Next Article in Journal
Targeting SOX18 Transcription Factor Activity by Small-Molecule Inhibitor Sm4 in Non-Small Lung Cancer Cell Lines
Next Article in Special Issue
Synthesis and Properties of (1R(S),5R(S),7R(S),8R(S))-1,8-Bis(hydroxymethyl)-6-azadispiro[4.1.4.2]tridecane-6-oxyl: Reduction-Resistant Spin Labels with High Spin Relaxation Times
Previous Article in Journal
Targeting Endothelial Necroptosis Disrupts Profibrotic Endothelial–Hepatic Stellate Cells Crosstalk to Alleviate Liver Fibrosis in Nonalcoholic Steatohepatitis
Previous Article in Special Issue
3-[N,N-Bis(sulfonyl)amino]isoxazolines with Spiro-Annulated or 1,2-Annulated Cyclooctane Rings Inhibit Reproduction of Tick-Borne Encephalitis, Yellow Fever, and West Nile Viruses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

N-N(+) Bond-Forming Intramolecular Cyclization of O-Tosyloxy β-Aminopropioamidoximes and Ion Exchange Reaction for the Synthesis of 2-Aminospiropyrazolilammonium Chlorides and Hexafluorophosphates

by
Lyudmila Kayukova
1,*,
Anna Vologzhanina
2,*,
Pavel Dorovatovskii
3,
Elmira Yergaliyeva
1,
Asem Uzakova
1 and
Aidana Duisenali
1
1
JSC A. B. Bekturov Institute of Chemical Sciences, 106 Shokan Ualikhanov St., Almaty 050010, Kazakhstan
2
A. N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 28 Vavilova St., B-334, Moscow 119334, Russia
3
NRC “Kurchatov Institute”, 1 Kurchatova Pl, Moscow 123098, Russia
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(14), 11315; https://doi.org/10.3390/ijms241411315
Submission received: 31 May 2023 / Revised: 20 June 2023 / Accepted: 22 June 2023 / Published: 11 July 2023

Abstract

:
Our research area is related to the spiropyrazolinium-containingcompounds, which are insufficiently studied compared with pyrazoline-containing compounds. Nitrogen-containing azoniaspiromolecules have also been well studied. In drug design and other areas, they are a priori important structures, since rigid spirocyclic scaffolds with the reduced conformational entropy are able to organize a closely spaced area. Azoniaspirostructures are currently of wide practical interest as ionic liquids, current sources (membranes), structure-directing agents in organocatalysis, and in the synthesis of ordered ceramics. Our goal was the synthesis of 2-aminospiropyrazolilammonium chlorides and hexafluorophosphates. Our methodology is based on the tosylation of β-aminopropioamidoximes with six-membered N-heterocycles (piperidine, morpholine, thiomorpholine, and phenylpiperazine) at the β-position. 2-Aminospiropyrazolilammonium chlorides and hexafluorophosphates were obtained by the reaction of double ion substitution in the reaction of toluenesulfonates of 2-aminospiropyrazolinium compounds with an ethereal solution of HCl in ethanol and with ammonium hexafluorophosphate in ethanol in quantitative yields of 5597%. The physicochemical characteristics of the synthesized compounds and their IR and NMR spectra are presented. The obtained salts were additionally characterized by the single-crystal XRD analysis. The presence of both axial and equatorial conformations of spirocations in solids was confirmed. 2-Aminospiropyrazolilammonium chlorides and hexafluorophosphates have weak in vitro antimicrobial activity on Gram-positive and Gram-negative bacterial lines.

1. Introduction

Our research area concerns spiropyrazolinium-containing compounds, which are insufficiently studied as compared with other azoniaspiroalkanes (Scheme 1). The latter possess numerous practical applications, which are described below. At the same time, pyrazoline derivatives, which belong to the family of nitrogen-containing heterocycles, have been the subject of a global diphenylpyrazoline market report by Market Strides (a global aggregator and publisher of market research reports). The diphenylpyrazoline market utilizes both pharmaceutical and industrial applications. The pharmaceutical industry is the largest end-user of diphenyl pyrazoline due to its various biological activities, including antioxidant, anti-inflammatory, anti-cancer, and anti-diabetic properties, as well as its ability to treat cardiovascular disorders. The industrial market covers textiles, detergents, paper production, cosmetics, and plastics. In addition, diphenyl pyrazoline is also used in the agrochemical industry as a pesticide and an insecticide [1].
It should be taken into account that the related family, azoniaspirocompounds, is a developed class of compounds with long-established synthetic methods and wide application areas. Thus, azoniaspiro compounds can be synthesized by reaction of the α,ω-dibromoalkane with a cyclic amine (e.g., 1,5-dibromopentane and piperidine to obtain 6-azonia-spiro-[5,5]-undecane [2]. Chiral quaternary N-spiroammonium bromides with 3′,4′-dihydro-1′H-spiro[isoindoline-2,2′-isoquinoline]skeleton were synthesized by reacting amino alcohols with the corresponding dibromide in acetonitrile or dichloromethane in the presence of DIPEA as a base at r.t. [3]. The spirocyclic ammonium derivatives contain a positively charged quaternary nitrogen atom, hence the fact that they can be isolated and characterized only in the form of salt [4]. Among spirocyclic nitrogen-containing compounds, small azoniaspiromolecules a priori represent important structures in drug design, since rigid spirocyclic scaffolds with reduced conformational entropy are able to selectively interact with active sites of pathogens and are expected to be selective drugs when bound to a target protein [5].
In this regard, the introduction mainly summarizes data on azoniaspirocompounds with parallel use of known data on spiropyrazolinium compounds, which may shed light on the possible areas of application of the 2-aminospiropyrazolilammoniums that we are developing.

1.1. Nitrogen-Containing Spiro-Heterocycles as Active Pharmaceutical Ingredients

Nitrogen-containing heterocyclic fragments provide biological activity to drug-like molecules. Most of the drugs on the pharmaceutical market include nitrogen-containing heterocyclic moieties [6]. The known nitrogen-containing spiro-heterocycles that are used in medical practice [7,8,9,10] include fenspiride, which is an anti-inflammatory, antiallergic, and antispasmodic agent, used for acute respiratory infections in most cases for children [7], and irbesartan, which is a selective angiotensin II receptor (type AT1) antagonist that reduces the concentration of aldosterone in the blood plasma, lowers peripheral vascular resistance, and lowers systemic arterial blood pressure (BP) [8]. Fluspirilene is an active antipsychotic agent with a pronounced antipsychotic effect, effective for hallucinations, delusions, and autism. It also calms emotional and psychomotor agitation [9]. Trospium is is a non-selective muscarinic receptor antagonist, which is used to treat the symptoms of an overactive bladder [10].
The results of the antibacterial activity study of chiral quaternary N-spiro ammonium bromides with 3′,4′-dihydro-1′H-spiro[isoindoline-2,2′-isoquinoline]skeleton showed that the Gram-negative bacteria were more susceptible towards the tested compounds than the Gram-positive bacteria. Their MICs on Gram-negative bacteria typically range from 12 to 100 mg/L, whilst those for Gram-positive bacteria vary from 50 to 200 mg/L [3].
As we have found, spiropyrazolilammonium compounds with the azoniaspiro moiety (Scheme 2) exhibit high antitubercular and antidiabetic activity, which is nearly as large as those for the reference drugs rifampicin and acarbose [11,12].

1.2. Azoniaspirocompounds as a Renewable Energy Sources

The stability of anion-exchange membranes (AEMs) for renewable energy sources and the search for durable materials for them is an important issue, and research is being conducted in this direction. Azoniaspirocompounds have been a popular subject of research that is devoted to components of anion exchange membranes [14,15,16,17]. In particular, 6-azonia-spiro[5.5]undecane exhibits a much longer half-life (87.3 h and 110 h) compared to the half-lives of tetramethyl ammonium cation (61.9 h) and benzyltrimethyl ammonium cation (4.2 h) under the same testing conditions (6 M NaOH, 160 °C) [16]. When studying the effect of the ring size on membrane properties, it was found that both 6-azonia-spiro[5.5]decane and 5-azonia-spiro[4.4]nonan-based AEMs exhibit excellent alkaline stability (around 95% retention of conductivity after immersion in 1 M KOH solution at 80 °C for 720 h), high conductivity (up to 85.7 mS cmat 80 °C), and feasible tensile strength and elongation at break [17].

