Next Article in Journal
Cobalt–Carbon Nanoparticles with Silica Support for Uptake of Cationic and Anionic Dyes from Polluted Water
Next Article in Special Issue
Synthesis and Antimicrobial Activity of δ-Viniferin Analogues and Isosteres
Previous Article in Journal
Machine-Learning-Enabled Virtual Screening for Inhibitors of Lysine-Specific Histone Demethylase 1
Previous Article in Special Issue
A Novel Phenylpyrrolidine Derivative: Synthesis and Effect on Cognitive Functions in Rats with Experimental Ishemic Stroke
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Monoterpene-Containing Substituted Coumarins as Inhibitors of Respiratory Syncytial Virus (RSV) Replication

by
Tatyana M. Khomenko
1,
Anna A. Shtro
2,
Anastasia V. Galochkina
2,
Yulia V. Nikolaeva
2,
Galina D. Petukhova
2,
Sophia S. Borisevich
3,
Dina V. Korchagina
1,
Konstantin P. Volcho
1,* and
Nariman F. Salakhutdinov
1
1
Department of Medicinal Chemistry, N.N. Vorozhtsov Novosibirsk Institute of Organic Chemistry, Acad. Lavrentjev Ave. 9, 630090 Novosibirsk, Russia
2
Laboratory of Chemotherapy for Viral Infections, Smorodintsev Research Intitute of Influenza, Professor Popova Str., 15/17, 197376 St. Petersburg, Russia
3
Laboratory of Physical Chemistry, Ufa Chemistry Institute of the Ufa Federal Research Center, 71 Octyabrya pr., 450054 Ufa, Russia
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(24), 7493; https://doi.org/10.3390/molecules26247493
Submission received: 11 November 2021 / Revised: 7 December 2021 / Accepted: 9 December 2021 / Published: 10 December 2021

Abstract

:
Respiratory syncytial virus (RSV) is a critical cause of infant mortality. However, there are no vaccines and adequate drugs for its treatment. We showed, for the first time, that O-linked coumarin–monoterpene conjugates are effective RSV inhibitors. The most potent compounds are active against both RSV serotypes, A and B. According to the results of the time-of-addition experiment, the conjugates act at the early stages of virus cycle. Based on molecular modelling data, RSV F protein may be considered as a possible target.

1. Introduction

Respiratory syncytial virus (RSV), which belongs to the Pneumoviridae family, is an enveloped negative-sense RNA virus with two major serotypes, A and B [1]. RSV infects the respiratory tract, causing annual epidemics during the cold season. Despite a relatively low variability of this virus, the immunity to it is unstable, which may cause repeated infections of the same individual throughout life. It usually causes a cold. Patients under the age of 2 years, especially those born prematurely or with a heart condition, as well as the elderly, develop different symptoms of respiratory syncytial infection [2,3,4]. These age groups present with involvement of not only the upper respiratory tract, but the lower respiratory tract as well, and develop severe bronchiolitis and pneumonia, which can lead to death. RSV is the most common cause of bronchiolitis and pneumonia in children younger than one year of age [5]. According to a meta-analysis carried out in 2010, the number of children who may die from this disease is estimated to be about 199,000 annually [6]. RSV-associated childhood respiratory illness has become a challenge since the summer of 2021, when the number of cases increased sharply, which may have been due to the relaxation of COVID-19 quarantine measures [7]. There is no vaccine for RSV. Therapy for respiratory syncytial infection is usually symptomatic. The only treatment option is a non-specific and poorly effective antiviral agent: ribavirin [8,9]. In addition, the monoclonal antibody Palivizumab is approved for prophylaxis, but it is expensive and only moderately effective at reducing hospitalization rates.
In recent decades, there has been significant progress in the identification of potential targets for RSV therapy and the search for low-molecular-weight inhibitors of RSV replication; the results of these studies were summarized in a review published in 2019 [10]. Of particular interest are new low-molecular-weight agents against RSV, such as (perylen-3-ylethynyl)uracil derivative 1 [11], furanoxazine-fused benzimidazole 2 [12], and benzamide 3 [13] (Figure 1), which inhibit RSV replication by acting on cellular targets, as well as JNJ-53718678 [14] and sisunatovir [15], which are under second phase of clinical trials and are highly effective inhibitors of the fusion (F) protein. The F protein is essential for viral entry into the host cell.
Many coumarin derivatives display a variety of biological activities [16,17,18,19,20], in particular antiviral activity [21,22,23,24,25,26,27]. Tetrahydroisoquinoline 4 is the only coumarin derivatives which exhibit anti-RSV activity [27]. Attachment of monoterpenoid fragments to parent molecules is known to significantly enhance their antiviral activity [28,29]. Recently, we have found that monoterpene-containing substituted 7-hydroxycoumarins effectively inhibit H1N1 influenza virus replication, with compound 5b being most active [30]. Conjugate 5a was found to exhibit the highest activity when added to infected cell culture at early stages of viral reproduction, probably due to the interaction with viral hemagglutinin. Replacement of the monoterpenoid moiety with a benzyl substituent led to a complete loss of antiviral activity. However, there are no data on RSV inhibitory activity of coumarin derivatives comprising a terpene moiety. In this study, we revealed the ability of monoterpene–coumarin conjugates to inhibit RSV replication and suggested a possible mechanism of their action.

2. Results and Discussion

2.1. Chemistry

Our earlier study on the activity of monoterpene–coumarin conjugates against the influenza virus revealed that the structure and absolute configuration of a monoterpene moiety and the size of an annulated aliphatic ring had a significant effect on the antiviral activity. In addition, as the length of an aliphatic bridge between monoterpene and coumarin moieties increases from one to two CH2-groups, the antiviral activity enhances [30]. Given these data, in this study we synthesized a library of compounds that included both previously obtained coumarin–monoterpenoid hybrids [30,31] and new compounds with coumarin and monoterpene bicyclic fragments separated from each other by three or four CH2-groups, as well as a number of nitrogen-containing coumarin derivatives with an NH2-group instead of an OH-group.
7-Hydroxycoumarin derivatives were prepared from commercially available umbelliferone 6 and 4-methyl-7-hydroxycoumarin 7, as well as coumarins 9 and 10 (Scheme 1) synthesized from resorcinol 8 according to [31].
Bromides 11ac were synthesized in the reaction between appropriate alcohols and PBr3 in accordance with [31], using geraniol, (–)-myrtenol, and (+)-myrtenol produced from (+)-α-pinene as starting monoterpenoids (Scheme 2). Nopol bromide 11d was previously obtained by the method presented in [31] in a low yield; therefore, the NBS/PPh3 system [32] was used to synthesize compound 11d.
Derivatives of (–)-α-pinene and its homologue (–)-nopol exhibited the highest activity against the influenza virus; therefore, we synthesized compounds containing an extended aliphatic chain and an (–)-α-pinene moiety. A homolog of nopol bromide 11d, which contained an additional CH2-group, was synthesized according to the following scheme: (+)-Trans-pinocarveol, which was prepared according to [33], reacted with triethyl orthoacetate in the presence of hexanoic acid to form ester 12, which was then reduced by LiAlH4 to alcohol 13 [34]. Bromide 11e was prepared by treating alcohol 13 with NBS/PPh3.
Another nopol bromide homolog containing two additional CH2-groups, compound 11j, was synthesized as follows (Scheme 3): (−)-β-Pinene reacted with acrolein in the presence of ZnBr2 to form aldehyde 14 (conversion, 55%; yield after chromatography, 25%), which was reduced by NaBH4 to alcohol 15 [35]. The interaction of alcohol 15 with NBS/PPh3 resulted in bromide 11f [35]. 1-(Bromomethyl)-3-methoxybenzene 11g was synthesized from 3-methoxybenzaldehyde; in addition, benzyl bromide 11h was used.
Aurapten 16a and other coumarin derivatives 1619 were prepared in the reaction of 7-hydroxycoumarins 610 with appropriate bromides 11 (Scheme 4), as described elsewhere [30]. The products were purified by recrystallization or column chromatography (yields 24–80%). The reaction of nopyl bromide 11d with methylcoumarin 7 failed due to the formation of a complex reaction mixture with a high resinification level.
7-Aminocoumarins 23 and 24 were synthesized according to the method presented in [36], starting from 3-aminophenol 20, via intermediate compound 21, and its interaction with the appropriate ketoesters (Scheme 5).
Furthermore, compound 23 reacted with (–)-nopol aldehyde, and on reduction with NaBH3CN, produced secondary amine 25. In addition, acylation of amine 23 with acetic anhydride led to acetamide 26 [37]. Similarly to compound 25, amine 27 was synthesized from amine 24 and (–)-myrtenal (Scheme 6).

2.2. Biology

2.2.1. Cytotoxicity Test

The cytotoxicity was tested using a standard MTT-test (detailed below in Section 3.2.1). A series of 2-fold dilutions of compounds was made; then, each dilution was added to 24 h old cell monolayer, after which, cells were incubated for 24 h. Cell viability was assessed by adding the MTT solution. Based on the data obtained, the CC50 was calculated.

