Next Article in Journal
Wood Extractives of Silver Fir and Their Antioxidant and Antifungal Properties
Next Article in Special Issue
New Generation of Meso and Antiprogestins (SPRMs) into the Osteoporosis Approach
Previous Article in Journal
FTIR Spectroscopy as a Tool to Study Age-Related Changes in Cardiac and Skeletal Muscle of Female C57BL/6J Mice
Previous Article in Special Issue
Peptaibol-Containing Extracts of Trichoderma atroviride and the Fight against Resistant Microorganisms and Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Facile Approach to Bis(isoxazoles), Promising Ligands of the AMPA Receptor

by
Dmitry A. Vasilenko
1,
Kirill S. Sadovnikov
1,
Kseniya N. Sedenkova
1,
Dmitry S. Karlov
1,
Eugene V. Radchenko
1,
Yuri K. Grishin
1,
Victor B. Rybakov
1,
Tamara S. Kuznetsova
1,
Vladimir L. Zamoyski
2,
Vladimir V. Grigoriev
1,2,
Vladimir A. Palyulin
1,* and
Elena B. Averina
1,*
1
Department of Chemistry, Lomonosov Moscow State University, 119991 Moscow, Russia
2
Institute of Physiologically Active Compounds, Russian Academy of Sciences, Chernogolovka, 142432 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(21), 6411; https://doi.org/10.3390/molecules26216411
Submission received: 24 September 2021 / Revised: 14 October 2021 / Accepted: 21 October 2021 / Published: 23 October 2021
(This article belongs to the Special Issue Medicinal Chemistry in Europe III)

Abstract

:
A convenient synthetic approach to novel functionalized bis(isoxazoles), the promising bivalent ligands of the AMPA receptor, was elaborated. It was based on the heterocyclization reactions of readily available electrophilic alkenes with the tetranitromethane-triethylamine complex. The structural diversity of the synthesized compounds was demonstrated. In the electrophysiological experiments using the patch clamp technique on Purkinje neurons, the compound 1,4-phenylenedi(methylene)bis(5-aminoisoxazole-3-carboxylate) was shown to be highly potent positive modulator of the AMPA receptor, potentiating kainate-induced currents up to 70% at 10−11 M.

Graphical Abstract

1. Introduction

An isoxazole ring occurs in various natural and synthetic compounds with a wide spectrum of biological activities [1]. Isoxazole derivatives are known as marketed drugs [2,3,4,5,6] and as promising candidates for the anticancer, antiviral, antifungal, antimicrobial, immunomodulating, analgesic, and antipsychotic drugs, among others [7,8,9,10,11,12,13]. One of the isoxazole-containing compounds, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA), is a specific agonist for the AMPA receptor, one of the most important subtypes of glutamate receptors. Involved in many significant neurophysiological processes related to the fast synaptic excitatory transmission, memory formation, etc., the AMPA receptor is a promising pharmacological target [14,15]. Its positive allosteric modulators (PAMs) are potential candidates for the development of cognition enhancers and drugs against the Alzheimer’s and Parkinson’s diseases, multiple sclerosis, soft cognitive disorders, age-related cognition and memory impairments, autism, depression, drug addiction, etc. [16,17,18,19,20,21,22,23,24]. They attract significant interest due to their neuroprotective effect and the lack of the excitotoxicity problems that result from an overdose of a direct agonist. Therefore, the creation of compounds acting as allosteric modulators of AMPA receptors is currently one of the most promising directions for the progress of neuropharmacology.
We were able to develop a series of novel AMPA receptor PAMs based on different scaffolds, which possessed very high levels of experimentally confirmed activity [25,26,27,28,29]. In the design of the compounds for the present study, we took into account that the dimeric and larger monomeric positive allosteric modulators of the AMPA receptor are known to bind more efficiently and demonstrate significantly higher potency in the nanomolar or picomolar concentration ranges [27,28,29,30,31,32], compared with their monomeric units or other small monomeric ligands. This is due to the U-shaped form of the AMPA receptor’s allosteric binding site, which allows both structural fragments in the molecule to interact with the receptor simultaneously. When selecting and prioritizing potential PAM chemotypes for the synthesis and biological evaluation, we considered the pharmacophore and QSAR models [33,34,35] that we have developed, the results of the molecular docking and molecular dynamics simulations, the predicted physico-chemical and ADMET (absorption, distribution, metabolism, excretion, toxicity) profiles, and our previous experience in designing the modulators with some related scaffolds and validating their activity [29,32]. In particular, the structures containing the phenylene-based linkers and aminoisoxazole moieties were identified among the most promising potential PAMs.
Recently, we have elaborated a novel preparative method for the synthesis of 3-EWG-5-nitroisoxazoles, based on the heterocyclization of electrophilic alkenes via treatment with the tetranitromethane-triethylamine (TNM-TEA) complex [36,37,38,39]. Until now, this method has been used to produce polysubstituted isoxazoles containing one heterocyclic fragment in a molecule. In this work, with a focus on the future exploitation of bis(isoxazoles) as bivalent ligands of the AMPA receptor, electrophilic dienes containing two electrophilic double bonds were studied for the first time in the simultaneous heterocyclization of both double bonds under the action of the TNM-TEA complex. We report the synthesis of the desired compounds, as well as the results of their electrophysiological evaluation, the molecular modeling aiming to elucidate the mechanism of action, and the prediction of their physico-chemical and ADMET properties.

2. Results and Discussion

2.1. Chemistry

Easily available unsaturated esters 2a-i containing two electrophilic double bonds were utilized in the heterocyclization reaction (Table 1).
The starting dienes 2a-i were synthesized from the diols 1a-i using two equivalents of acryl chloride and Et3N [40]. Initially, the reaction was performed in CH2Cl2; however, in the case of the diols 1a-f, the products were obtained in moderate yields due to the low solubility of the starting compounds 1a-f. We found that changing the solvent to CH3CN in the acylation reaction significantly improved the yield of 2a-f. For the synthesis of 2g and 2h, THF was used due to the low solubility of the corresponding diols 1g and 1h in the other solvents. Thus, under the modified conditions, the acrylic esters 2a-i bearing the linear, aromatic, and polycyclic moieties were obtained in good yields.
The electrophilic dienes 2a-i were involved in the heterocyclization reaction upon treatment with the in situ generated TNM-TEA complex in 1,4-dioxane at 70 °C to give bis(5-nitroisoxazoles) 3a-h in moderate to good yields. It was found that the yields of the heterocycles 3a-c with the linear linkers tended to increase with the elongation of the carbon chain. The bis(5-nitroisoxazoles) 3d-f bearing the benzyl ester fragments were obtained in modest yields due to the partial oxidation of the benzyl groups in the presence of tetranitromethane, which can act as an oxidizer [41]. It should be noted that the heterocyclization of the adamantane-containing diene 2g led to the target bis(isoxazole) 3g also giving a side product 4 (Scheme 1).
Surprisingly, we failed to involve the hydroquinone derivative, diene 2i with isolated double bonds, in the heterocyclization as only bis(4,4,4-trinitrobutanoic ester) 5a was isolated as the main product (Scheme 2). Trinitro-substituted adduct formation is an alternative pathway for the reaction of electrophilic alkenes with the TNM-TEA complex according to the previously suggested mechanism [36,38]. To affirm whether this route is general for the aryl acrylic esters, we obtained alkene 2j and employed it in the reaction with TNM-TEA, which also afforded only Michael adduct 5b under various conditions (Scheme 2).
All synthesized compounds were characterized and identified using 1H and 13C NMR spectroscopy (see Supplementary Materials). The structure of the obtained bis(5-nitroisoxazole) 3c was additionally confirmed by the X-ray analysis (Figure 1) (see Supplementary Materials).
According to molecular modeling, bis(5-aminoisoxazoles) can be of interest as bivalent modulators of the AMPA receptor and compound 6 is among the most promising ones based on the pharmacophore and QSAR models, the results of the molecular docking and molecular dynamics simulations, and the predicted physico-chemical and ADMET profiles. In view of this, we have studied the reduction of 5-nitro substituted heterocycle 3d in various conditions [12,13,42,43]. The attempt to reduce 5-nitroisoxazole 3d using SnCl2 in EtOH failed due to its poor solubility, yet the application of the system Na2S2O4/THF/H2O allowed us to obtain the target bis(5-aminoisoxazole) 6 in a satisfactory yield (Scheme 3).
Unfortunately, the attempts to obtain other bis(5-aminoisoxazoles) failed because of their extremely low solubility in all common solvents. In particular, the application of the different reduction conditions for bis(5-nitroisoxazole) 3c resulted in a negligibly small amount of corresponding 5-amino derivatives, and only acylated product 7 was successfully isolated in a good yield (Scheme 4).