1.3. Azoniaspirocompounds as Phase Transfer Catalysts

In organic synthesis, the necessity of carrying out reactions between water soluble and oil soluble reagents arises very often. Starks found that organic-soluble quaternary ammonium cations were suitable agents for the transport of anions from the aqueous phase to the organic phase. They act as powerful reaction accelerators [18]. Kitamura et al. reported C2-symmetric ammonium salts (Maruoka Phase Transfer Catalysts)catalyzing monoalkylation of glycine-derived Schiff bases with alkyl halides in order to synthesize 𝛼-alkyl-𝛼-amino acids under remarkably low catalyst loadings [19,20,21,22].

1.4. Azoniaspirocompounds as Structure-Directing Agents in Zeolite Synthesis

Azoniaspirocompounds have found application as structure-directing agents (SDA) in the synthesis of zeolites, which have enormous use in oil refining and petrochemistry due to their exceptional catalytic and selective adsorption properties in combination with thermal and chemical stability. Zeolites also found numerous applications in fine chemical synthesis and environmental catalysis, and they dominate in the global catalyst market [23,24].
Every year, due to many synthetic methods to obtain different structures with variations in pore size, surface area, pore volume, and physical properties, the number of new zeolite structures increases, although the overall number of known topologies stored in a database created by the International Zeolite Association is only 252 [25].
The use of organic SDAs is associated with the concept of the host-guest chemistry of organic molecules and inorganic systems during zeolite synthesis [26]. The addition of these organic molecules to the reaction mixtures provokes a particular ordering of inorganic units around them, which directs the crystallization pathway towards a unique zeolite framework. Their molecular size and shape, hydrophobicity, rigidity vs. flexibility, and hydrothermal stability all determine the structure-directing effect of these organic species [27]. One of the latest requests in material chemistry concerns attempts to transfer the asymmetric nature of organic chiral molecules used as SDA to the zeolite lattice in order to produce chiral enantioselective frameworks [28].
Organic SDAs must meet the appropriate criteria for hydrophobicity/hydrophilicity due to the C/N+ ratio, typically between 11 and 16, and this limits the number of organic SDA candidates that can be used [29].
Hydrothermal synthesis of zeolites was carried out using the 5-azoniaspiro[4,4]nonane as SDA in a fluoride-free medium. Germanosilicate zeolites have received great attention due to a number of extra-large pore systems. Phase pure IWW-type zeolite was obtained after hydrothermal treatment at 175 °C for 3 days with a Si/Ge ratio around 4.9 [30].
A number of azoniaspiro compounds, namely, 5-azoniaspiro-[4,5]-decane, 6-azoniaspiro-[5,5]-undecane, and 6-azonia-spiro-[5,6]-dodecane, have all contributed to the selective formation of pure zeolites with given topologies [31]. The extra-large-pore germanosilicates have been synthesized in basic media using a wide variety of sterically overloaded spiroazonia compounds as SDAs. The influence of the composition of the reaction mixture and the nature of the SDAs (structure, hydrophilicity/hydrophobicity balance, rigidity, and pKa) on the phase selectivity and the degree of crystallinity has been investigated [32].

1.5. Azoniaspirocompounds as Ionic Liquids

The ability of room temperature ionic liquids to melt at the room temperature often suffers from poor thermal stability, and these room temperature ionic liquids have limited utility at higher temperatures [33]. Fullyinorganic molten salts are stable at elevated temperatures, but they do have high melting points, which limit their utility at lower temperatures. It has been proposed that azoniaspiro salts may partially bridge this gap, creating a continuum of ionic liquids from room to elevated temperatures. Azoniaspiro salts have not yet been widely explored in the context of ionic liquid chemistry, though [3].
Based on the need to expand the temperature range in which ionic liquids can operate, a comparative study of the thermal stability of a series of alkylammonium chloride salts, incorporating one saturated ring of five, six, seven atoms and azoniaspirocompounds with two saturated rings of five, six, seven or eight atoms, was undertaken. By employing temperature-ramped thermogravimetric analysis (TGA), relative thermal stabilities of the chloride species were compared. The authors hypothesized that the inclusion of a second cyclic structure would enhance steric interactions around the ammonium nitrogen center, providing improved thermal stability of the azoniaspiroammonium salts.In fact, two-ring spirocyclic tetraalkylammonium chloride salts are, in general, more thermally stable than the single-ring analogues, exhibiting higher Tonset and Tstart values [33]. It is worth mentioning here that the hexafluorophosphate anion is a constituent part of a series of novel hexafluorophosphate salts that are based on N,N-dialkylimidazolium and substituted N-alkylpyridinium cations, which display the behavior of ionic liquids at temperatures above their melting point [34].

