxide (0.006%). Then, the reaction mixture was incubated in a thermostatic shaker Biosan ES-20 with slow stirring at 37 °C for 5 h. An aliquot was taken from the incubation medium every hour, centrifuged for 5 min (1600× g), and the degree of hemolysis was determined by the hemoglobin content in the supernatant on a ThermoSpectromic Genesys 20 spectrophotometer (Thermo Fisher Scientific, Waltham, MA, USA) at 524 nm [56]. The percentage of hemolysis was calculated in relation to the complete hemolysis of the sample. Each experiment was designed with 4–8 replicates. Statistical data processing and charting were performed using the Microsoft Office Excel 2007 software package (Microsoft, Redmond, WA, USA). The data represent the mean ± SE.

4.5. Antibacterial Activity

Minimum inhibitory concentrations (MICs) of compounds was determined by the broth microdilution assay in 96-well plates (Eppendorf, Hamburg, Germany) according to the EUCAST rules for antimicrobial susceptibility testing [57] in full Mueller-Hinton broth (MH). The concentrations of compounds to be tested ranged from 1 to 2048 µg/mL. The MIC was determined as the lowest concentration of an antibiotic for which no visible bacterial growth could be observed after 24 h of incubation.

4.6. Antifungal Activity

MICs on C. albicans was determined using the broth microdilution method in 96-well plates (Eppendorf) in MH broth as recommended in the protocol CLSI M27-A3 [58]. C. albicans was grown overnight and diluted with MH broth until optical density of 0.5 to obtain the working solution. Then, 2-fold serial dilutions of compounds in concentrations from 1 to 1024 μg/mL were prepared in MH broth and seeded with fungi (1% v/v of working solution) with subsequent incubation at 37 °C for 24 h. The MIC was defined as the lowest concentration of the compound at which no visible growth could be seen.

4.7. Mutagenicity and Cytotoxicity

The mutagenicity of compounds was evaluated in the Ames test with S. typhimurium TA98, TA100 and TA102 strains as described in [38]. The spot-test modification has been applied to avoid false-negative results driven with antibacterial activity of compounds. The tested compound was considered to be mutagenic if the number of revertant colonies increased more than 2 times close to the filter paper with compound.
Cytotoxicity of compounds was determined using the microtetrazolium test (MTT) on EBL cells. The cells were cultured in DMEM—Dulbecco’s Modified Eagle’s Medium (Sigma Aldrich, St. Louis, MO, USA) supplemented with 10% FBS, 2 mM L-glutamine, 100 µg/mL penicillin and 100 µg/mL streptomycin. Cells were seeded in 96-well plates with the density of 3000 cells per well and left overnight to allow for the attachment. Cells were next cultured at 37 °C and 5% CO2 in the presence of compounds of interest at various concentrations from 1.25 to 160 µg/mL. After 24 h of cultivation the cells were subjected to MTT-assay. The formazan was solubilized by DMSO and the optical density was measured on Tecan Infinite 200Pro at 570 nm. The concentration required to inhibit cellular dehydrogenase activity by 50% (CC50 value) was calculated by using GraphPad Prism 6.0 software (GraphPad software, San Diego, CA, USA).

4.8. Antiviral Activity Assessment

4.8.1. Toxicity Studies

MTT [59] was used to study cytotoxicity of the compounds. Briefly, series of three-fold dilutions of each compound (300–4 µg/mL) in MEM were prepared. MDCK cells (ATCC CCL-34) were incubated for 72 h at 36 °C in 5% CO2 in the presence of the dissolved substances. The degree of destruction of the cell monolayer was then evaluated in the MTT. The cells were washed twice with saline, and a solution of 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (0.5 mg/mL) in cell culture medium was added to the wells. After 1h incubation, the wells were washed and the formazan residue dissolved in DMSO (0.1 mL per well). The optical density of wells was then measured on a Multiscan FC photometer at wavelength of 540 nm and plotted against concentration of the compounds. Each concentration was tested in three parallels. The 50% cytotoxic dose (CC50) of each compound (i.e., the compound concentration that causes the death of 50% cells in a culture, or decreasing the optical density twice as compared to the control wells) was calculated from the data obtained. Values of CC50 obtained in µg/mL were then calculated into micromoles.

