Next Article in Journal
One Pot Synthesis, Surface, and Magnetic Properties of Ni–NiO@C Nanocomposites
Next Article in Special Issue
Glycerol as Ligand in Metal Complexes—A Structural Review
Previous Article in Journal
The Role of Boron Addition on Solidification Behavior and Microstructural Evolution of a High Niobium-Containing TiAl Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Structural Analysis of Chiral Bis-dihydro[1,3]-naphthoxazines and Imidazolidine Derivatives Prepared by Three-Component Mannich-Type Condensation

by
Maya Tavlinova-Kirilova
1,
Krasimira Dikova
1,
Maya K. Marinova
1,
Mariana Kamenova-Nacheva
1,
Rusi Rusew
2,
Hristina Sbirkova-Dimitrova
2,
Boris Shivachev
2,*,
Kalina Kostova
1 and
Vladimir Dimitrov
1,*
1
Institute of Organic Chemistry with Centre of Phytochemistry, Bulgarian Academy of Sciences, Acad. G. Bonchev Street, bl. 9, 1113 Sofia, Bulgaria
2
Institute of Mineralogy and Crystallography “Acad. Ivan Kostov”, Bulgarian Academy of Sciences, Acad. G. Bonchev Street, bl. 107, 1113 Sofia, Bulgaria
*
Authors to whom correspondence should be addressed.
Submission received: 20 September 2023 / Revised: 3 October 2023 / Accepted: 12 October 2023 / Published: 14 October 2023
(This article belongs to the Special Issue Coordination Complexes with Bio-Based Ligands)

Abstract

:
Enantiomerically pure (S)-1-phenylethan-1-amine has been applied in Mannich-type condensation between formaldehyde and naphthalenediols leading to the synthesis of chiral bis-dihydro[1,3]naphthoxazines in excellent yields. Salen-type structures have been synthesized, applying R,R- or S,S-cyclohexane-1,2-diamines in condensation with formaldehyde and naphthalene-2-ol. The obtained chiral imidazolidine derivatives of the type 1,1′-(((3a,7a)-hexahydro-1H-benzo[d]imidazole-1,3(2H)diyl)bis(methylene))bis(naphthalen-2-ol) were evaluated as pre-catalysts for the addition of diethyl zinc to aldehydes. The structures of the newly synthesized compounds were elucidated using 1D and 2D NMR experiments (COSY, HMBC, HSQS), elemental analysis, mass spectrometry (HRMS spectra) and single-crystal X-ray diffraction (SCXRD). The products were further characterized with powder X-ray diffraction (PXRD) and thermal analysis (DSC).

1. Introduction

Multicomponent reactions are an attractive tool for creating complex polyfunctional organic compounds [1,2]. The synthesis of dihydrooxazines by three-component Mannich-type condensation of naphthol, amine and formaldehyde has a high potential for obtaining an important class of heterocyclic compounds. Structurally diverse dihydro[1,3]oxazines have demonstrated a broad range of biological activities (bactericides, fungicides, anti-inflammatory, antitumor, anti-HIV) and have found applications in materials science [3,4].
The synthesis of dihydro[1,3]oxazines occurs most commonly through one-pot condensation between aromatic alcohol, formaldehyde and amine applying different solvents or in some examples under solvent-free conditions [3]. The use of readily available chiral amines offers the opportunity for synthesis of chiral dihydrooxazines. To date, there are only a limited number of examples describing the synthesis of enantiopure dihydro[1,3]naphthoxazine using chiral amines. These compounds are employed as intermediates in the synthesis of N-substituted aminobenzylnaphthols and demonstrate a significant activity in asymmetric catalysis [5]. In our previous studies, we demonstrated a practicable approach for the synthesis of 1,3-naphthoxazines through a three-component condensation involving naphthalene-2-ol, formaldehyde and a series of chiral amines. The subsequent reduction obtained the corresponding N-methyl derivatives in a convenient way [6]. In a similar synthetic approach, we applied 2,6-, 2,3- and 1,5-dihydroxy naphthalenes in the condensation reaction (Betti reaction) with aromatic aldehydes and (S)-(−)-1-phenylethane-1-amine, synthesizing chiral aminobenzylnaphthols in good yields [7].
In this study, we present the synthesis of bis-dihydro[1,3]naphthoxazines applying 2,3- and 2,6-dihydroxy naphthalenes, (S)-(−)-1-phenylethane-1-amine and formaldehyde. Of further interest was the condensation of cyclohexane-1,2-diamine, naphthalene-2-ol and formaldehyde due to the opportunity to obtain salen-type structures [8,9]. The structural elucidation of isolated chiral compounds applying advanced NMR and X-ray crystallographic experiments was of particular importance for understanding the possible catalyst applications of the compounds. Some of the synthesized structures were preliminarily evaluated in a model catalytic reaction.

2. Materials and Methods

2.1. Chemistry (General Methods)

All reagents were of commercial grade and were used without additional purification. Aluminum sheets precoated with silica gel 60 F254 (Merck, Darmstadt, Germany) were used for performing thin-layer chromatography (TLC) while silica gel 60 (230–400 mesh, Merck, Rahway, NJ, USA) was used for flash chromatography. Optical rotation α D 20  measurements were performed using a JASCO P2000 polarimeter, Jasco Inc., Hachioji, Japan. The NMR spectra were recorded on a Bruker Avance II+ 600 spectrometer (600 MHz for 1H, 150 MHz for 13C NMR) in CDCl3 with TMS as the internal standard for chemical shifts (δ, ppm). The NMR data for 1H and 13C were reported as follows: chemical shift, multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, br = broad, m = multiplet), integration, coupling constants (J, Hz) and identification. The assignment of the 1H and 13C NMR spectra was made on the basis of DEPT, COSY, HSQC, HMBC and NOESY experiments. All assignments marked with an asterisk are tentative. Mass spectra (MS) were recorded on a QSTAR pulsar I mass spectrometer (AB/MDS Sciex) by electrospray ionization (ESI) and were reported as fragmentation in m/z with relative intensities (%) in parentheses. The mass spectra were recorded on Thermo Scientific (Waltham, MA, USA) UHPLC system coupled with a Q Exactive Plus Hybrid Quadrupole-Orbitrap Mass Spectrometer Thermo Scientific (HESI HRMS). The spectra were processed by the Xcalibur Free Style program.