2.2. Electrophysiological Evaluation

Next, bis(5-aminoisoxazole) 6 was examined as a positive modulator of the AMPA receptor. The electrophysiological experiments were carried out using the patch clamp technique on freshly isolated Purkinje neurons, as described earlier [28,29]. The influence of the compound 6 on the kainate-induced currents is shown in Table 2. The potentiation of the kainate-induced AMPA receptor currents was observed in a wide concentration range (10–12–10–6 M) and had a bell-shaped concentration dependence with maximum potentiation at 10–11 M.
As can be seen, the activity of the compound 6 was comparable or superior, in terms of effective concentration and/or maximum potentiation, to that of some of the most potent known positive modulators of the AMPA receptor 8, 9, and was significantly higher than that of cyclothiazide (CTZ) (Figure 2) [29,32,44].

2.3. Molecular Modeling

In order to elucidate the probable mechanism of action of the positive allosteric modulator 6, its interaction with the dimeric ligand-binding domain of the GluA2 AMPA receptor was modeled using a molecular docking and molecular dynamics simulation. The compound’s binding mode in the PAM binding site, at the interface between ligand-binding domains, was stable over the entire course of the simulation (100 ns). Similarly to the other larger dimeric modulators [28], the 6 molecule attained a slightly unsymmetric position, occupying a central subpocket as well as one of the side subpockets of the symmetrical PAM binding site (Figure 3A,B). The binding was stabilized by steric fit, hydrophobic interactions, and a number of hydrogen bonds (Figure 3B,C). The plot of the root mean square deviations (RMSD) for the protein, glutamate, and ligand (6) heavy atoms (Figure 4) as well as the visual inspection of the trajectory confirm that the system stability was retained over the entire course of the production simulation (100 ns), although the ligand position was slightly adjusted compared to the docking pose. Overall, these results confirmed that the compound 6 could indeed act as a positive AMPA receptor modulator binding in the validated PAM binding site.

2.4. Prediction of ADMET, Physicochemical, and PAINS Profiles

Several ADMET and physicochemical properties for the compound 6 were calculated (Table 3). They demonstrated a high predicted value for the intestinal absorption, enabling its oral administration. The predicted lipophilicities and aqueous solubilities were also appropriate for potential drug-like compounds according to the commonly accepted rules of thumb. Due to the moderate predicted blood–brain barrier permeability (brain concentration is expected to be about 10% of the plasma concentration), acceptable CNS bioavailability could be anticipated. Both parameters of the cardiac toxicity risk (hERG pKi and pIC50) (4.1–6.0 log units) were in the lower or medium parts of their possible ranges (3–9 log units), indicating a likely absence of the hERG liabilities. The integral quantitative estimate of drug-likeness (QED) was greater than 0.6, confirming the favorable likely properties. The pan assay interference compounds (PAINS) filter check did not identify any alerts.
Overall, the predicted ADMET, physicochemical, and PAINS properties of the positive allosteric modulator 6 were quite acceptable for the potential lead compounds at the early drug development stages, although additional checks and structure optimization would likely be required.

3. Materials and Methods

3.1. Chemistry

The 1H and 13C NMR spectra were recorded on the 400 MHz spectrometers Bruker Avance 400 and Agilent 400-MR (400.0 MHz for 1H; 100.6 MHz for 13C) at room temperature. The chemical shifts δ were measured in ppm with respect to the solvent (1H: CDCl3, δ = 7.26 ppm, CD3OD, δ = 3.31 ppm, CD3C(O)CD3, δ = 2.05 ppm, (CD3)2SO, δ = 2.50 ppm; 13C: CDCl3, δ = 77.16 ppm, CD3OD, δ = 49.0 ppm, CD3C(O)CD3, δ = 29.84 ppm, (CD3)2SO δ = 39.52 ppm). Chemical shifts (δ) are given in ppm; J values are given in Hz. When necessary, the assignments of signals in the NMR spectra were made using 2D techniques. The accurate mass measurements (HRMS) were performed on a Bruker micrOTOF II instrument, using electrospray ionization (ESI). The measurements were performed in a positive ion mode (interface capillary voltage 4500 V), or in a negative ion mode (3200 V). The melting points (mp) were uncorrected. The analytical thin layer chromatography was carried out with Silufol silica gel plates (supported on aluminum), and the detection was performed by a UV lamp (254 and 365 nm) and chemical staining (5% aqueous solution of KMnO4). The column chromatography was performed on silica gel (230–400 mesh, Merck, Darmstadt, Germany).
Diols 1d [45], 1e [46], 1f [47], 1g [48] were synthesized by the described methods. Dienes 2a-f,i were also synthesized by the described method [40].
All other starting materials were commercially available.

3.1.1. General Procedure for the Preparation of Compounds 2g,h

Triethylamine (1.40 mL, 1.01 g, 10.0 mmol) was added to the solution of starting diol 1 (5.0 mmol) in THF (70 mL) under an Ar atmosphere. The mixture was cooled to (–10)–(–5) °C and the acryloyl chloride (0.81 mL, 0.91 g, 10.0 mmol) was added dropwise. The resulting mixture was stirred at this temperature for 3 h, and then at room temperature for 24 h. After cooling, the reaction mixture was filtered through a layer of Al2O3, which was washed with CH2Cl2. The filtrate was concentrated in vacuo, and the obtained residue was purified by silica-gel column chromatography to afford the desired product.
Adamantane-1,3-diylbis(methylene) diacrylate (2g) was isolated to a pure state as a colorless oil with a 46% (0.70 g) yield; Rf 0.28 (petroleum ether:EtOAc = 20:1); 1H NMR (CDCl3, 400 MHz): δ 1.32–1.38 (m, 2H, CH2), 1.42–1.56 (m, 8H, 4CH2), 1.58–1.66 (m, 2H, CH2), 2.04–2.12 (m, 2H, 2CH), 3.78 (s, 4H, 2CH2O), 5.80 (dd, 2J = 1.6 Hz, 3J = 10.6 Hz, 2H, 2CH2), 6.11 (dd, 3J = 10.6 Hz, 3J = 17.4 Hz, 2H, 2CHCO), 6.35 (dd, 2J = 1.6 Hz, 3J = 17.4 Hz, 2H, 2CH2); 13C NMR (CDCl3, 100.6 MHz): δ 28.0 (2CH), 33.8 (CH2), 36.3 (CH2), 38.8 (4CH2), 41.1 (2C), 73.6 (2CH2O), 128.6 (2CH=), 130.6 (2CH=), 166.3 (C=O); HRMS-ESI (M + Na+): Found: 327.1555. Calculated for C18H24NaO4+: 327.1567.
[(3s,7s)-7-(Acryloyloxy)bicyclo[3.3.1]non-3-yl]methyl acrylate (2h) was isolated to a pure state as colorless oil with a 54% (0.64 g) yield; Rf 0.49 (petroleum ether:EtOAc = 10:1); 1H NMR (CDCl3, 400 MHz): δ 1.10–1.20 (m, 1H, CH2), 1.37– 1.54(m, 2H, CH2), 1.69–1.81 (m, 2H, CH2), 1.82–1.92 (m, 2H, CH2), 1.90–2.10 (m, 4H, 2CH2), 2.10–2.20 (m, 2H, 2CH), 3.94 (d, 3J= 6.6 Hz, 2H, CH2O), 5.16–5.26 (m, 1H, CHO), 5.80 (dd, 2J = 1.5 Hz, 3J = 10.5 Hz, 1H, CH2), 5.82 (dd, 2J = 1.5 Hz, 3J = 10.5 Hz, 1H, CH2), 6.10 (dd, 3J = 10.5 Hz, 3J = 17.3 Hz, 2H, 2CH), 6.38 (dd, 2J = 1.5 Hz, 3J = 17.3 Hz, 2H, CH2); 13C NMR (CDCl3, 100.6 MHz): δ 23.5 (2CH), 28.1 (CH2), 29.4 (2CH2), 29.7 (CH), 37.3 (2CH2), 70.3 (CHO + CH2O), 128.7 (=CH), 129.4 (=CH), 130.6 (=CH2), 130.6 (=CH2), 165.6 (C=O), 166.4 (C=O); HRMS-ESI (M + Na+): Found: 301.1412. Calculated for C16H22NaO4+: 301.1410.