4.8.2. Cell Protection Assay

The compounds in appropriate concentrations were added to MDCK cells (0.1 mL per well). Cells were further infected with A/Puerto Rico/8/34 (H1N1) influenza virus (m.o.i 0.01). Plates were incubated for 72 h at 36 °C at 5% CO2. After that, cell viability was assessed by MTT test as described above. The cytoprotective activity of compounds was considered as their ability to increase the values of IC50 compared to control wells (with virus only, no drugs). Based on the results obtained, the values of IC50, i.e., concentration of compounds that results in 50% cells protection, were calculated using GraphPad Prism software. Values of IC50 obtained in μg/mL were then calculated into μM. For each compound, the value of selectivity index (SI) was calculated as ratio of CC50 to IC50. The compounds with SI of 10 and higher were considered as active [39].

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/molecules27165101/s1. The 1H NMR, 13C NMR and IR spectra of novel compounds.

Author Contributions

Conceptualization, S.A.R. and D.V.S.; methodology, D.V.S., A.V.B., R.V.R., O.G.S., A.R.K. and V.V.Z.; validation, D.V.S.; formal analysis, D.V.S.; investigation, Y.V.G., J.M.L., R.V.R., O.G.S., D.R.B., N.D.Z., E.O.S. and A.S.V.; data curation, Y.V.G., A.V.B., R.V.R., O.G.S., A.R.K., V.V.Z. and D.V.S.; writing—original draft preparation, Y.V.G. and D.V.S.; writing—review and editing, D.V.S.; visualization, D.V.S.; supervision, S.A.R.; project administration, S.A.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Russian Science Foundation, grant number 21-13-00245.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was supported by Kazan Federal University Strategic Academic Leadership Program (PRIORITY-2030) (antimicrobial part). This work was performed using equipment from the Center for Collective Use “Chemistry” of the Institute of Chemistry of the Komi Scientific Center, Ural Branch of the Russian Academy of Sciences. The study of antioxidant and membrane-protective activity was performed using the equipment of the Center of Collective Usage “Molecular Biology” of the Institute of Biology of the Komi Scientific Center of the Ural Branch of the Russian Academy of Sciences at the expense of the state assignment (project No. 122040600022-1). The used animals were taken from the scientific collection of experimental animals from the Institute of Biology of the Komi Scientific Center, Ural Branch of the Russian Academy of Sciences (http://www.ckp-rf.ru/usu/471933/, accessed on 28 July 2022). The X-ray investigation of compound 1c was performed using the equipment of the center for collective use “Analytical Center of the IOMC RAS” with the financial support of the grant “Ensuring the development of the material and technical infrastructure of the centers for collective use of scientific equipment” (Unique identifier RF-2296.61321X0017, Agreement Number 075-15-2021-670).

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors, but can be specially prepared upon prior request.