2.2. Synthesis of Bis-dihydro[1,3]naphthoxazines

2.2.1. Synthesis of 3,10-Bis((S)-1-phenylethyl)-2,3,4,9,10,11-hexa-hydronaphtho[1,2-e:4,3-e′]bis([1,3]oxazine) (3)

To a mixture of naphthalene-2,3-diol (0.100 g, 0.62 mmol, 1) and (S)-(−)-1-phenylethane-1-amine (0.159 g, 1.30 mmol, 2) in ethanol (5 mL), paraformaldehyde (0.094 g, 3.10 mmol) was added. The mixture was stirred at 50 °C for 3 h. After evaporation of the solvent, the crude product was column chromatographed (CH2Cl2/MeOH = 100:1) to provide 0.275 g (98%) of 3. mp 206–211 °C. α D 20 = +121 (c 1.014, CHCl3); 1H NMR (600 MHz, CDCl3) δ 1.53 (d, 6H, J = 6.6 Hz, H-14, H-14′), 4.07–4.12 (m, 4H, H-13, H-13′, H*-11), 4.39 (d, 2H, J = 16.4 Hz, H*-11′), 5.06* (d, 2H, J = 10.2 Hz, H-21), 5.34* (dd, 2H, J = 10.2, 0.8 Hz, H-21′), 7.24–7.27 (m, 4H, H-7, H-8, H-18, H-18′), 7.30–7.33 (m, 4H, H-17, H-17′, H-19, H-19′), 7.35–7.36 (m, 4H, H-16, H-16′, H-20, H-20′), 7.37–7.38 (m, 2H, H-6, H-9); 13C NMR (150.92 MHz, CDCl3) δ 21.68 (2q, C-14, C-14′), 46.00 (2t, C-11, C-11′), 58.26 (2d, C-13, C-13′), 80.37 (2t, C-21, C-21′), 111.70 (2s, C-1, C-4), 121.43 (2d, C-6, C-9), 123.94 (2d, C-7, C-8), 126.35 (2s, C-5, C-10), 127.19 (2d, C-18, C-18′), 127.22 (4d, C-16, C-16′, C-20, C-20′), 128.51 (4d, C-17, C-17′, C-19, C-19′), 143.52 (2s, C-2, C-3), 144.78 (2s, C-15, C-15′); MS (ESI) m/z (rel. int.): 451 ([M + H]+, 34). HRMS (ESI+) m/z calcd. for C30H30N2O2+ [M + H]+ 451.2381, found 451.2238. Anal. calcd. for C30H30N2O2 (450.58): C, 79.97; H, 6.71; N, 6.22. Found: C, 80.21; H, 6.82; N, 6.25.

2.2.2. Synthesis of 3,9-Bis((S)-1-phenylethyl)-2,3,4,8,9,10-hexahydronaphtho[1,2-e:5,6-e′]bis([1,3]oxazine) (5)

To a mixture of naphthalene-2,6-diol (0.100 g, 0.62 mmol, 4) and (S)-(−)-1-phenylethane-1-amine (0.159 g, 1.30 mmol, 2) in ethanol (3 mL), 37% aq. solution of formaldehyde (calculated to provide 5 equiv. of formaldehyde) was added. The mixture was stirred at 50 °C for 1h. After evaporation of the solvent, the crude product was column chromatographed (petroleum ether/MTBE = 20:1) to provide 0.265 g (94%) of 5. mp 183–185 °C. α D 20 = +29 (c 1.005, CHCl3); 1H NMR (600 MHz, CDCl3) δ 1.49 (d, 6H, J = 6.6 Hz, H-14, H-14′), 4.03 (q, 2H, J = 6.6 Hz, H-13, H-13′), 4.13* (d, 2H, J = 16.8 Hz, H-11), 4.38* (d, 2H, J = 16.8 Hz, H-11′), 4.88* (d, 2H, J = 10.1 Hz, H-21), 5.13* (dd, 2H, J = 10.1, 1.2 Hz, H-21′), 6.98 (d, 2H, J = 9.1 Hz, H-3, H-8), 7.24–7.36 (m, 12H, H-18, H-18′, H-4, H-9, H-17, H-17′, H-19, H-19′, H-16, H-16′, H-20, H-20′); 13C NMR (150 MHz, CDCl3) δ 21.63 (2q, C-14, C-14′), 46.19 (2t, C-11, C-11′), 58.18 (2d, C-13, C-13′), 79.79 (2t, C-21, C-21′), 113.38 (2s, C-1, C-6), 118.61 (2d, C-3, C-8), 120.82 (2d, C-4, C-9), 127.08 (2s, C-5, C-10), 127.22 (2d, C-18, C-18′), 127.27 (4d, C-16, C-16′, C-20, C-20′), 128.52 (4d, C-17, C-17′, C-19, C-19′), 144.72 (2s, C-15, C-15′), 150.91 (2s, C-2, C-7); HRMS (ESI+) m/z calcd. for C30H30N2O2+ [M + H]+ 451.2381, found 451.2238. Anal. calcd. for C30H30N2O2 (450.58): C, 79.97; H, 6.71; N, 6.22. Found: C, 80.18; H, 6.93; N, 6.04.

2.3. Synthesis of Cyclohexane-1,2-diamine Derivatives

2.3.1. Synthesis of 1,1′-(((3aR,7aR)-Hexahydro-1H-benzo[d]imidazole-1,3(2H)-diyl)-bis(methylene))bis(naphthalen-2-ol) ((R,R)-9)