2.2.2. Antiviral Activity

The antiviral activity against the respiratory syncytial virus was assessed by adding a series of 3-fold dilutions of test compounds, followed by addition of the virus in a series of 10-fold dilutions. Cells were incubated for 1 h; then, the virus was washed out and compounds were added again. Cell were incubated for 6 days, after which the viral presence was investigated, using the ELISA method.
The virus titer was calculated using the Reed and Muench method.
The obtained results are shown in Table 1. Compounds were considered promising with a selectivity index of 10 or higher.
At the first stage, all the prepared compounds were tested for their ability to inhibit the replication of RSV A. Aurapten 16a, containing an acyclic monoterpene substituent, exhibited high activity against RSV A in the lower micromolar range, but its selectivity index was less than 8 due to high cytotoxicity. Among compounds 16b,c containing an α-pinene fragment, only compound 16c, derived from (+)-α-pinene, exhibited noticeable activity, with the selectivity index exceeding 11. Elongation of the hydrocarbon chain from 16b to compounds 16df led to a nonlinear change in the biological properties: high antiviral activity was exhibited by compounds with one (16d) and three (16f), but not two (16e) additional CH2-groups. Due to its lower cytotoxicity, compound 16f displayed the best selectivity index, about 65, among monoterpene–coumarin conjugates of this type. Coumarin derivative 16g, containing an aromatic substituent, showed a high selectivity index (30), which was related to its low cytotoxicity, rather than its high activity.
Among 4-methylcoumarin derivatives 17, compound 17c, containing an (+)-α-pinene moiety, exhibited high activity in the submicromolar range, whereas its (–)-isomer was 20-fold less active. Compound 17c had a high selectivity index of 90. Elongation of the hydrocarbon chain (a compound with one additional CH2-group was not prepared due to the formation of a complex reaction mixture) did not affect the activity much; it was almost the same, similarly to toxicity. The less cytotoxic compound 17c had a selectivity index of 62. Compound 17g, containing an aromatic substituent, did not display significant antiviral activity.
Investigation of the biological properties of compound 18, comprising a cyclopentane ring annulated with a coumarin moiety, showed that (+)-α-pinene derivative 18c with a selectivity index of 100 exhibited the highest activity and moderate toxicity. The other compounds were either toxic (18a,b,eg) or inactive (18d); only compound 18e had a noticeable selectivity index of 18.
Among coumarin 19, containing an annulated cyclohexane ring, the highest activity was exhibited by compounds 19a and 19d, which were derived from geraniol and nopol, respectively. High selectivity indices were also displayed by compounds 19c and 19f, which were both less active and less toxic. Coumarin derivative 19g, containing a methoxybenzyl substituent, had a selectivity index of more than 10, which was due to its low cytotoxicity. Removal of the methoxy group from the aromatic ring in compound 19h led to a sharp increase in its toxicity.
Amine 25 showed significant activity and moderate toxicity, which resulted in a selectivity index of more than 10. Transition to acetamide 26 led to a loss of the antiviral effect. Compound 27 was almost one order of magnitude more active and significantly more toxic than amine 25, with the selectivity index being about 15. Compound 27 was much more active than its oxygen-containing analog 19b, which substantiates further research regarding the synthesis of N-linked monoterpene–coumarin conjugates and the investigation of their anti-RSV activity.
A significant portion of the prepared coumarin derivatives were tested for their ability to inhibit RSV replication. Among the compounds tested, 16c,g, 17c,f, 18c, 19a,c,d, and 19f had a selectivity index of more than 10, with the last two being most active. In general, activity of O-linked coumarin–monoterpene conjugates against RSV type B was similar to that against RSV type A, which indicates that the antiviral activity of these compounds lacks type specificity. At the same time, 7-aminocoumrine derivatives demonstrated some activity against RSV type A, but not type B. Among compounds with similar SI against both RSV types, 19c was chosen to investigate a possible action mechanism.
To elucidate action mechanisms of the most active compounds, we performed an experiment using the time-of-addition method. Compound 19c was added to the cells with replicating virus at different time points corresponding to different stages of viral life cycle. The results are shown in Figure 2.
According to the data presented in Figure 2, compound 19c reduces the virus titer at all time points, except for prophylactic use 2 h before virus entry into cells and late introduction after 24 h. This means that action of the compound occurred within 0–6 h after infection. A slight decrease in the virus titer was also noted at time points 2 and 4. Our data suggest that the target of 19c may be the surface F and/or G proteins, viroporin SH, and, probably, L protein (its fragment responsible for transcription).

2.3. Molecular Modeling Study

According to the results of the time-of-addition experiment, the surface F protein may be considered a potential biological target. In addition, the pharmacophore features of the inhibitor of F-protein sisunatovir and compounds 19c, 19f, and 19h are similar. In all cases, there are hydrophobic parts capable of hydrophobic intermolecular contacts, and donors and acceptors of hydrogen bonds (more details are presented in the Supplementary Materials).
We suggest that compounds 19c, 19f, and 19h can bind to a fusion peptide region [38,39] (Figure 3A), similarly to sisunatovir [15] (Figure 3B). Several functional amino acids are located in the binding site of potential entry inhibitors (Figure 3A). The red segment in Figure 3 is a part of the fusion peptide (137–140 amino acids), and the yellow segment corresponds to amino acids of the membrane anchor. Entry inhibitors can interact with these amino acids and inhibit the conformational transition from a pre- to post-fusion conformation. The coumarin moiety of compounds 19c, 19f, and 19h occurs in a cavity surrounded by phenylalanines and forms π–π stacking interactions with them (Figure 3C,D). The terpene fragment of 19c and 19f occurs in a hydrophobic cavity composed of amino acids of the fusion peptide region (Figure 3C,E). The aromatic part of 19h cannot be placed into the cavity due to a short linker between it and the coumarin fragments (Figure 3D). Compound 19h forms intramolecular interactions with amino acids of the membrane anchor.
Docking scores shown in Figure 3 may be considered as a parameter that characterizes the affinity of compounds to the binding site. The best positions are shown in Figure 3. The best ligand position was chosen based on the clustering energy and formation energy of the ligand–protein complex. For more energy parameters, see the Supplementary Materials. Affinity values of lead compounds have a slightly higher value than those for sisunatovir. Compound 19h had the lowest affinity. These results correlate with experimental data (Table S1). Thus, RSV F protein may be considered as a possible target.

3. Materials and Methods

3.1. Chemistry

3.1.1. General Chemical Methods

Reagents and solvents were purchased from commercial suppliers (Sigma-Aldrich, Acros, Japan) and used as received. GC-MS: Agilent 7890A gas chromatograph equipped with an Agilent 5975C quadrupole mass spectrometer as a detector; quartz column HP-5MS (copolymer 5%–diphenyl–95%–dimethylsiloxane) of length 30 m, internal diameter 0.25 mm and stationary phase film thickness 0.25 µm. Optical rotation: polAAr 3005 spectrometer. 1H and 13C NMR: Bruker DRX-500 apparatus at 500.13 MHz (1H) and 125.76 MHz (13C) and Bruker Avance—III 600 apparatus at 600.30 MHz (1H) and 150.95 MHz (13C), J in Hz; structure was determined by analyzing the 1H NMR spectra, including 1H–1H double resonance spectra and 1H–1H 2D homonuclear correlation (COSY, NOESY), J-modulated 13C NMR spectra (JMOD), and 13C–1H 2D heteronuclear correlation with one-bond (C–H COSY, 1J(C,H) = 160 Hz, HSQC, 1J(C,H) = 145 Hz) and long-range spin–spin coupling constants (C–H COSY, 1J(C,H) = 160 Hz, HSQC, 1J(C,H) = 145 Hz). HR-MS: DFS Thermo Scientific spectrometer in a full scan mode (15–500 m/z, 70 eV electron impact ionization, direct sample administration).
Spectral and analytical investigations were carried out at the Collective Chemical Service Center of the Siberian Branch of the Russian Academy of Sciences. All product yields are given for pure compounds purified by recrystallization from ethanol or isolated by column chromatography (SiO2; 60–200 μ; Macherey-Nagel). The purity of the target compounds was determined by GC-MS methods. All of the target compounds reported in this paper had a purity of no less than 95%.

3.1.2. Synthesis of Coumarins 9, 10

Syntheses were carried out from resorcinol 5 and appropriate β-ketoesters 6, 7, in accordance with [31]. Concentrated H2SO4 (5 mL, 94 mmol) was added dropwise to cooled (0–5 °C) solution of resorcinol 8 (45 mmol) and appropriate β-ketoesters (45 mmol) in dry ethanol (15 mL), with vigorous stirring. The mixture was stirred until it congealed, left overnight at r.t., and poured into ice water (150 mL). The resulting solid was filtered off and crystallized from ethanol–water (75%). The yields of 9, 10 were 64% and 70%, respectively.

3.1.3. Synthesis of Bromides 11a–c,g

(+)-Myrtenal was synthesized according to the procedure presented in [41] by the oxidation of (+)-α-pinene using a t-BuOOH/SeO2 system with a 57% yield.
(+)-Myrtenol was synthesized from the corresponding aldehyde via reduction to alcohols with NaBH4, as described above. NaBH4 (10.3 mmol) was added to a cooled (0–5 °C) solution of 10.3 mmol of the appropriate aldehyde in methanol (20 mL), and the reaction mixture was stirred for 3 h at room temperature. Then, 5% aqueous HCl was added to obtain a pH of 4–5. The solvent was distilled, and the product was extracted using ether and dried with Na2SO4. The solvent was evaporated; the resulting alcohol (54% yield) was used in the synthesis without purification.
(3-Methoxyphenyl)methanol was synthesized from 3-methoxybenzaldehyde via a reaction with NaBH4, as described above (yield: 34%).
Bromide 11a was synthesized from geraniol via the reaction with PBr3 [30].
PBr3 (8.9 mmol) was added to cooled (0–5 °C) solution of geraniol (26.7 mmol) in dry ether (30 mL), and the reaction mixture was stirred for 2 h at r.t. Saturated aqueous NaHCO3 was added, and the product was extracted with ether. The extracts were washed with brine, dried with Na2SO4, and evaporated.
Additional bromides 11b,c,g were synthesized as described above. Compounds 11a–c,g (with yields of 91%, 55%, 60% and 65%, respectively) were sufficiently pure and used for the next step without purification.

3.1.4. Synthesis of Bromide 11d

Bromide 11d was synthesized from (−)-nopol via reaction with NBS–PPh3, as described in [32].
Triphenylphosphine (2.0 equiv., 6.1 g, 23 mmol) was dissolved in dry DCM (23 mL) under argon. N-bromosuccinimide (NBS) (2.0 equiv., 4.2 g, 23 mmol) was added to this solution in small portions over 5 min in an ice-water bath. Subsequently, the resulting deep red mixture was stirred at room temperature for 30 min. Then, pyridine (1 mL) was added, and the color turned reddish-brown. (–)-Nopol (1.0 equiv., 2.0 mL, 12 mmol) was added to the mixture dropwise over 10 min. The reaction mixture was stirred overnight at room temperature. Later, the mixture was diluted with hexane (40 mL) and filtered through a silica gel plug. Then, the reaction flask was stirred with EtOAc–hexane (1:1, 40 mL) and filtered through the silica gel plug. Later, it was concentrated in vacuo and the crude residue was purified by chromatography on SiO2 (hexane) to obtain bromide 11d (2.3 g, 70% yield).