References

  1. Koval’, I.V. The Chemistry of Disulfides. Russ. Chem. Rev. 1994, 63, 735–750. [Google Scholar] [CrossRef]
  2. Heldreth, B.; Turos, E. Microbiological Properties and Modes of Action of Organosulfur-Based Anti-Infectives. Curr. Med. Chem. Anticancer Agents 2005, 4, 295–315. [Google Scholar] [CrossRef]
  3. Chung, J.; Chen, T.; Missiakas, D. Transfer of Electrons across the Cytoplasmic Membrane by DsbD, a Membrane Protein Involved in Thiol-Disulphide Exchange and Protein Folding in the Bacterial Periplasm. Mol. Microbiol. 2000, 35, 1099–1109. [Google Scholar] [CrossRef]
  4. Song, X.; Yue, Z.; Nie, L.; Zhao, P.; Zhu, K.; Wang, Q. Biological Functions of Diallyl Disulfide, a Garlic-Derived Natural Organic Sulfur Compound. J. Evid. Based Complement. Altern. Med. 2021, 2021, 1–13. [Google Scholar] [CrossRef]
  5. Koh, S.-H.; Kwon, H.; Park, K.H.; Ko, J.K.; Kim, J.H.; Hwang, M.S.; Yum, Y.N.; Kim, O.-H.; Kim, J.; Kim, H.-T.; et al. Protective Effect of Diallyl Disulfide on Oxidative Stress-Injured Neuronally Differentiated PC12 Cells. Mol. Brain Res. 2005, 133, 176–186. [Google Scholar] [CrossRef]
  6. Filomeni, G.; Aquilano, K.; Rotilio, G.; Ciriolo, M.R. Reactive Oxygen Species-Dependent c-Jun NH2-Terminal Kinase/c-Jun Signaling Cascade Mediates Neuroblastoma Cell Death Induced by Diallyl Disulfide. Cancer Res. 2003, 63, 5940–5949. [Google Scholar]
  7. Agassi, S.F.T.; Yeh, T.-M.; Chang, C.-D.; Hsu, J.-L.; Shih, W.-L. Potentiation of Differentiation and Apoptosis in a Human Promyelocytic Leukemia Cell Line by Garlic Essential Oil and Its Organosulfur Compounds. Anticancer Res. 2020, 40, 6345–6354. [Google Scholar] [CrossRef]
  8. Li, W.-R.; Ma, Y.-K.; Xie, X.-B.; Shi, Q.-S.; Wen, X.; Sun, T.-L.; Peng, H. Diallyl Disulfide from Garlic Oil Inhibits Pseudomonas Aeruginosa Quorum Sensing Systems and Corresponding Virulence Factors. Front. Microbiol. 2019, 9, 3222. [Google Scholar] [CrossRef]
  9. Tsao, S.-M.; Liu, W.-H.; Yin, M.-C. Two Diallyl Sulphides Derived from Garlic Inhibit Meticillin-Resistant Staphylococcus Aureus Infection in Diabetic Mice. J. Med. Microbiol. 2007, 56, 803–808. [Google Scholar] [CrossRef] [Green Version]
  10. Casella, S.; Leonardi, M.; Melai, B.; Fratini, F.; Pistelli, L. The Role of Diallyl Sulfides and Dipropyl Sulfides in the In Vitro Antimicrobial Activity of the Essential Oil of Garlic, Allium Sativum L., and Leek, Allium Porrum L. Phytother. Res. 2013, 27, 380–383. [Google Scholar] [CrossRef]
  11. Maldonado, P.D.; Chánez-Cárdenas, M.E.; Pedraza-Chaverrí, J. Aged Garlic Extract, Garlic Powder Extract, S-Allylcysteine, Diallyl Sulfide and Diallyl Disulfide Do Not Interfere with the Antibiotic Activity of Gentamicin. Phytother. Res. 2005, 19, 252–254. [Google Scholar] [CrossRef]
  12. Liu, W.; Hsu, C.; Yin, M. In Vitro Anti-Helicobacter Pylori Activity of Diallyl Sulphides and Protocatechuic Acid. Phytother. Res. 2008, 22, 53–57. [Google Scholar] [CrossRef]
  13. Chung, J.G.; Chen, G.W.; Wu, L.T.; Chang, H.L.; Lin, J.G.; Yeh, C.C.; Wang, T.F. Effects of Garlic Compounds Diallyl Sulfide and Diallyl Disulfide on Arylamine N-Acetyltransferase Activity in Strains of Helicobacter Pylori from Peptic Ulcer Patients. Am. J. Chin. Med. 1998, 26, 353–364. [Google Scholar] [CrossRef]