To a solution of naphthalene-2-ol (0.375 g, 2.20 mmol, 7) in methanol (5 mL), paraformaldehyde (0.071 g, 2.00 mmol) and R,R-cyclohexane-1,2-diamine (0.135 g, 1.00 mmol, 6) were added. The mixture was refluxed for 2 h, whereby a white precipitate was formed. The reaction mixture was cooled to room temperature and the product was collected by filtration and washed with MeOH to provide 0.245 g (47%) of (R,R)-9. mp 190–192 °C. α D 20 = +83 (c 1.00, CHCl3). 1H NMR (600 MHz, CDCl3) δ 1.25–1.35 (m, 2H, H-15, H-15′), 1.40–1.50 (m, 2H, H-14, H-14′), 1.82–1.90 (m, 2H, H-15, H-15′), 2.09–2.17 (m, 2H, H-14, H-14′), 2.48–2.56 (m, 2H, H-13, H-13′), 3.66 (s, 2H, H-16), 4.21 (d, 2H, J = 14.5 Hz, H-11, H-11′), 4.38 (d, 2H, J = 14.5 Hz, H-11, H-11′), 7.03 (d, 2H, J = 8.9 Hz, H-3, H-3′), 7.22–7.26 (m, 2H, H-7, H-7′), 7.38–7.44 (m, 2H, H-8, H-8′), 7.62 (d, 2H, J = 8.9 Hz, H-4, H-4′), 7.69 (d, 2H, J = 8.0 Hz, H-6, H-6′), 7.76 (d, 2H, J = 8.6 Hz, H-9, H-9′), 11.98 (br s, 2H, HO-(C-2, C-2′)); 13C NMR (600 MHz, CDCl3) δ 23.99 (2t, C-15, C-15′), 28.94 (2t, C-14, C-14′), 51.52 (2t, C-11, C-11′), 69.42 (2d, C-13, C-13′), 76.53 (t, C-16), 111.09 (2s, C-1, C-1′), 118.93, (2d, C-3, C-3′), 120.90 (2d, C-9, C-9′), 122.58 (2d, C-7, C-7′), 126.48 (2d, C-8, C-8′), 128.43 (2s, C-5, C-5′), 128.80 (2d, C-6, C-6′), 129.42 (2d, C-4, C-4′), 132.02 (2s, C-10, C-10′), 155.96 (2s, C-2, C-2′); MS (ESI) m/z (rel. int.): = 439 ([M + H]+, 67). HRMS (ESI+) m/z calcd. for C29H30N2O2+ [M + H]+ 439.2382, found 439.2376. Anal. Calcd. for C29H30N2O2 (438.57): C, 79.42; H, 6.89; N, 6.39. Found: C, 79.14; H, 6.88; N, 6.41

2.3.2. Synthesis of 1,1′-(((3aS,7aS)-Hexahydro-1H-benzo[d]imidazole-1,3(2H)-diyl)-bis(methylene))bis(naphthalen-2-ol) ((S,S)-9)

(1S,2S)-(−)-cyclohexane-1,2-diamine (as tartrate) (0.5 g, 1.9 mmol, 6) was mixed with 3 mL water and K2CO3 (0.525, 3.8 mmol). A mixture of paraformaldehyde (0.114 g, 3.8 mmol) and naphthalene-2-ol (0.6 g, 4.2 mmol, 7) in ethanol (10 mL) was added. The reaction mixture was stirred at 50 °C for 24 h. After evaporation of the solvent, the residue was dissolved in CH2Cl2 and washed with water (2 × 10 mL). The organic layer was dried over anhydrous Na2SO4, filtered and the solvent was evaporated under reduced pressure. The crude product was column chromatographed (petroleum ether/CH2Cl2 = 1:5) to provide 0.260 g (31%) of ((S,S)-9) mp 186–188 °C; α D 20 = −86 (c 1.00, CHCl3); The NMR data are identical to those of enantiomer (R,R)-9. HRMS (ESI+) m/z calcd. for C29H30N2O2+ [M + H]+ 439.2382, found 439.2379.

2.4. General Procedure for Enantioselective Addition of Diethylzinc to Aldehydes

To a solution of the corresponding ligand ((R,R)-9 or (S,S)-9) (3 mol %) in toluene (4 mL), Et2Zn (1.7 mmol, 1M solution in hexane) was added dropwise at 0 °C in an Ar atmosphere. The mixture was stirred for 30 min at 0 °C and then the corresponding aldehyde (1.0 mmol) was added at −20 °C. The reaction mixture was stirred at 20 °C and monitored by TLC (petroleum ether/MTBE = 5:1) until the aldehyde was consumed. The mixture was quenched (aq. NH4Cl), extracted with Et2O and dried over anhydrous Na2SO4. After evaporation of the solvent, the crude product was purified by column chromatography (petroleum ether/MTBE = 5:1).

2.5. Powder X-ray Diffraction (PXRD)

Powder XRD patterns were determined for the powdered compounds 3, (R,R)-9 and (S,S)-9 to establish crystalline properties, purity and eventual presence of polymorphs. Powder samples of the synthesized compounds were analyzed on an Empyrean Powder X-ray diffractometer (Malvern Panalytical, Almelo, The Netherlands) in the 2–50° 2θ range using Cu radiation (λ = 1.5406 Å) and a PIXcel3D detector. The diffraction patterns of the crystals were compared with those of starting compounds and the computed from SCXRD pattern to confirm the presence or absence of additional phases.

2.6. Single-Crystal X-ray Diffraction (SCXRD)

Single crystals of compounds 3 and (R,R)-9 with suitable size (~0.3 × 0.3 × 0.3 mm3) and diffracting quality were mounted on nylon loops. Diffraction data for 3 were collected on a SupernovaDual diffractometer equipped with an Atlas CCD detector while diffraction data for (R,R)-9 were collected on a Bruker D8 Venture diffractometer equipped with a PHOTON II CPAD detector. Both diffractometers operate with a micro-focus sealed X-ray source, generating MoKα radiation with λ = 0.71073 Å. The collected data were processed with CrysAlisPro [10] (for 3) or APEX4 [11] (for (R,R)-9) programs. The structures were solved with intrinsic phasing methods and refined by the full-matrix least-squares method of F2 using ShelxT [12]and ShelxL [13] as implemented in the OLEX2-ver.1.5 graphical interface [14]. All non-hydrogen atoms were located successfully from Fourier map and were refined anisotropically. Hydrogen atoms were placed on calculated positions riding on the parent carbon atoms (Ueq = 1.2 for C-Haromatic = 0.93 Å and C-Hmethylenic = 0.97 Å) while those riding on a heteroatoms were refined from electron density maps. Ortep-3v2 software [15] was used to prepare the figures visualizing the compounds in the asymmetric unit. Complete crystallographic data for the structure of compounds 3 and (R,R)-9 reported in this paper have been deposited in the CIF format with the Cambridge Crystallographic Data Center as 2294902 and 2294903, respectively. These data, last accessed on 14 September 2023, can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html.