3.1.5. Synthesis of Bromide 11e

Bromide 11e was synthesized from alcohol 13 via a reaction with NBS–PPh3, as described for 11d (the yield of 11e-41%). Alcohol 13 was obtained by the reduction of ether 12, synthesized from (+)-trans-pinocarveol [34] with LiAIH4.
A mixture of (+)-trans-pinocarveol (2.45 g, 16 mmol; obtained by the oxidation of β-pinene with t-BuOOH in the presence of SeO2, in accordance with [33]), triethyl orthoacetate (3.93 g, 24 mmol), and hexanoic acid (0.25 g, 2.4 mmol) was heated for 6 h at 150 °C. The mixture was then diluted with Et2O (150 mL) and washed successively with sat. aq. NaHCO3 soln. and H2O, dried with Na2SO4, and evaporated. Compound 12 was purified by column chromatography on SiO2, eluent–hexane–ethyl acetate (1.92 g, yield 54%).
A solution of 12 (2.22 g, 10 mmol) in Et2O (10 mL) was added dropwise to a suspension of LiAIH4 (2.1 g, 55 mmol) in refluxing Et2O (20 mL) . After 1 h at r.t., the mixture was cooled to 0 °C, and H2O (2 mL), 15% aq. NaOH soln. (2 mL), and then H2O (3 mL) were added. After 30 min, the mixture was filtered and evaporated. Compound 13 was purified by column chromatography on SiO2, eluent–hexane–ethyl acetate (1.40 g, yield 78%).

3.1.6. Synthesis of Bromide 11f

Aldehyde 14, obtained according to [35], was reduced to alcohol 15 by NaBH4 (yield: 86%). The OH group of alcohol 15 was replaced by Br using NBS–PPh3 (yield 11f: 84%). Then, 100 mL Et2O, acrolein (8.6 g, 154 mmol) and (−)-β-pinene (14.5 g, 107 mmol) were added to a solution of anhydrous ZnBr2 (4.3 g, 19 mmol). The solution was stirred for 48 h at 25 °C. Then, it was poured into water (200 mL) and filtered by suction to remove zinc salts. The ether layer was separated, and the aqueous layer was extracted with 2 × 50 mL Et2O. The combined ether layers were dried and evaporated. Compound 14 was purified by column chromatography on SiO, eluent–hexane–CHCl3 (30%) (14: 3.76 g, yield 25%, unreacted (−)-β-pinene-3.13 g).