2. Results and Discussion

2.1. Synthesis and Spectra

In order to obtain new compounds that potentially possess useful properties as active pharmaceutical ingredients and/or for possible industrial uses, such as phase transfer catalysts, ionic liquids, structure directing agents (SDAs), and components for ion exchange membranes, a number of 2-aminopyrazolilammonium salts have been obtained—namely, chlorides and hexafluorophosphates (9–16, Scheme 3).
The synthesis of 2-aminospiropyrazolilammonium chlorides 912 at the first stage included tosylation of four β-aminopropioamidoximes, where the β-amino group contains piperidin-1-yl (1), morpholin-1-yl (2), thiomorpholin-1-yl (3), or 4 -phenylpiperizin-1-yl (4) in CHCl3 at r.t. in the presence of diisopropylethylamine for 15–20 h with TLC monitoring. A series of toluenesulfochlorides (5–8) has been described by our team previously [35,36]. The tosylation products isolated as white precipitates were recrystallized from i-PrOH in 45–65% yields.
The second stage of the salts 9–12 synthesis consists in ethereal HCl action until pH = 2 is reached on tosylates 5–8 ethanol solutions. The reaction of exchange of tosylate anion for chloride anion proceeds instantaneously—after dropping an ether solution of HCl to the tosylates 5–8 ethanol solution, white flakes of chlorides 912 are immediately formed, and after recrystallization from i-PrOH, their yields were 55–97% (Scheme 3, Table 1).
It is worth noting that compounds 11 and 12, as shown by the X-ray diffraction data, crystallize as hydrates, and salt 12 contains the 2-aminospiropyrazolilammonium dication (Scheme 4, Section 2.3):
Using a shortened synthetic route, the formation of hexafluorophosphates 13–16, without isolating the intermediate 2-aminospiropyrazolilammonium tosylates 5–8, was carried out. In a one-pot reaction, a mixture of para-toluolsulfochloride, one of the substrates of β-aminopropioamidoximes (1–4), the Bu3N base, and ammonium hexafluorophosphate in CHCl3 solution (Scheme 5) were allowed to react. The reaction time to obtain products 1316 in a 55–90% quantity after recrystallization from i-PrOH was 24 h.
To our knowledge, information about 1,5-diazaspiro-1-en-5-ium spiropyrazolinium salts is limited by our publications. However, when using the tosylation reaction, stable indazole derivatives previously obtained—1,1–disubstituted indazolium hexafluorophosphates, which have a nitrogen atom at the head of the bridge [37].
It is assumed that 2aminospiropyrazolilammoniumtosylates, which are more thermodynamically favorable compared to O-sulfochlorination products, are first formed in the reaction mixture [38]. In this case, the intramolecular reaction of amination of intermediate products of O-tosylation of β-aminopropioamidoximes (Scheme 5, A) with nucleophilic attack of the lone electron pair of the amine nitrogen atom on the imine nitrogen atom of the amidoxime group occurs, followed by the elimination of the good-leaving tosylate group, and this leads to tosylates 58.
The oxime N-O bond break is accompanied by the N-N(+) bond formation. As a result of the exchange reaction of the tosylate group for the hexafluorophosphate anion in the reaction mixture 2-aminospiropyrazolilammonium hexafluorophosphates 1316 are formed. It is noteworthy that, as can be seen from Table 1, the melting points of a number of chlorides 912 are significantly higher than those of a number of corresponding hexafluorophosphates 1316(Δtm.p. 9 and 13: ~97 °C; Δtm.p. 10 and 14: ~78 °C;Δtm.p. 11 and 15: ~108 °C; Δtm.p. 12 and 16: ~48 °C).
It must be noted that compounds 10 and 11 were obtained earlier. First, 2-amino-8-oxa-1,5-diazaspiro[4.5]dec-1-en-5-ammonium chloride monohydrate (10·H2O) was obtained by the acid hydrolysis of 5-substituted phenyl-3-[β-(morpholin-1-yl)ethyl]-1,2,4-oxadiazoles, and it has an X-ray diffraction description (the structure was registered at the Cambridge Crystallographic Database under the number CCDC 2049801 [11]). Herein, we present the crystal structure of pure anhydrous 10, which has not yet been published. Second, 2-amino-8-thia-1,5-diazaspiro[4.5]dec-1-en-5-ammonium chloride (11) was precipitated for the first time by acid hydrolysis of 5-substituted phenyl-3-[β-(thiomorpholine-1-yl)ethyl]-1,2,4-oxadiazoles [39], and then through the reaction of double substitution of ions by the reaction of 2-amino-8-thia-1,5-diazaspiro[4.5]dec-1-en-5-ammonium toluenesulfonate with DIPEA hydrochloride [12]. Its structure was deposited to the Cambridge Crystallographic Databaseunder the number CCDC 2106798.
Thus, crystal structures 9, 10, and 1216 are new. In the IR spectra of 2-aminospiropyrazolilammonium salts (916), there are characteristic bands of stretching vibrations. The stretching vibrations of the NH2 group occur in the region of 3342–3500 cm−1, with those of bonds C=N – at 1642–1659 cm−1, and the hexafluorophosphate group has a characteristic stretching vibration band νP–F in the region of 836–839 cm−1.
In the 1H-NMR spectra of compounds 916, protons NH2 are in the field δ 7.21–7.50 ppm. The protons of the α- and β-methylene groups of compounds 916 give triplet signals in the regions, respectively, of δ 3.04–3.17 ppm and δ 3.70–3.95 ppm.
The diastereotopic nature of the geminal protons of the methylene groups that are located at the ammonium nitrogen atom manifests itself in the 1H-NMR spectra of compounds 9 and 1216 as pairs of multiplet signals with an intensity of two protons at δ: 3.35m (ax), 3.44m (eq) (9); 3.52m (ax), 3.98m (eq) (12); 3.38m (ax), 3.46m (eq) (13); 3.41m (ax), 3.65m (eq) (14); 3.64m (ax), 3.87m (eq) (15); and 3.49m (ax), 3.98m (eq) (16). This can be interpreted as the effect of the slow rotation of β-heterocycles, which allows the equatorial and axial protons to be noticed.
The signals of Csp3 and Csp2 carbon atoms in the 13C-NMR spectra of compounds 916 are also characteristic. Thus, the signals of carbon atoms of the C=N bond of the pyrazolinium ring in compounds 916 were recorded in the range δ 168.2–170.1 ppm, and signals of aromatic carbon atoms Csp2 of compounds 12 and 16 were found in the range δ 115.3–156.7 ppm.
The carbon atoms of the α- and β-methylene groups of compounds 9, 10, 1214, and 16 give resonance in the ranges of δ 31.5–32.5 ppm and 44.5–62.4 ppm, respectively, and the α-methylene group signal of compound 15 is due to the electron-donating nature of sulfur in the heterocycle at 23.1 ppm. For this reason, there is also a higher field position of the signals of methylene groups carbon atoms located at the sulfur atom of the thiomorpholine ring in compound 15 at δ 31.3 ppm compared with the position of the 13C NMR signals of methylene groups at the oxygen atom in compound 10 (60.9 ppm) andcompound 14 (62.2 ppm) as well as at the nitrogen atom in the compounds 12 (44.5 ppm) and 16 (45.7 ppm).
Carbon atoms of methylene groups at the ammonium spirocyclic nitrogen atom of compounds 9,10, and 1116 are presented in the range δ 62.9–64.7 ppm, and carbon atoms of the (CH2)3 group of piperidine derivatives 9 and 13 have signals at δ 21.0 and 21.9 ppm and 20.5 and 21.5 ppm.
It should be pointed out that the 13C NMR spectra of compounds 1316 have a characteristic feature—in the region of δ 38.4–39.5 ppm, a multiplet signal appears due to the presence of fluorine in the molecules.

2.2. The In Vitro Antibacterial Activity of 2-Amino-1,5-diazaspiro[4.5]dec-1-en-5-ammoniumchlorides (912) and Hexafluorophosphates (1316)

The antimicrobial activity of the samples of 2-amino-1,5-diazaspiro[4.5]dec-1-en-5-ammonium chlorides (912) and hexafluorophosphates (1316) was investigated against Gram-positive bacteria Staphylococcus aureus and Bacillus subtilis as well as Gram-negative Escherichia coli, Pseudomonas aeruginosa strain, and to the yeast fungus Candida albicans by diffusion into agar (wells).
The reference drugs were gentamicin for bacteria and nystatin for the yeast fungus Candida albicans. Table 2 contains data about the antimicrobial activity of the studied samples.
We discovered that samples 11, 13, and 15 exhibited weak antimicrobial activity against the Gram-positive Staphylococcus aureus cell line (11.2–13.7 mm), and that gentamicin is active on this strain at 24.5 mm. The samples of compounds 9, 11, 12, 14, and 15 also showed weak antimicrobial activity against the Gram-negative Escherichia coli strain (11.3–13.5 mm). For comparison, the drug gentamicin has high activity, and it inhibits bacteria on a disk with a diameter of 26.4 mm. At the same time, sample 11 also shows weak antibacterial activity against the Gram-negative test Pseudomonas aeruginosa strain (11.5 mm) (gentamicin—20.2 mm). None of the compounds showed antifungal activity against Candida albicans.
The remaining test samples do not show antimicrobial activity against the presented test strains of microorganisms.

2.3. X-ray Diffraction

Single-crystal XRD study of salts 9, 10, and 1216 obtained from the reaction mixtures confirmed the formation of spiropyrazolinium salts in all of the cases. Salts (C7H14N3S)Cl·H2O (11) and (C13H20N4)Cl2·H2O (12) crystallized as hydrates. Asymmetric units of other salts (9, 13–16) contained only one cation and one anion. Salt (C13H20N4)Cl2·H2O (12) is the first representative of spiropyrazolinium dications characterized by X-ray diffraction, and it contains a protonated nitrogen atom in position 8, which was clearly seen on the residual density map. Asymmetric units of all of the known salts are depicted in Figure 1.
In all of the solids, the five-membered ring is found in the envelope conformation with the C(1) atom situated 0.05(1)–0.39(1) Å above the mean plane formed by four other atoms. The six-membered heterocycle realized the chair conformation. However, overall spiropyrazolinium conformations are different. In particular, the N(2) atom of the five-membered pyrazolinium cycle in chloride 9 and hexafluorophosphates 1316 realized more thermodynamically stable axial conformation in respect to the six-membered cycle, while in the other solids, it realizes the equatorial conformation. The phenyl ring in 12 is nearly coplanar with the N(1)-N(2)-C(3)=C(4) plane; the angle between them is 1.82(6)° only. For the two symmetrically independent cations in 16,the corresponding angle values are equal to 47.42(8) and 74.39(8)°.
Crystal structures of chlorides and hexafluorophosphates indicate the low likelihood of H-bond formation with watermolecules, as proposed by Delori et al. [40]. The likelihood of hydrogen-bond formation was estimated using the Mercury package [41]), as described by Galek et al. [42]) and Vologzhanina [43]). In Table 3, data for a pure 2-aminospiropyrazolinium chlorides and hexafluorophosphates are compared for anhydrous salts and their hydrates. The N–H…Hal bond is the most expected one for all cases. Its propensity is much higher than that for the N–H…O(water) bond, which is in accordance with rare occurrence of hydrates for this family of salts. However, the high propensity of O–H…Hal bonds demonstrates that water molecules can act as linkers between anions.
Experimentally observed H-bonded associates are in accord with results of these calculations. All anions take part in H-bonding in order to form the associates depicted in Figure 2. H-bonded tetramers were found in salts 9, 15, and 16. Infinite chains were observed in solid 10, 12, 14, and 16. A total of 13 infinite layers were found, which realize the square-lattice topology of their underlying net. We should note that 16 contains two types of associates. For all cases, the amino group takes part in two hydrogen bonds, and chloride anions act as acceptors of one (9), two (10), or three (12) H-bonds. In addition to H-bonds, halogen atoms take part in weak C-H…Hal interactions and halogen bonds.

