Next Article in Journal
Do Curdlan Hydrogels Improved with Bioactive Compounds from Hop Exhibit Beneficial Properties for Skin Wound Healing?
Previous Article in Journal
Extracellular Vesicles and Cx43-Gap Junction Channels Are the Main Routes for Mitochondrial Transfer from Ultra-Purified Mesenchymal Stem Cells, RECs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Allosteric Modulators of AMPA Receptors: Synthesis and Study of Their Functional Activity by Radioligand-Receptor Binding Analysis

by
Elena A. Golubeva
1,
Mstislav I. Lavrov
1,
Polina N. Veremeeva
1,
Tatiana V. Vyunova
2,
Konstantin V. Shevchenko
2,
Maxim A. Topchiy
3,
Andrey F. Asachenko
3 and
Vladimir A. Palyulin
1,*
1
Department of Chemistry, Lomonosov Moscow State University, 119991 Moscow, Russia
2
Laboratory of Molecular Pharmacology of Peptides, Institute of Molecular Genetics, National Research Centre Kurchatov Institute, 123182 Moscow, Russia
3
A.V. Topchiev Institute of Petrochemical Synthesis, Russian Academy of Sciences, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(12), 10293; https://doi.org/10.3390/ijms241210293
Submission received: 29 April 2023 / Revised: 9 June 2023 / Accepted: 15 June 2023 / Published: 18 June 2023
(This article belongs to the Section Molecular Neurobiology)

Abstract

:
The synthetic approaches to three new AMPA receptor modulators—derivatives of 1,11-dimethyl-3,6,9-triazatricyclo[7.3.1.13,11]tetradecane-4,8,12-trione—had been developed and all steps of synthesis were optimized. The structures of the compounds contain tricyclic cage and indane fragments necessary for binding with the target receptor. Their physiological activity was studied by radioligand-receptor binding analysis using [3H]PAM-43 as a reference ligand, which is a highly potent positive allosteric modulator of AMPA receptors. The results of radioligand-binding studies indicated the high potency of two synthesized compounds to bind with the same targets as positive allosteric modulator PAM-43 (at least on AMPA receptors). We suggest that the Glu-dependent specific binding site of [3H]PAM-43 or the receptor containing this site may be one of the targets of the new compounds. We also suggest that enhanced radioligand binding may indicate the existence of synergistic effects of compounds 11b and 11c with respect to PAM-43 binding to the targets. At the same time, these compounds may not compete directly with PAM-43 for its specific binding sites but bind to other specific sites of this biotarget, changing its conformation and thereby causing a synergistic effect of cooperative interaction. It can be expected that the newly synthesized compounds will also have pronounced effects on the glutamatergic system of the mammalian brain.

1. Introduction

Synaptic transmission is the main pathway of neural communication, which forms the basis for the analysis, systematization, and memorization of information in the central nervous system (CNS) [1,2,3]. Rapid excitatory synaptic transmission in the CNS is mainly mediated via the AMPA subtype of glutamate receptors [4,5], and that makes these receptors key targets for the regulation of synaptic excitation.
When AMPA receptors are activated by glutamate, they rapidly desensitize [6,7,8,9]. To date, a number of allosteric modulators of these receptors are known, which by themselves do not activate AMPA receptors but increase the current induced by the agonist, thus slowing down the rate of desensitization [10,11,12,13]. This fact largely determines a wide range of possible therapeutic applications of such modulators. One of the significant neurophysiological actions of allosteric modulators of AMPA receptors is synaptic plasticity [14,15,16], playing a key role in the mechanism of neural memory. The important therapeutic potential of AMPA receptor allosteric modulators is also based on their ability to significantly increase the expression of neurotrophic factors, namely, nerve growth factor (NGF) and brain-derived neurotrophic factor (BDNF), which in turn lead to the regeneration of nervous tissue and protection of neurons from neurotoxic damage [17,18,19,20,21].
Thus, the above facts make AMPA receptor allosteric modulators promising candidates for the development of drugs for the treatment of cognitive disorders (including early stages of Alzheimer’s disease), depression, and a number of other CNS pathologies [4,5], attention-deficit/hyperactivity disorder (ADHD) and mood disorders [22,23], as well as drug-induced respiratory depression [24]. The key advantage of allosteric modulators of AMPA receptors is that (unlike agonists or antagonists) they potentiate the receptor only in the presence of an endogenous ligand and, therefore, cause fewer side effects [25,26].
However, despite the encouraging preclinical results, most AMPA receptor modulators to date have not progressed beyond early clinical development [27]. These problems are associated with insufficiently relevant in vitro and in vivo models of neurological and neuropsychiatric disorders, issues of pharmacodynamics and pharmacokinetics, as well as with a limited understanding of the processes underlying the activity and specificity of the action of modulators: despite significant progress in the development of cell models [28] and animal models [29,30,31,32] of neurodegenerative diseases, none of them is yet able to fully capture all the factors associated with these complex pathologies. Nevertheless, the development of new chemotypes of AMPA receptor modulators has become a priority for many pharmaceutical companies [27].
Earlier extensive research in the field of molecular modeling and molecular dynamics studies of complexes of various ligands with AMPA receptors was carried out, and it was experimentally shown that some bispidine derivatives are highly active with respect to the glutamatergic system [33,34]. As a result of these studies, new potential allosteric modulators based on a bispidine moiety-containing tricyclic scaffold were developed. In particular, the derivatives of 1,11-dimethyl-3,6,9-triazatricyclo[7.3.1.13,11]tetradecane-4,8,12-trione in electrophysiological studies using the patch clamp technique, showed pronounced modulatory effects (both positive and negative) in subnanomolar concentrations with respect to AMPA receptors [35,36,37,38,39]. An important advantage of both positive and negative allosteric modulators of AMPA receptors is that they suppress the process of desensitization—the inability of the receptor to re-transit into an activated state, despite the presence of an agonist.
Nowadays, the radioligand-receptor binding analysis using tritium-labeled molecules is widely and successfully applied to study the molecular mechanisms of action of biologically active compounds and is one of the best, fast, and most reliable techniques [40]. This approach makes it possible to identify molecular targets—specific receptors through which the substance affects cell metabolism—and to characterize the interactions of ligands with detected receptors qualitatively and quantitatively. It also allows one to conduct qualitative and quantitative studies of the spectrum of interactions of the developed new drugs with membrane receptors of various types [41], to compile a receptor profile of a biologically active compound [42], to study the relationships between the structure of ligands and their action [43], to conduct a comparative assessment of compounds by their interaction with the biological targets [44,45], and to analyze pharmacokinetics [46].
In this article, we have developed and optimized the synthetic approaches to several derivatives of 1,11-dimethyl-3,6,9-triazatricyclo[7.3.1.13,11]tetradecane-4,8,12-trione as new potential AMPA receptor modulators and studied their functional activity by radioligand-receptor binding analysis using tritiated compound PAM-43 (an active, positive AMPA receptor modulator) as a reference ligand.