3.1.7. Synthesis of Compounds 16ag, 17ac,eg, 18ag and 19ag

Compounds 16ad,g, were synthesized from coumarin 6 and the corresponding bromides 11ad,g with the use of DBU and DMF, in accordance with [30].
DBU (1.0 mmol) and the corresponding bromides 11ad,g (0.75 mmol) were added to compound 1 (0.5 mmol) in dry DMF (5 mL) at r.t. under stirring. The reaction mixture was stirred at r.t. for 15 min, and then heated at 60 °C for 5 h. H2O (15 mL) was added, and the product was extracted with ethyl acetate. The extracts were washed with brine, dried with Na2SO4, and evaporated.
Compounds 16e,f, 16e,f, 17e,f and 18e,f were synthesized as above.
Compounds 16a, 17a, and 18a were synthesized from coumarins 6,7,9 and geranyl bromide 11a with the use K2CO3 and ethanol, in accordance with [31].
K2CO3 (1.0 mmol) and geranyl bromide 11a (0.75 mmol) were added to corresponding compounds 5–7 (0.5 mmol) in dry ethanol (5 mL) at r.t. under stirring. The reaction mixture was stirred at r.t. for 15 min, and then heated at 60 °C for 5 h. The hot solution was filtered; the filtrate was kept at −18 °C for 48 h.
Compounds 17ac,g, 18ad,g, and 19ad,g were synthesized from coumarins 7,9,10 and corresponding bromides 11ad,g with the use K2CO3 and ethanol, as described above.
1H NMR spectra of 16ad,g, 17ac,g, 18ad,g, and 19ad,g coincided with the corresponding spectra published in the literature [30,31].
Products 16ag, 17ac,eg, 18ag, and 19ag were isolated in the individual form: (a) by recrystallization from ethanol; or (b) by column chromatography on a silica gel, eluent–hexane, solution containing from 25% to 100% chloroform in hexane, ethanol.
7-(3-((1R,5S)-6,6-Dimethylbicyclo[3.1.1]hept-2-en-2-yl)propoxy)-2H-chromen-2-one (16e). Yield: 58%, method b. M.p. 75 °C. α 589 24.5 = −19.4 (c = 1.40, CHCl3). HRMS: 324.1714 ([M]+, m/z calcd for C21H24O3 324.1720). 1H-NMR (CDCl3, δH): 0.80 (s, 3H-C(21)); 1.11 (d, 1H, 2J = 8.5, Hanti H–C(19)); 1.24 (s, 3H-C(20)); 1.77–1.89 (m, 2H, 2H-C(11)); 2.00 (ddd, 1H, J18,16 = J18,19sin = 5.6, J18,14 = 1.4, H-C(18)); 2.03–2.11 (m, 3H, 2H-C(12)), 1H-C(16)); 2.15 (dm, 1H, 2J = 17.5, H-C(15)); 2.22 (dm, 1H, 2J = 17.5, H’-C(15)); 2.34 (ddd, 1H, 2J = 8.5, J19sin,16 = J19sin,18 = 5.6, Hsin-C(19)); 3.94–3.98 (m, 2H, 2H-C(10)); 5.18–5.22 (m, 1H, H-C(14)); 6.19 (d, 1H, J3,4 = 9.5, H-C(3)); 6.75 (d, 1H, J9,7 = 2.4, H-C(9)); 6.79 (dd, 1H, J7,6 = 8.5, J7.9 = 2.4, H-(C-7)); 7.32 (d, 1H, J6.7 = 8.5, H-6); 7.59 (d, 1H, J4,3 = 9.5, H-4). 13C-NMR (CDCl3, δC): 155.74 (s, C(1)); 161.05 (s, C(2)); 112.72 (d, C(3)); 143.27 (d, C(4)); 112.21 (s, C(5)); 128.53 (d, C(6)); 112.78 (d, C(7)); 162.24 (s, C(8)); 101.15 (d, C(9)); 68.11 (t, C(10)); 26.42 (t, C(11)); 32.83 (t, C(12)); 146.90 (s, C(13)); 116.61 (d, C(14)); 31.09 (t, C-(15)); 40.66 (d, C(16)); 37.78 (s, C(17)); 45.58 (d, C(18)); 31.53 (t, C(19)); 26.14 (q, C(20)); 21.02 (q, C(21)).
7-(4-((1R,5S)-6,6-Dimethylbicyclo[3.1.1]hept-2-en-2-yl)butoxy)-2H-chromen-2-one (16f). Yield: 80%, method b. α 589 22 = −9.0 (c = 1.78, CHCl3). HRMS: 337.1793 ([M-H]+, m/z calcd for C22H25O3 337.1798). 1H-NMR (CDCl3, δH): 0.80 (s, 3H-C(22)); 1.11 (d, 1H, 2J = 8.5, Hanti-C(20)); 1.24 (s, 3H-C(21)); 1.42–1.56 (m, 2H, 2H-C(12)); 1.74–1.81 (m, 2H, 2H-C(11)); 1.95–2.02 (m, 3H, 2H-C(13)), 1H-C(19)); 2.02–2.08 (m, 1H, H-C(17)); 2.15 (dm, 1H, 2J = 17.4, H-C(16)); 2.22 (dm, 1H, 2J = 17.4, H’-C(16)); 2.32 (ddd, 1H, 2J = 8.5, J20sin,17 = J20sin,19 = 5.6, Hsin-C(20)); 3.98 (t, 2H, J10,11 = 6.5, 2H-C(10)); 5.16–5.20 (m, 1H, H-C(15)); 6.20 (d, 1H, J3,4 = 9.5, H-C(3)); 6.76 (d, 1H, J9,7 = 2.4, H-C(9)); 6.79 (dd, 1H, J7,6 = 8.6, J7.9 = 2.4, H-C(7)); 7.32 (d, 1H, J6.7 = 8.6, H-C(6)); 7.59 (d, 1H, J4,3 = 9.5, H-C(4)). 13C-NMR (CDCl3, δC): 155.77 (s, C(1)); 161.07 (s, C(2)); 112.74 (d, C(3)); 143.26 (d, C(4)); 112.22 (s, C(5)); 128.53 (d, C(6)); 112.79 (d, C(7)); 162.26 (s, C(8)); 101.21 (d, C(9)); 68.38 (t, C(10)); 28.56 (t, C(11)); 23.32 (t, C(12)); 36.31 (t, C(13)); 147.68 (s, C(14)); 116.14 (d, C(15)); 31.12 (t, C(16)); 40.73 (d, C(17)); 37.78 (s, C(18)); 45.60 (d, C(19)); 31.52 (t, C(20)); 26.19 (q, C(21)); 21.05 (q, C(22)).
7-(3-((1R,5S)-6,6-Dimethylbicyclo[3.1.1]hept-2-en-2-yl)propoxy)-4-methyl-2H-chromen-2-one (17e). Yield: 49%, method a. M.p. 86 °C. α 589 24.5 = −19.5 (c = 0.65, CHCl3). HRMS: 337.1796 ([M-H]+, m/z calcd for C22H25O3 337.1798). 1H-NMR (CDCl3, δH): 0.81 (s, 3H-C(22)); 1.13 (d, 1H, 2J = 8.5, Hanti-C(20)); 1.25 (s, 3H-C(21)); 1.78–1.90 (m, 2H, 2H-C(12)); 2.01 (ddd, 1H, J19,17 = J19,20sin = 5.6, J19,15 = 1.3, H-C(19)); 2.04–2.12 (m, 3H, 2H-C(13)), 1H-C(17)); 2.17 (dm, 1H, 2J = 17.5, H-C(16)); 2.23 (dm, 1H, 2J = 17.5, H’-C(16)); 2.35 (ddd, 1H, 2J = 8.5, J20sin,17 = J20sin,19 = 5.6, Hsin-C(20)); 2.37 (d, 3H, J10,3 = 0.8, 3H-C(10)); 3.97 (t, 2H, J11,12 = 6.5, 2H-C(11)); 5.20–5.23 (m, 1H, H-C(15)), 6.09 (q, 1H, J3,10 = 0.8, H-C(3)); 6.77 (d, 1H, J9,7 = 2.5, H-C(9)); 6.82 (dd, 1H, J7,6 = 8.8, J7,9 = 2.5, H-C(7)); 7.46 (d, 1H, J6,7 = 8.8, H-C(6)). 13C-NMR (CDCl3, δC): 155.16 (s, C(1)); 161.24 (s, C(2)); 111.66 (d, C(3)); 152.46 (s, C(4)); 113.28 (s, C(5)); 125.32 (d, C(6)); 112.55 (d, C(7)); 162.09 (s, C(8)); 101.19 (d, C(9)); 18.53 (q, C(10)); 68.10 (t, C(11)); 26.47 (t, C(12); 32.88 (t, C(13)); 146.96 (s, C(14)); 116.64 (d, C(15)); 31.13 (t, C(16)); 40.69 (d, C(17)); 37.83 (s, C(18)); 45.60 (d, C(19)); 31.57 (t, C(20)); 26.18 (q, C(21)); 21.06 (q, C(22)).
7-(4-((1R,5S)-6,6-Dimethylbicyclo[3.1.1]hept-2-en-2-yl)butoxy)-4-methyl-2H-chromen-2-one (17f). Yield: 24%, method a. M.p. 77 °C. α 589 26 = −20.3 (c = 0.60, CHCl3). HRMS: 351.1954 ([M-H]+, m/z calcd for C23H27O3 351.1955). 1H-NMR (CDCl3, δH): 0.81 (s, 3H-C(23)); 1.11 (d, 1H, 2J = 8.5, Hanti-C(21)); 1,24 (s, 3H-C(22)); 1.43–1.57 (m, 2H, 2H-C(13)); 1.74–1.82 (m, 2H, 2H-C(12)); 1.96–2.02 (m, 3H, 2H-C(14)), 1H-C(20)); 2.03–2.08 (m, 1H, H-C(18)); 2.16 (dm, 1H, 2J = 17.5, H-C(17)); 2.23 (dm, 1H, 2J = 17.5, H’-C(17)); 2.33 (ddd, 1H, 2J = 8.5, J21sin,18 = J21sin,20 = 5.6, Hsin-C(21)); 2.37 (d, 3H, J10,3 = 1.2, 3H-C(10)); 3.98 (t, 2H, J11,12 = 6.5, 2H-C(11)); 5.16–5.20 (m, 1H, H-C(16)); 6.10 (q, 1H, J3,10 = 1.2, H-C(3)); 6.77 (d, 1H, J9,7 = 2.5, H-C(9)); 6.82 (dd, 1H, J7,6 = 8.8, J7,9 = 2.5, H-C(7)); 7.45 (d, 1H, J6,7 = 8.8, H-C(6)). 13C-NMR (CDCl3, δC): 155.14 (s, C(1)); 161.25 (s, C(2)); 111.65 (d, C(3)); 152.46 (s, C(4)); 113.24 (s, C(5)); 125.31 (d, C(6)); 112.53 (d, C(7)); 162.06 (s, C(8)); 101.18 (d, C(9)); 18.55 (q, C(10)); 68.33 (t, C(11)); 28.59 (t, C(12)); 23.33 (t, C(13); 36.35 (t, C(14)); 147.70 (s, C(15)); 116.13 (d, C(16)); 31.12 (t, C(17)); 40.70 (d, C(18)); 37.80 (s, C(19)); 45.55 (d, C(20)); 31.54 (t, C(21)); 26.20 (q, C(22)); 21.07 (q, C(23)).
7-(3-((1R,5S)-6,6-Dimethylbicyclo[3.1.1]hept-2-en-2-yl)propoxy)-2,3-dihydrocycl penta[c]chromen-4(1H)-one (18e) Yield 47%, method a. M.p. 68 °C. α 589 24.5 = −18.6 (c = 0.88, CHCl3). HRMS: 364.2038 ([M]+, m/z calcd for C24H28O3 364.2033). 1H-NMR (CDCl3, δH): 0.81 (s, 3H-C(24)); 1.12 (d, 1H, 2J = 8.5, Hanti-C(22)); 1.25 (s, 3H-C(23)); 1.77–1.90 (m, 2H, 2H-C(14)); 2.01 (ddd, 1 H, J21.19 = J21,22sin5.6, J21,17 = 1.4, H-C(21)); 2.04–2.12 (m, 3H, 2H-C(15), 1H-C(19)); 2.13–2.19 (m, 3H, 2H-C(11), 1H-C(18)); 2.23 (dm, 1H, 2J = 17.6, H’-C(18)); 2.34 (ddd, 1H, 2J = 8.5, J22s,19 = J22s,21 = 5.6, Hsin-C(22)); 2.83–2.88 (m, 2H, 2H-C(10)); 2.98–3.03 (m, 2H, 2H-C(12)); 3.96 (t, 2H, J13,14 = 6.5, 2H-C(13)); 5.19–5.23 (m, 1H, H-C(17)); 6.78–6.82 (m, 2H,H-C(7), H-C(9)); 7.27–7.32 (m, 1H, H-C(6)). 13C-NMR (CDCl3, δC): 155.64 (s, C(1)); 160.46 (s, C(2)); 124.16 (s, C(3)); 156.28 (s, C(4)); 112.11 (s, C(5)); 125.38 (d, C(6)); 112.45 (d, C(7)); 161.43 (s, C(8)); 101.10 (d, C(9)); 30.20 (t, C(10)); 22.43 (t, C(11)); 31.90 (t, C(12)); 68.04 (t, C(13)); 26.50 (t, C(14)); 32.90 (t, C(15)); 146.99 (s, C(16)); 116.59 (d, C(17)); 31.12 (t, C(18)); 40.68 (d, C(19)); 37.81 (s, C(20)); 45.59 (d, C(21)); 31.56 (t, C(22)); 26.17 (q, C(23)); 21.05 (q, C(24)).
7-(4-((1R,5S)-6,6-Dimethylbicyclo[3.1.1]hept-2-en-2-yl)butoxy)-2,3-dihydrocyclopenta[c]chromen-4(1H)-one (18f). Yield: 53%, method a. M.p. 61 °C. α 589 26 = −17.0 (c = 1.00, CHCl3). HRMS: 377.2110 ([M-H]+, m/z calcd for C25H29O3 377.2111). 1H-NMR (CDCl3, δH): 0.80 (s, 3H-C(25)); 1.11 (d, 1H, 2J = 8.5, Hanti-C(23)); 1.24 (s, 3H-C(24)); 1.43–1.56 (m, 2H, 2H-C(15)); 1.73–1.81 (m, 2H, 2H-C(14)); 1.95–2.01 (m, 3H, 2H-C(16), H-C(22)); 2.02–2.07 (m, 1H, H-C(20); 2.12–2.19 (m, 3H, 2H-C(11), 1H-C(19)); 2.22 (dm, 1H, 2J = 17.5, H’-C(19)); 2.32 (ddd, 1H, 2J = 8.5, J23sin,20 = J23sin,22 = 5.6, Hsin-C(23)); 2.82–2.88 (m, 2H, 2H-C(10)); 2.97–3.03 (m, 2H, 2H-C(12)); 3.97 (t, 2H, J13,14 = 6.5, 2H-C(13)); 5.16–5.20 (m, 1H, H-C(18)); 6.78–6.82 (m, 2H, H-C(7), H-C(9)); 7.26–7.32 (m, 1H, H-C(6)). 13C-NMR (CDCl3, δC): 155.62 (s, C(1)); 160.48 (s, C(2)); 124.14 (s, C(3)); 156.29 (s, C(4)); 112.07 (s, C(5)); 125.37 (d, C(6)); 112.42 (d, C(7)); 161.40 (s, C(8)); 101.10 (d, C(9)); 30.19 (t, C(10)); 22.41 (t, C(11)); 31.90 (t, C(12)); 68.26 (t, C(13)); 28.62 (t, C(14)); 23.34(t, C(15)); 36.35 (t, C(16)); 147.72 (s, C(17)); 116.10 (d, C(18)); 31.11 (t, C(19)); 40.69 (d, C(20)); 37.79 (s, C(21)); 45.54 (d, C(22)); 31.53 (t, C(23)); 26.19 (q, C(24)); 21.06 (q, C(25)).
3-(3-((1R,5S)-6,6-Dimethylbicyclo[3.1.1]hept-2-en-2-yl)propoxy)-7,8,9,10-tetrahydro-6H-benzo[c]chromen-6-one (19e). Yield: 33%, method a. M.p. 70 °C. α 589 24.5 = −17.8 (c = 0.73, CHCl3). HRMS: 377.2110 ([M-H]+, m/z calcd for C25H29O3 377.2111). 1H-NMR (CDCl3, δH): 0.81 (s, 3H-C(25)); 1.12 (d, 1H, 2J = 8.6, Hanti-C(23)); 1.25 (s, 3H-C(24)); 1.73–1.89 (m, 6H, 2H-C(11), 2H-C(12), 2H-C(15)); 2.01 (ddd, 1 H, J22.20 = J22,23sin = 5.6, J22,18 = 1.4, H-C(22)); 2.04–2.12 (m, 3H, 2H-C(16), 1H-C(20)); 2.16 (dm, 1H, 2J = 17.5, H-C(19)); 2.23 (dm, 1H, 2J = 17.5, H’-C(19)); 2.34 (ddd, 1H, 2J = 8.6, J23sin,20 = J23sin,22 = 5.6, Hsin-C(23)); 2.50–2.55 (m, 2H, 2H-C(10)); 2.69–2.74 (m, 2H, 2H-C(13)); 3.96 (t, 2H, J14,15 = 6.5, 2H-C(14)); 5.19–5.23 (m, 1H, H-C(18)); 6.75 (d, 1H, J9,7 = 2.5, H-C(9)); 6.80 (dd, 1H, J7,6= 8.8, J7,9 = 2.5, H-C(7)); 7.41 (d, 1H, J6,7 = 8.8, H-C(6)). 13C-NMR (CDCl3, δC): 153.38 (s, C(1)); 162.09 (s, C(2)); 120.22 (s, C(3)); 147.19 (s, C(4)); 113.46 (s, C(5)); 123.90 (d, C(6)); 112.23 (d, C(7)); 160.85 (s, C(8)); 100.97 (d, C(9)); 23.69 (t, C(10)); 21.57 (t, C(11)); 21.27 (t, C(12)); 25.08 (t, C(13)); 67.99 (t, C(14)); 26.52 (t, C(15)); 32.91 (t, C(16)); 147.03 (s, C(17)); 116.57 (d, C(18)); 31.13 (t, C(19)); 40.70(d, C(20)); 37.82 (s, C(21)); 45.61 (d, C(22)); 31.57 (t, C(23)); 26.18 (q, C(24)), 21.06 (q, C(25)).
3-(4-((1R,5S)-6,6-Dimethylbicyclo[3.1.1]hept-2-en-2-yl)butoxy)-7,8,9,10-tetrahydro-6H-benzo[c]chromen-6-one (19f). Yield: 36%, method a. M.p. 69 °C. α 589 27 = −17.6 (c = 0.90, CHCl3). HRMS: 391.2265 ([M-H]+, m/z calcd for C26H31O3 391.2268). 1H-NMR (CDCl3, δH): 1H-NMR (CDCl3, δH): 0.81 (s, 3H-C(26)); 1.11 (d, 1H, 2J = 8.5, Hanti-C(24)); 1.24 (s, 3H-C(25)); 1.42–1.56 (m, 2H, 2H-C(16)); 1.73–1.86 (m, 6H, 2H-C(11), 2H-C(12), 2H-C(15)); 1.95–2.02 (m, 3H, 2H-C(17), H-C(23)); 2.02–2.08 (m, 1H, H-C(21)); 2.16 (dm, 1H, 2J = 17.4, H-C(20)); 2.22 (dm, 1H, 2J = 17.4, H’-C(20)); 2.33 (ddd, 1H, 2J = 8.5, J24sin,21 = J24sin,23 = 5.6, Hsin-C(24)); 2.49–2.55 (m, 2H, 2H-C(10));)); 2.68–2.75 (m, 2H, 2H-C(13)); 3.96 (t, 2H, J14,15 = 6.5, 2H-C(14)); 5.16–5.20 (m, 1H, H-C(19)); 6.75 (d, 1H, J9,7 = 2.5, H-C(9)); 6.79 (dd, 1H, J7,6= 8.8, J7,9 = 2.5, H-C(7)); 7.41 (d, 1H, J6,7 = 8.8, H-C(6)). 13C-NMR (CDCl3, δC): 153.36 (s, C(1)); 162.10 (s, C(2)); 120.19 (s, C(3)); 147.19 (s, C(4)); 113.42 (s, C(5)); 123.89 (d, C(6)); 112.19 (d, C(7)); 160.82 (s, C(8)); 100.95 (d, C(9)); 23.68 (t, C(10)); 21.56 (t, C(11)); 21.25 (t, C(12)); 25.08 (t, C(13)); 68.20 (t, C(14)); 28.64 (t, C(15)); 23.35 (t, C(16)); 36.37(t, C(17)); 147.74 (s, C(18)); 116.09 (d, C(19)); 31.12 (t, C(20)); 40.70 (d, C(21)); 37.80 (s, C(22)); 45.55 (d, C(23)); 31.54 (t, C(24)); 26.20 (q, C(25)), 21.07 (q, C(26)).