3. Materials and Methods

3.1. Synthesis

The reagents were purchased from different chemical suppliers and were purified before use. FTIR spectra were obtained on a Thermo Scientific Nicolet 5700 FTIR instrument (Thermo Fisher Scientific, Inc., Waltham, MA, USA) in KBr pellets. The 1H- and 13C NMR spectra were recorded in DMSO-d6.The 1H- and 13CNMR spectra of compounds 916 were acquired on a Bruker Avance III 500 MHz NMR spectrometer (Bruker, BioSpin GMBH, Rheinstetten, Germany). The signals of the residual undeuterated solvents were used as a reference for the 1H-NMR (2.50 ppm) and 13C-NMR (39.5 ppm) spectra.
Elemental analysis was carried out on a CE440 elemental analyzer (Exeter Analytical, Inc., Shanghai, China). The melting points were determined in glass capillaries on a PTP(M) apparatus (Khimlabpribor, Klin, Russia). The reaction progress and purity of the obtained products were controlled using Sorbfifil (Sorbpolymer, Krasnodar, Russia) TLC plates coated with CTX-1A silica gel, with grain size 5–17 µm, containing UV-254 indicator. The eluent for TLC analysis was a mixture of benzene–EtOH, 1:3. The solvents for the synthesis, recrystallization, and TLC analysis (ethanol, 2-PrOH, benzene, DMF, and acetone) were purified according to the standard techniques.

3.1.1. General Procedure for the Preparation of 2-Amino-1,5-diazaspiro[4.5]dec-1-en-5-ammonium Chlorides (9, 10, 12)

Ethereal HCl was added dropwise until pH = 2 was reached to a solution of 0.0015 mol of tosylates 5, 8 in 10 mL of anhydrous ethanol. Then, 20 mL of absolute ether was poured. The resulting white precipitates of chlorides 9, 10, 12 were obtained after recrystallization from i-PrOH in 68 and 70% yields.
2-Amino-1,5-diazaspiro[4.5]dec-1-en-5-ium chloride (9). To a solution of 0.49 g (0.0015 mol) of 2-amino-1,5-diazaspiro[4.5]dec-1-en-5-ium 4-methylbenzenesulfonate (5) in 10 mL of ethanol ethereal HCl was added until pH = 2 was reached, and then 20 mL of absolute ether was poured. After recrystallization of the resulting white precipitate from i-PrOH, we obtained 0.19 g (68%) of chloride 9, m.p. 288–290 °C, Rf 0.01. IR (KBr, ν, cm–1): 1663 (C=N); 2848 (Csp3−H); 3341, 3400 (N–H)2. 1H-NMR (500 MHz, DMSO-d6): 1.59 m, 1.80 m, 1.81 m, [6H, (CH2)3], 3.12 (t, J = 7.0 Hz, 2H, α-CH2), 3.80 (t, J = 7.0 Hz, 2H, β-CH2), 3.35 [m, 2H, N(+)(CHax)2] and 3.44 [m, 2H, N(+)(CHeq)2], 7.24 (s, 2H, NH2). 13C-NMR (126 MHz, DMSO-d6): 21.0, 21.9, 31.5, 60.7, 64.3, 168.5. Anal. Calcd for C8H16ClN3, %: C, 50.66; H, 8.50. Found, %: C, 50.27; H, 8.96.
Amino-8-oxa-1,5-diazaspiro[4.5]dec-1-en-5-ium chloride (10). To a solution of 0.49 g (0.0015 mol) of 2-amino-8-oxa-1,5-diazaspiro[4.5]dec-1-en-5-ium 4-methylbenzenesulfonate (6) in 10 mL of ethanol ethereal HCl was added until pH = 2 was reached, and then 20 mL of absolute ether was poured. After recrystallization of the resulting white precipitate from i-PrOH, we obtained 0.19 g (68%) of chloride 10, m.p. 271 °C, Rf 0.01. IR (KBr, ν, cm–1): 1648 (C=N); 2927, 2957 (Csp3−H); 3438 (N–H)2. 1H-NMR (500 MHz, DMSO-d6): 3.04 (t, J = 7.0 Hz, 2H, α-CH2), 3.60 [m, 4H, O(CH2)2], 3.92 [m, 6H, β-CH2, N(+)(CH2)2], 7.47 (s, 2H, NH2). 13C-NMR (126 MHz, DMSO-d6):32.0, 60.9, 64.5, 170. Anal. Calcd for C7H14ClN3O (191.66), %: C, 48.34; H, 8.11. Found, %: C, 48.20; H, 7.75.
Mono(2-amino-8-phenyl-1,5,8-triazaspiro[4.5]dec-1-en-5-ium) dichloride hydrate (12). To a solution of 0.6 g (0.0015 mol) 2-amino-8-phenyl-1,5,8-triazaspiro[4.5]dec-1-en-5-ium 4-methylbenzenesulfonate (8) in 10 mL of ethanol ethereal HCl was added until pH = 2 was reached, and then 20 mL of absolute ether was poured in the reaction mixture. After recrystallization of the resulting white precipitate from i-PrOH, we obtained 0.30 g (70%) of chloride 12; m.p. 277–280 °C, Rf 0.01. IR (KBr, ν, cm–1): 1651 (C=N), 1600 (C=C), 2839, 2981, 3002 (Csp3−H), 3303 (Csp2−H), 3308, 3400 [N(-H)2]. 1H-NMR (500 MHz, DMSO-d6): 3.17 (t, J = 7.0 Hz, 2H, α-CH2), 3.95 (t, J = 7.0 Hz, 2H, β-CH2), 3.58 [m, 4H, N(CH2)2], 3.52 [m, 2H, N(+)(CHax)2] and 3.98 [m, 2H, N(+)(CHeq)2], 7.41 (s, 2H, NH2), 7.81−8.53 (m, 5H, C(sp2)H). 13C-NMR (126 MHz, DMSO-d6): 31.5, 44.5, 61.5, 62.9, 115.3, 120.4, 123.8, 156.7, 168.6. Anal. Calcd for C13H22N4Cl2O (320,24): C, 48.76; H, 6.61. Found: C, 48.63; H, 6.36.

3.1.2. General Procedure for the Preparation of 2-Amino-1,5-diazaspiro[4.5]dec-1-ene-5-ammonium Hexafluorophosphates (1316)