  14. Alam, M.; Zubair, S.; Farazuddin, M.; Ahmad, E.; Khan, A.; Zia, Q.; Malik, A.; Mohammad, O. Development, Characterization and Efficacy of Niosomal Diallyl Disulfide in Treatment of Disseminated Murine Candidiasis. Nanomed. Nanotechnol. Biol. Med. 2013, 9, 247–256. [Google Scholar] [CrossRef]
  15. Kocić-Tanackov, S.; Dimić, G.; Lević, J.; Tanackov, I.; Tepić, A.; Vujičić, B.; Gvozdanović-Varga, J. Effects of Onion (Allium cepa L.) and Garlic (Allium sativum L.) Essential Oils on the Aspergillus Versicolor Growth and Sterigmatocystin Production. J. Food Sci. 2012, 77, M278–M284. [Google Scholar] [CrossRef]
  16. Shoji, S.; Furuishi, K.; Yanase, R.; Miyazaka, T.; Kino, M. Allyl Compounds Selectively Killed Human Immunodeficiency Virus (Type 1)-Infected Cells. Biochem. Biophys. Res. Commun. 1993, 194, 610–621. [Google Scholar] [CrossRef]
  17. Hall, A.; Troupin, A.; Londono-Renteria, B.; Colpitts, T. Garlic Organosulfur Compounds Reduce Inflammation and Oxidative Stress during Dengue Virus Infection. Viruses 2017, 9, 159. [Google Scholar] [CrossRef]
  18. Nicolaou, K.C.; Hughes, R.; Pfefferkorn, J.A.; Barluenga, S.; Roecker, A.J. Combinatorial Synthesis through Disulfide Exchange: Discovery of Potent Psammaplin a Type Antibacterial Agents Active against Methicillin-ResistantStaphylococcus Aureus (MRSA). Chem. Eur. J. 2001, 7, 4280–4295. [Google Scholar] [CrossRef]
  19. Jing, Q.; Hu, X.; Ma, Y.; Mu, J.; Liu, W.; Xu, F.; Li, Z.; Bai, J.; Hua, H.; Li, D. Marine-Derived Natural Lead Compound Disulfide-Linked Dimer Psammaplin A: Biological Activity and Structural Modification. Mar. Drugs 2019, 17, 384. [Google Scholar] [CrossRef] [Green Version]
  20. Moriarty, R.M.; Roll, D.M.; Ku, Y.-Y.; Nelson, C.; Ireland, C.M. A Revised Structure for the Marine Bromoindole Derivative Citorellamine. Tetrahedron Lett. 1987, 28, 749–752. [Google Scholar] [CrossRef]
  21. Roll, D.M.; Ireland, C.M. Citorellamine, a New Bromoindole Derivative from Polycitorella-Mariae. Tetrahedron Lett. 1985, 26, 4303–4306. [Google Scholar] [CrossRef]
  22. Long, T.E. Repurposing Thiram and Disulfiram as Antibacterial Agents for Multidrug-Resistant Staphylococcus Aureus Infections. Antimicrob. Agents Chemother. 2017, 61, e00898-17. [Google Scholar] [CrossRef] [Green Version]
  23. Sheppard, J.G.; Frazier, K.R.; Saralkar, P.; Hossain, M.F.; Geldenhuys, W.J.; Long, T.E. Disulfiram-Based Disulfides as Narrow-Spectrum Antibacterial Agents. Bioorg. Med. Chem. Lett. 2018, 28, 1298–1302. [Google Scholar] [CrossRef] [Green Version]
  24. Potula, H.-H.S.K.; Shahryari, J.; Inayathullah, M.; Malkovskiy, A.V.; Kim, K.-M.; Rajadas, J. Repurposing Disulfiram (Tetraethylthiuram Disulfide) as a Potential Drug Candidate against Borrelia Burgdorferi In Vitro and In Vivo. Antibiotics 2020, 9, 633. [Google Scholar] [CrossRef]
  25. Liegner, K.B. Disulfiram (Tetraethylthiuram Disulfide) in the Treatment of Lyme Disease and Babesiosis: Report of Experience in Three Cases. Antibiotics 2019, 8, 72. [Google Scholar] [CrossRef] [Green Version]