2.7. Thermal Analysis (DSC)

Differential scanning calorimetry was conducted to estimate thermotropic properties and thermal behaviors of the new compounds. DSC analyses were performed on a Discovery DSC250 (TA instruments, New Castle, DE, USA). Samples weighing between 1 and 5 mg were heated in aluminum pans from 30 to 250 °C (10 °C·min−1) in nitrogen (flow rate 25 mL·min−1). Endothermic/exothermic effects of solvent evaporation, melting point and decomposition of the synthesized compounds were determined from the DSC experiments.

3. Results and Discussion

3.1. Synthesis

The condensation reactions of naphthalene-2,3-diol (1) or naphthalene-2,6-diol (4), (S)-(−)-1-phenylethane-1-amine (2) and formaldehyde (applied as formalin, 37% aq. solution or paraformaldehyde) were performed in ethanol at 50 °C as shown in Scheme 1a. The products 3 and 5 were isolated in high yields (98% and 94%, respectively) after short reaction times (1–3 h) and purification by column chromatography. As the next step, preparation the corresponding N-methyl derivatives through reduction with LiAlH4 was attempted [6]. Surprisingly, all reduction experiments applying different conditions and reducing agents (LiAlH4 or NaBH4) were unsuccessful and even traces of the desired N-methylated compounds could not be detected.
The condensation of cyclohexane-1,2-diamine (6), naphthalene-2-ol (7) and formaldehyde (Scheme 1b) was of particular interest, since the formation of a salen-type structure was expected. Salen ligands are described as efficient catalysts in certain enantioselective processes [8]. The initial condensation experiments were performed with racemic cyclohexane-1,2-diamine (6) in ethanol as a solvent at room temperature (Scheme 1b). An aqueous solution of formaldehyde (37%) was used in large excess and the reaction outcome was monitored by TLC. The formation of the expected bis-dihydro[1,3]naphthoxazine 8 was recognized in the crude reaction mixture by NMR spectroscopy (Figure 1, Table 1). The amine 6 was consumed over the course of a 3 h reaction time (Table 2, entry 1). Besides this, an additional compound with similar structure was identified and determined as imidazolidine 9 (Scheme 1b). The experiments for separation and isolation (column chromatography, crystallization) of products 8 and 9 in pure form were unsuccessful. However, it was possible to distinguish the two compounds in a mixture and to determine undoubtedly the structures by using NMR and HRMS experiments. In Figure 1, representative 1H NMR spectra of 8 and 9 are shown and in Table 1 the characteristic chemical shifts of the individual compounds are introduced.
In order to optimize the formation of compound 8 preferentially, a series of experiments have been performed (Table 2). In general, prolonged reaction time and heating up to 50 °C promotes the formation of compound 9 (compare entries 1 to 6). If the mixture is refluxed for 2.5 h, compound 9 is formed preferentially (entry 7). The result introduced in entry 5 needs some comment. In fact, the ratio 8:9 after 18 h at 50 °C was 10:7. After washing the crude mixture with acetonitrile, the ratio was shifted in favor of 9 (8:9 = 20:19). However, pure 8 could not be isolated from the filtrate (due to stability problems, vide infra).
When toluene was used as solvent and the mixture heated at 50 °C for 24 h, the ratio of the two products 8 and 9 formed was 5:1 (entry 8, Table 1). Even at this ratio, the isolation of bis-dihydro[1,3]naphthoxazine 8 in pure form was impossible. All attempts to purify 8 by crystallization using various solvents and mixtures of solvents did not lead to a positive result. Furthermore, 8 is unstable under column chromatography conditions even when the silica gel is deactivated with Et3N. On standing, bis-dihydro[1,3]naphthoxazine 8 is transformed slowly into 9, detected by its increased fraction in the ratio. Pure 9 is stable at room temperature; however, if mixed with formaldehyde, the formation of 8 was observed in a short time, forming nearly a 1:1 ratio of 8:9. It is important to emphasize that reduction experiments for treatment of mixtures of 8 and 9 with LiAlH4 or NaBH4 led to destruction of 8, whereas 9 remained almost unchanged.
It is interesting to note that the synthesis of compound 8 was recently described (racemic diamine 6 was used) [9]. The structural proof was reported based on elemental analysis and IR spectroscopy and no NMR data were presented. However, the compound described as 8 was further used as a ligand to prepare salen-type complexes with the transition metal Mn, which were characterized by single-crystal X-ray crystallography. Consequently, based on the published results of [9] and our results, it is possible to assume that salen-type metal complexes could be formed from mixtures of 8 and 9 under suitable conditions.
In further experiments, the formation of 9 could be optimized (entries 9–12). In refluxing mixtures of diamine 6, naphthalene-2-ol and formaldehyde (37 aq. solution or paraformaldehyde, 2 equivalent), the reaction outcome was exclusively shifted to the formation of compound 9. Conveniently, the imidazolidine 9 crystalize in the reaction mixture and could be isolated by filtration. The highest yield of pure 9 (63% after column chromatography) was realized by heating the reaction mixture at 50 °C for 24 h (entry 12).
The optimized conditions (entry 12) were applied for the synthesis of the corresponding chiral derivatives (Figure 2) applying (R,R)-6 (in free base form) and (S,S)-6 (as tartrate salt). The yields of pure compounds we could realize were 47% for (R,R)-9 (isolated by crystallization) and 31% for (S,S)-9 (isolated by column chromatography).

3.2. Enantioselective Addition of Et2Zn to Aldehydes Catalyzed by Ligands (R,R)-9 and (S,S)-9

The imidazolidine derivatives (R,R)-9 and (S,S)-9 possess suitable structure and properties to be used as ligands in the model reaction for the enantioselective addition of diethylzinc to aldehydes by using a standard procedure (Scheme 2). The imidazolidine derivatives of 9 were applied in 3 mol % quantity by using of standard conditions described in the literature (see Section 2) [7,16,17,18]. A series of aromatic aldehydes were tested in the addition reaction, leading to the chiral secondary alcohols of type 11. The yields obtained (Table 3) ranged from moderate to very good; however, in several cases (entries 4–6 and 10), this was after prolonged reaction times. The enantioselectivities achieved were moderate. The prolonged reaction times indicates the relatively low efficiency of the chiral catalyst formed in situ in respect to the enantioselectivity, although the chemical yields were acceptable and, in some cases, even very high.
The best asymmetric induction was achieved by the addition reaction of diethylzinc to 2-methoxybenzaldehyde catalyzed by ligands (R,R)-9 and (S,S)-9 (53% and 58% ee, respectively, Table 3).