To a solution of 0.0020 mol of β-aminopropioamidoximes 14 in 20 mL of CH2Cl2 0.0020 mol of Bu3N in 5 mL of CH2Cl2 and 0.0020 mol of ammonium hexafluorophosphate were added, one after the other. At r.t., with stirring, a solution of 0.0020 mol of para-toluenesulfonyl chloride in 10 mL of CH2Cl2 was dripped. The reaction mixture was stirred for 24 h at r.t. The course of the reaction was monitored by TLC. The formed white precipitates of hexafluorophosphates 1316 after recrystallization from i-PrOH were isolated in a 55–90% yield.
2-Amino-1,5-diazaspiro[4.5]dec-1-en-5-ium hexafluorophosphate(13). To a solution of 0.34 g (0.002 mol) of β-(piperidin-1-yl)propioamidoxime(1) in 20 mL of CH2Cl2 0.37 g (0.002 mol) of Bu3N in 5 mL of CH2Cl2 and 0.33 g (0.002 mol) of NH4F6 all of these elements were added, one after the other. Following stirring for 0.5 h, 0.38 g (0.002 mol) of para-touenesulfochloride in 10 mL of CH2Cl2 was dripped. The reaction mixture was stirred for 24 h at r.t. The progress of the reaction was monitored by TLC. After recrystallization of the resulting white precipitate from i-PrOH, 0.33 g (55%) of transparent crystals of hexafluorophosphate 13 were obtained, m.p.192–193 °C, Rf 0.05. IR (KBr, ν, cm–1): 837 (P-F), 1651 (C=N), 2949(Csp3−H), 3399, 3503 [N(-H)2]. 1H-NMR(500 MHz, DMSO-d6): 1.61 m, 1.75 m, 1.89 m, [6H, (CH2)3], 3.14(t, J = 7.0 Hz, 2H, α-CH2), 3.38 [m, 2H, N(+)(CHax)2] and 3.46 [m, 2H, N(+)(CHeq)2], 3.76(t, J = 7.0 Hz, 2H, β-CH2), 7.21(s, 2H, NH2). 13C-NMR (126 MHz, DMSO-d6): 20.5, 21.5, 31.11, 38.9 (m, F), 60.3, 64.1, 168.2.Anal. Calcd for C8H16F6N3P (299,20) C, 32.11; H, 5.39.Found: C, 32.63; H, 5.45.
2-Amino-8-oxa-1,5-diazaspiro[4.5]dec-1-en-5-ium hexafluorophosphate (14). To a solution of 0.35 g (0.002 mol) of β-(morpholin-1-yl)propioamidoxime(2) in 20 mL of CH2Cl2 0.37 g (0.002 mol) of Bu3N in 5 mL of CH2Cl2 and 0.33 g (0.002 mol) of NH4F6 all of these elements were added, one after the other. Following stirring for 0.5 h, 0.38 g (0.002 mol) of para-touenesulfochloride in 10 mL of CH2Cl2 was added. The reaction mixture was stirred for 24 h at r.t.The progress of the reaction was monitored by TLC. After recrystallization of the resulting white precipitate from i-PrOH, 0.54 g (90%) of transparent crystals of hexafluorophosphate 14 were obtained, m.p.191–193 °C, Rf 0.08. IR (KBr, ν, cm–1): 839 (P-F), 1119 (C-O), 1645 (C=N),2957(Csp3−H), 3399, 3505 [N(-H)2]. 1H-NMR(500 MHz, DMSO-d6): 3.14(t, J = 7.0 Hz, 2H, α-CH2), 3.76(t, J = 7.0 Hz, 2H, β-CH2), 3.38 [m, 4H, O(CH2)2], 3.41 [m, 2H, N(+)(CHax)2] and 3.65[m, 2H, N(+)(CHeq)2], 7.30 (s, 2H, NH2). 13C-NMR (126 MHz, DMSO-d6): 31.3, 38.4 (m, F), 62.2, 63.2, 169.0.Anal. Calcd for C7H14F6N3OP (301,17) C, 27.92; H, 4.69.Found: C, 28.15; H, 4.96.
Amino-8-thia-1,5-diazaspiro[4.5]dec-1-en-5-ium hexafluorophosphate (15). To a solution of 0.38 g (0.002 mol) of β-(thiomorpholin-1-yl)propioamidoxime(3) in 20 mL of CH2Cl20.37 g (0.002 mol) of Bu3N in 5 mL of CH2Cl2 and 0.33 g (0.002 mol) of NH4F6 all of these elements were added, one after the other. Following stirring for 0.5 h, 0.38 g (0.002 mol) of para-touenesulfochloride in 10 mL of CH2Cl2 was added. The reaction mixture was then stirred for 24 h at r.t. The progress of the reaction was monitored by TLC. After recrystallization of the resulting white precipitate from i-PrOH, 0.64 g (85%) of transparent crystals of hexafluorophosphate 15 were obtained, m.p.180–182 °C, Rf 0.01. IR (KBr, ν, cm–1): 841 (P-F),1221 (C-S), 1647 (C=N), 3189 (Csp3−H), 3397, 3503 [N(-H)2]. 1H-NMR(500 MHz, DMSO-d6): 3.10 (t, J = 7.0 Hz, 2H, α-CH2), 3.70(t, J = 7.0 Hz, 2H, β-CH2), 2.97 [m, 2H, S(CHeq)2] and 3.11 [m, 2H, SCH(ax)2], 3.64 [m, 2H, N(+)(CHeq)2] and 3.87 [m, 2H, N(+)(CHax)2], 7.40 (s, 2H, NH2). 13C-NMR (126 MHz, DMSO-d6): 23.1, 31.3, 39.1 (m, F), 62.4, 64.7, 168.9.Anal. Calcd for C7H14F6N3PS (317,23) C, 26.50; H, 4.45.Found: C, 26.91; H, 4.25.
2-Amino-8-phenyl-1,5,8-triazaspiro[4.5]dec-1-en-5-ium hexafluorophosphate (16). To a solution of 0.50 g (0.002 mol) of β-(4-phenylpiperazin-1-yl)propioamidoxime(4) in 20 mL of CH2Cl2 0.37 g (0.002 mol) of Bu3N in 5 mL of CH2Cl2 and 0.33 g (0.002 mol) of NH4F6 for 0.5 h, 0.38 g (0.002 mol) of para-toluenesulfochloride in 10 mL of CH2Cl2 was added. The reaction mixture was then stirred for 24 h at r.t. The progress of the reaction was monitored by TLC. After recrystallization of the resulting white precipitate from i-PrOH, 0.64 g (85%) of transparent crystals of hexafluorophosphate (16) were obtained, m.p. 240–242 °C, Rf 0.01. IR (KBr, ν, cm–1): 836 (P-F), 1250 (C-N), (1597 (C=C), 1642 (C=N),2863 (Csp3−H), 3313 (Csp2−H), 3411, 3515 [N(-H)2]. 1H-NMR (500 MHz, DMSO-d6): 3.17 (t, J = 7.0 Hz,2H, α-CH2), 3.95(t, J = 7.0 Hz, 2H, β-CH2), 3.56 [m, 4H, N(CH2)2], 3.49 [m, 2H, N(+)(CHax)2] and 3.98[m, 2H, N(+)(CHeq)2], 7.29 (s, 2H, NH2), 6.91−7.13 (m, 5H, C(sp2)H). 13C-NMR (126 MHz, DMSO-d6): 32.5, 39.5 (m, F), 45.7, 62.6, 64.10, 117.5, 121.7, 130.7, 151.0, 170.1. Anal. Calcd for C13H19F6N4P (376,29): C, 41.50; H, 5.09. Found, %: C, 41.92; H, 5.51.

3.2. In Vitro Evaluation of Antimicrobial and Antifungal Activity of the Compounds 916

In vitro antimicrobial activity of eight samples 916 against Staphylococcus aureus, Bacillus subtilis strains of Gram-positive bacteria, Escherichia coli and Pseudomonas aeruginosa of Gram-negative strains, and Candida albicans yeast fungus was carried outby the agar diffusion method (wells). The reference drugs were gentamicin for bacteria and nystatin for the yeast fungus Candida albicans [44].
The cultures were grown in a liquid medium at pH 7.3 ± 0.2 at a temperature of 30 to 35 °C for 18–20 h. The cultures were diluted 1:1000 in a sterile 0.9% isotonic sodium chloride solution, 1 mLin cups with appropriate elective, nutrient media for the studied test strains, and inoculated according to the «solid lawn» method. After drying, wells 6.0 mm in size were formed on the agar surface, into which solutions of the studied samples, gentamicin, and nystatin were all added. In the control, distilled water was used in equivolume amounts.
The studied samples were tested in the amount of 1 μg. The inoculations were incubated at 37°C, and the growing cultures were counted after 24 h. The antimicrobial activity of the samples was assessed by the diameter of the growth inhibition zones of the test strains (mm). The diameter of growth inhibition zones less than 10 mm and continuous growth in the cup was assessed as the absence of antibacterial activity, 10–15 mm being weak activity, 15–20 mm being moderately pronounced activity, and more than 20 mm being pronounced activity. Each sample was tested in three parallel experiments. Statistical processing was carried out by parametric statistics methods with the calculation of the arithmetic mean and standard error.