2. Results and Discussion

2.1. Synthesis of Target Compounds 11ac 

The key tricyclic compounds 11ac were synthesized from indan-2-one 1 according to Scheme 1. Oxime 2, obtained from indan-2-one 1, was reduced to amine 3 hydrochloride. Then, bromo-derivative 4 was synthesized by electrophilic aromatic bromination. It underwent Pd-catalyzed cyanation after preliminary protection of the NH2-group. Further, deprotection gave the key intermediate 7 for obtaining a series of amides 8ac. The reduction led to amines 9ac, which were alkylated with dichloro derivative 10 to provide corresponding tricyclic compounds 11ac.
For the obtaining of indan-2-one oxime 2, a well-known method [47,48] was modified. The synthesis was carried out by mixing a solution of indan-2-one 1 and hydroxylamine hydrochloride in the EtOH/H2O system (1:1). The proposed changes make it possible to avoid the addition of pyridine and thus facilitate the isolation of the product. For synthesis of indan-2-ylamine hydrochloride 3 a widely used method based on heterogeneous hydrogenation [47,48] was modified. It was shown that the addition of PdCl2 was not required for the reaction. Conducting the reaction at a hydrogen pressure of 5 atm (in an autoclave) increases the yield, reduces the reaction time, and prevents the formation of byproducts. That leads to a clean product without the need for additional purification, thus making it possible to obtain indan-2-ylamine hydrochloride 3 with the yield of 91%.
Selective bromination of indan-2-ylamine hydrochloride 3 presents certain difficulties since the rates of monobromination reaction and dibromination reaction in an aqueous solution are similar. As a result of the procedure optimization, it was found that the key factor was the amount of water. The optimal H2O/indan-2-ylamine hydrochloride ratio of 4.5 mL/g allows the reaction to be carried out selectively in good yield due to the fact that the monobromo derivative formed in the course of the reaction precipitates and does not undergo the second bromination.
The cyano-derivative 6 was obtained from the Boc-protected compound 5 by a cross-coupling reaction. First, the reaction with CuCN was conducted under various conditions, but the yields were low. Then, palladium catalysts were tested. As a result of the procedure optimization, it was found that the key factor was the addition of the reducing agent PMHS (polymethylhydrosiloxane), which significantly increased the yield. The best results were achieved using dppf (1,1′-bis(diphenylphosphino)ferrocene) as a ligand, Zn(CN)2 as a source of CN and PMHS. Further, deprotection provided the key intermediate 7 for obtaining a series of amides 8ac. Despite two additional stages of protection and deprotection, this way is preferable in comparison with acylationcyanation, since it allows branching of the synthetic scheme at a later stage and does not require the cyanation technique optimization for each amide.
Compounds 8a,c were synthesized by acylation of amine 7 with commercially available acids in the presence of CDI. It should be noted that an alternative method was used for compound 8b using the corresponding acyl chloride, since the reaction with the acid and CDI did not proceed.
Amines 9ac were obtained by the reduction of corresponding nitriles 8ac. The key problem of the CN group reduction is to ensure selectivity in the presence of an amide group. That makes it impossible to use the classical methods of reduction by complex hydrides.
Heterogeneous hydrogenation of the corresponding nitriles was carried out in an autoclave using 10% Pd/C as a catalyst. By varying the reaction conditions (time, pressure, and concentration of HCl), it was found that the key factor for the reaction is the addition of 1.5 eq. concentrated HCl. The developed method has been successfully applied to the synthesis of amines 9a,b in the form of hydrochlorides with high yields. It should be noted that in the case of compound 8b, such conditions result in the reduction in both the CN group and benzene ring [49]; therefore, nitrile 8b was used for the synthesis of amine 9b with the alicyclic ring. Amine 9c could not be obtained by heterogeneous hydrogenation. Therefore, the method of nitrile reduction using NaBH4 and CoCl2·6H2O was used [50,51,52]. This method has been modified for amine 9c. The main factor is the reduction time. With a short reaction time, the reduction did not proceed completely, and the product was isolated as a mixture with the initial nitrile 8c. If the reaction is carried out for too long, side processes occur that significantly reduce the yield of the final compound. The best results were achieved when the reaction was carried out for 4 h.
The key tricyclic compounds 11ac were synthesized by double alkylation of the corresponding amines 9ac with dichloro derivative 10. Compound 10 was obtained using a modified method for the acylation of 5,7-dimethyl-1,3-diazaadamantan-6-one in a two-phase CHCl3/H2O system, which makes it easier to isolate the product and increases the yield [36].
The reaction of alkylation was optimized by varying the base, solvent, and temperature. K2CO3 and Cs2CO3 as bases gave the best results; finally, K2CO3 was used in the reaction. NMP (N-methyl-2-pyrrolidone) was used as a solvent because it completely dissolves the starting amines and has a rather high boiling point to keep the necessary reaction temperature.

2.2. Biological Studies

To make a conclusion about the biological activity of synthesized potential AMPA receptor modulators, we performed a selective functional screening of the ligand-receptor binding. At this stage, we tried to avoid commonly used AMPA-receptor-specific radioligands, such as [3H]AMPA, [3H]fluorowillardiine, or similar, which have many different receptor targets as a rule, and many contradictory effects have been observed for various allosteric AMPA receptor modulators when these kinds of radioligands were used (see, e.g., [53,54,55]). That is why we have chosen the ligand whose targets are most likely to coincide (or overlap most tightly) with the targets of the synthesized compounds. The tritium-labeled compound PAM-43 ([3H]PAM-43 [56]) was used as a basic ligand. PAM-43 (a derivative of 3,7-diazabicyclo[3.3.1]nonane [57], Figure 1) is a positive AMPA receptor modulator, a proposed new drug candidate that has a stimulating effect on memory, learning, and integrative brain functions. PAM-43 (in a dose range of 0.05–0.5 mg/kg) stimulates memory and cognitive abilities in rats [58]. The chemical structure of PAM-43 is similar to that of some previously known synthetic allosteric modulators of AMPA receptors, and the molecular modeling also confirmed the ability of PAM-43 to act as an allosteric modulator on AMPA receptors [59]. It has also been shown that PAM-43 (as an allosteric modulator) is able to potentiate ionotropic GluR currents in rat cerebellar Purkinje neurons in a concentration-dependent manner [57]. The radioligand-receptor binding studies revealed the existence of at least two different sites of [3H]PAM-43 specific bindings localized on rat brain cells’ plasmatic membranes. These sites are different: one of them is Glu-independent and characterized by a lower affinity, and the other has a higher affinity but appears only in the presence of free Glu [57]. Theoretically, the molecular targets of PAM-43 and synthesized novel potential AMPA receptor modulators are tightly close or overlap. Thus, we assume that the existence of noticed effects of low concentrations of synthesized modulators on [3H]PAM-43 specific bindings will likely indicate the presence of functional and biological activities in tested substances.
The effect of compounds 11ac on specific binding of [3H]PAM-43 was tested for both its sites—high and low affine (Figure 2 and Figure 3). All the compounds potentiated the binding of [3H]PAM-43 on its high-affinity Glu-dependent site, but with different efficiency (Figure 2). Compound 11b was the most active and increased the binding almost twice in the concentration range of 0.1–10 nM. Compound 11c showed less efficacy, and compound 11a showed no more than a 20% increase in [3H]PAM-43 binding in all the tested concentrations of the compound.
So, the data obtained show that the Glu-dependent site of [3H]PAM-43 specific bindings, or a receptor containing this site, may be one of the targets of compounds 11b and 11c. We also assume that the increase in binding of the radioligand may be related to the effects of cooperativity of specific binding. The compounds 11b and 11c may bind first but facilitate subsequent binding of PAM-43 or increase its retention time on the receptor. In this light, we may expect the same or higher functional and, probably, the biological activity of compounds 11b and 11c.
The second site of [3H]PAM-43 specific binding is Glu-independent but possesses low affinity. Compounds 11b and 11c also had a positive effect on this site—they doubled the binding (Figure 3). The compound 11a was inactive, but we cannot exclude its efficacy at some higher concentrations.
So, both sites of [3H]PAM-43 specific bindings showed high sensitivity to low concentrations of compounds 11b and 11c, and the effect in both cases was twofold. Both compounds have demonstrated a significant increase in PAM-43 binding. According to the available data, AMPA receptors have multiple binding sites [60], so we can suppose that the binding sites of compounds 11b and 11c in the receptors at least partially do not coincide with the binding sites of PAM-43, however binding of these compounds leads to a pronounced synergistic increase in PAM-43 binding. That confirms that biotargets of new compounds and biotargets of PAM-43 are interrelated. Thus, we can expect the biological effects of new compounds to be similar to those of PAM-43. The compound 11a almost had no effects, but we cannot exclude that 11a may target the same as PAM-43 sites but with lower affinity. This is a matter of separate follow-up studies.