  26. Turos, E.; Revell, K.D.; Ramaraju, P.; Gergeres, D.A.; Greenhalgh, K.; Young, A.; Sathyanarayan, N.; Dickey, S.; Lim, D.; Alhamadsheh, M.M. Unsymmetric Aryl–Alkyl Disulfide Growth Inhibitors of Methicillin-Resistant Staphylococcus Aureus and Bacillus Anthracis. Bioorg. Med. Chem. 2008, 16, 6501–6508. [Google Scholar] [CrossRef] [Green Version]
  27. Branowska, D.; Ławecka, J.; Sobiczewski, M.; Karczmarzyk, Z.; Wysocki, W.; Wolińska, E.; Olender, E.; Mirosław, B.; Perzyna, A.; Bielawska, A.; et al. Synthesis of Unsymmetrical Disulfanes Bearing 1,2,4-Triazine Scaffold and Their in Vitro Screening towards Anti-Breast Cancer Activity. Monatsh. Chem. 2018, 149, 1409–1420. [Google Scholar] [CrossRef] [Green Version]
  28. Wang, L.; Bao, B.-B.; Song, G.-Q.; Chen, C.; Zhang, X.-M.; Lu, W.; Wang, Z.; Cai, Y.; Li, S.; Fu, S.; et al. Discovery of Unsymmetrical Aromatic Disulfides as Novel Inhibitors of SARS-CoV Main Protease: Chemical Synthesis, Biological Evaluation, Molecular Docking and 3D-QSAR Study. Eur. J. Med. Chem. 2017, 137, 450–461. [Google Scholar] [CrossRef]
  29. Toropov, A.A.; Toropova, A.P.; Veselinović, A.M.; Leszczynska, D.; Leszczynski, J. SARS-CoV M pro Inhibitory Activity of Aromatic Disulfide Compounds: QSAR Model. J. Biomol. Struct. Dyn. 2022, 40, 780–786. [Google Scholar] [CrossRef]
  30. Pestova, S.V.; Izmest’ev, E.S.; Shevchenko, O.G.; Rubtsova, S.A.; Kuchin, A.V. Synthesis and Membranoprotective Properties of New Disulfides with Monoterpene and Carbohydrate Fragments. Russ. Chem. Bull. 2015, 64, 723–731. [Google Scholar] [CrossRef]
  31. Gyrdymova, Y.V.; Sudarikov, D.V.; Shevchenko, O.G.; Rubtsova, S.A.; Slepukhin, P.A.; Kutchin, A.V. Caryophyllane Thiols, Vinyl Thioethers, Di- and Bis-Sulfides: Antioxidant and Membrane Protective Activities. Chem. Biodivers. 2017, 14, e1700296. [Google Scholar] [CrossRef]
  32. Gyrdymova, Y.V.; Sinegubova, E.O.; Muryleva, A.S.; Zarubaev, V.V.; Rubtsova, S.A. Correction to: Anti-Influenza Activity of Several Caryophyllane Thiosesquiterpenoids. Chem. Nat. Compd. 2020, 56, 191. [Google Scholar] [CrossRef] [Green Version]
  33. Startseva, V.A.; Kuznetsov, I.V.; Nikitina, L.E.; Osmanov, V.K.; Borisov, A.V.; Matsulevich, Z.V.; Klochkov, V.V. Hetarenesulfenyl(Selenyl) Chlorination of (+)-Camphene. Chem. Nat. Compd. 2015, 51, 671–674. [Google Scholar] [CrossRef]
  34. Pugh, K.C.; Gera, L.; Stewart, J.M. Synthesis and Stability of 3-Nitro-2-Pyridinesulfenyl Chloride (NpysCl). Int. J. Pept. Protein Res. 2009, 42, 159–164. [Google Scholar] [CrossRef]
  35. Sagara, T.; Okamura, M.; Shimohigashi, Y.; Ohno, M.; Kanematsu, K. Specific Affinity Labeling of μ Opioid Receptors in Rat Brain by S-Activated Sulfhydryldihydromorphine Analogs. Bioorg. Med. Chem. Lett. 1995, 5, 1609–1614. [Google Scholar] [CrossRef]
  36. Senter, P.D.; Pearce, W.E.; Greenfield, R.S. Development of a Drug-Release Strategy Based on the Reductive Fragmentation of Benzyl Carbamate Disulfides. J. Org. Chem. 1990, 55, 2975–2978. [Google Scholar] [CrossRef]
  37. Perlikowska, W.; Gouygou, M.; Mikolajczyk, M.; Daran, J.-C. Enantiomerically Pure Disulfides: Key Compounds in the Kinetic Resolution of Chiral PIII-Derivatives with Stereogenic Phosphorus. Tetrahedron Asymmetry 2004, 15, 3519–3529. [Google Scholar] [CrossRef]