3.3. Crystal Structure Description

Single crystals of compounds 3 and (R,R)-9 were obtained by slow evaporation from a hot methanol solution. Compounds 3 and (R,R)-9 crystallize in non-centrosymmetric monoclinic P21 and orthorhombic P212121 space groups with one molecule in the asymmetric unit (ASU) and two and four molecules in the unit cell, respectively (Table S1). The ORTEP view of the molecules in the ASU of compounds 3 and (R,R)-9 with appropriate labeling scheme is shown in Figure 3. Selected bond lengths and angles from the single-crystal X-ray diffraction experiment are provided in Table 4 and they are comparable with other similar structures in the Cambridge structural database (ref. codes PAFCEH [19], OWEHEE [20], VESHAD [21], SOVYOR [22], HAJWUO [23], CAXJEU [24] and AFIHII [25]).
The molecule of compound 3 (Figure 3a) can be divided into two main fragments—a central 2,3,4,9,10,11-hexahydronaphtho[1,2-e:4,3-e′]bis([1,3]oxazine) core covalently bonded with two ethyl benzene side chains. Furthermore, the 2,3,4,9,10,11-hexahydronaphtho[1,2-e:4,3-e′]bis([1,3]oxazine) core can be divided into a fused naphthalene and two 1,3-oxazinane rings sharing the common C1–C2 and C3–C4 carbon atoms. The naphthalene ring is planar with an rmsd of 0.012 Å while the two 1,3-oxazinane rings are non-planar (rmsd ~0.200 Å) and have a “half-chair” conformation. The two N atoms from the 1,3-oxazinane rings (namely N12 and N22) are covalently bonded to the α-carbon atoms—labeled C13 or C24—of the ethylbenzene moieties. The angle between the normal of the mean planes of the benzene rings C15–C16–C17–C18–C19–C20 and C26–C27–C28–C29–C30–C31 from the ethylbenzene fragments and the naphthalene fragment are 135.77(13)° and 140.33(14)°, respectively. The twist/fold angles between the C15–C16–C17–C18–C19–C20/naphthalene and C26–C27–C28–C29–C30–C31/naphthalene fragments are 126.02(16)°/112.35(19)° and 133.67(17)°/117.20(20)°, respectively.
The molecule of compound (R,R)-9 (Figure 3b) consists of a central octahydro-1H-benzo[d]imidazole fragment that is connected with two 1-methylnaphthalen-2-ol side chains via N–C covalent bonds (N3–C7 and N2–C5). Furthermore, the octahydro-1H-benzo[d]imidazole moiety (rmsd = 0.207 Å) is built up by fused six-membered cyclohexane and five-membered imidazolidine aliphatic rings sharing the C15 and C16 carbon atoms. The cyclohexane and the imidazolidine rings are non-planar (rmsd = 0.239 Å and 0.200 Å) and have “chair” and “half-chair” conformation, respectively. The 1-methylnaphthalen-2-ol side chains are planar with an rmsd of 0.040 Å (for the N3-bonded one) and 0.015 Å (for N2-bonded one). The angles between the normal of the mean planes of the N3-bonded and N2-bonded 1-methylnaphthalen-2-ol fragments and the central octahydro-1H-benzo[d]imidazole fragment are 81.99° and 100.24° and the twist/fold angles are 82.19°/168.6° and 98.85°/29.75°, respectively.
The molecule of 3 lacks proton-donating groups (like C–OH, NH, NH2); therefore, typical hydrogen bonding interactions are not expected. The packing of the molecules in the crystal structure is governed by weak C–H…O and C–H…π interactions. On the other hand, the molecule of (R,R)-9 has two proton-donating hydroxyl groups (–OH) and two N atoms from the imidazole moiety acting as proton acceptors. Thus, formation of intramolecular hydrogen bonding interactions of the O–H…N type is expected and detected (Figure 4, Table 5). The crystal structure of (R,R)-9 is further stabilized by weak C-H…O interactions (π...π interactions are not detected).
The thermal stability of compounds 3, 5, (R,R)-9 and (S,S)-9 was evaluated using differential scanning calorimetry in the temperature region of 30–250 °C. The DSC thermograms of compound 3, (R,R)-9 and (S,S)-9 (Figure 5a,c,d) disclose a similar thermal comportment, e.g., a sharp endothermic effect followed by sharp exothermic effect. For all 3, (R,R)-9 and (S,S)-9 compounds, the endothermic effect is related to the melting of the compounds while the immediately following exothermic effect is associated with the decomposition of the compounds. According to the DSC data, the melting behavior for (R,R)-9 and (S,S)-9 is similar, with maximums at 191 and 188 °C, respectively (Figure 5c,d). The thermal comportment of 5 features also an endothermic and an exothermic effect (Figure 5b). For compound 5, the melting is observed at ~185 °C while the decomposition is delayed with ~40 °C after the melting has finished.
Overall, the DSC characterization of compounds 3, 5, (R,R)-9 and (S,S)-9 reveals their high purity and stability up to and above 180 °C (no decomposition or phase transitions are observed). The purity and the lack of polymorphs or isomerization in compounds 3 and (R,R)-9 are also confirmed by comparing the powder diffraction patterns of the bulk samples (in blue, Figure 6) and those generated from the single-crystal structural analysis (in red, Figure 6).