3. Materials and Methods

3.1. Chemistry

General information: The chemicals (indan-2-one, benzo[c][1,2,5]oxadiazole-5-carboxylic acid, carbonyldiimidazole,) were purchased from the commercial sources (Sigma-Aldrich, St. Louis, MO, USA) and used as received. 3,7-bis(chloroacetyl)-1,5-dimethyl-3,7-diazabicyclo[3.3.1]nonan-9-one were prepared according to the published methods [36] and characterized by NMR spectra. NMR spectra were recorded on the Agilent 400-MR (Bruker, Karlsruhe, Germany) spectrometer (400.0 MHz for 1H; 100.6 MHz for 13C) at room temperature; chemical shifts (δ) were measured with reference to the solvents: CDCl3 for 1H (δ = 7.26 ppm), 13C (δ = 77.16 ppm), DMSO-d6 for 1H (δ = 2.50 ppm), 13C (δ = 39.50 ppm) and CD3OD for 1H (δ = 3.31 ppm), 13C (δ = 49.00 ppm). Chemical shifts (δ) are given in ppm; J values are given in Hz. Accurate mass measurements (HRMS) were performed on a Bruker micrOTOF II instrument (Bruker, Karlsruhe, Germany) using electrospray ionization (ESI). The measurements were done in a positive ion mode (interface capillary voltage 4500 V). Analytical thin layer chromatography was carried out with Merck Silica Gel 60 plates (supported on aluminum, Sigma-Aldrich, St. Louis, MO, USA); the detection was done by UV lamp (254 and 365 nm) and chemical staining (solution of ninhydrin in EtOH). Column chromatography was performed on Merck Silica Gel 60 (Sigma-Aldrich, St. Louis, MO, USA). Original 1H and 13C NMR spectra for compounds 8ac, 9ac, and 11ac are given in Supplementary Materials.
Indan-2-one oxime (2). A mixture of indan-2-one 1 (26.42 g, 0.200 mol) and NH2OH·HCl (17.32 g, 0.250 mol) in EtOH/H2O (1:1, 1000 mL) was stirred first for 1.5 h, then under heating at 50 °C for additional 1 h. The reaction mixture was cooled to room temperature. The precipitate formed was filtered off and washed with distilled water. The product was recrystallized from EtOH (150 mL). This yielded 20.89 g (0.141 mol, 71%) of pure indan-2-one oxime 2 as needle crystals. mp: 150–153 °C [47], spectral data [61].
Indan-2-ylamine hydrochloride (3). A mixture of indan-2-one oxime 2 (5.40 g, 0.037 mol) and 10% Pd/C (3.62 g, 9 mol. %) in 1M HCl/MeOHabs (110 mL) was placed into a glass autoclave, and hydrogen was passed through (p = 5 atm) while stirring for 6 h. The reaction mixture was filtered through celite, washed with MeOH, and the solvent was distilled off. This yielded 5.64 g (0.033 mol, 91%) of pure indan-2-ylamine hydrochloride 3 as a white crystalline product. mp: 244–245 °C [47], spectral data [62].
5-Bromoindan-2-ylamine hydrobromide (4). To a solution of indan-2-ylamine hydrochloride 3 (2.82 g, 16.6 mmol) in distilled water (12.6 mL), bromine (940 μL, 18.3 mmol) was added dropwise while stirring. The reaction mixture was stirred for 9 h. After that, HBrconc (ω = 48%, 2.82 mL, 24.9 mmol) was added, and the reaction mixture was stirred for additional 3 h. The precipitate formed was filtered off and washed with distilled water (4.5 mL). The product was dissolved in EtOH (37 mL), precipitated with Et2O and filtered off. This yielded 3.15 g (13.3 mmol, 65%) of pure 5-bromoindan-2-ylamine hydrobromide 4 as a crystalline product. mp: >300 °C [63], spectral data [63].
Tert-butyl(5-bromoindan-2-yl)carbamate (5). A 20% aqueous solution of KOH was added to a suspension of 5-bromoindan-2-ylamine hydrobromide 4 (3.11 g, 10.6 mmol) in distilled water (pH 10). The resulting mixture was extracted with CH2Cl2, the combined organic phases were dried over anhydrous Na2SO4, filtered, and then the solvent was distilled off. The product was dissolved in absolute tetrahydrofuran (40 mL), then Boc2O (2.43 g, 11.1 mmol) was added, and the reaction mixture was stirred for 1.5 h. After that, the solvent was distilled off. The resulting product was treated with n-hexane (9 mL) and filtered off. This yielded 3.27 g (10.5 mmol, 99%) of pure tert-butyl(5-bromoindan-2-yl)carbamate 5 as a white solid. mp: 167–169 °C, spectral data [64].
Tert-butyl(5-cyanoindan-2-yl)carbamate (6). A mixture of tert-butyl(5-bromoindan-2-yl)carbamate 5 (936 mg, 3 mmol), Pd(OAc)2 (13.5 mg, 0.06 mmol, 2 mol. %), and dppf (33.3 mg, 0.06 mmol, 2 mol. %) in N-methyl-2-pyrrolidone/H2O (10:1 by volume, 3.3 mL) was stirred for 1 min. Then, PMHS (100 mg) and Zn(CN)2 (180 mg, 1.53 mmol) were added, and the reaction mixture was stirred under heating at 120 °C for 12 h. Afterward, the reaction mixture was cooled to room temperature, and the solvent was distilled off. The resulting product was purified by flash chromatography (eluents CH2Cl2, CH2Cl2/light petroleum ether (1:1)). This yielded 541 mg (2.1 mmol, 70%) of pure tert-butyl(5-cyanoindan-2-yl)carbamate 6 as a white solid. mp: 161–162 °C, spectral data [64].
2-Aminoindan-5-carbonitrile (7). To a solution of tert-butyl(5-cyanoindan-2-yl)carbamate 6 (980 mg, 3.79 mmol) in CH2Cl2 (25 mL), trifluoroacetic acid (2.82 mL, 36.9 mmol) was added, and the reaction mixture was stirred for 2 h. Afterward, distilled water (10 mL) and a 20% aqueous solution of KOH were added to the reaction mixture (pH = 10). The organic phase was separated; the aqueous phase was extracted with CH2Cl2. The combined organic phases were dried over anhydrous Na2SO4, then filtered, and the solvent was distilled off. This yielded 594 mg (3.75 mmol, 99%) of pure 2-aminoindan-5-carbonitrile 7 as a colorless oily product. Spectral data (in the form of hydrochloride) [64].
N-(5-Cyanoindan-2-yl)benzo[c][1,2,5]oxadiazole-5-carboxamide (8b). To a solution of benzo[c][1,2,5]oxadiazole-5-carboxylic acid chloride (242 mg, 1.33 mmol) in absolute acetonitrile, K2CO3 (350 mg, 2.53 mmol) and solution of 2-aminoindan-5-carbonitrile 7 (200 mg, 1.26 mmol) in absolute acetonitrile were added. The resulting mixture was stirred under heating at 40 °C until complete consumption of the starting amine (TLC control Rf = 0.07 in CHCl3/EtOH (20:1) system). Afterward, the reaction mixture was cooled to room temperature and filtered. The solid residue was washed with absolute acetonitrile until no more product was washed out (TLC control). The filtrates were combined, and the solvent was distilled off. This yielded pure N-(5-cyanoindan-2-yl)benzo[c][1,2,5]oxadiazole-5-carboxamide 8b (95%) as a light brown crystalline product. Rf = 0.69 (CHCl3/EtOH (20:1)). mp: 187–189 °C. 1H NMR (DMSO-d6), δ: 3.00–3.09 (m, 2H), 3.31–3.40 (m, 2H), 4.74 (m, 1H), 7.46 (d, 1H, 3J 7.8 Hz), 7.62 (dd, 1H, 3J 7.8 Hz, 4J 1.4 Hz), 7.71 (s, 1H), 7.94 (dd, 1H, 3J 9.4 Hz, 4J 1.3 Hz), 8.11 (dd, 1H, 3J 9.4 Hz, 5J 1.0 Hz), 8.52 (t, 1H, J 1.2 Hz), 9.10 (d, 1H, 3J 6.7 Hz). 13C NMR (DMSO-d6), δ: 38.58, 39.12, 50.68, 109.18, 115.81, 116.28, 119.34, 125.74, 128.17, 130.76, 131.56, 137.81, 142.94, 147.66, 148.81, 149.09, 164.55. HRMS (ESI), m/z 305.1032 (calc. C17H12N4O2 [M + H]+, m/z: 305.1033).
General procedure for acylation of 2-aminoindan-5-carbonitrile7. A mixture of corresponding acid (1 eq.) and carbonyldiimidazole (1.05 eq.) in absolute acetonitrile was stirred until the formation of corresponding imidazolide (TLC control). Then, a solution of 2-aminoindan-5-carbonitrile 7 (1 eq.) in absolute acetonitrile was added, and the reaction mixture was stirred until complete consumption of the starting amine (TLC control Rf = 0.07 in CHCl3/EtOH (20:1) system). After that, the solvent was distilled off.
N-(5-Cyanoindan-2-yl)spiro[1,3-benzodioxole-2,1′-cyclohexane]-5-carboxamide (8a). The resulting product was treated with Et2O and filtered off. This yielded pure N-(5-cyanoindan-2-yl)spiro[1,3-benzodioxole-2,1′-cyclohexane]-5-carboxamide 8a (81%) as a light brown solid. Rf = 0.79 (CHCl3/EtOH (20:1)). mp: 149–150 °C. 1H NMR (CDCl3), δ: 1.50 (m, 2H), 1.72 (m, 4H), 1.89 (dist. t, 4H), 2.93–3.00 (m, 2H), 3.40–3.48 (m, 2H), 4.93 (m, 1H), 6.17 (d, 1H, 3J 7.3 Hz), 6.71 (d, 1H, 3J 8.1 Hz), 7.14 (d, 1H, 4J 1.8 Hz), 7.20 (dd, 1H, 3J 8.1 Hz, 4J 1.8 Hz), 7.34 (d, 1H, 3J 7.8 Hz), 7.50 (br. d, 1H, 3J 7.8 Hz), 7.53 (s, 1H). 13C NMR (CDCl3), δ: 22.66, 23.99, 34.74, 39.24, 39.91, 50.51, 106.88, 107.36, 110.13, 118.80, 119.62, 120.60, 125.22, 127.08, 127.91, 130.63, 141.91, 146.46, 147.43, 150.05, 166.65. HRMS (ESI), m/z 375.1699 (calc. C23H22N2O3 [M + H]+, m/z: 375.1703).