  38. McCann, J.; Ames, B.N. A Simple Method for Detecting Environmental Carcinogens as Mutagens. Ann. N. Y. Acad. Sci. 1976, 271, 5–13. [Google Scholar] [CrossRef]
  39. Smee, D.F.; Hurst, B.L.; Evans, W.J.; Clyde, N.; Wright, S.; Peterson, C.; Jung, K.-H.; Day, C.W. Evaluation of Cell Viability Dyes in Antiviral Assays with RNA Viruses That Exhibit Different Cytopathogenic Properties. J. Virol. Methods 2017, 246, 51–57. [Google Scholar] [CrossRef]
  40. SAINT. Data Reduction and Correction Program; Bruker AXS: Madison, WI, USA, 2014. [Google Scholar]
  41. Sheldrick, G.M. SHELXT—Integrated Space-Group and Crystal-Structure Determination. Acta Crystallogr. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  42. Sheldrick, G.M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  43. Krause, L.; Herbst-Irmer, R.; Sheldrick, G.M.; Stalke, D. Comparison of Silver and Molybdenum Microfocus X-ray Sources for Single-Crystal Structure Determination. J. Appl. Crystallogr. 2015, 48, 3–10. [Google Scholar] [CrossRef] [Green Version]
  44. Singh, W.M.; Baruah, J.B. Rosolic Acid–Catalyzed Conversion of Thiols to Disulfides. Synth. Commun. 2008, 39, 325–331. [Google Scholar] [CrossRef]
  45. Stoyanov, S.; Petkov, I.; Antonov, L.; Stoyanova, T.; Karagiannidis, P.; Aslanidis, P. Thione–Thiol Tautomerism and Stability of 2- and 4-Mercaptopyridines and 2-Mercaptopyrimidines. Can. J. Chem. 1990, 68, 1482–1489. [Google Scholar] [CrossRef]
  46. Izmest’ev, E.S.; Sudarikov, D.V.; Rubtsova, S.A.; Slepukhin, P.A.; Kuchin, A.V. Asymmetric Synthesis of New Optically Active Sulfinamides of Menthane Series and Their Derivatives. Russ. J. Org. Chem. 2012, 48, 184–192. [Google Scholar] [CrossRef]
  47. Banach, A.; Ścianowski, J.; Ozimek, P. The Use of Sulfides Derived from Carane, P-Menthane, Pinane, and Bornane in the Synthesis of Optically Active Epoxides. Phosphorus Sulfur Silicon Relat. Elem. 2014, 189, 274–284. [Google Scholar] [CrossRef]
  48. Martıínez-Ramos, F.; Vargas-Dıíaz, M.E.; Chacón-Garcıía, L.; Tamariz, J.; Joseph-Nathan, P.; Zepeda, L.G. Highly Diastereoselective Nucleophilic Additions Using a Novel Myrtenal-Derived Oxathiane as a Chiral Auxiliary. Tetrahedron Asymmetry 2001, 12, 3095–3103. [Google Scholar] [CrossRef]
  49. Banina, O.A.; Sudarikov, D.V.; Krymskaya, Y.V.; Frolova, L.L.; Kuchin, A.V. Synthesis of Chiral Hydroxythiols Based on Oxygen-Containing α- and β-Pinene Derivatives. Chem. Nat. Compd. 2015, 51, 261–265. [Google Scholar] [CrossRef]
  50. Hayashi, K.; Ichimaru, Y.; Sugiura, K.; Maeda, A.; Harada, Y.; Kojima, Y.; Nakayama, K.; Imai, M. Efficiency of Lithium Cations in Hydrolysis Reactions of Esters in Aqueous Tetrahydrofuran. Chem. Pharm. Bull. 2021, 69, 581–584. [Google Scholar] [CrossRef]
  51. Acker, C.I.; Brandão, R.; Rosário, A.R.; Nogueira, C.W. Antioxidant Effect of Alkynylselenoalcohol Compounds on Liver and Brain of Rats in Vitro. Environ. Toxicol. Pharmacol. 2009, 28, 280–287. [Google Scholar] [CrossRef]
  52. Stefanello, S.T.; Prestes, A.S.; Ogunmoyole, T.; Salman, S.M.; Schwab, R.S.; Brender, C.R.; Dornelles, L.; Rocha, J.B.T.; Soares, F.A.A. Evaluation of in Vitro Antioxidant Effect of New Mono and Diselenides. Toxicol. In Vitro 2013, 27, 1433–1439. [Google Scholar] [CrossRef] [Green Version]