4. Conclusions

The three-component Mannich-type condensation of naphthalene-2-ol and isomeric naphthalenediols, formaldehyde and chiral amines is an excellent tool for the synthesis in excellent yields of enantiomerically pure substituted bis-dihydro[1,3]naphthoxazines (compounds 3 and 5). The properly optimized condensation reaction of naphthalene-2-ol, formaldehyde and chiral cyclohexane-1,2-diamine (both enantiomers, R,R- and S,S-configuration) leads to the unusual structure of 1,1′-(((3a,7a)-hexahydro-1H-benzo[d]imidazole-1,3(2H)-diyl)-bis(methylene))bis(naphthalen-2-ol), compound 9, of which the two enantiomers were isolated. The latter were evaluated as pre-catalysts for enantioselective addition of diethylzinc to aldehydes showing moderate asymmetric induction.
In compound 3, the naphthalene ring is planar, while the 1,3-oxazinane rings have a “half-chair” conformation. In (R,R)-9, the cyclohexane and imidazolidine rings have a non-planar “chair” and “half-chair” conformation, respectively. Compound 3 lacks proton-donating groups and, therefore, typical hydrogen bonding interactions are not expected in its crystal structure. In contrast, (R,R)-9 has two proton-donating hydroxyl groups (–OH) and two N atoms from the imidazole moiety acting as proton acceptors; thus intramolecular hydrogen bonding interactions of the O–H…N type stabilize the molecular and crystal structure.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/cryst13101495/s1, Figures S1–S6. The NMR spectra of compound 3; Figure S7. The HRMS spectrum of compound 3; Figures S8–S13. The NMR spectra of compound 5; Figure S14. The HRMS spectrum of compound 5; Figure S15. The 1H NMR spectrum of compound 8; Figure S16. The 1H NMR spectrum of a mixture of compounds 8 and 9 (1:0.95 ratio); Figure S17. The HRMS spectrum of a mixture of compounds 8 and 9; Figures S18–S23. The NMR spectra of compound (R,R)-9; Figure S24. The HRMS spectrum of compound (R,R)-9; Table S1. Most important data collection and refinement parameters for compounds 3 and (R,R)-9.

Author Contributions

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

Funding

This research received funding by Operational Program “Science and Education for Smart Growth” 2014–2020, co-financed by European Union through the European Structural and Investment Funds, Grant BG05M2OP001-1.002-0012, project “Sustainable utilization of bio-resources and waste of medicinal and aromatic plants for innovative bioactive products” and technical support from the project PERIMED BG05M2OP001-1.002-0005/29.03.2018 (2018–2023).