N-(5-Cyanoindan-2-yl)benzo[b]thiophene-5-carboxamide (8c). The resulting product was recrystallized from acetonitrile. This yielded pure N-(5-cyanoindan-2-yl)benzo[b]thiophene-5-carboxamide 8c (85%) as a white solid. Rf = 0.78 (CHCl3/EtOH (20:1)). mp: 185–187 °C. 1H NMR (CDCl3), δ: 2.99–3.07 (m, 2H), 3.43–3.51 (m, 2H), 5.00 (m, 1H), 6.50 (d, 1H, 3J 6.6 Hz), 7.34–7.38 (m, 2H), 7.49–7.53 (m, 3H), 7.69 (dd, 1H, 3J 8.4 Hz, 4J 1.3 Hz), 7.90 (d, 1H, 3J 8.4 Hz), 8.22 (d, 1H, 4J 1.0 Hz). 13C NMR (CDCl3), δ: 39.20, 39.88, 50.59, 110.13, 118.81, 122.04, 122.17, 122.20, 123.74, 125.21, 127.59, 127.90, 130.06, 130.63, 138.98, 141.88, 142.37, 146.42, 167.32. HRMS (ESI), m/z 319.0894 (calc. C19H14N2OS [M + H]+, m/z: 319.0900).
General procedure for reduction in nitriles8a,b. A mixture of corresponding nitrile 8 (1 eq.), 10% Pd/C (5–10 mol. %), and HClconc (1.5 eq.) in EtOH (3 mL) was placed into a glass autoclave, and hydrogen was passed through (p = 5 atm) while stirring until complete consumption of the starting nitrile (TLC control). The reaction mixture was filtered through celite, washed with EtOH, and the solvent was distilled off.
General procedure for reduction in nitriles8a,b. A mixture of corresponding nitrile 8 (1 eq.), 10% Pd/C (5–10 mol. %), and HClconc (1.5 eq.) in EtOH (3 mL) was placed into a glass autoclave, and hydrogen was passed through (p = 5 atm) while stirring until complete consumption of the starting nitrile (TLC control). The reaction mixture was filtered through celite, washed with EtOH, and the solvent was distilled off.
N-(5-(Aminomethyl)indan-2-yl)spiro[1,3-benzodioxole-2,1′-cyclohexane]-5-carboxamide (9a) hydrochloride. This yielded pure N-(5-(aminomethyl)indan-2-yl)spiro[1,3-benzodioxole-2,1′-cyclohexane]-5-carboxamide 9a hydrochloride (83%) as a light brown solid. mp: 209–211 °C. 1H NMR (CD3OD), δ: 1.53 (m, 2H), 1.74 (m, 4H), 1.91 (dist. t, 2H), 3.02 (m, 2H), 3.34 (m, 2H), 4.08 (s, 2H), 4.78 (m, 1H), 6.77 (d, 1H, 3J 8.3 Hz), 7.23–7.37 (m, 5H), 7.59 (s, 1H), 8.93 (br. s, 1H). 13C NMR (CD3OD), δ: 22.46, 23.72, 34.33, 38.02, 38.26, 42.87, 51.07, 106.69, 106.99, 119.14, 121.25, 124.50, 124.58, 126.96, 127.00, 131.04, 141.87, 141.96, 147.17, 149.93, 167.90. HRMS (ESI), m/z 379.2011 (calc. C23H26N2O3 [M + H]+, m/z: 379.2016).
N-(5-(aminomethyl)indan-2-yl)-4,5,6,7-tetrahydrobenzo[c][1,2,5]oxadiazole-5-carboxamide (9b) hydrochloride. This yielded pure N-(5-(aminomethyl)indan-2-yl)-4,5,6,7-tetrahydrobenzo[c][1,2,5]oxadiazole-5-carboxamide 9b hydrochloride (83%) as a brown solid. mp: 240–241 °C (decomp.). 1H NMR (CD3OD), δ: 1.95 (m, 1H), 2.12 (m, 1H), 2.79–3.03 (m, 7H), 3.28 (m, 2H), 4.08 (s, 2H), 4.60 (m, 1H), 7.25–7.33 (m, 3H). 13C NMR (CD3OD), δ: 18.03, 21.98, 25.49, 38.27, 38.50, 38.99, 42.83, 50.37, 124.58, 124.65, 127.00, 131.11, 141.74, 141.79, 150.77, 150.95, 174.21. HRMS (ESI), m/z 313.1661 (calc. C17H20N4O2 [M + H]+, m/z: 313.1659).
N-(5-(Aminomethyl)indan-2-yl)benzo[b]thiophene-5-carboxamide (9c). A solution of N-(5-cyanoindan-2-yl)benzo[b]thiophene-5-carboxamide 8c (172 mg, 0.54 mmol) in absolute methanol (20 mL) was cooled to 0 °C. Then, CoCl2·6H2O (257 mg, 1.08 mmol) and NaBH4 (204 mg, 5.40 mmol) were added under an inert atmosphere (argon). The reaction mixture was stirred under cooling at 0 °C for 30 min and then at room temperature for 4 h. After that, distilled water (2 mL) and 20% aqueous solution of KOH were added to the reaction mixture (pH = 9). The precipitate formed was filtered off and washed with methanol. The solvent was distilled off from the combined filtrate. The product was dissolved in distilled water, and the solution was extracted with CH2Cl2. The combined organic phases were dried over anhydrous Na2SO4, then filtered, and the solvent was distilled off. The resulting product was purified by column chromatography (eluents CHCl3, CHCl3/EtOH (100:1), CHCl3/EtOH (50:1), CHCl3/EtOH (20:1), CHCl3/EtOH (5:1)). This yielded 80 mg (0.25 mmol, 46%) of pure N-(5-(aminomethyl)indan-2-yl)benzo[b]thiophene-5-carboxamide 9c as a white solid. mp: 223–225 °C. 1H NMR (CD3OD), δ: 2.95–3.06 (m, 2H), 3.27–3.37 (m, 2H), 3.80 (s, 2H), 4.84 (m, 1H), 7.15 (d, 1H, 3J 7.5 Hz), 7.19–7.23 (m, 2H), 7.44 (d, 1H, 3J 5.5 Hz), 7.65 (d, 1H, 3J 5.5 Hz), 7.79 (dd, 1H, 3J 8.5 Hz, 4J 1.7 Hz), 7.96 (d, 1H, 3J 8.5 Hz), 8.32 (s, 1H). 13C NMR (CD3OD), δ: 37.99, 38.29, 44.47, 51.16, 121.58, 122.11, 122.18, 123.14, 123.52, 123.87, 125.54, 127.27, 130.18, 138.78, 139.06, 139.70, 141.10, 142.31, 168.68. HRMS (ESI), m/z 323.1208 (calc. C19H18N2OS [M + H]+, m/z: 323.1213).
General procedure for alkylation of amines9a,b,c. To a solution of 3,7-bis(chloroacetyl)-1,5-dimethyl-3,7-diazabicyclo[3.3.1]nonan-9-one 10 (1 eq.) and the corresponding amine 9 (1 eq.) in NMP, K2CO3 (5 eq.) was added, and the resulting mixture was stirred under heating at 75 °C for 15 h. After that, the reaction mixture was cooled to room temperature, filtered off, and the solid residue was washed with NMP until no more product was washed out (TLC control). The filtrates were combined, and the solvent was distilled off with an oil pump. The solid residue was treated with Et2O and filtered off. The resulting product was purified by column chromatography (eluents CHCl3, CHCl3/EtOH (100:1), CHCl3/EtOH (50:1)).
N-(5-((1,11-Dimethyl-4,8,12-trioxo-3,6,9-triazatricyclo[7.3.1.13,11]tetradec-6-yl)methyl)indan-2-yl)spiro[1,3-benzodioxole-2,1′-cyclohexane]-5-carboxamide (11a). This yielded pure N-(5-((1,11-dimethyl-4,8,12-trioxo-3,6,9-triazatricyclo[7.3.1.13,11]tetradec-6-yl)methyl)indan-2-yl)spiro[1,3-benzodioxole-2,1′-cyclohexane]-5-carboxamide 11a (63%) as a white solid. Rf = 0.43 (CHCl3/EtOH (20:1)). mp: 147–149 °C. 1H NMR (CDCl3), δ: 0.99 (s, 3H), 1.08 (s, 3H), 1.50 (m, 2H), 1.72 (m, 4H), 1.89 (m, 4H), 2.72 (d, 2J 13.5 Hz), 2.88–2.94 (m, 2H), 3.01 (d, 2H, 2J 13.2 Hz), 3.18–3.23 (m, 2H), 3.35–3.44 (m, 2H), 3.58 (s, 2H), 3.75–3.79 (m, 2H), 4.79 (m, 1H), 4.90–4.93 (m, 4H), 6.24 (d, 1H, 3J 7.3 Hz), 6.71 (d, 1H, 3J 8.1 Hz), 7.13–7.25 (m, 5H). 13C NMR (CDCl3), δ: 15.82, 16.21, 23.08, 24.40, 35.16, 39.79, 39.97, 45.36, 45.78, 51.22, 54.34, 55.43, 60.32, 60.45, 62.42, 107.25, 107.79, 120.01, 120.97, 125.10, 126.44, 127.76, 128.63, 134.45, 141.45, 141.86, 147.80, 150.36, 166.94, 168.15, 210.91. HRMS (ESI), m/z 627.3179 (calc. C36H42N4O6 [M + H]+, m/z: 627.3177).
N-(5-((1,11-Dimethyl-4,8,12-trioxo-3,6,9-triazatricyclo[7.3.1.13,11]tetradec-6-yl)methyl)indan-2-yl)-4,5,6,7-tetrahydrobenzo[c][1,2,5]oxadiazole-5-carboxamide (11b). This yielded pure N-(5-((1,11-dimethyl-4,8,12-trioxo-3,6,9-triazatricyclo[7.3.1.13,11]tetradec-6-yl)methyl)indan-2-yl)-4,5,6,7-tetrahydrobenzo[c][1,2,5] oxadiazole-5-carboxamide 11b (53%) as a white solid. Rf = 0.35 (CHCl3/EtOH (20:1)). mp: 169–171°C. 1H NMR (CDCl3), δ: 1.00 (s, 3H), 1.09 (s, 3H), 2.00 (m, 1H), 2.19 (m, 1H), 2.56 (m, 1H), 2.73–2.87 (m, 5H), 3.04 (d, 2H, 2J 13.6 Hz), 3.11–3.23 (m, 5H), 3.33–3.41 (m, 2H), 3.61 (s, 2H), 3.73–3.78 (m, 2H), 4.80 (m, 1H), 4.91 (m, 4H), 6.05 (d, 1H, 3J 7.2 Hz), 7.14–7.29 (m, 3H). 13C NMR (CDCl3), δ: 15.87, 16.22, 19.08, 23.29, 26.05, 39.83, 39.91, 40.50, 45.35, 45.82, 50.82, 54.34, 55.49, 60.18, 60.35, 62.34, 125.13, 126.45, 128.76, 134.36, 141.17, 141.58, 150.76, 150.99, 168.12, 172.70, 210.82. HRMS (ESI), m/z 561.2824 (calc. C30H36N6O5 [M + H]+, m/z: 561.2820).
N-(5-((1,11-Dimethyl-4,8,12-trioxo-3,6,9-triazatricyclo[7.3.1.13,11]tetradec-6-yl)methyl)indan-2-yl)benzo[b]thiophene-5-carboxamide (11c). This yielded pure N-(5-((1,11-dimethyl-4,8,12-trioxo-3,6,9-triazatricyclo[7.3.1.13,11]tetradec-6-yl)methyl)indan-2-yl)benzo[b]thiophene-5-carboxamide 11c (56%) as a white solid. Rf = 0.31 (CHCl3/EtOH (20:1)). mp: 162–164 °C. 1H NMR (CDCl3), δ: 1.00 (s, 3H), 1.09 (s, 3H), 2.74 (d, 2H, 2J 13.5 Hz), 2.96–3.03 (m, 4H), 3.21–3.26 (m, 2H), 3.44–3.51 (m, 2H), 3.60 (s, 2H), 3.78 (d, 2H, 2J 14.2 Hz), 4.93 (d, 4H, 2J 13.5 Hz), 5.02 (m, 1H), 6.45 (d, 1H, 3J 6.6 Hz), 7.17 (d, 1H, 3J 7.7 Hz), 7.22 (s, 1H), 7.30 (d, 1H, 3J 7.7 Hz), 7.40 (d, 1H, 3J 5.6 Hz), 7.53 (d, 1H, 3J 5.6 Hz), 7.71 (dd, 1H, 3J 8.4 Hz, 4J 1.5 Hz), 7.92 (d, 1H, 3J 8.4 Hz), 8.25 (d, 1H, 4J 1.5 Hz). 13C NMR (CDCl3), δ: 15.84, 16.21, 39.87, 40.03, 45.37, 45.83, 51.35, 54.34, 55.45, 60.34, 60.43, 62.44, 122.42, 122.57, 122.61, 124.19, 125.15, 126.49, 127.98, 128.69, 130.73, 134.53, 139.44, 141.42, 141.84, 142.72, 167.58, 168.15, 210.94. HRMS (ESI), m/z 571.2369 (calc. C32H34N4O4S [M + H]+, m/z: 571.2374).