3.1.2. Synthesis of Compound 2j

Triethylamine (0.70 mL, 0.51 g, 5.0 mmol) was added to the solution of 1j (1.14 g, 5.0 mmol) in CH2Cl2 (10.00 mL) under an Ar atmosphere. The mixture was cooled to (–10)–(–5) °C and the acryloyl chloride (0.40 mL, 0.46 g, 5.0 mmol) was added dropwise. The resulting mixture was stirred at this temperature for 3 h, and then at room temperature for 24 h. The reaction mixture was poured into water (15 mL) and extracted with dichloromethane (4 × 20 mL). The combined organic layer was washed with brine solution (3 × 20 mL) and was dried over anhydrous Mg2SO4. The solvent was evaporated in vacuo, and the residue was purified by flash chromatography on silica gel to afford the desired product.
4-(Adamantan-1-yl)phenyl acrylate (2j) was isolated to a pure state as a colorless solid with a 95% (1.34 g) yield; mp. = 98–99 °C; 1H NMR (CDCl3, 400 MHz): δ 1.71–1.84 (m, 6H, 3CH2(Ad)), 1.88–1.97 (m, 6H, 3CH2(Ad)), 2.07–2.15 (m, 3H, 3CH(Ad)), 5.99 (dd, 2J = 1.2 Hz, 3J = 10.6 Hz, 1H, CH2), 6.32 (dd, 3J = 10.6 Hz, 3J = 17.6 Гц, 1H, CHCO), 6.60 (dd, 2J = 1.2 Hz, 3J = 17.6 Hz, 1H, CH2), 7.05–7.11 (m, 2H, 2CH(Ar)), 7.35–7.41 (m, 2H, 2CH(Ar)). 13C NMR (CDCl3, 100.6 MHz): δ 29.0 (3CH), 36.1 (C) 36.8 (3CH2), 43.3 (3CH2), 120.9 (2CH(Ar)), 126.0 (2CH(Ar)), 128.2 (=CH), 132.4 (=CH2), 148.3 (C), 149.0 (C), 164.8 (C=O); HRMS-ESI (M + H+): Found: 283.1691. Calculated for C19H23O2+: 283.1693.

3.1.3. General Procedure for the Preparation of Bis(isoxazoles) 3a-h, 4, 5a

Triethylamine (0.56 mL, 0.40 g, 4.0 mmol) was added dropwise to a solution of tetranitromethane (0.60 mL, 0.98 g, 5.0 mmol) in 1,4-dioxane (6.00 mL) at 0 °C (ice bath). The reaction mixture was stirred for 5 min, and alkene (1.0 mmol) in 2.00 mL of 1,4-dioxane was added dropwise. Then, the cooling factor was removed and the mixture was stirred at 70 °C for 2 h. The solvent was removed under reduced pressure, and the product was isolated by column chromatography.
Ethane-1,2-diyl bis(5-nitroisoxazole-3-carboxylate) (3a) was isolated to a pure state as a colorless solid with a 46% (0.16 g) yield; mp.= 138–141 °C; Rf 0.24 (petroleum ether:EtOAc = 3:1); 1H NMR (CDCl3, 400 MHz): δ 4.81 (s, 4H, 2CH2O), 7.42 (s, 2H, 2CH); 13C NMR (CDCl3, 100.6 MHz): δ 64.0 (2CH2O), 102.4 (2CH), 157.6 (2C), 157.7 (2C), 166.0 (2CNO2); Anal. Calculated for C6H6N2O5: C, 35.10; H, 1.77; N, 16.37. Found: C, 35.12; H, 1.79; N, 16.09.
Propane-1,3-diyl bis(5-nitroisoxazole-3-carboxylate) (3b) was isolated to a pure state as a colorless solid with a 55% (0.20 g) yield; mp. = 75–77 °C; Rf 0.14 (petroleum ether:EtOAc = 4:1). 1H NMR (CDCl3, 400 MHz): δ 2.35 (quint., 3J = 6.1 Hz, 2H, CH2), 4.59 (t, 3J = 6.1 Hz, 4H, 2CH2O), 7.43 (s, 2H, 2CH); 13C NMR (CDCl3, 100.6 MHz): δ 27.7 (CH2), 63.7 (2CH2O), 102.5 (2CH), 157.7 (2C), 158.0 (2C), 165.8 (2CNO2); Anal. Calculated for C6H6N2O5: C, 37.09; H, 2.26; N, 15.73. Found: C, 37.57; H, 2.07 N, 15.21.
Butane-1,4-diyl bis(5-nitroisoxazole-3-carboxylate) (3c) was isolated to a pure state as a colorless solid with a 64% (0.24 g) yield; mp. = 124 °C; Rf 0.16 (petroleum ether:EtOAc = 4:1). 1H NMR (CDCl3, 400 MHz): δ 1.95–2.03 (m, 4H, 2CH2), 4.48–4.56 (m, 4H, 2CH2O), 7.41 (s, 2H, 2CH); 13C NMR (CDCl3, 100.6 MHz): δ 25.2 (2CH2), 66.5 (2CH2O), 102.4 (2CH), 157.8 (2C=O), 158.2 (2C), 165.9 (2CNO2).
1,4-Phenylenedi(methylene) bis(5-nitroisoxazole-3-carboxylate) (3d) was isolated to a pure state as a colorless solid with a 32% (0.13 g) yield; mp. = 150–151 °C; Rf 0.24 (petroleum ether:EtOAc = 4:1); 1H NMR (CDCl3, 400 MHz): δ 5.46 (s, 4H, 2CH2O), 7.40 (s, 2H, 2CH), 7.50 (s, 4H, 4CH(Ar)); 13C NMR (CDCl3, 100.6 MHz): δ 68.3 (2CH2O), 102.5 (2CH), 129.4 (4CH(Ar)), 135.1 (2C(Ar)), 157.6 (2C=O), 158.1 (2C), 165.8 (2CNO2); HRMS-ESI (M + Na+): Found: 441.0296. Calculated for C16H10N4NaO10+: 441.0289.
1,3-Phenylenedi(methylene) bis(5-nitroisoxazole-3-carboxylate)(3e) was isolated to a pure state as a colorless solid with a 40% (0.17 g) yield; mp. = 108–110 °C; Rf 0.22 (petroleum ether:EtOAc = 3:1). 1H NMR (CDCl3, 400 MHz): δ 5.47 (s, 4H, 2CH2O), 7.41 (s, 2H, 2CH), 7.45–7.50 (m, 3H, 3CH(Ar)), 7.53–7.56 (m, 1H, CH(Ar)); 13C NMR (CDCl3, 100.6 MHz): δ 68.3 (2CH2O), 102.5 (2CH), 129.1 (CH(Ar)), 129.52 (2CH(Ar)), 129.59 (CH(Ar)), 134.9 (2C(Ar)), 157.6 (2C=O), 158.1 (2C), 165.8 (2CNO2); HRMS-ESI (M + Na+): Found: 441.0283. Calculated for C16H10N4NaO10+: 441.0289.
1,2-Phenylenedi(methylene) bis(5-nitroisoxazole-3-carboxylate) (3f) was isolated to a pure state as a colorless solid with a 22% (92 mg) yield; mp. = 112–114 °C; Rf 0.56 (petroleum ether:EtOAc = 3:1); 1H NMR (CDCl3, 400 MHz): δ 5.62 (s, 4H, 2CH2O), 7.40 (s, 2H, 2CH), 7.43–7.49 (m, 2H, 2CH(Ar)), 7.52–7.59 (m, 2H, 2CH(Ar)); 13C NMR (CDCl3, 100.6 MHz): δ 66.3 (2CH2O), 102.5 (2CH), 120.1 (2CH(Ar)), 131.3 (2CH(Ar)), 133.3 (2C(Ar)), 157.5 (2C=O), 158.0 (2C), 165.8 (2CNO2); HRMS-ESI (M + Na+): Found: 441.0285. Calculated for C16H10N4NaO10+: 441.0289.
Adamantane-1,3-diylbis(methylene) bis(5-nitroisoxazole-3-carboxylate) (3g) was isolated to a pure state as a colorless solid with a 51% (0.24 g) yield; mp. = 127–129 °C; Rf 0.27 (petroleum ether:EtOAc = 5:1); 1H NMR (CDCl3, 400 MHz): δ 1.46–1.53 (m, 2H, CH2(Ad)), 1.54–1.67 (m, 8H, 4CH2(Ad)), 1.67–1.73 (m, 2H, CH2(Ad)), 2.13–2.24 (m, 2H, 2CH(Ad)), 4.09 (m, 4H, 2CH2O), 7.39 (s, 2H, 2CH); 13C NMR (CDCl3, 100.6 MHz): δ 27.7 (2CH(Ad)), 34.1 (2C(Ad)), 36.0 (CH2(Ad)), 38.5 (4CH2(Ad)), 40.7 (CH2(Ad)), 75.7 (2CH2O), 102.4 (2CH), 157.9 (2C=O), 158.3 (2C), 165.8 (2CNO2); HRMS-ESI (M + Na+): Found: 499.1065. Calculated for C20H20N4NaO10+: 499.1072.
(3-(((Cyanocarbonyl)oxy)methyl)adamantan-1-yl)methyl 5-nitroisoxazole-3-carboxylate (4) was isolated to a pure state as a colorless solid with a 28% (0.11 g) yield; mp. = 67–68 °C; Rf 0.47 (petroleum ether:EtOAc = 5:1); 1H NMR (CDCl3, 400 MHz): δ 1.41–1.45 (m, 2H, CH2), 1.47–1.72 (m, 10H, 5CH2), 2.15–2.22 (m, 2H, CH), 3.98 (s, 2H, CH2O), 4.09 (s, 2H, CH2O), 7.40 (s, 1H, CH); 13C NMR (CDCl3, 100.6 MHz): δ 27.6 (2CH), 33.9 (C), 34.0 (C), 35.9 (CH2), 38.2 (2CH2), 38.4 (2CH2), 40.5 (CH2), 75.6 (CH2O), 77.4 (CH2O), 102.5 (CH), 109.4 (C), 144.5 (C), 157.9 (C), 158.2 (C); HRMS-ESI (M + K+): Found: 428.0850. Calculated for C18H19N3KO7+: 428.0855.
((3s,7s)-7-{[(5-nitroisoxazol-3-yl)carbonyl]oxy}bicyclo[3.3.1]non-3-yl)methyl 5-nitroisoxazole-3-carboxylate (3h) was isolated to a pure state as a light yellow oil with a 65% (0.29 g) yield; Rf 0.21 (petroleum ether:EtOAc = 7:1); 1H NMR (CDCl3, 400 MHz): δ 1.19–1.27 (m, 1H, CH2), 1.53–1.63 (m, 2H, 2CH2), 1.82–1.91 (m, 2H, 2CH2), 1.91–1.97 (m, 1H, CH2), 1.97–2.13 (m, 4H, 4CH2), 2.15–2.29 (m, 3H, 3CH), 4.35 (d, 3J= 7.1 Hz, 2H, CH2O), 5.45–5.54 (m, 1H, CHO), 7.39 (s, H, CH), 7.44 (s, H, CH); 13C NMR (CDCl3, 100.6 MHz): δ 23.1 (2CH), 27.7 (CH2), 28.9 (2CH2), 29.6 (CH), 37.2 (2CH2), 72.4 (CH2O), 74.0 (CHO), 102.50 (CH), 102.54 (CH), 157.2 (C=O), 157.7 (C=O), 158.5 (C), 158.8 (C), 165.7 (2CNO2); HRMS-ESI (M + K+): Found: 489.0660. Calculated for C18H18KN4O10+: 489.0655.
1,4-Phenylene bis(4,4,4-trinitrobutanoate) (5a) was isolated to a pure state as a colorless solid with a 54% (0.28 g) yield; mp. = 150–151 °C; Rf 0.53 (petroleum ether:EtOAc = 5:1); 1H NMR (CD3C(O)CD3, 400 MHz): δ 3.24 (t, 3J = 7.5 Hz, 2H, CH2), 3.87 (br.t, 3J = 7.5 Hz, 2H, CH2), 7.22 (c, 4H, 4CH); 13C NMR (CD3C(O)CD3, 100.6 MHz): δ 29.1 (2CH2), 29.9 (2CH2), 123.4 (4CH(Ar)), 131.1 (2C(NO2)3), 149.2 (2C(Ar)), 169.5 (2C=O); HRMS-ESI (M + NH4+): Found: 538.0643. Calculated for C14H16N7O16+: 538.0648.