Data Availability Statement

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

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. Rodrigues, M.O.; Eberlin, M.N.; Neto, B.A. How and why to investigate multicomponent reactions mechanisms? A critical review. Chem. Rec. 2021, 21, 2762–2781. [Google Scholar] [CrossRef] [PubMed]
  2. Insuasty, D.; Castillo, J.; Becerra, D.; Rojas, H.; Abonia, R. Synthesis of biologically active molecules through multicomponent reactions. Molecules 2020, 25, 505. [Google Scholar] [CrossRef] [PubMed]
  3. Lathwal, A.; Mathew, B.P.; Nath, M. Syntheses, biological and material significance of dihydro[1,3]oxazine derivatives: An overview. Curr. Org. Chem. 2021, 25, 133–174. [Google Scholar] [CrossRef]
  4. Shen, S.B.; Ishida, H. Synthesis and characterization of polyfunctional naphthoxazines and related polymers. J. Appl. Polym. Sci. 1996, 61, 1595–1605. [Google Scholar] [CrossRef]
  5. Liu, D.-X.; Zhang, L.-C.; Wang, Q.; Da, C.-S.; Xin, Z.-Q.; Wang, R.; Choi, M.C.; Chan, A.S. The application of chiral aminonaphthols in the enantioselective addition of diethylzinc to aryl aldehydes. Org. Lett. 2001, 3, 2733–2735. [Google Scholar] [CrossRef]
  6. Tavlinova-Kirilova, M.; Marinova, M.; Angelova, P.; Kamenova-Nacheva, M.; Kostova, K.; Dimitrov, V. Three component condensation of a Betti-type–efficient tool for synthesis of chiral naphthoxazines and aminobenzylnaphthols for enantioselective diethylzinc addition to aldehydes. Bulg. Chem. Commun. 2016, 48, 705–712. [Google Scholar]
  7. Marinova, M.; Kostova, K.; Tzvetkova, P.; Tavlinova-Kirilova, M.; Chimov, A.; Nikolova, R.; Shivachev, B.; Dimitrov, V. Synthesis of 1, 3-aminonaphthols by diastereoselective Betti-type aminoalkylation of dihydroxy naphthalenes; diastereoselectivity, absolute configuration, and application. Tetrahedron Asymmetry 2013, 24, 1453–1466. [Google Scholar] [CrossRef]
  8. Gualandi, A.; Calogero, F.; Potenti, S.; Cozzi, P.G. Al (salen) metal complexes in stereoselective catalysis. Molecules 2019, 24, 1716. [Google Scholar] [CrossRef]
  9. Bikas, R.; Emami, M.; Ślepokura, K.; Noshiranzadeh, N. Preparing Mn (III) salen-type Schiff base complexes using 1, 3-oxazines obtained by Mannich condensation: Towards removing ortho-hydroxyaldehydes. New J. Chem. 2017, 41, 9710–9717. [Google Scholar] [CrossRef]
  10. CrysAlis PRO; Rigaku Oxford Diffraction Ltd., UK Ltd.: Yarnton, UK, 2021.
  11. Bruker, S. APEX3 V2016. 9-0, SAINT V8. 37A; Bruker AXS Inc.: Madison, WI, USA, 2016; Volume 2013, p. 2014. [Google Scholar]
  12. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. A 2008, 64, 112–122. [Google Scholar] [CrossRef]
  13. Sheldrick, G. Crystal structure refinement with SHELXL. Acta Crystallogr. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  14. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  15. Farrugia, L. WinGX and ORTEP for Windows: An update. J. Appl. Crystallogr. 2012, 45, 849–854. [Google Scholar] [CrossRef]
  16. Kitamura, M.; Suga, S.; Kawai, K.; Noyori, R. Catalytic asymmetric induction. Highly enantioselective addition of dialkylzincs to aldehydes. J. Am. Chem. Soc. 1986, 108, 6071–6072. [Google Scholar] [CrossRef]
  17. Noyori, R.; Kitamura, M. Enantioselective addition of organometallic reagents to carbonyl compounds: Chirality transfer, multiplication, and amplification. Angew. Chem. Int. Ed. Engl. 1991, 30, 49–69. [Google Scholar] [CrossRef]
  18. Genov, M.; Dimitrov, V.; Ivanova, V. New δ-aminoalcohol for the enantioselective addition of dialkylzincs to aldehydes. Tetrahedron Asymmetry 1997, 8, 3703–3706. [Google Scholar] [CrossRef]
  19. Ramachandran, G.; Sathiyanarayanan, K.I.; Sathishkumar, M.; Rathore, R.S.; Giridharan, P. Dual behavior of ammonium acetate for the synthesis of diverse symmetrical/unsymmetrical bis[1,3]oxazines possessing anticancer activity. Synth. Commun. 2015, 45, 2227–2239. [Google Scholar] [CrossRef]
  20. Ejfler, J.; Krauzy-Dziedzic, K.; Szafert, S.; Lis, T.; Sobota, P. Novel chiral and achiral benzoxazine monomers and their thermal polymerization. Macromolecules 2009, 42, 4008–4015. [Google Scholar] [CrossRef]
  21. Buckley, B.R.; Boxhall, J.Y.; Page, P.C.B.; Chan, Y.; Elsegood, M.R.; Heaney, H.; Holmes, K.E.; McIldowie, M.J.; McKee, V.; McGrath, M.J. Mannich and O-alkylation reactions of tetraalkoxyresorcin[4]arenes–the use of some products in ligand-assisted reactions. Eur. J. Org. Chem. 2006, 22, 5117–5134. [Google Scholar] [CrossRef]
  22. Barluenga, J.; Joglar, J.; González, F.J.; Fustero, S.; Krüger, C.; Tsay, Y.-H. Synthesis of 1, 3-amino alcohols from 2-aza-1, 3-dienes by reduction of 5, 6-dihydro-2H-1, 3-oxazines. Synthesis 1991, 387–393. [Google Scholar] [CrossRef]
  23. Hua, B.; Ding, Y.; Alimi, L.O.; Moosa, B.; Zhang, G.; Baslyman, W.S.; Sessler, J.; Khashab, N.M. Tuning the porosity of triangular supramolecular adsorbents for superior haloalkane isomer separations. Chem. Sci. 2021, 12, 12286–12291. [Google Scholar] [CrossRef] [PubMed]
  24. Ding, Y.; Alimi, L.O.; Du, J.; Hua, B.; Dey, A.; Yu, P.; Khashab, N.M. Pillar[3]trianglamines: Deeper cavity triangular macrocycles for selective hexene isomer separation. Chem. Sci. 2022, 13, 3244–3248. [Google Scholar] [CrossRef] [PubMed]
  25. Rivera, A.; Quiroga, D.; Ríos-Motta, J.; Kučeraková, M.; Dušek, M. meso-4, 4′-Dimethoxy-2, 2′-{[(3aR, 7aS)-2, 3, 3a, 4, 5, 6, 7, 7a-octahydro-1H-benzimidazole-1, 3-diyl] bis (methylene)} diphenol. Acta Crystallogr. E Crystallogr. Commun. 2013, 69, o1057–o1058. [Google Scholar] [CrossRef]
Scheme 1. Condensation of (a) naphthalene-diols 1 or 4 and (S)-(−)-1-phenylethane-1-amine (2) and (b) cyclohexane-1,2-diamine (6) and naphthalene-2-ol (7) with formaldehyde.
Scheme 1. Condensation of (a) naphthalene-diols 1 or 4 and (S)-(−)-1-phenylethane-1-amine (2) and (b) cyclohexane-1,2-diamine (6) and naphthalene-2-ol (7) with formaldehyde.
Crystals 13 01495 sch001
Figure 1. Comparison of 1H NMR spectra (CDCl3 at 298 K) of (a) a mixture of 8 (blue) and 9 (red) (20:19 ratio); (b) crude product 8 (contains 16% of 9) and (c) pure compound 9.
Figure 1. Comparison of 1H NMR spectra (CDCl3 at 298 K) of (a) a mixture of 8 (blue) and 9 (red) (20:19 ratio); (b) crude product 8 (contains 16% of 9) and (c) pure compound 9.
Crystals 13 01495 g001
Figure 2. Structures and configuration of (R,R)-9 and (S,S)-9.
Figure 2. Structures and configuration of (R,R)-9 and (S,S)-9.
Crystals 13 01495 g002
Scheme 2. The enantioselective addition of Et2Zn to aldehydes.
Scheme 2. The enantioselective addition of Et2Zn to aldehydes.
Crystals 13 01495 sch002
Figure 3. ORTEP [15] view of the molecules present in the ASU of (a) 3, (b) (R,R)-9, with appropriate labeling scheme. The thermal ellipsoids are given with 50% probability level.
Figure 3. ORTEP [15] view of the molecules present in the ASU of (a) 3, (b) (R,R)-9, with appropriate labeling scheme. The thermal ellipsoids are given with 50% probability level.
Crystals 13 01495 g003
Figure 4. Visualization of the detected hydrogen-bonding interactions of the O–H…N type (red dots), weak C–H…O interactions (the strongest one is labeled, e.g., C5–H5A…O1, 2.495 Å) and short contacts stabilizing the crystal structure of compound (R,R)-9.
Figure 4. Visualization of the detected hydrogen-bonding interactions of the O–H…N type (red dots), weak C–H…O interactions (the strongest one is labeled, e.g., C5–H5A…O1, 2.495 Å) and short contacts stabilizing the crystal structure of compound (R,R)-9.
Crystals 13 01495 g004
Figure 5. DSC thermograms of (a) 3, (b) 5, (c) (R,R)-9, (d) (S,S)-9.
Figure 5. DSC thermograms of (a) 3, (b) 5, (c) (R,R)-9, (d) (S,S)-9.
Crystals 13 01495 g005
Figure 6. Comparison between the collected powder X-ray diffraction data (in blue color) and that calculated from single crystal (in red color) for (a) compound 3 and (b) compound (R,R)-9.
Figure 6. Comparison between the collected powder X-ray diffraction data (in blue color) and that calculated from single crystal (in red color) for (a) compound 3 and (b) compound (R,R)-9.
Crystals 13 01495 g006
Table 1. Characteristic chemical shifts (in ppm) of selected protons allowing one to distinguish between compounds 8 and 9 in mixtures.
Table 1. Characteristic chemical shifts (in ppm) of selected protons allowing one to distinguish between compounds 8 and 9 in mixtures.
Some Characteristic
Protons
89
1H NMR
(Number of Protons)
1H NMR
(Number of Protons)
C(H)-114.36 (4H)4.21 (2H)
4.39 (2H)
C(H)-133.01–3.16 (2H)2.45–2.60 (2H)
C(H)-164.98 (4H)3.66 (2H)
Table 2. Optimization the condensation of cyclohexane-1,2-diamine (6) in racemic form, naphthalen-2-ol (7) and formaldehyde with the purpose to alter the ratio of 8:9.
Table 2. Optimization the condensation of cyclohexane-1,2-diamine (6) in racemic form, naphthalen-2-ol (7) and formaldehyde with the purpose to alter the ratio of 8:9.
HCHO
(Equivalent)
SolventTime (h)T (°C)Molar Ratio 8:9 3Yield 5 of 9 (%)
110 1EtOH3205:2-
210 1EtOH5202:1-
310 1EtOH1505:2-
410 1EtOH6502:1-
510 1EtOH185020:19 4-
610 2EtOH24507:1034 6
710 1EtOH2.5reflux2:5-
810 2toluene24505:1-
92 1MeOH2reflux0:150 7
102 2MeOH2reflux0:146 7
112 2EtOH2reflux0:134 7
122 2EtOH24500:163 6
1 A 37% aq. solution of formaldehyde; 2 paraformaldehyde; 3 molar ratio determined by 1H NMR spectroscopy of crude mixture; 4 ratio determined after washing the crude reaction mixture with acetonitrile; 5 yields of isolated compound; 6 after column chromatography; 7 after filtration of the crystallized 9 from the reaction mixture.
Table 3. Catalytic activity and efficiency of (R,R)-9 and (S,S)-9 as ligands in the enantioselective addition of Et2Zn to aldehydes.
Table 3. Catalytic activity and efficiency of (R,R)-9 and (S,S)-9 as ligands in the enantioselective addition of Et2Zn to aldehydes.
EntryAldehydeLigandsTime (h)Yield 11 (%) 1ee (%)
Configuration
12-Methoxybezaldehyde(R,R)-9726653 (R) 2
22-Methoxybezaldehyde(S,S)-9248058 (S) 2
31-Naphthaldehyde(R,R)-9524539 (R) 3
41-Naphthaldehyde(S,S)-91205430 (S) 3
54-Chlorobenzaldehyde(R,R)-91449049 (R) 3
64-Chlorobenzaldehyde(S,S)-91203050 (S) 3
7Cyclohexanecaboxaldehyde(R,R)-9244638 (S) 2
8Cyclohexanecaboxaldehyde(S,S)-9244940 (R) 2
9Ferrocenecarboxaldehyde(R,R)-9487938 (R) 3
10Ferrocenecarboxaldehyde(S,S)-91207442 (S) 3
1 The yield of isolated products in pure form after column chromatography; 2 enantiomeric excess was determined by GC analyses; 3 enantiomeric excess was determined by HPLC analyses.
Table 4. Selected bond lengths and angles for compounds 3 and (R,R)-9.
Table 4. Selected bond lengths and angles for compounds 3 and (R,R)-9.
Compound 3Compound (R,R)-9
Bond Lengths, ÅBond Lengths, Å
AtomAtomÅAtomAtomÅAtomAtomÅAtomAtomÅ
O1C31.373(5)N22C231.466(5)O1C251.360(4)C1C51.510(3)
O2C21.363(5)N12C131.491(6)O2C261.357(3)C7C81.511(3)
C3C21.420(6)N22C241.490(6)N2C51.468(3)C9C101.422(4)
O1C211.456(6)C3C41.362(6)N3C71.475(3)C3C41.419(4)
O2C321.456(6)C2C11.367(6)N2C61.476(3)C15C161.508(3)
N12C321.429(6)C7C81.381(9)N3C61.468(3)C19C201.523(5)
N12C111.467(6)C24C251.528(7)N3C151.467(3)C17C181.527(4)
N22C211.431(5)C13C141.521(6)N2C161.471(3)C18C191.524(5)
Bond Angles, °Bond Angles, °
AtomAtomAtom° AtomAtomAtom°
C2O2C32113.6(4) C6N3C7113.0(2)
C3O1C21114.0(3) C6N2C5113.1(2)
N12C32O2113.7(4) N3C6N2106.1(2)
N12C11C1111.9(4) N3C15C20117.3(2)
C32N12C11108.3(3) N2C16C17118.0(2)
N12C13C15109.0(4) C25C8C7121.6(3)
N22C23C4111.2(3) C26C1C5118.7(2)
C21N22C23107.9(3) C16C17C18107.6(3)
N22C24C26110.3(4) C15C20C19108.8(3)
Table 5. Observed hydrogen bonds and weak interactions in (R,R)-9.
Table 5. Observed hydrogen bonds and weak interactions in (R,R)-9.
DHAd(D-H)/Åd(H-A)/Åd(D-A)/ÅD-H-A/°
O1H1N30.821.862.596(3)148.4
O2H2N20.821.932.650(3)146.4
C5H5AO2 10.972.493.223(3)131.7
Symmetry operation: 1 −1/2 + x, 1/2y, 1 − z.
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