3.2. Radioligand Binding

Synthesis of Tritium-Labeled PAM-43. Tritium-labeled PAM-43 (3,7-bis(2,3-dihydro-1-benzofuran-5-ylcarbonyl)-1,5-dimethyl-3,7-diazabicyclo[3.3.1]-nonan-9-one) was synthesized at Laboratory of Molecular Pharmacology of Peptides, Institute of Molecular Genetics, National Research Centre «Kurchatov Institute» (Moscow, Russia). A solid-phase method [56] was used. The resulting compound [3H]PAM-43 was characterized by the specific activity of 132 Ci/mmol and a chemical purity of 98.7%.
Radioligand Binding. The radioligand-receptor assay was performed on a special unit, and the MultiScreenHTS 96-Well Filter Plates (MultiScreen System, EMD Millipore, Darmstadt, Germany) were used. As a main radioligand, we used tritium-labeled PAM-43 ([3H]PAM-43 with a specific radioactivity of 132 Ci/mmol and radiochemical purity of over 95%). Plasmatic membranes of rat brain cells were the target tissue. The incubation of the reaction mixture was performed directly in the wells of the standard 96-well plates with GF/B filters (Millipore). The reaction mixture (final volume, 200 μL) contained: 50 μL of the radioactively labeled ligand in buffer solution and 50 μL of buffer (containing either unlabeled ligand PAM-43 (250 μM), or the compound under study, depending on the experiment point), and a membrane protein solution. The reaction was initiated by the addition of 100 μL of membrane protein solution (whose final concentration in the incubation mixture was 0.2 mg/mL) dissolved in Buffer B (50 mM Tris-HCl, 1 mM CaCl2, 0.003% BSA, pH 7.4 at 30 °C for incubation or pH 7.4 at 4 °C for plates washing) with an added cocktail of inhibitors (100 μM PMSF + 10 μM Bacitracin + 5 μM Pepstatin A). In some experiments, the reaction mixture contained 250 μM of Glu. L-Glu was preliminarily added to plasmatic membranes in the concentration of 500 μM (30 min before the incubation with the radioligand). The plates were kept at 30 °C with continuous shaking for 20 min. After the incubation, the plates were air-dried, filters were detached, and transferred into scintillation vials, each containing 4 mL of the liquid scintillator (Unisolve 100; Koch-Light, Haverhill, UK); radioactivity was measured using a Tri-Carb 2100R liquid scintillation counter (Packard BioScience, USA). Mathematical processing of results was conducted using the SigmaPlot 10.0 software suite (Systat Software Inc., San Jose, CA, USA).
Experimental design. There were three types of the reaction mixture on a standard 96-well plate. The first type contained the radioactively labeled ligand and a membrane protein solution. The second type contained the radioactively labeled ligand, a membrane protein solution, and the unlabeled ligand PAM-43. Additionally, the third type contained the radioactively labeled ligand, a membrane protein solution, and the compound under study (in different concentrations). Specific binding was determined as the difference in radioactivity measured in samples of the first and second types. A value of 100% on the graphs corresponds to [3H]PAM-43 specific binding determined in the presence of unlabeled PAM-43. The data measured in the samples of the third type were subjected to statistical and mathematical processing, taking into account the value of nonspecific binding. The influence of the compounds under study on [3H]PAM-43-specific binding was represented as a proportion (in %) of [3H]PAM-43-specific binding.
Statistical Analysis. Data from the radioligand-receptor binding assay were determined using non-linear regression analysis (five-parameter logistic curve) included in the Pharmacology module of the SigmaPlot 10 software suite (Systat Software Inc., San Jose, CA, USA). Values on the graphs represent the mean ± S.E. of 3 independent experiments. Each experiment was conducted on isolated plasmatic membranes of different groups of rats; each mean in the experiment was obtained as average from 6 separated experimental volumes (wells of the standard 96-well plates with GF/B filters (Millipore)).
Animal Management. Male adult albino outbred rats (mean body weight, 180–200 g) were housed in plastic cages under standard laboratory conditions, which included a controlled ambient temperature (22–25 °C), a 12 h light/dark cycle, and 60 ± 10% humidity during all experiments. Before the start of an experiment, animals were allowed access to standard laboratory rat pellet chow and water ad libitum.
Isolation of Plasma Membranes. Rat brain cells’ plasmatic membranes were isolated at 4 °C. The rats were decapitated; their brains were washed with cold PBS, and brain structures (cortex and hippocampus) were isolated and added to Buffer A1 (10 mM Tris-HCl, pH 7.4 at 4 °C, saccharose 0.32 M, 1 mM EDTA, 1 mM benzamidine, 0.1 mM PMSF). The resulting samples were homogenized in 10 volumes of the buffer using a Teflon-in-glass homogenizer, and then the homogenate was centrifuged at 1000× g for 20 min, the sediment removed, and the supernatant centrifuged at 40,000× g for 30 min. The dense brown mitochondria-rich sediment at the bottom of the tube was removed, and the less-dense translucent sediment of membranes was resuspended in Buffer A1, transferred to a clean tube, and centrifuged again at 40,000× g for 30 min. The sediment was resuspended in Buffer A2 (10 mM Tris-HCl, pH 7.4 at 4 °C, 0.22 M saccharose), divided into portions, frozen in liquid nitrogen, and stored no longer than 30 days at −70 °C. Protein concentration in membrane samples was measured according to the Hartree–Lowry method.