3.1.4. Synthesis of Compound 5b

Triethylamine (0.28 mL, 0.20 g, 2.0 mmol) was added dropwise to a solution of tetranitromethane (0.30 mL, 0.49 g, 2.5 mmol) in 1,4-dioxane (4.00 mL) at 0 °C (ice bath). The reaction mixture was stirred for 5 min, and alkene 2j (282 mg, 1.0 mmol) in 2.00 mL of 1,4-dioxane was added dropwise. Then, the cooling factor was removed, and the mixture was stirred at 70 °C for 2 h. The solvent was removed under reduced pressure, and the product was isolated by column chromatography.
4-(Adamantan-1-yl)phenyl 4,4,4-trinitrobutanoate (5b) was isolated to a pure state as a colorless solid with a 68% (0.59 g) yield; mp. = 125–127 °C; Rf 0.72 (petroleum ether:EtOAc = 10:1); 1H NMR (CDCl3, 400 MHz): δ 1.70–1.84 (m, 6H, 3CH2(Ad)), 1.86–1.95 (m, 6H, 3CH2(Ad)), 2.06–2.15 (m, 3H, 3CH(Ad)), 3.03 (t, 3J = 7.5 Hz, 2H, CH2), 3.50 (br. t, 3J = 7.5 Hz, 2H, CH2), 7.00–7.06 (m, 2H, 2CH(Ar)), 7.35–7.41 (m, 2H, 2CH(Ar)). 13C NMR (CDCl3, 100.6 MHz): δ 28.7 (CH2), 29.0 (3CH(Ad)), 29.8 (CH2), 36.1 (C(Ad)), 36.8 (3CH2(Ad)), 43.3 (3CH2(Ad)), 120.6 (2CH(Ar)), 126.2 (2CH(Ar)), 128.9 (C(NO2)3), 148.0 (C), 149.7 (C), 168.3 (C=O); HRMS-ESI (M + NH4+): Found: 451.1828. Calculated for C20H27N4O4+: 451.1823.

3.1.5. Synthesis of Compound 6

Na2S2O4 (1.05 g, 6.0 mmol) was added to a solution of 1,4-phenylenedi(methylene) bis(5-nitroisoxazole-3-carboxylate) (418 mg, 1.0 mmol) in a mixture of THF (8.00 mL) and H2O (8.00 mL). The resulting mixture was stirred at 90 °C for 1 h and then cooled down to room temperature. Then, H2O (8.00 mL) and HCl conc. (3.70 mL) were added. The resulting mixture was stirred at 60 °C for 15 min, and then cooled down to room temperature and quenched with NaHCO3. The mixture was poured into brine (20 mL) and extracted with EtOAc (4 × 20 mL). The combined organic layers were washed with brine (20 mL) and dried (anhyd. MgSO4). The solvent was evaporated in vacuo, and the title product was isolated by column chromatography.
1,4-Phenylenedi(methylene) bis(5-aminoisoxazole-3-carboxylate)(6) was isolated to a pure state as a colorless solid with a 32% (0.12 g) yield; mp. = 150–151 °C; Rf 0.24 (petroleum ether:EtOAc = 4:1); 1H NMR (CD3OD, 400 MHz): δ 5.35 (c, 4H, 2CH2O), 5.38 (c, 2H, 2CH), 7.47 (c, 4H, 4CH(Ar)). 13C NMR (CD3OD, 100.6 MHz): δ 67.8 (2CH2O), 79.2 (2CH), 129.7 (4CH(Ar)), 137.2 (2C(Ar)), 158.1 (2C=O), 161.7 (2C), 173.8 (2C); HRMS-ESI (M + Na+): Found: 381.0802. Calculated for C16H14N4NaO6+: 381.0806.

3.1.6. Synthesis of Compound 7

Na2S2O4 (1.05 g, 6.0 mmol) was added to a solution of butane-1,4-diyl bis(5-nitroisoxazole-3-carboxylate) (0.37 mg, 1.0 mmol) in a mixture of THF (8.00 mL) and H2O (8.00 mL). The resulting mixture was stirred at 90 oC for 1 h and then cooled down to room temperature. Then, the resulting mixture was concentrated in vacuo. The residue was suspended in Ac2O (5.00 mL, 5.40 g, 53.0 mmol), and cerium (III) chloride (493 mg, 2.0 mmol) was added. The reaction mixture was stirred for 24 h, and then poured into water and neutralized with a solution of NaHCO3 (up to pH 7–8). The product was extracted by CH2Cl2 (4 × 20 mL). The combined organic extract was dried over MgSO4. The solvent was removed under reduced pressure and the product was isolated by column chromatography.
Butane-1,4-diyl bis(5-acetamidoisoxazole-3-carboxylate) (7) was isolated to a pure state as a colorless solid with a 46% (0.18 g) yield; mp. = 225–226 oC with decomposition; Rf 0.34 (CH2Cl2:MeOH = 20:1). 1H NMR (DMSO, 400 MHz): δ 1.77–1.85 (m, 2H, 2CH2), 2.12 (s, 6H, 2CH3), 4.30–4.41 (m, 2H, 2CH2O), 6.47 (c, 2H, 2CH), 11.94 (c, 2H, 2NH). 13C NMR (CDCl3-CD3OD, 100.6 MHz): δ 23.0 (2CH3), 25.1 (2CH2), 65.4 (2CH2O), 89.0 (2CH), 157.0 (2C), 159.9 (2C=O), 162.9 (2C), 167.6 (2C=O); HRMS-ESI (M + H+): Found: 395.1192. Calculated for C16H19N4O8+: 395.1197.