Tavlinova-Kirilova, M.; Dikova, K.; Marinova, M.K.; Kamenova-Nacheva, M.; Rusew, R.; Sbirkova-Dimitrova, H.; Shivachev, B.; Kostova, K.; Dimitrov, V. Synthesis and Structural Analysis of Chiral Bis-dihydro[1,3]-naphthoxazines and Imidazolidine Derivatives Prepared by Three-Component Mannich-Type Condensation. Crystals 2023, 13, 1495. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst13101495

AMA Style

Tavlinova-Kirilova M, Dikova K, Marinova MK, Kamenova-Nacheva M, Rusew R, Sbirkova-Dimitrova H, Shivachev B, Kostova K, Dimitrov V. Synthesis and Structural Analysis of Chiral Bis-dihydro[1,3]-naphthoxazines and Imidazolidine Derivatives Prepared by Three-Component Mannich-Type Condensation. Crystals. 2023; 13(10):1495. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst13101495

Chicago/Turabian Style

Tavlinova-Kirilova, Maya, Krasimira Dikova, Maya K. Marinova, Mariana Kamenova-Nacheva, Rusi Rusew, Hristina Sbirkova-Dimitrova, Boris Shivachev, Kalina Kostova, and Vladimir Dimitrov. 2023. "Synthesis and Structural Analysis of Chiral Bis-dihydro[1,3]-naphthoxazines and Imidazolidine Derivatives Prepared by Three-Component Mannich-Type Condensation" Crystals 13, no. 10: 1495. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst13101495

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