3.1.8. Synthesis of 7-aminocoumarines 23 and 24

7-Aminocoumarins 23 and 24 were synthesized from m-aminophenol 20, in accordance with [37].
Methoxycarbonyl chloride (3.6 mL 47 mmol) was added dropwise to a cooled (5–10 °C) suspension of m-aminophenol 20 (4.4 g, 40 mmol) and K2CO3 (3.5 g) in 35 mL of ethyl acetate and 3 mL of water, with vigorous stirring. The mixture was stirred for 1 h; then, 10 mL of water was added, and the mixture was stirred for another 3 h. The product was extracted with ethyl acetate. The extracts were washed with water, 1 M H2SO4, and brine, dried with Na2SO4, and evaporated. The resulting solid was crystallized from benzene to give 5.7 g of 21 (77%).
A mixture of compound 21 (4.6 g, 28 mmol) and 5.5 mL acetoacetic ester was added dropwise to 12 mL H2SO4 with vigorous stirring. The mixture was stirred for 2 h and diluted with 50 mL of ice-water. The precipitate was removed by filtration, washed with water, MeOH, and ether, and dried to give 4,0 g of 22 (61%).
A suspension of 2.8 g (12 mmol) of compound 22 in 6 mL of 45% KOH solution was stirred at 90 °C for 0.5 h until the solution formed. The mixture was cooled and diluted with water and acidified with concentrated HCl to pH 5–6. A solution of alkali was added to the suspension, to obtain pH 8. The mixture was stirred until crystallization ceased. The precipitate was removed by filtration, washed with water, MeOH, ether and dried to give 1.53 g of 23 (73%).
Similarly, compound 24 was synthesized from compound 21 with a yield of 64%.

3.1.9. Synthesis of 7-aminocoumarines 2527

Amine 25 was obtained by the interaction of compound 23 and (−)-nopinal and subsequent reduction with NaBH3CN, in accordance with [42].
Compound 25 (0.097 g,1.0 mmol), (−)-nopinal (0.112 g 1.2 mmol) (synthesized by the oxidation of (−)-nopol with IBX according to the procedure [43]) and acetic acid (100 μL) were dissolved in methanol (5 mL) and stirred at room temperature for 2.5 h. NaBH3CN (0.110 g, 2.0 mmol) was added, and the reaction mixture was stirred at room temperature for 1.5 h. Methanol was evaporated and the reaction mixture was extracted with CH2Cl2. The organic layer was washed with brine, dried over anhydrous Na2SO4, filtered, and evaporated. The residue was crystallized from ethanol to give 0.115 g of 25 (64%).
Similarly, compound 27 was synthesized from amine 24 and (−)-myrtenal (yield-57%).
7-N-acetylaminocoumarin 26 was synthesized from 7-aminocoumarin 23, in accordance with [38].
A mixture of 7-aminocoumarin 23 (0.200 g, 1.1 mmol) and DMAP (20 mg) was dissolved in 1 mL CH2Cl2. Acetic anhydride (0.2 mL, 2.1 mmol) was added, and the mixture was stirred at room temperature for 24 h. On completion of the reaction, 10 mL of ice-cold water was added. The precipitate was filtered, washed with water, and dried. The resulting solid was crystallized from ethanol to give 0.176 g of 26 (71%). NMR spectrum 26 coincided with the corresponding spectrum published in the literature [38].
7-(2-((1R,5S)-6,6-Dimethylbicyclo[3.1.1]hept-2-en-2-yl)ethylamino)-4-methyl-2H-chromen-2-one (25). Yield: 64% M.p. 131 °C. α 589 27 = −19.7 (c = 0.95, CHCl3). HRMS: 323.1878 ([M]+, m/z calcd for C21H25O2N1 323.1880). 1H-NMR (CDCl3, δH): 0.81 (s, 3H-C(21)); 1.11 (d, 1H, 2J = 8.6, Hanti-C(19)); 1.25 (s, 3H-C(20)); 2.04 (ddd, 1 H, J18.16 = J18,19sin = 5.6, J18,14 = 1.4, H-C(18)); 2.07–2.11 (m, 1H, H-C(16)); 2.21 (dm, 1H, 2J = 17.6, H-C(15)); 2.25–2.33 (m, 3H, 22H-C(12), 1H’-C(15)); 2.31 (d, 3H, J10,3 = 1.2, 3H-C(10)); 2.36 (ddd, 1H, 2J = 8.6, J19sin,16 = J19sin,18 = 5.6, Hsin-C(19)); 3.10–3.19 (m, 2H, 2H-C(11)); 5.32–5.36 (m, 1H, H-C(14)); 5.95 (q, 1H, J3,10 = 1.2, H-C(3)); 6.44 (d, 1H, J9,7 = 2.4, H-C(9)); 6.49 (dd, 1H, J7,6 = 8.7, J7,9 = 2.4, H-C(7)); 7.32 (d, 1H, J6,7 = 8.7, H-C(6)).
13C-NMR (CDCl3, δC): 155.83 (s, C(1)); 161.80 (s, C(2)); 109.39 (d, C(3)); 152.79 (s, C(4)); 110.64 (s, C(5)); 125.34 (d, C(6)); 110.57 (d, C(7)); 150.96 (s, C(8)); 98.32 (d, C(9)); 18.40 (q, C(10)); 40.89 (t, C(11)); 35.79 (t, C(12)); 144.85 (s, C(13)); 119.21 (d, C(14)); 31.24 (t, C(15)); 40.59 (d, C(16)); 37.86 (s, C(17)); 45.13 (d, C(18)); 31.64 (t, C(19)); 26.08 (q, C(20)); 21.10 (q, C(21)).
3-(((1R,5S)-6,6-Dimethylbicyclo[3.1.1]hept-2-en-2-yl)methylamino)-7,8,9,10-tetrahydro-6H-benzo[c]chromen-6-one (27). Yield: 57%. α 589 27 = −16.3 (c = 0.74, CHCl3). HRMS: 349.2040 ([M]+, m/z calcd for C23H27O2N1 349.2036). 1H-NMR (CDCl3, δH): 0.79 (s, 3H-C(23)); 1.12 (d, 1H, 2J = 8.6, Hanti-C(21); 1.25 (s, 3H-C(22)); 1.71–1.82 (m, 4H, 2H-C(11), 2H-C(12)); 2.05–2.10 (m, 2H, H-C(18), H-C(20)); 2.19 (dm, 1H, 2J = 17.8, H-C(17)); 2.26 (dm, 1H, 2J = 17.8, H’-C(17)); 2.36 (ddd, 1H, 2J = 8.6, J21sin,18 = J21sin,20 = 5.6, Hsin-C(21)); 2.47–2.52 (m, 2H, 2H-C(10)); 2.64–2.70 (m, 2H, 2H-C(13)); 3.62–3.66 (m, 2H, 2H-C(14)); 5.41–5.45 (m, 1H, H-C(16)); 6.45 (d, 1H, J9,7 = 2.4, H-C(9)); 6.50 (dd, 1H, J7,6= 8.7, J7,9 = 2.4, H-C(7)); 7.28 (d, 1H, J6,7 = 8.7, H-C(6)).13C-NMR (CDCl3, δC): 153.92 (s, C(1)); 162.57 (s, C(2)); 117.92 (s, C(3)); 147.63 (s, C(4)); 110.83 (s, C(5)); 123.77 (d, C(6)); 110.51 (d, C(7)); 149.93 (s, C(8)); 98.57 (d, C(9)); 23.62 (t, C(10)); 21.73 (t, C(11)); 21.38 (t, C(12)); 24.99 (t, C(13)); 48.54 (t, C(14)); 144.05 (s, C(15)); 118.81 (d, C(16)); 31.01 (t, C(17)); 40.71 (d, C(18)); 38.00 (s, C(19)); 43.87 (d, C(20)); 31.46 (t, C(21)); 26.01 (q, C(22)), 21.00 (q, C(23)).