3.2. Electrophysiological Evaluation

In vitro electrophysiological experiments were carried out using a patch clamp technique with the local fixation of potential, as described earlier [28,29]. Freshly isolated single Purkinje neurons from the cerebellum of 12–15-day old Wistar rats were used as a test system. Transmembrane currents were induced by the activation of the AMPA receptors with a solution of their partial agonist kainic acid, using a fast superfusion of solutions, where 30 μL of the agonist buffer (the agonist concentration was varied in the range of 10–6–10–4 M) were added to the constant flow of the neuron-washing buffer. The applications for the control and for each concentration of a compound were performed in triplicate. The transmembrane currents for the individual neurons were recorded using 2.5–5.5 MΩ borosilicate microelectrodes in a whole-cell configuration with an EPC-9 device from HEKA, Germany. The data were processed by the Pulsfit program from HEKA, Germany. Cyclothiazide (CTZ), as a well-known positive allosteric modulator of AMPA receptors, was used as a reference ligand. The experimental results for compound 6 are presented in Table 2.

3.3. Molecular Modeling

The structure of the dimeric ligand-binding domain of the GluA2 AMPA receptor was obtained from the Protein Data Bank (PDB: 4FAT) [49]. Upon the removal of the ions and small molecules (except for the two receptor-bound glutamate agonist molecules), the protein was allowed to relax during the molecular dynamics simulation for 100 ns (see below for the simulation protocol). The most frequently occurring structure was identified by the clustering of the frames in the stable part of the trajectory (40–100 ns). The ligand structure was converted to 3D and preoptimized in the MMFF94 force field using Avogadro 1.2.0 software [50], and then the ligand and protein structures were prepared for molecular docking using AutoDock Tools 1.5.6 [51]. The molecular docking to the positive allosteric modulator binding site was performed with AutoDock Vina 1.1.2 software [52] (grid box size 22 Å × 29 Å × 40 Å, exhaustiveness = 16). The pose with the best scoring function value and ligand position was selected, and the complex model was built using the USCF Chimera 1.15 software [53].
The molecular dynamics simulations were performed using the CHARMM36/CGenFF 4.4 force field [54,55] in the GROMACS 2021.2 software [56]. The initial models of the systems were built using the Ligand Reader & Modeler and Solution Builder modules of the CHARMM-GUI web service [57,58]. The protein molecule was inserted into a rectangular box of water in the TIP3P model; the distance from the protein to the box border was no less than 10 Å. Individual, randomly selected water molecules were replaced with potassium and chlorine ions to ensure the electrical neutrality of the system and the total concentration of KCl about 0.15 M. For each system, the molecular mechanics minimization (up to 5000 steps) was performed on the CPU, followed by equilibration for 125 ps at the temperature of 300 K and a constant volume using the v-rescale thermostat on the NVIDIA GeForce RTX 3080 GPU. The production simulation was performed on the GPU at the constant pressure of 1 atm and the temperature of 300 K using the v-rescale thermostat and the Parrinello–Rahman barostat. The hydrogen atom movements were constrained using the LINCS algorithm. For the analysis and visualization of the results, the CPPTRAJ software [59] in the AmberTools 21 package [60] and UCSF Chimera were used.

3.4. Prediction of ADMET, Physicochemical, and PAINS Profiles

The lipophilicity (LogPow) and aqueous solubility (pSaq) were estimated by the ALogPS 3.0 neural network model implemented in the OCHEM platform [61]. Human intestinal absorption (HIA) [62], blood–brain barrier permeability (LogBB) [63,64], and hERG-mediated cardiac toxicity risk (channel affinity pKi and inhibitory activity pIC50) [65] were estimated using the integrated online service for ADMET properties prediction (ADMET Prediction Service) [66]. This server implements predictive QSAR models based on accurate and representative training sets, fragmental descriptors, and artificial neural networks. The quantitative estimate of drug-likeness (QED) values [67] were calculated and the pan assay interference compounds (PAINS) alerts were checked using RDKit version 2020.03.4 software [68].

4. Conclusions

In conclusion, we have developed an efficient, simple, and concise approach to the previously unknown functionalized bis(isoxazoles), based on the heterocyclization of readily available electrophilic dienes upon treatment with a TNM-TEA complex. The reaction tolerates a variety of linkers in bis(heterocycles) including aliphatic, aromatic, and polycyclic moieties. The remarkable activity of bis(isoxazoles) as positive allosteric modulators of the AMPA receptor was demonstrated by the example of bis(5-aminoisoxazole) 6, which caused a potentiation of the kainate-induced AMPA receptor currents in a wide concentration range (10–12–10–6 M). The molecular modeling confirmed that this compound could interact with the validated PAM binding site. Its predicted ADMET, physicochemical, and PAINS properties were quite acceptable for potential lead compounds at the early drug development stages. These results allow us to expect further beneficial applications of this structural type of heterocycles in the search for new compounds with nootropic and neuroprotective properties.

Supplementary Materials

Copies of the NMR spectra and the X-ray data are available online.

Author Contributions

Conceptualization, D.S.K., V.A.P. and E.B.A.; software, D.S.K. and E.V.R.; investigation, D.A.V., K.S.S., E.V.R., Y.K.G., V.B.R., T.S.K., V.L.Z. and V.V.G.; writing—original draft preparation, D.A.V., K.S.S., and E.V.R.; writing—review and editing, E.V.R., K.N.S., V.A.P. and E.B.A.; visualization, K.N.S., V.B.R. and E.V.R.; supervision, T.S.K., V.A.P. and E.B.A.; funding acquisition, V.A.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation, grant no. 17-15-01455. The study of CTZ was supported by the State Assignment of IPAC RAS, topic no. 0090_2019_0005.

Institutional Review Board Statement

The study was conducted according to the guidelines of the Declaration of Helsinki, and approved by the Institutional Animal Review Board of IPAC RAS.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data of the compounds are available from the authors.

Acknowledgments

The study was fulfilled using NMR spectrometer Agilent 400-MR and diffractometer STADI VARI Pilatus-100K purchased by the MSU Development Program.

Conflicts of Interest

The authors declare no conflict 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 are available from the authors.