3.3. X-ray Diffraction

Single crystals precipitated from reaction mixtures. Intensity data for 15 were measured at 100.0(2) K using a 1-axis MarDTB goniometerequipped with Rayonix SX165 CCD detector at the “Belok/XSA” beamline of the Kurchatov Synchrotron Radiation Source [45,46]. The direct geometry and φ-scanning with the detector plane perpendicular to the beam (λ = 0.745 Å) were used. The data were indexed and integrated using the XDS, ver. 2023 software [47]. The intensities of reflections for the other samples were measured with a Bruker Apex II DUO CCD diffractometer (λ(MoKα) = 0.71073Å, graphite monochromator) at 100.0(2) K (12·H2O), 140.0(2) K (9, 10, 13, 14), or 295.0(2) K (16). The structures were solved using the SHELXT program [48] and refined against F2 using SHELXL-2018 [49] and OLEX2 [50] program packages. Non-hydrogen atoms were refined in an anisotropic approximation. Positions of H(C) atoms were calculated and H(O) atoms were localized on difference Fourier maps. Hydrogen atoms were refined isotropically.The riding model was applied with Uiso(H) = 1.5Ueq(O), 1.2Ueq(N) and 1.2Ueq(C). Experimental details and crystal parameters are listed in Tables S1 and S2. Crystallographic information files are available from the Cambridge Crystallographic Data Center upon request (http://www.ccdc.cam.ac.uk/structures, accessed on 20 May 2023).

4. Conclusions

Synthetic paths to 2-aminospiropyrazolilammonium chlorides and hexafluorophosphates have been attested. Target compounds were obtained with good yield and characterized by physical-chemical and spectral methods. XRD data demonstrated that the compounds are able to form different conformations, and that they readily take part in H-bonding through the amino-group. Hexafluorophosphates are characterized by their relatively low melting temperatures, which makes them potentially active as ionic liquid salts. The in vitro screening of the salts also revealed weak antimicrobial activity against the Gram-positive Staphylococcus aureus strain for 11, 13, and 15, and against the Gram-negative test Escherichia coli strainfor salts 9, 11, 12, 14, and 15.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241411315/s1.

Author Contributions

Conceptualization, L.K. and A.V.; methodology, A.U.; software, A.V.; validation, L.K., A.V. and A.U.; formal analysis, A.D. and E.Y.; investigation, A.U., A.D., E.Y. and P.D.; resources, A.D.; data curation, L.K.; writing—original draft preparation, L.K., A.V.; writing—review and editing, L.K. and A.V.; visualization, L.K. and A.V.; supervision, L.K.; project administration, L.K.; funding acquisition, L.K. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Committee of Science of the Ministry of Science & High Education of the Republic of Kazakhstan (grant IRN AP14870011).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article.