3.2. Biology

3.2.1. Cytotoxicity Test

The compounds were weighed in amounts of 2 mg and dissolved in 100 μL of DMSO. Then, the resulting solution was adjusted with the medium to a concentration of 1000 μg/mL, and a series of twofold dilutions was prepared from it. One-day culture of HEp2 cells, grown in 96-well plates, cell concentration 3 × 105/well of the plate, was checked visually in an inverted microscope for the integrity of the monolayer. Plates were selected for work where the cell closure was 60–80%.
Dilutions of the compounds at the appropriate concentration were added to the plate in a volume of 100 μL in each well in two replicates for each tested concentration. The plates were incubated for 24 h at 37 °C in the presence of 5% CO2. Cell viability was assessed using the MTT assay.
The MTT solution was prepared on a maintenance medium at a concentration of 0.5 mg/mL. Then, 0.1 mL of MTT solution was added to each well. After 1.5 h of MTT contact at 37 °C at a CO2 concentration of 5%, MTT was discarded with the cells of the well and 0.1 mL of ethyl alcohol 96% was poured, after which the optical density in the wells was measured at a wavelength of 535 nm. Based on the data obtained, the CC50 was calculated.

3.2.2. Antiviral Activity

The antiviral activity against the respiratory syncytial virus (RSV A—strain A2, RSV B—strain 9320) was assessed in a series of threefold dilutions of test compounds, starting from ½CC50, which were added to HEp-2 cell culture at a double concentration, at 100 μL per well, followed by the addition of 100 μL of the virus in a series of 10-fold dilutions. Cells were incubated at 37 °C and 5% CO2 for 1 h. Then, the virus was washed out, and the compounds were again added at a single concentration and incubated at 37 °C and 5% CO2 for 6 days. For the enzyme-linked immunosorbent assay (ELISA), cell culture was fixed with cold 80% acetone at –20 °C for 15 min and then washed with phosphate-buffered saline containing 0.05% Tween 20. Next, a solution of primary mouse anti-RSV F protein antibodies was added to the culture and incubated at room temperature under continuous stirring for 2 h. Then, cells were again washed with buffer, secondary anti-mouse antibodies were added, and the cells were incubated under continuous stirring for 2 h. Then, the antibodies were washed off, and a substrate–chromogenic mixture with tetramethylbenzidine was added. After 5 min, the reaction was stopped with 0.1 M sulfuric acid, and the optical density of the solution was measured at a wavelength of 450 nm. Wells with absorbance values twofold or greater than the cell control were considered contaminated. The virus titer was calculated using the Reed and Muench method. All experiments were made in triplicate.

3.2.3. Time-of-Addition Assay

Compound 19c was added at different time points before, after or simultaneously with the introduction of the virus. The time of addition of the compound was counted from point 0—the time of entry of the virus into the cell. During the period (𢄤1)–0, the cells together with the virus were incubated at 40 °C. All other experiments were carried out at 37 °C. RSV virus A 2 mL was added to the cells at a time that was conventionally designated as point −1, after which the cells were kept for an hour at a temperature of 40 °C. Then, at point 0, the virus was unbound. The cells were transferred to a thermostat at 37 °C, where they were incubated for 25 h. After this period, the medium was taken from each well and a series of ten-fold dilutions were made on a fresh cell culture and incubated for 6 days. For each compound, 2 repetitions were made by different operators. The virus titer was estimated by ELISA. The compound was added at the following times relative to the addition of the virus: point −2—the compound was introduced one hour before cell infection (prophylactic regimen); point 0—at the moment of temperature change; points 1, 2, 4, 6, 24—after 1, 2, 4, 6 and 24 h after the temperature change, respectively. In the wells marked (−2) − (25), the compound was kept throughout the experiment, starting from point -2 and until the end of the experiment −25 h. No compound was added to the control wells; instead, a similar volume of medium was added.

3.3. Molecular Modeling

3.3.1. Receptor and Ligand Preparation

Crystallographic structures of the RSV F protein (PDB codes 7LVW [39] and 7KQD [15]) were downloaded from the Protein Data Bank database [40]. The code of the full-length protein trimer was 7LVW; 7KQD is the code of a ligand–protein complex, but only one protein from three trimeric forms. Model protein structures were prepared using the Schrodinger Protein Preparation Wizard tool (Schrodinger Suite Software): hydrogen atoms were added and minimized; missing amino acid side chains were added; bond multiplicities were restored; solvent molecules were removed; and the entire structure was optimized in the OPLS3e force field [44] at a physiological pH value. For a correct calculation procedure, we used binding site alignment procedures 7LVW and 7KQD. As results, full-size proteins and ligands (sisunatovir) in complex were obtained.
The geometric parameters of ligands (coumarin derivatives) and sisunatovir (RSV F inhibitor) were also optimized, taking into account all permissible conformations.

3.3.2. Molecular Docking

Molecular docking was performed using Schrodinger Suite (Release 2020-4) software. Coumarin derivatives were docked using the forced ligand positioning protocol (IFD) under the following conditions: flexible protein and ligand; grid matrix size of 15 Å; and amino acids (within a radius of 5 Å from the ligand) restrained and optimized, taking into account the influence of a ligand. Docking solutions were ranked by evaluating the following calculation parameters: docking score (based on GlideScore minus penalties); ligand efficiency (LE, which takes into account an atomic distribution of the scoring function); a model energy value (Emodel), including a GlideScore value, energy-unrelated interactions, and parameters of the energy spent in positioning of the ligand in the binding site.

Supplementary Materials

NMR 1H and 13C spectra of compounds 1619, 25, and 27; Energy parameters of the docking study procedure; Figure S1—best position of Sisunatovir; Figure S2—best position of 19c; Figure S3—best position of 19f; Figure S4—best position of 19h; Figure S5—the pharmacophore features of known F-protein inhibitor of sisunatovir and compounds 19c, 19f, and 19h; 1. Figure S6—Dose-response curve and half maximal inhibitory concentration (IC50) values of active compounds in Hep-2 cells against RSV A and B. Table S1—Energy parameters of docking study procedure.

Author Contributions

Conceptualization, N.F.S., A.A.S., and K.P.V.; methodology, N.F.S., A.A.S., S.S.B., and K.P.V.; investigation, T.M.K., A.A.S., A.V.G., Y.V.N., G.D.P., S.S.B., and D.V.K.; writing—original draft preparation, T.M.K., A.A.S., S.S.B., and K.P.V.; writing—review and editing, N.F.S.; supervision, N.F.S.; project administration, K.P.V. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Russian Science Foundation (Moscow, Russia) grant 21-13-00026.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

The authors would like to acknowledge the Multi-Access Chemical Research Center SB RAS for spectral and analytical measurements.

Conflicts of Interest

The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Sample Availability

Samples of the compounds presented in this study are available on request from the authors.