References

  1. Sysak, A.; Obmińska-Mrukowicz, B. Isoxazole ring as a useful scaffold in a search for new therapeutic agents. Eur. J. Med. Chem. 2017, 137, 292–309. [Google Scholar] [CrossRef] [PubMed]
  2. Fricke, J.; Varkalis, J.; Zwillich, S.; Adler, R.; Forester, E.; Recker, D.P.; Verburg, K.M. Valdecoxib is more efficacious than rofecoxib in relieving pain associated with oral surgery. Am. J. Ther. 2002, 9, 89–97. [Google Scholar] [CrossRef] [PubMed]
  3. Molina, C.; Modesto, C.; Martín-Begué, N.; Arnal, C. Leflunomide, a valid and safe drug for the treatment of chronic anterior uveitis associated with juvenile idiopathic arthritis. Clin. Rheumatol. 2013, 32, 1673–1675. [Google Scholar] [CrossRef]
  4. Davidson, J.R.; Giller, E.L.; Zisook, S.; Overall, J.E. An efficacy study of isocarboxazid and placebo in depression, and its relationship to depressive nosology. Arch. Gen. Psychiatry 1988, 45, 120–127. [Google Scholar] [CrossRef] [PubMed]
  5. Doi, Y.; Chambers, H.F. Penicillins and β-lactamase inhibitors. In Mandell, Douglas, and Bennett’s Principles and Practice of Infectious Diseases; Bennett, J.E., Dolin, R., Blaser, M.J., Eds.; W.B. Saunders: Philadelphia, PA, USA, 2015; pp. 263–277. ISBN 978-1-4557-4801-3. [Google Scholar] [CrossRef]
  6. Farrington, M. Antibacterial drugs. In Clinical Pharmacology; Bennett, P.N., Brown, M.J., Sharma, P., Eds.; Churchill Livingstone: Oxford, UK, 2012; pp. 173–190. ISBN 978-0-7020-4084-9. [Google Scholar] [CrossRef]
  7. Kumbhare, R.M.; Kosurkar, U.B.; Janaki Ramaiah, M.; Dadmal, T.L.; Pushpavalli, S.N.C.V.L.; Pal-Bhadra, M. Synthesis and biological evaluation of novel triazoles and isoxazoles linked 2-phenyl benzothiazole as potential anticancer agents. Bioorg. Med. Chem. Lett. 2012, 22, 5424–5427. [Google Scholar] [CrossRef]
  8. Chernysheva, N.B.; Maksimenko, A.S.; Andreyanov, F.A.; Kislyi, V.P.; Strelenko, Y.A.; Khrustalev, V.N.; Semenova, M.N.; Semenov, V.V. Regioselective synthesis of 3,4-diaryl-5-unsubstituted isoxazoles, analogues of natural cytostatic combretastatin A4. Eur. J. Med. Chem. 2018, 146, 511–518. [Google Scholar] [CrossRef]
  9. Averina, E.B.; Vasilenko, D.A.; Gracheva, Y.A.; Grishin, Y.K.; Radchenko, E.V.; Burmistrov, V.V.; Butov, G.M.; Neganova, M.E.; Serkova, T.P.; Redkozubova, O.M.; et al. Synthesis and biological evaluation of novel 5-hydroxylaminoisoxazole derivatives as lipoxygenase inhibitors and metabolism enhancing agents. Bioorg. Med. Chem. 2016, 24, 712–720. [Google Scholar] [CrossRef]
  10. Azzali, E.; Machado, D.; Kaushik, A.; Vacondio, F.; Flisi, S.; Cabassi, C.S.; Lamichhane, G.; Viveiros, M.; Costantino, G.; Pieroni, M. Substituted N-phenyl-5-(2-(phenylamino)thiazol-4-yl)isoxazole-3-carboxamides are valuable antitubercular candidates that evade innate efflux machinery. J. Med. Chem. 2017, 60, 7108–7122. [Google Scholar] [CrossRef]
  11. Zimecki, M.; Bąchor, U.; Mączyński, M. Isoxazole derivatives as regulators of immune functions. Molecules 2018, 23, 2724. [Google Scholar] [CrossRef] [Green Version]
  12. Vasilenko, D.A.; Averina, E.B.; Zefirov, N.A.; Wobith, B.; Grishin, Y.K.; Rybakov, V.B.; Zefirova, O.N.; Kuznetsova, T.S.; Kuznetsov, S.A.; Zefirov, N.S. Synthesis and antimitotic activity of novel 5-aminoisoxazoles bearing alkoxyaryl moieties. Mend. Comm. 2017, 27, 228–230. [Google Scholar] [CrossRef]
  13. Vasilenko, D.A.; Dueva, E.V.; Kozlovskaya, L.I.; Zefirov, N.A.; Grishin, Y.K.; Butov, G.M.; Palyulin, V.A.; Kuznetsova, T.S.; Karganova, G.G.; Zefirova, O.N.; et al. Tick-borne flavivirus reproduction inhibitors based on isoxazole core linked with adamantane. Bioorg. Chem. 2019, 87, 629–637. [Google Scholar] [CrossRef]
  14. Traynelis, S.F.; Wollmuth, L.P.; McBain, C.J.; Menniti, F.S.; Vance, K.M.; Ogden, K.K.; Hansen, K.B.; Yuan, H.; Myers, S.J.; Dingledine, R. Glutamate receptor ion channels: Structure, regulation, and function. Pharmacol. Rev. 2010, 62, 405–496. [Google Scholar] [CrossRef] [Green Version]
  15. Lee, K.; Goodman, L.; Fourie, C.; Schenk, S.; Leitch, B.; Montgomery, J.M. AMPA receptors as therapeutic targets for neurological disorders. In Ion Channels as Therapeutic Targets, Part A; Donev, R., Ed.; Advances in Protein Chemistry and Structural Biology; Academic Press: Cambridge, MA, USA, 2016; Volume 103, pp. 203–261. [Google Scholar] [CrossRef]
  16. Mendez-David, I.; Guilloux, J.-P.; Papp, M.; Tritschler, L.; Mocaer, E.; Gardier, A.M.; Bretin, S.; David, D.J. S47445 produces antidepressant- and anxiolytic-like effects through neurogenesis dependent and independent mechanisms. Front. Pharmacol. 2017, 8, 462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Mozafari, N.; Shamsizadeh, A.; Fatemi, I.; Allahtavakoli, M.; Moghadam-Ahmadi, A.; Kaviani, E.; Kaeidi, A. CX691, as an AMPA receptor positive modulator, improves the learning and memory in a rat model of Alzheimer’s disease. Iran J. Basic Med. Sci. 2018, 21, 724–730. [Google Scholar] [CrossRef]
  18. Gordillo-Salas, M.; Pascual-Antón, R.; Ren, J.; Greer, J.; Adell, A. Antidepressant-like effects of CX717, a positive allosteric modulator of AMPA receptors. Mol. Neurobiol. 2020, 57, 3498–3507. [Google Scholar] [CrossRef] [PubMed]
  19. Lauterborn, J.C.; Palmer, L.C.; Jia, Y.; Pham, D.T.; Hou, B.; Wang, W.; Trieu, B.H.; Cox, C.D.; Kantorovich, S.; Gall, C.M.; et al. Chronic ampakine treatments stimulate dendritic growth and promote learning in middle-aged rats. J. Neurosci. 2016, 36, 1636–1646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Brogi, S.; Campiani, G.; Brindisi, M.; Butini, S. Allosteric modulation of ionotropic glutamate receptors: An outlook on new therapeutic approaches to treat central nervous system disorders. ACS Med. Chem. Lett. 2019, 10, 228–236. [Google Scholar] [CrossRef]
  21. Partin, K.M. AMPA receptor potentiators: From drug design to cognitive enhancement. Curr. Opin. Pharmacol. 2015, 20, 46–53. [Google Scholar] [CrossRef] [Green Version]
  22. Pirotte, B.; Francotte, P.; Goffin, E.; de Tullio, P. AMPA receptor positive allosteric modulators: A patent review. Expert Opin. Ther. Pat. 2013, 23, 615–628. [Google Scholar] [CrossRef]
  23. Ward, S.E.; Pennicott, L.E.; Beswick, P. AMPA receptor-positive allosteric modulators for the treatment of schizophrenia: An overview of recent patent applications. Future Med. Chem. 2015, 7, 473–491. [Google Scholar] [CrossRef]
  24. Reuillon, T.; Ward, S.E.; Beswick, P. AMPA receptor positive allosteric modulators: Potential for the treatment of neuropsychiatric and neurological disorders. Curr. Top. Med. Chem. 2016, 16, 3536–3565. [Google Scholar] [CrossRef]
  25. Lavrov, M.I.; Grigor’ev, V.V.; Bachurin, S.O.; Palyulin, V.A.; Zefirov, N.S. Novel bivalent positive allosteric modulators of AMPA receptor. Dokl. Biochem. Biophys. 2015, 464, 322–324. [Google Scholar] [CrossRef]
  26. Lavrov, M.I.; Veremeeva, P.N.; Karlov, D.S.; Zamoyski, V.L.; Grigoriev, V.V.; Palyulin, V.A. Tricyclic derivatives of bispidine as AMPA receptor allosteric modulators. Mend. Comm. 2019, 29, 619–621. [Google Scholar] [CrossRef]
  27. Lavrov, M.I.; Karlov, D.S.; Palyulin, V.A.; Grigoriev, V.V.; Zamoyski, V.L.; Brkich, G.E.; Pyatigorskaya, N.V.; Zapolskiy, M.E. Novel positive allosteric modulator of AMPA-receptors based on tricyclic scaffold. Mend. Comm. 2018, 28, 311–313. [Google Scholar] [CrossRef]
  28. Lavrov, M.I.; Karlov, D.S.; Voronina, T.A.; Grigoriev, V.V.; Ustyugov, A.A.; Bachurin, S.O.; Palyulin, V.A. Novel positive allosteric modulators of AMPA receptors based on 3,7-diazabicyclo[3.3.1]nonane scaffold. Mol. Neurobiol. 2020, 57, 191–199. [Google Scholar] [CrossRef]