4. Conclusions

In this paper, we described the development and optimization of synthetic approaches to three novel compounds designed as potential allosteric modulators of AMPA receptors. For the synthesized compounds, the radioligand-receptor binding analysis was performed using [3H]PAM-43, a synthetic positive allosteric modulator of AMPA receptors described earlier, as a standard ligand. According to the published data, there exist multiple binding sites in the AMPA receptor. The simultaneous action of a ligand on several sites can change the conformation of the receptor and lead to the net effect of either positive or negative allosteric modulation.
As a result of radioligand-receptor binding analysis, it was found that two synthesized compounds 11b and 11c bind with the same targets as positive allosteric modulator PAM-43 (at least on AMPA receptors) but with different binding sites, leading to a significant increase in the [3H]PAM-43 response. For compound 11a our results may indicate the selectivity with respect to the PAM-43 binding site, but its effect is weaker. We also hypothesize that the enhancement of radioligand binding may be due to specific binding cooperativity effects, and further optimization of the structures described here can lead to broad-spectrum drug candidates for the treatment and prevention of diseases associated with impaired functions of the central nervous system.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241210293/s1, 1H NMR and 13C NMR spectra for all isolated compounds.

Author Contributions

Conceptualization, methodology, V.A.P., M.I.L. and T.V.V.; investigation, E.A.G., P.N.V., M.I.L., T.V.V., K.V.S., M.A.T., A.F.A. and V.A.P.; writing—original draft preparation, E.A.G., P.N.V., M.I.L. and T.V.V.; writing—review and editing, V.A.P., M.I.L., P.N.V. and E.A.G.; supervision, V.A.P.; funding acquisition, V.A.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Russian Science Foundation grant number 22-15-00041. The development of general techniques of radioligand-receptor analysis was supported by the Program of Scientific Research of the National Research Center “Kurchatov Institute”.

Institutional Review Board Statement

All experimental animal procedures were approved by the local Ethics Committee of the Experimental Animal Care Society of the Institute of Molecular Genetics (no. 9 of 18 October 2019) and were conducted in accordance with the Russian law “On Establishing Rules of Good Laboratory Practice” (23 September 2010, no.708n). All methods used were in compliance with the requirements of EU Directive 2010/63/EU (used for scientific purposes with effect from 1 January 2013) for animal experiments. All efforts were made to minimize animal suffering and to reduce the number of animals used.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data of the current study are available from the corresponding author upon reasonable request.

Acknowledgments

The synthetic procedure to obtain compound 6 from 5 was developed at A. V. Topchiev Institute of Petrochemical Synthesis, RAS.

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.