References

  1. Afonso, C.L.; Amarasinghe, G.K.; Bányai, K.; Bào, Y.; Basler, C.F.; Bavari, S.; Bejerman, N.; Blasdell, K.R.; Briand, F.; Briese, T.; et al. Taxonomy of the order Mononegavirales: Update 2016. Arch. Virol. 2016, 161, 2351–2360. [Google Scholar] [CrossRef] [PubMed]
  2. Karron, R.A. Preventing respiratory syncytial virus (RSV) disease in children. Science 2021, 372, 686–687. [Google Scholar] [CrossRef] [PubMed]
  3. Lozano, R.; Naghavi, M.; Foreman, K.; Lim, S.; Shibuy, K.; Aboyans, V.; Abraham, J.; Adair, T.; Aggarwal, R.; Ahn, S.Y.; et al. Global and regional mortality from 235 causes of death for 20 age groups in 1990 and 2010: A systematic analysis for the Global Burden of Disease Study 2010. Lancet 2012, 380, 2095–2128. [Google Scholar] [CrossRef]
  4. Falsey, A.R.; Hennessey, P.A.; Formica, M.A.; Cox, C.; Walsh, E.E. Respiratory syncytial virus infection in elderly and high-risk adults. N. Engl. J. Med. 2005, 352, 1749–1759. [Google Scholar] [CrossRef] [PubMed]
  5. Huang, H.; Chen, S.; Zhang, X.; Hong, L.; Zeng, Y.; Wu, B. Detection and clinical characteristics analysis of respiratory viruses in hospitalized children with acute respiratory tract infections by a GeXP-based multiplex-PCR assay. J. Clin. Lab. Anal. 2020, 34, e23127. [Google Scholar] [CrossRef] [Green Version]
  6. Nair, H.; Nokes, D.J.; Gessner, B.D.; Dherani, M.; Madhi, S.A.; Singleton, R.J.; O’Brien, K.L.; Roca, A.; Wright, P.F.; Bruce, N.; et al. Global burden of acute lower respiratory infections due to respiratory syncytial virus in young children: A systematic review and meta-analysis. Lancet 2010, 375, 1545–1555. [Google Scholar] [CrossRef] [Green Version]
  7. Available online: http://www.pharmafile.com/news/583523/parents-warned-child-respiratory-illness-rise-summer (accessed on 20 September 2021).
  8. Wainwright, C. Acute viral bronchiolitis in children—A very common condition with few therapeutic options. Paediatr. Respir. Rev. 2010, 11, 39–45. [Google Scholar] [CrossRef] [PubMed]
  9. Empey, K.M.; Peebles, R.S., Jr.; Kolls, J.K. Pharmacologic advances in the treatment and prevention of respiratory syncytial virus. Clin. Infect. Dis. 2010, 50, 1258–1267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Cockerill, G.S.; Good, J.A.D.; Mathews, N. State of the Art in Respiratory Syncytial Virus Drug Discovery and Development. J. Med. Chem. 2019, 62, 3206–3227. [Google Scholar] [CrossRef] [PubMed]
  11. Nikolayeva, Y.V.; Ulashchik, E.A.; Chekerda, E.V.; Galochkina, A.V.; Slesarchuk, N.A.; Chistov, A.A.; Nikitin, T.D.; Korshun, V.A.; Shmanai, V.V.; Ustinov, A.V.; et al. 5-(Perylen-3-ylethynyl)uracil Derivatives Inhibit Reproduction of Respiratory Viruses. Russ. J. Bioorg. Chem. 2020, 46, 315–320. [Google Scholar] [CrossRef]
  12. Huo, X.; Hou, D.; Wang, H.; He, B.; Fang, J.; Meng, Y.; Liu, L.; Wei, Z.; Wang, Z.; Liu, F.-W. Design, synthesis, in vitro and in vivo anti-respiratory syncytial virus (RSV) activity of novel oxizine fused benzimidazole derivatives. Eur. J. Med. Chem. 2021, 224, 113684. [Google Scholar] [CrossRef]
  13. Xu, J.; Wu, W.; Chen, H.; Xue, Y.; Bao, X.; Zhou, J. Substituted N-(4-amino-2-chlorophenyl)-5-chloro-2-hydroxybenzamide analogues potently inhibit respiratory syncytial virus (RSV) replication and RSV infection-associated inflammatory responses. Bioorg. Med. Chem. 2021, 39, 116157. [Google Scholar] [CrossRef]
  14. Vendeville, S.; Tahri, A.; Hu, L.; Demin, S.; Cooymans, L.; Vos, A.; Kwanten, L.; Van den Berg, J.; Battles, M.B.; McLellan, J.S.; et al. Discovery of 3-({5-Chloro-1-[3-(methylsulfonyl)propyl]-1H-indol-2-yl}methyl)-1-(2,2,2-trifluoroethyl)-1,3-dihydro-2H-imidazo[4,5-c]pyridin-2-one (JNJ-53718678), a Potent and Orally Bioavailable Fusion Inhibitor of Respiratory Syncytial Virus. J. Med. Chem. 2020, 63, 8046–8058. [Google Scholar] [CrossRef]
  15. Cockerill, G.S.; Angell, R.M.; Bedernjak, A.; Chuckowree, I.; Fraser, I.; Gascon-Simorte, J.; Gilman, M.S.A.; Good, J.A.D.; Harland, R.; Johnson, S.M.; et al. Discovery of Sisunatovir (RV521), an Inhibitor of Respiratory Syncytial Virus Fusion. J. Med. Chem. 2021, 64, 3658–3676. [Google Scholar] [CrossRef] [PubMed]
  16. Annunziata, F.; Pinna, C.; Dallavalle, S.; Tamborini, L.; Pinto, A. An Overview of Coumarin as a Versatile and Readily Accessible Scaffold with Broad-Ranging Biological Activities. Int. J. Mol. Sci. 2020, 21, 4618. [Google Scholar] [CrossRef] [PubMed]
  17. Khomenko, T.M.; Zakharenko, A.L.; Chepanova, A.A.; Ilina, E.S.; Zakharova, O.D.; Kaledin, V.I.; Nikolin, V.P.; Popova, N.A.; Korchagina, D.V.; Reynisson, J.; et al. New Promising Inhibitors of Tyrosyl-DNA Phosphodiesterase I (Tdp 1) Combining 4-Arylcoumarin and Monoterpenoid Moieties as Components of Complex Antitumor Therapy. Int. J. Mol. Sci. 2020, 21, 126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Al-Warhi, T.; Sabt, A.; Elkaeed, E.B.; Eldehna, W.M. Recent advancements of coumarin-based anticancer agents: An up-to-date review. Bioorg. Chem. 2020, 103, 104163. [Google Scholar] [CrossRef] [PubMed]
  19. Küpeli Akkol, E.; Genç, Y.; Karpuz, B.; Sobarzo-Sánchez, E.; Capasso, R. Coumarins and Coumarin-Related Compounds in Pharmacotherapy of Cancer. Cancers 2020, 12, 1959. [Google Scholar] [CrossRef] [PubMed]
  20. Carneiro, A.; Matos, M.J.; Uriarte, E.; Santana, L. Trending Topics on Coumarin and Its Derivatives in 2020. Molecules 2021, 26, 501. [Google Scholar] [CrossRef] [PubMed]
  21. Mishra, S.; Pandey, A.; Manvati, S. Coumarin: An emerging antiviral agent. Heliyon 2020, 6, e03217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Gao, Y.; Cheng, H.; Khan, S.; Xiao, G.; Rong, L.; Bai, C. Development of coumarine derivatives as potent anti-filovirus entry inhibitors targeting viral glycoprotein. Eur. J. Med. Chem. 2020, 204, 112595. [Google Scholar] [CrossRef]
  23. Zhong, Z.J.; Cheng, L.P.; Pang, W.; Zheng, X.S.; Fu, S.K. Design, synthesis and biological evaluation of dihydrofurocoumarin derivatives as potent neuraminidase inhibitors. Bioorg. Med. Chem. Lett. 2021, 37, 127839. [Google Scholar] [CrossRef] [PubMed]
  24. Lai, Y.; Han, T.; Zhan, S.; Jiang, Y.; Liu, X.; Li, G. Antiviral Activity of Isoimperatorin against Influenza a Virus in vitro and its Inhibition of Neuraminidase. Front. Pharmacol. 2021, 12, 65782. [Google Scholar] [CrossRef] [PubMed]
  25. Li, Z.; Kong, D.; Liu, Y.; Liu, Y.; Li, M. Pharmacological perspectives and molecular mechanisms of coumarin derivatives against virus disease. Genes Dis. 2021. [Google Scholar] [CrossRef]
  26. Hu, Y.; Shan, L.; Qiu, T.; Liu, L.; Chen, J. Synthesis and biological evaluation of novel coumarin derivatives in rhabdoviral clearance. Eur. J. Med. Chem. 2021, 223, 113739. [Google Scholar] [CrossRef]
  27. Mazzei, M.; Nieddu, E.; Miele, M.; Balbi, A.; Ferrone, M.; Fermeglia, M.; Mazzei, M.T.; Pricl, S.; La Colla, P.; Marongiu, F.; et al. Activity of Mannich bases of 7-hydroxycoumarin against Flaviviridae. Bioorg. Med. Chem. 2008, 16, 2591–2605. [Google Scholar] [CrossRef] [PubMed]
  28. Yarovaya, O.I.; Salakhutdinov, N.F. Mono- and sesquiterpenes as a starting platform for the development of antiviral drugs. Russ. Chem. Rev. 2021, 90, 488–510. [Google Scholar] [CrossRef]
  29. Salakhutdinov, N.F.; Volcho, K.P.; Yarovaya, O.I. Monoterpenes as a renewable source of biologically active compounds. Pure Appl. Chem. 2017, 89, 1105–1117. [Google Scholar] [CrossRef]
  30. Khomenko, T.M.; Zarubaev, V.V.; Orshanskaya, I.R.; Kadyrova, R.A.; Sannikova, V.A.; Korchagina, D.V.; Volcho, K.P.; Salakhutdinov, N.F. Anti-influenza activity of monoterpene-containing substituted coumarins. Bioorg. Med. Chem. Lett. 2017, 27, 2920–2925. [Google Scholar] [CrossRef]
  31. Khomenko, T.; Zakharenko, A.; Odarchenko, T.; Arabshahi, H.J.; Sannikova, V.; Zakharova, O.; Korchagina, D.; Reynisson, J.; Volcho, K.; Salakhutdinov, N.; et al. New inhibitors of tyrosyl-DNA phosphodiesterase I (Tdp 1) combining 7-hydroxycoumarin and monoterpenoid moieties. Bioorg. Med. Chem. 2016, 24, 5573–5581. [Google Scholar] [CrossRef]
  32. Akgun, B.; Hall, D.G. Fast and tight boronate formation for click bioorthogonal conjugation. Angew. Chem. Int. Ed. 2016, 55, 3909–3913. [Google Scholar] [CrossRef] [PubMed]
  33. Il’ina, I.V.; Volcho, K.P.; Korchagina, D.V.; Barkhash, V.A.; Salakhutdinov, N.F. Reactions of Allyl Alcohols of the Pinane Series and of Their Epoxides in the Presence of Montmorillonite Clay. Helv. Chim. Acta 2007, 90, 353–368. [Google Scholar] [CrossRef]
  34. Chapuis, C.; Brauchli, R. Preparation of campholenal analogues: Chirons for the lipophilic moiety of sandalwood-like odorant alcohols. Helv. Chim. Acta 1992, 75, 1527–1546. [Google Scholar] [CrossRef]