  29. Temnyakova, N.S.; Vasilenko, D.A.; Lavrov, M.I.; Karlov, D.S.; Grishin, Y.K.; Zamoyski, V.L.; Grigoriev, V.V.; Averina, E.B.; Palyulin, V.A. Novel bivalent positive allosteric AMPA receptor modulator of bis-amide series. Mend. Comm. 2021, 31, 216–218. [Google Scholar] [CrossRef]
  30. Drapier, T.; Geubelle, P.; Bouckaert, C.; Nielsen, L.; Laulumaa, S.; Goffin, E.; Dilly, S.; Francotte, P.; Hanson, J.; Pochet, L.; et al. Enhancing action of positive allosteric modulators through the design of dimeric compounds. J. Med. Chem. 2018, 61, 5279–5291. [Google Scholar] [CrossRef] [PubMed]
  31. Goffin, E.; Drapier, T.; Larsen, A.P.; Geubelle, P.; Ptak, C.P.; Laulumaa, S.; Rovinskaja, K.; Gilissen, J.; de Tullio, P.; Olsen, L.; et al. 7-Phenoxy-substituted 3,4-dihydro-2H-1,2,4-benzothiadiazine 1,1-dioxides as positive allosteric modulators of α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors with nanomolar potency. J. Med. Chem. 2018, 61, 251–264. [Google Scholar] [CrossRef] [PubMed]
  32. Nazarova, A.A.; Sedenkova, K.N.; Karlov, D.S.; Lavrov, M.I.; Grishin, Y.K.; Kuznetsova, T.S.; Zamoyski, V.L.; Grigoriev, V.V.; Averina, E.B.; Palyulin, V.A. Bivalent AMPA receptor positive allosteric modulators of the bis(pyrimidine) series. Med. Chem. Commun. 2019, 10, 1615–1619. [Google Scholar] [CrossRef] [PubMed]
  33. Karlov, D.S.; Lavrov, M.I.; Palyulin, V.A.; Zefirov, N.S. Pharmacophore analysis of positive allosteric modulators of AMPA receptors. Russ. Chem. Bull. 2016, 65, 581–587. [Google Scholar] [CrossRef]
  34. Radchenko, E.V.; Karlov, D.S.; Lavrov, M.I.; Palyulin, V.A. Structural requirements for molecular design of positive allosteric modulators of AMPA receptor. Mend. Comm. 2017, 27, 623–625. [Google Scholar] [CrossRef]
  35. Karlov, D.S.; Lavrov, M.I.; Palyulin, V.A.; Zefirov, N.S. MM-GBSA and MM-PBSA performance in activity evaluation of AMPA receptor positive allosteric modulators. J. Biomol. Struct. Dyn. 2018, 36, 2508–2516. [Google Scholar] [CrossRef] [PubMed]
  36. Volkova, Y.A.; Averina, E.B.; Vasilenko, D.A.; Sedenkova, K.N.; Grishin, Y.K.; Bruheim, P.; Kuznetsova, T.S.; Zefirov, N.S. Unexpected heterocyclization of electrophilic alkenes by tetranitromethane in the presence of triethylamine. Synthesis of 5-nitroisoxazoles. J. Org. Chem. 2019, 84, 3192–3200. [Google Scholar] [CrossRef] [PubMed]
  37. Vasilenko, D.A.; Sedenkova, K.N.; Kuznetsova, T.S.; Averina, E.B. Synthetic approaches to nitro-substituted isoxazoles. Synthesis 2019, 51, 1516–1528. [Google Scholar] [CrossRef]
  38. Averina, E.B.; Samoilichenko, Y.V.; Volkova, Y.A.; Grishin, Y.K.; Rybakov, V.B.; Kutateladze, A.G.; Elyashberg, M.E.; Kuznetsova, T.S.; Zefirov, N.S. Heterocyclization of electrophilic alkenes with tetranitromethane revisited: Regiochemistry and the mechanism of nitroisoxazole formation. Tetrahedron Lett. 2012, 53, 1472–1475. [Google Scholar] [CrossRef]
  39. Vasilenko, D.A.; Dronov, S.E.; Parfiryeu, D.U.; Sadovnikov, K.S.; Sedenkova, K.N.; Grishin, Y.K.; Rybakov, V.B.; Kuznetsova, T.S.; Averina, E.B. 5-Nitroisoxazoles in SNAr reactions: Access to polysubstituted isoxazole derivatives. Org. Biomol. Chem. 2021, 19, 6447–6454. [Google Scholar] [CrossRef] [PubMed]
  40. Sun, M.; Hong, C.-Y.; Pan, C.-Y. A unique aliphatic tertiary amine chromophore: Fluorescence, polymer structure, and application in cell imaging. J. Am. Chem. Soc. 2012, 134, 20581–20584. [Google Scholar] [CrossRef]
  41. Nielsen, A.T. Nitrocarbons, 1st ed.; Wiley-VCH: New York, NY, USA, 1996; ISBN 978-0-471-18603-8. [Google Scholar]
  42. Averina, E.B.; Vasilenko, D.A.; Samoilichenko, Y.V.; Grishin, Y.K.; Rybakov, V.B.; Kuznetsova, T.S.; Zefirov, N.S. Chemoselective reduction of functionalized 5-nitroisoxazoles: Synthesis of 5-amino- and 5-[hydroxy(tetrahydrofuran-2-yl)amino]isoxazoles. Synthesis 2014, 46, 1107–1113. [Google Scholar] [CrossRef] [Green Version]
  43. Sadovnikov, K.S.; Vasilenko, D.A.; Sedenkova, K.N.; Rybakov, V.B.; Grishin, Y.K.; Alferova, V.A.; Kuznetsova, T.S.; Averina, E.B. Straightforward chemoselective 4-nitration of 5-aminoisoxazoles. Mend. Comm. 2020, 30, 487–489. [Google Scholar] [CrossRef]
  44. Radchenko, E.V.; Tarakanova, A.S.; Karlov, D.S.; Lavrov, M.I.; Palyulin, V.A. Ligands of the AMPA-subtype glutamate receptors: Mechanisms of action and novel chemotypes. Biomed. Khim. 2021, 67, 187–200. [Google Scholar] [CrossRef]
  45. Borioni, J.L.; Cavallaro, V.; Murray, A.P.; Peñéñory, A.B.; Puiatti, M.; García, M.E. Design, synthesis and evaluation of cholinesterase hybrid inhibitors using a natural steroidal alkaloid as precursor. Bioorg. Chem. 2021, 111, 104893. [Google Scholar] [CrossRef]
  46. Naveen Kumar, D.; Sudhakar, N.; Rao, B.V.; Kishore, K.H.; Murty, U.S. Synthesis of trans-1,8,12,13-tetraoxadispiro[4.1.4.2]tridecanes—A new class of peroxides. Tetrahedron Lett. 2006, 47, 771–774. [Google Scholar] [CrossRef]
  47. Yu, Z.; Eno, M.S.; Annis, A.H.; Morken, J.P. Enantioselective hydroformylation of 1-alkenes with commercial Ph-BPE ligand. Org. Lett. 2015, 17, 3264–3267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Averina, E.B.; Sedenkova, K.N.; Bakhtin, S.G.; Grishin, Y.K.; Kutateladze, A.G.; Roznyatovsky, V.A.; Rybakov, V.B.; Butov, G.M.; Kuznetsova, T.S.; Zefirov, N.S. symm-Tetramethylenecyclooctane: En route to polyspirocycles. J. Org. Chem. 2014, 79, 8163–8170. [Google Scholar] [CrossRef] [PubMed]
  49. Harms, J.E.; Benveniste, M.; Maclean, J.K.F.; Partin, K.M.; Jamieson, C. Functional analysis of a novel positive allosteric modulator of AMPA receptors derived from a structure-based drug design strategy. Neuropharmacology 2013, 64, 45–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Hanwell, M.D.; Curtis, D.E.; Lonie, D.C.; Vandermeersch, T.; Zurek, E.; Hutchison, G.R. Avogadro: An advanced semantic chemical editor, visualization, and analysis platform. J. Cheminform. 2012, 4, 17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Morris, G.M.; Huey, R.; Lindstrom, W.; Sanner, M.F.; Belew, R.K.; Goodsell, D.S.; Olson, A.J. AutoDock4 and AutoDockTools4: Automated docking with selective receptor flexibility. J. Comput. Chem. 2009, 30, 2785–2791. [Google Scholar] [CrossRef] [Green Version]
  52. Trott, O.; Olson, A.J. AutoDock Vina: Improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J. Comput. Chem. 2010, 31, 455–461. [Google Scholar] [CrossRef] [Green Version]
  53. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Couch, G.S.; Greenblatt, D.M.; Meng, E.C.; Ferrin, T.E. UCSF Chimera—A visualization system for exploratory research and analysis. J. Comput. Chem. 2004, 25, 1605–1612. [Google Scholar] [CrossRef] [Green Version]
  54. Huang, J.; MacKerell, A.D. CHARMM36 all-atom additive protein force field: Validation based on comparison to NMR data. J. Comput. Chem. 2013, 34, 2135–2145. [Google Scholar] [CrossRef] [Green Version]
  55. Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; et al. CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields. J. Comput. Chem. 2010, 31, 671–690. [Google Scholar] [CrossRef] [Green Version]
  56. Abraham, M.J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J.C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1–2, 19–25. [Google Scholar] [CrossRef] [Green Version]
  57. Jo, S.; Kim, T.; Iyer, V.G.; Im, W. CHARMM-GUI: A web-based graphical user interface for CHARMM. J. Comput. Chem. 2008, 29, 1859–1865. [Google Scholar] [CrossRef]
  58. Lee, J.; Cheng, X.; Swails, J.M.; Yeom, M.S.; Eastman, P.K.; Lemkul, J.A.; Wei, S.; Buckner, J.; Jeong, J.C.; Qi, Y.; et al. CHARMM-GUI input generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM simulations using the CHARMM36 additive force field. J. Chem. Theory Comput. 2016, 12, 405–413. [Google Scholar] [CrossRef]