  53. Bellé, N.A.V.; Dalmolin, G.D.; Fonini, G.; Rubin, M.A.; Rocha, J.B.T. Polyamines Reduces Lipid Peroxidation Induced by Different Pro-Oxidant Agents. Brain Res. 2004, 1008, 245–251. [Google Scholar] [CrossRef]
  54. Chawla, R.; Arora, R.; Kumar, R.; Sharma, A.; Prasad, J.; Singh, S.; Sagar, R.; Chaudhary, P.; Shukla, S.; Kaur, G.; et al. Antioxidant Activity of Fractionated Extracts of Rhizomes of High-Altitude Podophyllum Hexandrum: Role in Radiation Protection. Mol. Cell. Biochem. 2005, 273, 193–208. [Google Scholar] [CrossRef]
  55. Asakawa, T.; Matsushita, S. Coloring Conditions of Thiobarbituric Acid Test for Detecting Lipid Hydroperoxides. Lipids 1980, 15, 137–140. [Google Scholar] [CrossRef]
  56. Takebayashi, J.; Chen, J.; Tai, A. A Method for Evaluation of Antioxidant Activity Based on Inhibition of Free Radical-Induced Erythrocyte Hemolysis. In Advanced Protocols in Oxidative Stress II; Armstrong, D., Ed.; Humana Press: Totowa, NJ, USA, 2010; Volume 594, pp. 287–296. ISBN 978-1-60761-410-4. [Google Scholar]
  57. Leclercq, R.; Cantón, R.; Brown, D.F.J.; Giske, C.G.; Heisig, P.; MacGowan, A.P.; Mouton, J.W.; Nordmann, P.; Rodloff, A.C.; Rossolini, G.M.; et al. EUCAST Expert Rules in Antimicrobial Susceptibility Testing. Clin. Microbiol. Infect. 2013, 19, 141–160. [Google Scholar] [CrossRef] [Green Version]
  58. Rex, J.H. Reference Method for Broth Dilution Antifungal Susceptibility Testing of Yeats: Approved Standard, 3rd ed.; Clinical and Laboratory Standards Institute (CLSI): Wayne, PA, USA, 2008; pp. 1–25. [Google Scholar]
  59. Mosmann, T. Rapid Colorimetric Assay for Cellular Growth and Survival: Application to Proliferation and Cytotoxicity Assays. J. Immunol. Methods 1983, 65, 55–63. [Google Scholar] [CrossRef]
Scheme 1. General scheme for the synthesis of unsymmetrical disulfides.
Scheme 1. General scheme for the synthesis of unsymmetrical disulfides.
Molecules 27 05101 sch001
Figure 1. Molecular structure of methyl 2-[[(1S,2S,5R)-2-isopropyl-5methylcyclohexyl]disulfanyl]pyridine-3-carboxylate (1c) with thermal ellipsoids drawn at the 30% probability level.
Figure 1. Molecular structure of methyl 2-[[(1S,2S,5R)-2-isopropyl-5methylcyclohexyl]disulfanyl]pyridine-3-carboxylate (1c) with thermal ellipsoids drawn at the 30% probability level.
Molecules 27 05101 g001
Scheme 2. Hydrolysis of esters 1c–4c.
Scheme 2. Hydrolysis of esters 1c–4c.
Molecules 27 05101 sch002
Figure 2. Effect of unsymmetrical disulfides and Trolox (T) at concentrations of 100 μM and 1 mM on the content of TBA-RS in the brain homogenate in an hour after the initiation of LPO (ascorbate/Fe2+ initiators). C—control sample containing no test compounds; I—intact sample in which LPO was not initiated.
Figure 2. Effect of unsymmetrical disulfides and Trolox (T) at concentrations of 100 μM and 1 mM on the content of TBA-RS in the brain homogenate in an hour after the initiation of LPO (ascorbate/Fe2+ initiators). C—control sample containing no test compounds; I—intact sample in which LPO was not initiated.
Molecules 27 05101 g002