Acknowledgments

L.K. acknowledges to Z.T. Shulgau for conducting of antibacterial screening at the RSE National Center for Biotechnology, Science Committee of MS&HE RK. A.V. is grateful for support from the Ministry of Science and Higher Education of the Russian Federation (Agreement No. 075-03-2023-642).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Diphenyl Pyrazoline Market–Global Outlook and Forecast 2023–2030. 86p. Available online: https://www.24marketreports.com/chemicals-and-materials/global-diphenyl-pyrazoline-2022-553 (accessed on 13 April 2023).
  2. Millini, R.; Perego, C.; Frigerio, F.; Carluccio, L.; Bellussi, G.L. Azonia-spiro compounds as structure directing agents: A computation study. In Studies in Surface Science and Catalysis; van Steen, E., Claeys, I.M., Callanan, L.H., Eds.; Elsevier Science: Amsterdam, The Netherlands, 2004; Volume 154, pp. 275–282. [Google Scholar] [CrossRef]
  3. Bielawski, K.; Leszczyńska, K.; Kałuża, Z.; Bielawska, A.; Michalak, O.; Daniluk, T.; Staszewska-Krajewska, O.; Czajkowska, A.; Pawłowska, N.; Gornowicz, A. Synthesis and antimicrobial activity of chiral quaternary N-spiroammonium bromides with 3′,4′-dihydro-1′H-spiro[isoindoline-2,2′-isoquinoline] skeleton. Drug Des. Devel. Ther. 2017, 11, 2015–2028. [Google Scholar] [CrossRef] [PubMed]
  4. Cuevas-Yañez, E.; BasavanagUnnamatla, M.V.; García-Eleno, M.A.; Chakroborty, S. 12.20-Systems with a Spirocyclic Heteroatom. In Comprehensive Heterocyclic Chemistry IV; Elsevier Science: Amsterdam, The Netherlands, 2022; Volume 12, pp. 621–632. [Google Scholar] [CrossRef]
  5. Zheng, Y.; Tice, C.M.; Singh, S.B. The use of spirocyclic scaffolds in drug discovery. Bioorg. Med. Chem. Lett. 2014, 24, 3673–3682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Kerru, N.; Gummidi, L.; Maddila, S.; Gangu, K.K.; Jonnalagadda, S.B. A Review on Recent Advances in Nitrogen-Containing Molecules and Their Biological Applications. Molecules 2020, 25, 1909. [Google Scholar] [CrossRef] [Green Version]
  7. Kolosova, N.G.; Shakhnazarova, M.D. Fenpirideuse in acute respiratory infections in children. Meditsinsky Sovet 2019, 2, 120–123. [Google Scholar] [CrossRef]
  8. Forni, V.; Wuerzner, G.; Pruijm, M.; Burnier, M. Long-term use and tolerability of irbesartan for control of hypertension. Integr. Blood Press. Control 2011, 4, 17–26. [Google Scholar] [CrossRef] [Green Version]
  9. Vardanyan, R.S. 6-Antipsychotics (Neuroleptics). In Synthesis of Essential Drugs; Vardanyan, R.S., Hruby, V.J., Eds.; Elsevier Science: Amsterdam, The Netherlands, 2006; pp. 83–101. [Google Scholar] [CrossRef]
  10. Wuest, M. xPharm: The Comprehensive Pharmacology Reference; Enna, S.J., Bylund, D.B., Eds.; Elsevier Science: Amsterdam, The Netherlands, 2008; pp. 1–6. [Google Scholar] [CrossRef]
  11. Kayukova, L.; Vologzhanina, A.; Praliyev, K.; Dyusembaeva, G.; Baitursynova, G.; Uzakova, A.; Bismilda, V.; Chingissova, L.; Akatan, K. Boulton-Katritzky Rearrangement of 5-Substituted Phenyl-3-[2-(morpholin-1-yl) ethyl]-1, 2, 4-oxadiazoles as a Synthetic Path to Spiropyrazoline Benzoates and Chloride with Antitubercular Properties. Molecules 2021, 26, 967–982. [Google Scholar] [CrossRef] [PubMed]
  12. Kayukova, L.; Vologzhanina, A.; Dorovatovskii, P.; Baitursynova, G.; Yergaliyeva, E.; Kurmangaliyeva, A.; Shulgau, Z.; Adekenov, S.; Shaimerdenova, Z.; Akatan, K. Reaction Products o fβ-Aminopropioamidoximes Nitrobenzenesulfochlorination:Linear and Rearranged to Spiropyrazolinium Salts with Antidiabetic Activity. Molecules 2022, 27, 2181. [Google Scholar] [CrossRef]
  13. JSC «A.B. Bekturov Institute of Chemical Sciences». The Usage of a Products of Arylsulfochlorination of Beta-Aminopropioamidoximes as a Compounds with Antidiabetic Activity. Kazakhstan Patent for Useful Model No. 6926, 25 February 2022.
  14. Varcoe, J.R.; Atanassov, P.; Dekel, D.R.; Herring, A.M.; Hickner, M.A.; Kohl, P.A.; Kucernak, A.R.; Mustain, W.E.; Nijmeijer, K.; Scott, K.; et al. Anion-exchange membranes in electrochemical energy systems. Energy Environ. Sci. 2014, 7, 3135–3191. [Google Scholar] [CrossRef] [Green Version]
  15. Marino, M.G.; Kreuer, K.D. Alkaline stability of quaternary ammonium cations for alkaline fuel cell membranes and ionic liquids. ChemSusChem 2015, 8, 513–523. [Google Scholar] [CrossRef]
  16. Xue, J.; Zhang, J.; Liu, X.; Huang, T.; Jiang, H.; Yin, Y.; Qin, Y.; Guiver, M.D. Toward alkaline-stable anion exchange membranes in fuel cells: Cycloaliphatic quaternary ammonium-based anion conductors. Electrochem. Energy Rev. 2022, 5, 348–400. [Google Scholar] [CrossRef]
  17. Badwal, S.P.; Giddey, S.S.; Munnings, C.; Bhatt, A.I.; Hollenkamp, A.F. Emerging electrochemical energy conversion and storage technologies. Front Chem. 2014, 2, 79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Starks, C.M. Phase-Transfer Catalysis I. Heterogonous Reactions Involving Anion Transfer by Quaternary Ammonium and Phosponium Salts. J. Am. Chem. Soc. 1971, 93, 195–199. [Google Scholar] [CrossRef]
  19. Kitamura, M.; Shirakawa, S.; Maruoka, K. Powerful chiral phase-transfer catalysts for the asymmetric synthesis of alphaalkyl- and alpha,alpha-dialkyl-alpha-amino acids. Angew. Chem. Int. Ed. 2005, 44, 1549–1551. [Google Scholar] [CrossRef] [PubMed]
  20. Ooi, T.; Arimura, Y.; Hiraiwa, Y.; Yuan, L.; Kano, T.; Inoue, T.; Matsumoto, J.; Maruoka, K. Highly Enantioselective Monoalkylation of p-Chlorobenzaldehyde Imine of Glycine tert-Butyl Ester under Mild Phase-Transfer Conditions. TetrahedronAsymmetry 2006, 17, 603–606. [Google Scholar] [CrossRef]
  21. Ooi, T.; Maruoka, K. Development and applications of C2-symmetric, chiral, phase-transfer catalysts. Aldrichimica Acta 2007, 40, 77–86. [Google Scholar] [CrossRef]
  22. Lee, H.-J.; Maruka, K. Recent Asymmetric Phase-Transfer Catalysis with Chiral Binaphthyl-Modified and Related Phase-Transfer Catalysts over the Last 10 Years. Chem. Rec. 2023, e202200286. [Google Scholar] [CrossRef]
  23. Přech, J.; Pizarro, P.; Serrano, D.P.; Čejka, J. From 3D to 2D zeolite catalytic materials. Chem. Soc. Rev. 2018, 47, 8263–8306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Zeolites Market Forecast, Trend Analysis & Competition Tracking—Global Market Insights. Available online: www.factmr.com (accessed on 25 April 2023).
  25. Baerlocher, C.; McCusker, L.B. Database of Zeolite Structures. Available online: https://www.izastructure.org/databases/ (accessed on 20 May 2020).
  26. Chao, S.; Li, L.; Yang, L.; Li, Y. Molecular simulations of host-guest interactions between zeolite framework STW and its organic structure-directing agents. Chin. Chem. Lett. 2020, 31, 1951–1955. [Google Scholar] [CrossRef]
  27. Gómez-Hortigüela, L.; Camblor, M.Á. Introduction to the Zeolite Structure-Directing Phenomenon by Organic Species: General Aspects. In Insights into the Chemistry of Organic Structure-Directing Agents in the Synthesis of Zeolitic Materials; Gómez-Hortigüela, L., Ed.; Springer: Cham, Switzerland, 2017; Structure and Bonding; Volume 175. [Google Scholar] [CrossRef] [Green Version]
  28. Gómez-Hortigüela, L.; Mayoral, Á.; Liu, H.; Sierra, L.; Vaquerizo, L.; Mompeána, C.; Pérez-Pariente, J. Synthesis of large-pore zeolites from chiral structure-directing agents with two L-prolinol units. Dalton Trans. 2020, 49, 9618. [Google Scholar] [CrossRef]
  29. Lobo, R.F.; Zones, S.I.; Davis, M.E. Structure-direction in zeolite synthesis. J. Incl. Phenom. Mol. Recognit. Chem. 1995, 21, 47–78. [Google Scholar]
  30. Yuan, R.; Claes, N.; Verheyen, E.; Tuel, A.; Bals, S.; Breynaert, E.; Martens, J.A.; Kirschhock, C.E.A. Synthesis of IWW-type germanosilicate zeolite using 5-azoniaspiro[4,4]nonane as structure directing agent. New J. Chem. 2016, 40, 4319–4324. [Google Scholar] [CrossRef]
  31. Millini, R.; Carluccio, L.; Frigerio, F.; O’Neil Parker, W., Jr.; Bellussi, G. Zeolite synthesis in the presence of azonia-spiro compounds as structure-directing agents. Microporous Mesoporous Mater. 1998, 24, 199–211. [Google Scholar] [CrossRef]
  32. Shvets, O.V.; Kasian, N.; Zukal, A.; Pinkas, J.; Cejka, J. The Role of Template Structure and Synergism between Inorganic and Organic Structure Directing Agents in the Synthesis of UTL Zeolite. Chem. Mater. 2010, 22, 3482–3495. [Google Scholar] [CrossRef]
  33. Clough, M.T.; Geyer, K.; Hunt, P.A.; McIntosh, A.J.; Rowe, R.; Welton, T.; White, A.J. Azoniaspiro salts: Towards bridging the gap between room-temperature ionic liquids and molten salts. Phys. Chem. Chem. Phys. 2016, 18, 3339–3351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Gordon, C.M.; Holbrey, J.D.; Kennedy, A.R.; Seddon, K.R. Ionic liquid crystals: Hexafluorophosphate salts. J. Mater. Chem. 1998, 8, 2627–2636. [Google Scholar] [CrossRef]
  35. Kayukova, L.A.; Baitursynova, G.P.; Yergaliyeva, E.M.; Zhaksylyk, B.A.; Yelibayeva, N.S.; Kurmangaliyeva, A.B. Arylsulphonates of spiropyrazolines and O-tosilate-β-(benzimidazol-1-yl)propioamidoxime as the products of β-aminopropioamidoximestosylation. Chem. J. Kaz. 2021, 2, 22–32. [Google Scholar] [CrossRef]
  36. Kayukova, L.A.; Vologzhanina, A.V.; Yergaliyeva, E.M.; Baitursynova, G.P.; Shulgau, Z.T.; Sergazy, S. Crystal structure and anti diabetic activity of 2-amino spiro pyrazolinium tosyl ates and the product of O-tosyl ation of β-(benzimidazol-1-yl)propio amidoxime. Acta Cryst. 2022, C78, 542–551. [Google Scholar] [CrossRef]
  37. Ning, Y.; Kawahata, M.; Yamaguchi, K.; Otani, Y.; Ohwada, T. Synthesis, Structure and N-N Bonding Character of 1,1-Disubstituted Indazolium Hexaflfluorophosphate. Chem. Commun. 2018, 54, 1881–1884. [Google Scholar] [CrossRef]
  38. Yergaliyeva, E.M.; Kayukova, L.A.; Bazhykova, K.B.; Gubenko, M.A.; Langer, P. Computational studies of the products of tosylation and para-nitrobenzenesulfochlorination. J. Struct. Chem. 2021, 62, 1969–1975. [Google Scholar] [CrossRef]
  39. Kayukova, L.A.; Orazbaeva, M.A.; Gapparova, G.I.; Beketov, K.M.; Espenbetov, A.A.; Faskhutdinov, M.F.; Tashkhodjaev, B.T. Rapid acid hydrolysis of 5-aryl-3-(β-thiomorpholinoethyl)-1,2,4-oxadiazoles. Chem. Heterocycl. Compd. 2010, 46, 879–886. [Google Scholar] [CrossRef]
  40. Delori, A.; Galek, P.T.A.; Pidcock, E.; Jones, W. Quantifying Homo- and Heteromolecular Hydrogen Bonds as a Guide for Adduct Formation. Chem. Eur. J. 2012, 18, 6835–6846. [Google Scholar] [CrossRef] [PubMed]
  41. Macrae, C.F.; Sovago, I.; Cottrell, S.J.; Galek, P.T.A.; McCabe, P.; Pidcock, E.; Platings, M.; Shields, G.P.; Stevens, J.S.; Towler, M.; et al. Mercury 4.0: From visualization to analysis, design and prediction. J. Appl. Cryst. 2020, 53, 226–235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Galek, P.T.A.; Allen, F.H.; Fábián, L.; Feeder, N. Knowledge-based H-bond prediction to aid experimental polymorph screening. CrystEngComm 2009, 11, 2634–2639. [Google Scholar] [CrossRef]
  43. Vologzhanina, A.V. Intermolecular Interactions in Functional Crystalline Materials: From Data to Knowledge. Crystals 2019, 9, 478. [Google Scholar] [CrossRef] [Green Version]
  44. Navashin, S.M. Rational Antibiotic Therapy; Navashin, S.M., Fomina, I.P., Eds.; Meditsina: Moscow, Russia, 1982; 496p. [Google Scholar]
  45. Lazarenko, V.A.; Dorovatovskii, P.V.; Zubavichus, Y.V.; Burlov, A.S.; Koshchienko, Y.V.; Vlasenko, V.G.; Khrustalev, V.N. High-Throughput Small-Molecule Crystallography at the ‘Belok’ Beamline of the Kurchatov Synchrotron Radiation Source: Transition Metal Complexes with Azomethine Ligands as a Case Study. Crystals 2017, 7, 325. [Google Scholar] [CrossRef] [Green Version]
  46. Svetogorov, R.D.; Dorovatovskii, P.V.; Lazarenko, V.A. Belok/XSA Diffraction Beamline for Studying Crystalline Samples at Kurchatov Synchrotron Radiation Source. Cryst. Res. Technol. 2020, 55, 1900184. [Google Scholar] [CrossRef]
  47. Kabsch, W. XDS. Acta Cryst. 2010, D66, 125–132. [Google Scholar] [CrossRef] [Green Version]
  48. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. 2015, A71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  49. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  50. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
Scheme 1. Representation of 1,5–diazaspiro–1–en–5-ammonium and azoniaspiroalkanes.
Scheme 1. Representation of 1,5–diazaspiro–1–en–5-ammonium and azoniaspiroalkanes.
Ijms 24 11315 sch001
Scheme 2. 2-Amino-1,5-diazathiospiro[4.5]-dec-1-ene-5-ammonium-containing salts with anti-tubercular [11] and anti-diabetic activities [12,13].
Scheme 2. 2-Amino-1,5-diazathiospiro[4.5]-dec-1-ene-5-ammonium-containing salts with anti-tubercular [11] and anti-diabetic activities [12,13].
Ijms 24 11315 sch002
Scheme 3. Synthesis of 2aminospiropyrazolilammonium chlorides (9–12) and hexafluorophosphates (13–16).
Scheme 3. Synthesis of 2aminospiropyrazolilammonium chlorides (9–12) and hexafluorophosphates (13–16).
Ijms 24 11315 sch003
Scheme 4. Schematic representation of monohydrates of salts (11) and (12) precipitated from reaction mixtures.
Scheme 4. Schematic representation of monohydrates of salts (11) and (12) precipitated from reaction mixtures.
Ijms 24 11315 sch004
Scheme 5. Formation of 2aminospiropyrazolilammonium hexafluorophosphates (1316) without isolation of intermediate 2aminospiropyrazolilammonium tosylates (58).
Scheme 5. Formation of 2aminospiropyrazolilammonium hexafluorophosphates (1316) without isolation of intermediate 2aminospiropyrazolilammonium tosylates (58).
Ijms 24 11315 sch005
Figure 1. Asymmetric units of salts 9, 10, 1216 in representation of atoms with thermal ellipsoids (drawn at p = 50%).Color code: C—grey; Cl—dark green; F—green; H—white; N—blue; O—red; P—violet; S—yellow.
Figure 1. Asymmetric units of salts 9, 10, 1216 in representation of atoms with thermal ellipsoids (drawn at p = 50%).Color code: C—grey; Cl—dark green; F—green; H—white; N—blue; O—red; P—violet; S—yellow.
Ijms 24 11315 g001aIjms 24 11315 g001b
Figure 2. H-bonded associates in solid 9, 10, 1216. H-bonds are depicted as dotted lines. Color code: C––grey; Cl––dark green; F––green; H––white; N––blue; O––red; P––violet; S––yellow.
Figure 2. H-bonded associates in solid 9, 10, 1216. H-bonds are depicted as dotted lines. Color code: C––grey; Cl––dark green; F––green; H––white; N––blue; O––red; P––violet; S––yellow.
Ijms 24 11315 g002
Table 1. Physicochemical characteristics of chlorides and hexafluorophosphates of 2aminospiropyrazolilammonium (9–16).
Table 1. Physicochemical characteristics of chlorides and hexafluorophosphates of 2aminospiropyrazolilammonium (9–16).
CompdXYield, %t, minm.p., °CRfCompdXYield, %t, h m.p., °CRf
9CH2685 288–2900.0113CH25524 192–1930.05
10O975 2710.0114O9024 191–1930.08
11S555 2900.0815S8024180–1820.01
12PhN705 288–2900.0116PhN8524240–2420.01
Table 2. In vitro antimicrobial activity of 2-amino-1,5-diazaspiro[4.5]dec-1-en-5-ammonium chlorides (912) and hexafluorophosphates (1316), mm *.
Table 2. In vitro antimicrobial activity of 2-amino-1,5-diazaspiro[4.5]dec-1-en-5-ammonium chlorides (912) and hexafluorophosphates (1316), mm *.
CompdStaphylococcusaureusBacillus subtilisEscherichia coliPseudomonasaeruginosaCandida albicans
9--12.5 ± 0.1--
10-----
1113.7 ± 0.1-12.6 ± 0.211.5 ± 0.2-
12--11.3 ± 0.1--
1312.5 ± 0.2----
14--11.6 ± 0.2--
1511.2 ± 0.2-13.5 ± 0.1--
16-----
Gentamicin24.5 ± 0.121.6 ± 0.226.4 ± 0.120.02± 0.122.7± 0.2
Nystatin----21.9 ± 0.2
*—significance of differences p < 0.05 compared with the comparison group.
Table 3. Predicted and experimentally observed hydrogen-bonded motifs in 2-aminospiropyrazolinium spiropyrazolinium chlorides and hexafluorophosphates for anhydrous salts and their hydrates.
Table 3. Predicted and experimentally observed hydrogen-bonded motifs in 2-aminospiropyrazolinium spiropyrazolinium chlorides and hexafluorophosphates for anhydrous salts and their hydrates.
DonorAcceptorPropensity aObserved at ChloridesObserved at Hexafluorophosphates
N–HHal1.00/0.999, 10, 10·H2O, 11·H2O, 12·H2O1316
O (water)0/0.7110·H2O, 11·H2O, 12·H2O
N20.44/0.549
Het<0.08/<0.20-15
O–HHal0/0.9610·H2O, 11·H2O, 12·H2O
O (water)0/0.48
N20/0.31
Het0/<0.1
a Valuesfor anhydrous salts and their hydrates are divided by a slash.
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