  59. Roe, D.R.; Cheatham, T.E. PTRAJ and CPPTRAJ: Software for processing and analysis of molecular dynamics trajectory data. J. Chem. Theory Comput. 2013, 9, 3084–3095. [Google Scholar] [CrossRef]
  60. Salomon-Ferrer, R.; Case, D.A.; Walker, R.C. An overview of the Amber biomolecular simulation package. WIREs Comput. Mol. Sci. 2013, 3, 198–210. [Google Scholar] [CrossRef]
  61. Sushko, I.; Novotarskyi, S.; Körner, R.; Pandey, A.K.; Rupp, M.; Teetz, W.; Brandmaier, S.; Abdelaziz, A.; Prokopenko, V.V.; Tanchuk, V.Y.; et al. Online chemical modeling environment (OCHEM): Web platform for data storage, model development and publishing of chemical information. J. Comput.-Aided Mol. Des. 2011, 25, 533–554. [Google Scholar] [CrossRef] [Green Version]
  62. Radchenko, E.V.; Dyabina, A.S.; Palyulin, V.A.; Zefirov, N.S. Prediction of human intestinal absorption of drug compounds. Russ. Chem. Bull. 2016, 65, 576–580. [Google Scholar] [CrossRef]
  63. Dyabina, A.S.; Radchenko, E.V.; Palyulin, V.A.; Zefirov, N.S. Prediction of blood-brain barrier permeability of organic compounds. Dokl. Biochem. Biophys. 2016, 470, 371–374. [Google Scholar] [CrossRef] [PubMed]
  64. Radchenko, E.V.; Dyabina, A.S.; Palyulin, V.A. Towards deep neural network models for the prediction of the blood-brain barrier permeability for diverse organic compounds. Molecules 2020, 25, 5901. [Google Scholar] [CrossRef]
  65. Radchenko, E.V.; Rulev, Y.A.; Safanyaev, A.Y.; Palyulin, V.A.; Zefirov, N.S. Computer-aided estimation of the hERG-mediated cardiotoxicity risk of potential drug components. Dokl. Biochem. Biophys. 2017, 473, 128–131. [Google Scholar] [CrossRef] [PubMed]
  66. ADMET Prediction Service. Available online: http://qsar.chem.msu.ru/admet/ (accessed on 10 October 2021).
  67. Bickerton, G.R.; Paolini, G.V.; Besnard, J.; Muresan, S.; Hopkins, A.L. Quantifying the chemical beauty of drugs. Nat. Chem. 2012, 4, 90–98. [Google Scholar] [CrossRef] [Green Version]
  68. RDKit: Open-Source Cheminformatics Software. Available online: https://www.rdkit.org/ (accessed on 10 October 2021).
Scheme 1. Heterocyclization of the adamantane-containing diene 2g.
Scheme 1. Heterocyclization of the adamantane-containing diene 2g.
Molecules 26 06411 sch001
Scheme 2. Reaction of the aryl acrylic esters 2i and 2j with TNM-TEA.
Scheme 2. Reaction of the aryl acrylic esters 2i and 2j with TNM-TEA.
Molecules 26 06411 sch002
Figure 1. The molecular structure of compound 3c. Displacement ellipsoids are drawn at the 50% probability level; H atoms are presented as spheres with an arbitrary radius.
Figure 1. The molecular structure of compound 3c. Displacement ellipsoids are drawn at the 50% probability level; H atoms are presented as spheres with an arbitrary radius.
Molecules 26 06411 g001
Scheme 3. Synthesis of bis(5-aminoisoxazole) 6.
Scheme 3. Synthesis of bis(5-aminoisoxazole) 6.
Molecules 26 06411 sch003
Scheme 4. Reduction of bis(5-nitroisoxazole) 3c.
Scheme 4. Reduction of bis(5-nitroisoxazole) 3c.
Molecules 26 06411 sch004
Figure 2. The examples of the known positive modulators of the AMPA receptor.
Figure 2. The examples of the known positive modulators of the AMPA receptor.
Molecules 26 06411 g002
Figure 3. Binding mode of the PAM 6 refined using molecular dynamics simulation. (A) General view of the dimeric ligand-binding domain of AMPA receptor (GluA2) and location of the binding site. (B) Binding pockets in the protein molecular surface colored by local hydrophobicity (brown for hydrophobic and blue for hydrophilic). (C) Detailed view of the binding site. The ligand is represented by a grey ball-and-stick model, the amino acid residues located within 3 Å of it are represented by beige stick models. Hydrogen bonds are shown as cyan lines.
Figure 3. Binding mode of the PAM 6 refined using molecular dynamics simulation. (A) General view of the dimeric ligand-binding domain of AMPA receptor (GluA2) and location of the binding site. (B) Binding pockets in the protein molecular surface colored by local hydrophobicity (brown for hydrophobic and blue for hydrophilic). (C) Detailed view of the binding site. The ligand is represented by a grey ball-and-stick model, the amino acid residues located within 3 Å of it are represented by beige stick models. Hydrogen bonds are shown as cyan lines.
Molecules 26 06411 g003
Figure 4. RMSD of the protein, glutamate, and ligand (6) heavy atoms during molecular dynamics simulation of the PAM complex with the dimeric ligand-binding domain of the GluA2 AMPA receptor.
Figure 4. RMSD of the protein, glutamate, and ligand (6) heavy atoms during molecular dynamics simulation of the PAM complex with the dimeric ligand-binding domain of the GluA2 AMPA receptor.
Molecules 26 06411 g004
Table 1. Synthesis of electrophilic dienes 2 and bis(5-nitroisoxazoles) 3.
Table 1. Synthesis of electrophilic dienes 2 and bis(5-nitroisoxazoles) 3.
Molecules 26 06411 i001
EntryCompoundLinkerSolvent 1Yield, % 2
23
1a–(CH2)2MeCN7346
2b–(CH2)3MeCN7655
3c–(CH2)4MeCN8164
4d Molecules 26 06411 i002MeCN8532
5e Molecules 26 06411 i003MeCN8440
6f Molecules 26 06411 i004MeCN8222
7g Molecules 26 06411 i005THF4651
8h Molecules 26 06411 i006THF5465
9i Molecules 26 06411 i007CH2Cl299
1 For the acylation of diols 1a-i; 2 Isolated yields.
Table 2. The effect of various concentrations of compounds on the kainate-induced AMPA receptor currents in rat cerebellum Purkinje cells.
Table 2. The effect of various concentrations of compounds on the kainate-induced AMPA receptor currents in rat cerebellum Purkinje cells.
CompoundNumber of NeuronsCompound Concentration/M, Current Amplitude (% to Control, ±SD)
10–1210–1110–10109108107106
64141 ± 7172 ± 9152 ± 7144 ± 5129 ± 4113 ± 4105 ± 3
8 [29]5 106 ± 6118 ± 7140 ± 8128 ± 5 119 ± 4 119 ± 3
9 [32]7108 ± 5132 ± 5143 ± 9170 ± 11123 ± 8 85 ± 6 78 ± 4
CTZ8 100 ± 3145 ± 11
Table 3. Predicted physicochemical and ADMET profiles of compound 6.
Table 3. Predicted physicochemical and ADMET profiles of compound 6.
CompoundMWLogPowpSaqLogBBHIAhERG pKihERG pIC50QED
6358.311.923.22−0.98995.984.130.62
Note: MW—molecular weight, LogPow—octanol-water partition coefficient, pSaq—aqueous solubility [–log(M)], LogBB—blood–brain barrier permeability, HIA—human intestinal absorption [%], hERG pKi—hERG potassium channel affinity [–log(M)], hERG pIC50—hERG potassium channel inhibitory activity [–log(M)], QED—quantitative estimate of drug-likeness.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Vasilenko, D.A.; Sadovnikov, K.S.; Sedenkova, K.N.; Karlov, D.S.; Radchenko, E.V.; Grishin, Y.K.; Rybakov, V.B.; Kuznetsova, T.S.; Zamoyski, V.L.; Grigoriev, V.V.; et al. A Facile Approach to Bis(isoxazoles), Promising Ligands of the AMPA Receptor. Molecules 2021, 26, 6411. https://doi.org/10.3390/molecules26216411

AMA Style

Vasilenko DA, Sadovnikov KS, Sedenkova KN, Karlov DS, Radchenko EV, Grishin YK, Rybakov VB, Kuznetsova TS, Zamoyski VL, Grigoriev VV, et al. A Facile Approach to Bis(isoxazoles), Promising Ligands of the AMPA Receptor. Molecules. 2021; 26(21):6411. https://doi.org/10.3390/molecules26216411

Chicago/Turabian Style

Vasilenko, Dmitry A., Kirill S. Sadovnikov, Kseniya N. Sedenkova, Dmitry S. Karlov, Eugene V. Radchenko, Yuri K. Grishin, Victor B. Rybakov, Tamara S. Kuznetsova, Vladimir L. Zamoyski, Vladimir V. Grigoriev, and et al. 2021. "A Facile Approach to Bis(isoxazoles), Promising Ligands of the AMPA Receptor" Molecules 26, no. 21: 6411. https://doi.org/10.3390/molecules26216411

Article Metrics

Back to TopTop