Figure 3. Hemolytic activity of unsymmetrical disulfides at the concentration of 100 μM (A) and 1 μM (B) after 1, 3 and 5 h of incubation.
Figure 3. Hemolytic activity of unsymmetrical disulfides at the concentration of 100 μM (A) and 1 μM (B) after 1, 3 and 5 h of incubation.
Molecules 27 05101 g003aMolecules 27 05101 g003b
Figure 4. Effect of unsymmetrical disulfides and Trolox (T) at a concentration of 1 μM on the level of H2O2-induced hemolysis of erythrocytes after 1–5 h of incubation.
Figure 4. Effect of unsymmetrical disulfides and Trolox (T) at a concentration of 1 μM on the level of H2O2-induced hemolysis of erythrocytes after 1–5 h of incubation.
Molecules 27 05101 g004
Table 1. Antibacterial, antifungal activity, cytotoxicity and mutagenicity of unsymmetrical disulfides.
Table 1. Antibacterial, antifungal activity, cytotoxicity and mutagenicity of unsymmetrical disulfides.
CompoundMIC, µg/mLCC50
EBL, μg/mL
Mutagenicity in the Ames Test
S. aureus ATCC 29213 (MSSA)S. aureus MRSA Clinical IsolateP. aeruginosa ATCC 27853C. albicans 703 Clinical Isolate
1. 1a16161651211.7TA 100
2. 1b16323251214.0TA 102
3. 1c32323251218.3NF 2
4. 1d326410246447.8NF
5. 2a128128641024104.3NF
6. 2b166432>102413.6NF
7. 2c>10246410241024122.6NF
8. 2d32512102412818.1NF
9. 3a32326410247.1NF
10. 3b323210245128.2NF
11. 3c6432102410248.1NF
12. 3d5121024102432117.7NF
13. 4a256>10241024>1024ND 1ND
14. 4b>1024>1024>1024>1024NDND
15. 4c1024>1024>10241024NDND
16. 4d5126410241611.14NF
Amikacin444NDNDND
FluconazoleNDNDND16NDND
1 ND—not determined; 2 NF—not found.
Table 2. Cytotoxicity and antiviral activity of monoterpenylhetaryl disulfides.
Table 2. Cytotoxicity and antiviral activity of monoterpenylhetaryl disulfides.
CompoundCC50, µMIC50, µMSI
1. 1a36 ± 35 ± 18
2. 1b29 ± 213 ± 22
3. 1c31 ± 212 ± 23
4. 1d38 ± 3>341
5. 2a99 ± 513 ± 38
6. 2b35 ± 313 ± 23
7. 2c104 ± 819 ± 36
8. 2d60 ± 45 ± 112
9. 3a30 ± 39 ± 34
10. 3b83 ± 416 ± 35
11. 3c>283255 ± 311
12. 3d285 ± 14132 ± 162
13. 4a68 ± 6>312
14. 4b34 ± 212 ± 33
15. 4c40 ± 234 ± 41
16. 4d178 ± 11>972
Rimantadine367 ± 2857 ± 86
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sudarikov, D.V.; Gyrdymova, Y.V.; Borisov, A.V.; Lukiyanova, J.M.; Rumyantcev, R.V.; Shevchenko, O.G.; Baidamshina, D.R.; Zakarova, N.D.; Kayumov, A.R.; Sinegubova, E.O.; et al. Synthesis and Biological Activity of Unsymmetrical Monoterpenylhetaryl Disulfides. Molecules 2022, 27, 5101. https://doi.org/10.3390/molecules27165101

AMA Style

Sudarikov DV, Gyrdymova YV, Borisov AV, Lukiyanova JM, Rumyantcev RV, Shevchenko OG, Baidamshina DR, Zakarova ND, Kayumov AR, Sinegubova EO, et al. Synthesis and Biological Activity of Unsymmetrical Monoterpenylhetaryl Disulfides. Molecules. 2022; 27(16):5101. https://doi.org/10.3390/molecules27165101

Chicago/Turabian Style

Sudarikov, Denis V., Yulia V. Gyrdymova, Alexander V. Borisov, Julia M. Lukiyanova, Roman V. Rumyantcev, Oksana G. Shevchenko, Diana R. Baidamshina, Nargiza D. Zakarova, Airat R. Kayumov, Ekaterina O. Sinegubova, and et al. 2022. "Synthesis and Biological Activity of Unsymmetrical Monoterpenylhetaryl Disulfides" Molecules 27, no. 16: 5101. https://doi.org/10.3390/molecules27165101

Article Metrics

Back to TopTop