References

  1. Reiner, A.; Levitz, J. Glutamatergic signaling in the central nervous system: Ionotropic and metabotropic receptors in concert. Neuron 2018, 98, 1080–1098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. 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] [PubMed]
  3. Hansen, K.B.; Wollmuth, L.P.; Bowie, D.; Furukawa, H.; Menniti, F.S.; Sobolevsky, A.I.; Swanson, G.T.; Swanger, S.A.; Greger, I.H.; Nakagawa, T.; et al. Structure, function, and pharmacology of glutamate receptor ion channels. Pharmacol. Rev. 2021, 73, 298–487. [Google Scholar] [CrossRef]
  4. Partin, K.M. AMPA receptor potentiators: From drug design to cognitive enhancement. Curr. Opin. Pharmacol. 2015, 20, 46–53. [Google Scholar] [CrossRef] [Green Version]
  5. 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] [Green Version]
  6. Twomey, E.C.; Sobolevsky, A.I. Structural mechanisms of gating in ionotropic glutamate receptors. Biochemistry 2018, 57, 267–276. [Google Scholar] [CrossRef] [Green Version]
  7. Chen, S.; Gouaux, E. Structure and mechanism of AMPA receptor—Auxiliary protein complexes. Curr. Opin. Struct. Biol. 2019, 54, 104–111. [Google Scholar] [CrossRef]
  8. Coombs, I.D.; Soto, D.; McGee, T.P.; Gold, M.G.; Farrant, M.; Cull-Candy, S.G. Homomeric GluA2(R) AMPA receptors can conduct when desensitized. Nat. Commun. 2019, 10, 4312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Salazar, H.; Mischke, S.; Plested, A.J.R. Measurements of the timescale and conformational space of AMPA receptor desensitization. Biophys. J. 2020, 119, 206–218. [Google Scholar] [CrossRef] [PubMed]
  10. 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] [PubMed] [Green Version]
  11. 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] [PubMed]
  12. 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] [PubMed]
  13. 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]
  14. Matthews, P.M.; Pinggera, A.; Kampjut, D.; Greger, I.H. Biology of AMPA receptor interacting proteins—From biogenesis to synaptic plasticity. Neuropharmacology 2021, 197, 108709. [Google Scholar] [CrossRef]
  15. Díaz-Alonso, J.; Nicoll, R.A. AMPA receptor trafficking and LTP: Carboxy-termini, amino-termini and TARPs. Neuropharmacology 2021, 197, 108710. [Google Scholar] [CrossRef]
  16. Wu, Q.-L.; Gao, Y.; Li, J.-T.; Ma, W.-Y.; Chen, N.-H. The role of AMPARs composition and trafficking in synaptic plasticity and diseases. Cell Mol. Neurobiol. 2022, 42, 2489–2504. [Google Scholar] [CrossRef] [PubMed]
  17. Cavalleri, L.; Merlo Pich, E.; Millan, M.J.; Chiamulera, C.; Kunath, T.; Spano, P.F.; Collo, G. Ketamine enhances structural plasticity in mouse mesencephalic and human iPSC-derived dopaminergic neurons via AMPAR-driven BDNF and mTOR signaling. Mol. Psychiatry 2018, 23, 812–823. [Google Scholar] [CrossRef] [Green Version]
  18. 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]
  19. 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] [Green Version]
  20. Pilar-Cuellar, F.; Castro, E.; Bretin, S.; Mocaer, E.; Pazos, Á.; Díaz, Á. S 47445 counteracts the behavioral manifestations and hippocampal neuroplasticity changes in bulbectomized mice. Prog. Neuropsychopharmacol. Biol. Psychiatry 2019, 93, 205–213. [Google Scholar] [CrossRef]
  21. Kunugi, A.; Tajima, Y.; Kuno, H.; Sogabe, S.; Kimura, H. HBT1, a novel AMPA receptor potentiator with lower agonistic effect, avoided bell-shaped response in in vitro BDNF production. J. Pharmacol. Exp. Ther. 2018, 364, 377–389. [Google Scholar] [CrossRef] [PubMed]
  22. Sanacora, G.; Zarate, C.A.; Krystal, J.H.; Manji, H.K. Targeting the glutamatergic system to develop novel, improved therapeutics for mood disorders. Nat. Rev. Drug Discov. 2008, 7, 426–437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Adler, L.A.; Kroon, R.A.; Stein, M. A translational approach to evaluate the efficacy and safety of the novel AMPA receptor positive allosteric modulator org 26576 in adult attention-deficit/hyperactivity disorder. Biol Psychiatry 2012, 72, 971–977. [Google Scholar] [CrossRef] [PubMed]
  24. Xiao, D.; Xie, F.; Xu, X.; Zhou, X. The impact and mechanism of ampakine CX1739 on protection against respiratory depression in rats. Future Med. Chem. 2020, 12, 2093–2104. [Google Scholar] [CrossRef] [PubMed]
  25. Suzuki, A.; Kunugi, A.; Tajima, Y.; Suzuki, N.; Suzuki, M.; Toyofuku, M.; Kuno, H.; Sogabe, S.; Kosugi, Y.; Awasaki, Y.; et al. Strictly regulated agonist-dependent activation of AMPA-R is the key characteristic of TAK-653 for robust synaptic responses and cognitive improvement. Sci. Rep. 2021, 11, 14532. [Google Scholar] [CrossRef]
  26. Suzuki, A.; Hara, H.; Kimura, H. Role of the AMPA receptor in antidepressant effects of ketamine and potential of AMPA receptor potentiators as a novel antidepressant. Neuropharmacology 2022, 222, 109308. [Google Scholar] [CrossRef]
  27. Kadriu, B.; Musazzi, L.; Johnston, J.N.; Kalynchuk, L.E.; Caruncho, H.J.; Popoli, M.; Zarate, C.A. Positive AMPA receptor modulation in the treatment of neuropsychiatric disorders: A long and winding road. Drug Discov. Today 2021, 26, 2816–2838. [Google Scholar] [CrossRef]
  28. Cetin, S.; Knez, D.; Gobec, S.; Kos, J.; Pišlar, A. Cell models for Alzheimer’s and Parkinson’s disease: At the interface of biology and drug discovery. Biomed. Pharmacother. 2022, 149, 112924. [Google Scholar] [CrossRef]
  29. Banerjee, R.; Rai, A.; Iyer, S.M.; Narwal, S.; Tare, M. Animal models in the study of Alzheimer’s disease and Parkinson’s disease: A historical perspective. Anim. Model. Exp. Med. 2022, 5, 27–37. [Google Scholar] [CrossRef]
  30. Yang, W.; Chen, X.; Li, S.; Li, X.-J. Genetically modified large animal models for investigating neurodegenerative diseases. Cell Biosci. 2021, 11, 218. [Google Scholar] [CrossRef]
  31. Rapaka, D.; Adiukwu, P.C.; Bitra, V.R. Experimentally induced animal models for cognitive dysfunction and Alzheimer’s disease. MethodsX 2022, 9, 101933. [Google Scholar] [CrossRef] [PubMed]
  32. Białoń, M.; Wąsik, A. Advantages and limitations of animal schizophrenia models. Int. J. Mol. Sci. 2022, 23, 5968. [Google Scholar] [CrossRef] [PubMed]
  33. Karlov, D.S.; Lavrov, M.I.; Palyulin, V.A.; Zefirov, N.S. MMGBSA 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]
  34. 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]
  35. 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. Mendeleev Commun. 2018, 28, 311–313. [Google Scholar] [CrossRef]
  36. Lavrov, M.I.; Stroganov, O.V.; Zamoyski, V.L.; Grigoriev, V.V.; Zapolskiy, M.E.; Sysolyatin, S.V.; Malykhin, V.V.; Surmachev, V.N.; Palyulin, V.A. Synthesis of an allosteric modulator of ionotropic glutamate receptors. Mendeleev Commun. 2020, 30, 156–158. [Google Scholar] [CrossRef]
  37. Grigoriev, V.V.; Lavrov, M.I.; Zamoyski, V.L.; Garibova, T.L.; Palyulin, V.A.; Bachurin, S.O. New Positive Allosteric Modulator of AMPA Receptors: In vitro and in vivo Studies. Dokl. Biochem. Biophys. 2019, 488, 304–306. [Google Scholar] [CrossRef]
  38. Zapolsky, M.E.; Zefirov, N.S.; Palyulin, V.A.; Lavrov, M.I. Tricyclic Derivatives of N,N′-Substituted 3,7-diazabicyclo[3.3.1]nonanes and Drugs Based Thereon. U.S. Patent 9440985 B2, 13 September 2016. [Google Scholar]
  39. Golubeva, E.A.; Lavrov, M.I.; Veremeeva, P.N.; Bovina, E.M.; Radchenko, E.V.; Topchiy, M.A.; Asachenko, A.F.; Zamoiski, V.L.; Grigoriev, V.V.; Palyulin, V.A. New 1,11-dimethyl-3,6,9-triazatricyclo[7.3.1.13,11]tetradecane-4,8,12-trione derivative as an allosteric modulator of the glutamatergic system. Mendeleev Commun. 2023, 33, 70–72. [Google Scholar] [CrossRef]
  40. Gasperi, V.; Savini, I.; Catani, M.V. Assay of CB1 Receptor Binding. Methods Mol. Biol. 2023, 2576, 95–109. [Google Scholar] [CrossRef]
  41. Arai, K.; Homma, T.; Morikawa, Y.; Ubukata, N.; Tsuruoka, H.; Aoki, K.; Ishikawa, H.; Mizuno, M.; Sada, T. Pharmacological profile of CS-3150, a novel, highly potent and selective non-steroidal mineralocorticoid receptor antagonist. Eur. J. Pharmacol. 2015, 761, 226–234. [Google Scholar] [CrossRef]
  42. Subramanian, N.; Kalkman, H.O. Receptor profile of P88-8991 and P95-12113, metabolites of the novel antipsychotic iloperidone. Prog. Neuropsychopharmacol. Biol. Psychiatry 2002, 26, 553–560. [Google Scholar] [CrossRef]
  43. Adamska-Bartłomiejczyk, A.; Lipiński, P.F.J.; Piekielna-Ciesielska, J.; Kluczyk, A.; Janecka, A. Pharmacological Profile and Molecular Modeling of Cyclic Opioid Analogues Incorporating Various Phenylalanine Derivatives. ChemMedChem 2020, 15, 1322–1329. [Google Scholar] [CrossRef] [PubMed]