  35. Cointeaux, L.; Berrien, J.-F.; Mayrargue, J. Synthesis of cardamom peroxide analogues by radical cyclization of hydroperoxyalkenes. Tetrahedron Lett. 2002, 43, 6275–6277. [Google Scholar] [CrossRef]
  36. Pozdnev, V.F. Improved method for synthesis of 7-amino-4-methylcoumarin. Chem. Heterocycl. Compd. 1990, 26, 264–265. [Google Scholar] [CrossRef]
  37. Kathuria, A.; Priya, N.; Chand, K.; Singh, P.; Gupta, A.; Jalal, S.; Gupta, S.; Raj, H.G.; Sharma, S.K. Substrate specificity of acetoxy derivatives of coumarins and quinolones towards Calreticulin mediated transacetylation: Investigations on antiplatelet function. Bioorg. Med. Chem. 2012, 20, 1624–1638. [Google Scholar] [CrossRef]
  38. Collins, P.L.; Fearns, R.; Graham, B.S. Respiratory Syncytial Virus: Virology, Reverse Genetics, and Pathogenesis of Disease. In Challenges and Opportunities for Respiratory Syncytial Virus Vaccines; Anderson, L., Graham, B., Eds.; Current Topics in Microbiology and Immunology; Springer: Berlin/Heidelberg, Germany, 2013; Volume 372. [Google Scholar] [CrossRef] [Green Version]
  39. Rossey, I.; Hsieh, C.-L.; Sedeyn, K.; Ballegeer, M.; Schepens, B.; McLellan, J.S.; Saelens, X. A vulnerable, membrane-proximal site in human respiratory syncytial virus F revealed by a prefusion-specific single-domain antibody. J. Virol. 2021, 95, e02279-20. [Google Scholar] [CrossRef] [PubMed]
  40. Berman, H.M.; Westbrook, J.; Feng, Z.; Gilliland, G.; Bhat, T.N.; Weissig, H.; Shindyalov, I.N.; Bourne, P.E. The Protein Data Bank. Nucleic Acids Res. 2000, 28, 235–242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Kiesgen de Richter, R.; Bonato, M.; Follet, M.; Kamenka, J.M. The (+)- and (−)-[2-(1,3-dithianyl0]myrtanylborane. Solid and stable monoalkylboranes for asymmetric hydroboration. J. Org. Chem. 1990, 55, 2855–2860. [Google Scholar] [CrossRef]
  42. El-Haggar, R.; Al-Wabli, R.I. Anti-inflammatory screening and molecular modeling of some novel coumarin derivatives. Molecules 2015, 20, 5374–5391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Ponomarev, K.; Pavlova, A.; Suslov, E.; Ardashov, O.; Korchagina, D.; Nefedov, A.; Tolstikova, T.; Volcho, K.; Salakhutdinov, N. Synthesis and analgesic activity of new compounds combining azaadamantane and monoterpene moieties. Med. Chem. Res. 2015, 24, 4146–4156. [Google Scholar] [CrossRef]
  44. Shivakumar, D.; Harder, E.; Damm, W.; Friesner, R.A.; Sherman, W. Improving the Prediction of Absolute Solvation Free Energies Using the Next Generation OPLS Force Field. J. Chem. Theory Comput. 2012, 8, 2553–2558. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Known compounds with anti-RSV activity.
Figure 1. Known compounds with anti-RSV activity.
Molecules 26 07493 g001
Scheme 1. Synthesis of 7-hydroxycoumarins.
Scheme 1. Synthesis of 7-hydroxycoumarins.
Molecules 26 07493 sch001
Scheme 2. Synthesis of monoterpenoid bromides 11ad.
Scheme 2. Synthesis of monoterpenoid bromides 11ad.
Molecules 26 07493 sch002
Scheme 3. Synthesis of bromides 11eh.
Scheme 3. Synthesis of bromides 11eh.
Molecules 26 07493 sch003
Scheme 4. Synthesis of substituted coumarins 1619.
Scheme 4. Synthesis of substituted coumarins 1619.
Molecules 26 07493 sch004
Scheme 5. Synthesis of aminocoumarins 23 and 24.
Scheme 5. Synthesis of aminocoumarins 23 and 24.
Molecules 26 07493 sch005
Scheme 6. Synthesis of aminocoumarin derivatives 2527.
Scheme 6. Synthesis of aminocoumarin derivatives 2527.
Molecules 26 07493 sch006
Figure 2. Activity of 19c against RSV A according to time-of-addition experiment.
Figure 2. Activity of 19c against RSV A according to time-of-addition experiment.
Molecules 26 07493 g002
Figure 3. Structure of the RSV F-protein: (A) structure corresponds to PDB [40] code 7LVW [39]: fusion peptide region is presented in violet secondary structure (137–157 amino acids); functional residues (137–140) are shown in red; amino acids of the membrane anchor are shown in yellow (amino acids 454–499); the green molecule is sisunatovir [15]; (B) sisunatovir in the binding site; (CE) positions of possible entry inhibitors (19c, 19f, and, 19h). π–π stacking interactions are denoted by blue dotted lines; H-bonds and salt bridges are denoted by yellow and violet lines, respectively.
Figure 3. Structure of the RSV F-protein: (A) structure corresponds to PDB [40] code 7LVW [39]: fusion peptide region is presented in violet secondary structure (137–157 amino acids); functional residues (137–140) are shown in red; amino acids of the membrane anchor are shown in yellow (amino acids 454–499); the green molecule is sisunatovir [15]; (B) sisunatovir in the binding site; (CE) positions of possible entry inhibitors (19c, 19f, and, 19h). π–π stacking interactions are denoted by blue dotted lines; H-bonds and salt bridges are denoted by yellow and violet lines, respectively.
Molecules 26 07493 g003
Table 1. Antiviral activity and cytotoxicity of compounds 1618, 2527 against RSV A and B.
Table 1. Antiviral activity and cytotoxicity of compounds 1618, 2527 against RSV A and B.
CompoundRCC50 a, µMRSV ARSV B
IC50 b, µMSIcIC50, µMSI c
16a Molecules 26 07493 i001 Molecules 26 07493 i00258.7 ± 7.66.7 ± 0.87.910.7±2.15.5
16b Molecules 26 07493 i003113 ± 25.6>111<134.4 ± 1.73.3
16c Molecules 26 07493 i004307.1 ± 34.627 ± 1.811.418.5±0.916.8
16d Molecules 26 07493 i00525.8 ± 1.62.9 ± 0.678.923.5±0.71.1
16e Molecules 26 07493 i00624.7 ± 6.7>25<114.8 ± 0.91.7
16f Molecules 26 07493 i007121.1 ± 24.52.5 ± 0.8764.6d.n.t.d-
16g Molecules 26 07493 i0081204.4 ± 20441.6 ± 3.229.937.6 ± 7.432
17a Molecules 26 07493 i009 Molecules 26 07493 i010483.4 ± 23.646.7 ± 8.710.352.4±3.49.2
17b Molecules 26 07493 i011795.7 ± 28.929 ± 3.227.4241 ± 413.3
17c Molecules 26 07493 i012116 ± 23.61.3 ± 0.75904.7±0.924.7
17e Molecules 26 07493 i01326.6 ± 12.312.4 ± 2.12.15 ± 1.75.3
17f Molecules 26 07493 i014658.2 ± 54.613.9 ± 1.662.120.1±3.232.7
17g Molecules 26 07493 i015172 ± 21.1>60.73>185.6<1.2
18a Molecules 26 07493 i016 Molecules 26 07493 i01726.6 ± 12.3>26<115.1±3.41.8
18b Molecules 26 07493 i01813.4 ± 4.6>13<1>130.9
18c Molecules 26 07493 i019416.2 ± 32.94.1 ± 0.781005.1±1.081.6
18d Molecules 26 07493 i020408.1 ± 56.1>400<1408.6<1
18e Molecules 26 07493 i02166.7 ± 23.14.1 ± 0.718.324.7 ± 5.32.7
18f Molecules 26 07493 i02247.3 ± 12.750.2 ± 13.2110 ± 1.74.7
18g Molecules 26 07493 i02337.2 ± 9.8>37<121.7 ± 2.71.8
19a Molecules 26 07493 i024 Molecules 26 07493 i02551.1 ± 11.30.57 ± 0.2900.6±0.1.85.2
19b Molecules 26 07493 i026333.9 ± 56.8107.6 ± 13.53.160.5 ± 11.65.5
19c Molecules 26 07493 i027379.5 ± 29.45.1 ± 1.1774.9 ± 0.678.2
19d Molecules 26 07493 i02821.9 ± 11.80.6 ± 0.13400.82±0.126.7
19e Molecules 26 07493 i02913.2 ± 4.86.1 ± 2.12.2>13<1
19f Molecules 26 07493 i030718.4 ± 21.915.3 ± 1.857.98.2 ± 0.8388.1
19g Molecules 26 07493 i031930.5 ± 56.981.8 ± 3.311.419.8±1.746.9
19h Molecules 26 07493 i03231.3 ± 7.4>31<15.2 ± 0.696
25 Molecules 26 07493 i033355,6 ± 12,733.7 ± 7.610.5114.9 ± 16.73.1
26 Molecules 26 07493 i03484,3 ± 11,76.2 ± 1.214.5>84<0.9
27 Molecules 26 07493 i03553,5 ± 10,951.5 ± 5.41>53<0.9
Ribavirin >409531.1 ± 6.7131.654.5±5.975.1
a CC50 is the median cytotoxic concentration, i.e., the concentration causing 50% cell death. b IC50 is the 50% inhibiting concentration, i.e., the concentration causing a 50% decrease in virus replication. c SI is the selectivity index, which is the CC50/IC50 ratio. CC50’s and IC50’s are presented as mean ± standard deviation. The values are calculated from three independent experiments. d d.n.t.—did not test.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Khomenko, T.M.; Shtro, A.A.; Galochkina, A.V.; Nikolaeva, Y.V.; Petukhova, G.D.; Borisevich, S.S.; Korchagina, D.V.; Volcho, K.P.; Salakhutdinov, N.F. Monoterpene-Containing Substituted Coumarins as Inhibitors of Respiratory Syncytial Virus (RSV) Replication. Molecules 2021, 26, 7493. https://doi.org/10.3390/molecules26247493

AMA Style

Khomenko TM, Shtro AA, Galochkina AV, Nikolaeva YV, Petukhova GD, Borisevich SS, Korchagina DV, Volcho KP, Salakhutdinov NF. Monoterpene-Containing Substituted Coumarins as Inhibitors of Respiratory Syncytial Virus (RSV) Replication. Molecules. 2021; 26(24):7493. https://doi.org/10.3390/molecules26247493

Chicago/Turabian Style

Khomenko, Tatyana M., Anna A. Shtro, Anastasia V. Galochkina, Yulia V. Nikolaeva, Galina D. Petukhova, Sophia S. Borisevich, Dina V. Korchagina, Konstantin P. Volcho, and Nariman F. Salakhutdinov. 2021. "Monoterpene-Containing Substituted Coumarins as Inhibitors of Respiratory Syncytial Virus (RSV) Replication" Molecules 26, no. 24: 7493. https://doi.org/10.3390/molecules26247493

Article Metrics

Back to TopTop