Kayukova, L.; Vologzhanina, A.; Dorovatovskii, P.; Yergaliyeva, E.; Uzakova, A.; Duisenali, A. N-N(+) Bond-Forming Intramolecular Cyclization of O-Tosyloxy β-Aminopropioamidoximes and Ion Exchange Reaction for the Synthesis of 2-Aminospiropyrazolilammonium Chlorides and Hexafluorophosphates. Int. J. Mol. Sci. 2023, 24, 11315. https://doi.org/10.3390/ijms241411315

AMA Style

Kayukova L, Vologzhanina A, Dorovatovskii P, Yergaliyeva E, Uzakova A, Duisenali A. N-N(+) Bond-Forming Intramolecular Cyclization of O-Tosyloxy β-Aminopropioamidoximes and Ion Exchange Reaction for the Synthesis of 2-Aminospiropyrazolilammonium Chlorides and Hexafluorophosphates. International Journal of Molecular Sciences. 2023; 24(14):11315. https://doi.org/10.3390/ijms241411315

Chicago/Turabian Style

Kayukova, Lyudmila, Anna Vologzhanina, Pavel Dorovatovskii, Elmira Yergaliyeva, Asem Uzakova, and Aidana Duisenali. 2023. "N-N(+) Bond-Forming Intramolecular Cyclization of O-Tosyloxy β-Aminopropioamidoximes and Ion Exchange Reaction for the Synthesis of 2-Aminospiropyrazolilammonium Chlorides and Hexafluorophosphates" International Journal of Molecular Sciences 24, no. 14: 11315. https://doi.org/10.3390/ijms241411315

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