  44. Dambrova, M.; Zvejniece, L.; Liepinsh, E.; Cirule, H.; Zharkova, O.; Veinberg, G.; Kalvinsh, I. Comparative pharmacological activity of optical isomers of phenibut. Eur. J. Pharmacol. 2008, 583, 128–134. [Google Scholar] [CrossRef] [PubMed]
  45. Zeilinger, M.; Pichler, F.; Nics, L.; Wadsak, W.; Spreitzer, H.; Hacker, M.; Mitterhauser, M. New approaches for the reliable in vitro assessment of binding affinity based on high-resolution real-time data acquisition of radioligand-receptor binding kinetics. EJNMMI Res. 2017, 7, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Garzone, P.D.; Kroboth, P.D. Pharmacokinetics of the newer benzodiazepines. Clin. Pharmacokinet. 1989, 16, 337–364. [Google Scholar] [CrossRef] [PubMed]
  47. Cannon, J.G.; Perez, J.A.; Pease, J.P. Comparison of biological effects of N-alkylated congeners of β-phenethylamine derived from 2-aminotetralin, 2-aminoindan, and 6-aminobenzocycloheptene. J. Med. Chem. 1980, 23, 745–749. [Google Scholar] [CrossRef]
  48. Levin, N.; Graham, B.E.; Kolloff, H.G. Physiologically active indanamines. J. Org. Chem. 1944, 9, 380–391. [Google Scholar] [CrossRef]
  49. Cere, V.; Dal Monte, D.; Sandri, E. Catalytic Hydrogenation of Benzo-2,l, 3.-oxadiazoles. Tetrahedron 1972, 28, 3271–3276. [Google Scholar] [CrossRef]
  50. Heinzman, S.W.; Ganem, B. Mechanism of sodium borohydride-cobaltous chloride reductions. J. Am. Chem. Soc. 1982, 104, 6801–6802. [Google Scholar] [CrossRef]
  51. Itsuno, S.; Sakurai, Y.; Ito, K. Reduction of some functional groups with zirconium tetrachloride/sodium borohydride. Synth. Commun. 1988, 12, 995–997. [Google Scholar] [CrossRef]
  52. Satoh, T.; Suzuki, S. Reduction of organic compounds with sodium borohydride-transition metal salt systems (1). Reduction of organic nitride, nitro and amide compounds to primary amines. Tetrahedron Lett. 1969, 52, 4555–4558. [Google Scholar] [CrossRef]
  53. Yamada, K.A.; Turetsky, D.M. Allosteric interactions between cyclothiazide and AMPA/kainate receptor antagonists. Br. J. Pharmacol. 1996, 117, 1663–1672. [Google Scholar] [CrossRef] [Green Version]
  54. Kessler, M.; Arai, A.; Quan, A.; Lynch, G. Effect of cyclothiazide on binding properties of AMPA-type glutamate receptors: Lack of competition between cyclothiazide and GYKI 52466. Mol. Pharmacol. 1996, 49, 123–131. [Google Scholar]
  55. Lindén, A.M.; Yu, H.; Zarrinmayeh, H.; Wheeler, W.J.; Skolnick, P. Binding of an AMPA receptor potentiator ([3H]LY395153) to native and recombinant AMPA receptors. Neuropharmacology 2001, 40, 1010–1018. [Google Scholar] [CrossRef] [PubMed]
  56. Nagaev, I.Y.; Shevchenko, K.V.; Shevchenko, V.P.; Myasoedov, N.F.; Grigoriev, V.V.; Lavrov, M.I.; Bondarenko, E.V.; Kalashnikova, E.E. Synthesis of tritium-labeled PAM-43. Mendeleev Commun. 2018, 28, 64–65. [Google Scholar] [CrossRef]
  57. Vyunova, T.V.; Andreeva, L.A.; Shevchenko, K.V.; Grigoriev, V.V.; Palyulin, V.A.; Lavrov, M.I.; Bondarenko, E.V.; Kalashnikova, E.E.; Myasoedov, N.F. Characterization of a New Positive Allosteric Modulator of AMPA Receptors PAM-43: Specific Binding of the Ligand and its Ability to Potentiate AMPAR Currents. Curr. Mol. Pharmacol. 2020, 13, 217–223. [Google Scholar] [CrossRef] [PubMed]
  58. Bachurin, S.O.; Grigorev, V.V.; Palyulin, V.A.; Lavrov, M.I.; Zefirov, N.S.; Garibova, T.L.; Voronina, T.A.; Roziev, R.A. N,N′-Substituted 3,7-diazabicyclo[3.3.1]nonanes, Pharmaceutical Compositions Based Thereon and Use Thereof. RU Patent 2613071, 29 October 2017. [Google Scholar]
  59. Bachurin, S.O.; Grigorev, V.V.; Zefirov, N.S.; Lavrov, M.I.; Lapteva, V.L.; Palyulin, V.A. N,N′-Substituted 3,7-diazabicyclo[3,3,1]nonanes with Pharmacological Effect, Pharmacological Compositions on Their Base, and Application Method. RU Patent 2333211, 10 September 2008. [Google Scholar]
  60. Golubeva, E.A.; Lavrov, M.I.; Radchenko, E.V.; Palyulin, V.A. Diversity of AMPA receptor ligands: Chemotypes, binding modes, mechanisms of action, and therapeutic effects. Biomolecules 2023, 13, 56. [Google Scholar] [CrossRef]
  61. Lyle, R.E.; Walsh, D.A. An improved synthesis of 1,4-dihydro-3-[2H]-isoquinolone. Org. Prep. Proced. Int. 1973, 5, 299. [Google Scholar] [CrossRef]
  62. Göksu, S.; Seçen, H. Concise syntheses of 2-aminoindans via indan-2-ol. Tetrahedron 2005, 61, 6801–6807. [Google Scholar] [CrossRef]
  63. Prashad, M.; Hu, B.; Har, D.; Repic, O.; Blacklock, T.J.; Acemoglub, M. Efficient and Practical Syntheses of (R)-(5-Amino-2,3-dihydro-1H-inden-2-yl)-carbamic Acid Methyl Ester. Adv. Synth. Catal. 2001, 343, 461–472. [Google Scholar] [CrossRef]
  64. Barvian, K.K.; Speake, J.D.; Cowan, D.J.; Larkin, A.L.; Szewczyk, J.R. Novel Heterocycle Compounds. U.S. Patent 20090054431A1, 26 February 2009. [Google Scholar]
Scheme 1. Synthesis of target compounds 11ac.
Scheme 1. Synthesis of target compounds 11ac.
Ijms 24 10293 sch001
Figure 1. PAM-43 structure.
Figure 1. PAM-43 structure.
Ijms 24 10293 g001
Figure 2. Specific binding of [3H]PAM-43 (50 nM, L-Glu saturation) in the presence of different concentrations of synthesized compounds to corresponding receptors on rat cerebral cortex neurons’ plasmatic membranes. The abscissa represents the concentration of a compound in the incubation mixture, nmol per liter. The ordinate represents the proportion of [3H]PAM-43 (50 nM) specific binding, %. A value of 100% corresponds to [3H]PAM-43 (50 nM) specific binding determined in the presence of unlabeled PAM-43 (250 μM). L-Glu (500 μM) was preliminarily added to plasmatic membranes 30 min before the incubation with radioligand and compounds; the final concentration of L-Glu in the binding assay was 250 μM.
Figure 2. Specific binding of [3H]PAM-43 (50 nM, L-Glu saturation) in the presence of different concentrations of synthesized compounds to corresponding receptors on rat cerebral cortex neurons’ plasmatic membranes. The abscissa represents the concentration of a compound in the incubation mixture, nmol per liter. The ordinate represents the proportion of [3H]PAM-43 (50 nM) specific binding, %. A value of 100% corresponds to [3H]PAM-43 (50 nM) specific binding determined in the presence of unlabeled PAM-43 (250 μM). L-Glu (500 μM) was preliminarily added to plasmatic membranes 30 min before the incubation with radioligand and compounds; the final concentration of L-Glu in the binding assay was 250 μM.
Ijms 24 10293 g002
Figure 3. Specific binding of [3H]PAM-43 (100 nM) in the presence of different concentrations of synthesized compounds to corresponding receptors on rat cerebral cortex neurons’ plasmatic membranes. The abscissa represents the concentration of a compound in the incubation mixture, nmol per liter. The ordinate represents the proportion of [3H]PAM-43 (100 nM) specific binding, %. A value of 100% corresponds to [3H]PAM-43 (100 nM) specific binding determined in the presence of unlabeled PAM-43 (250 μM).
Figure 3. Specific binding of [3H]PAM-43 (100 nM) in the presence of different concentrations of synthesized compounds to corresponding receptors on rat cerebral cortex neurons’ plasmatic membranes. The abscissa represents the concentration of a compound in the incubation mixture, nmol per liter. The ordinate represents the proportion of [3H]PAM-43 (100 nM) specific binding, %. A value of 100% corresponds to [3H]PAM-43 (100 nM) specific binding determined in the presence of unlabeled PAM-43 (250 μM).
Ijms 24 10293 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Golubeva, E.A.; Lavrov, M.I.; Veremeeva, P.N.; Vyunova, T.V.; Shevchenko, K.V.; Topchiy, M.A.; Asachenko, A.F.; Palyulin, V.A. New Allosteric Modulators of AMPA Receptors: Synthesis and Study of Their Functional Activity by Radioligand-Receptor Binding Analysis. Int. J. Mol. Sci. 2023, 24, 10293. https://doi.org/10.3390/ijms241210293

AMA Style

Golubeva EA, Lavrov MI, Veremeeva PN, Vyunova TV, Shevchenko KV, Topchiy MA, Asachenko AF, Palyulin VA. New Allosteric Modulators of AMPA Receptors: Synthesis and Study of Their Functional Activity by Radioligand-Receptor Binding Analysis. International Journal of Molecular Sciences. 2023; 24(12):10293. https://doi.org/10.3390/ijms241210293

Chicago/Turabian Style

Golubeva, Elena A., Mstislav I. Lavrov, Polina N. Veremeeva, Tatiana V. Vyunova, Konstantin V. Shevchenko, Maxim A. Topchiy, Andrey F. Asachenko, and Vladimir A. Palyulin. 2023. "New Allosteric Modulators of AMPA Receptors: Synthesis and Study of Their Functional Activity by Radioligand-Receptor Binding Analysis" International Journal of Molecular Sciences 24, no. 12: 10293. https://doi.org/10.3390/ijms241210293

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop