Next Article in Journal
GGA and GGA + U Study of ThMn2Si2 and ThMn2Ge2 Compounds in a Body-Centered Tetragonal Ferromagnetic Phase
Next Article in Special Issue
Reagent-Controlled Highly Stereoselective Difluoromethylation: Efficient Access to Chiral α-Difluoromethylamines from Ketimines
Previous Article in Journal
Chemical Composition and In Vitro Antioxidant Activity of Sida rhombifolia L. Volatile Organic Compounds
Previous Article in Special Issue
Regioselective Synthesis of 5-Trifluoromethyl 1,2,4-Triazoles via [3 + 2]-Cycloaddition of Nitrile Imines with CF3CN
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Trifluoromethylated Monoterpene Amino Alcohols

by
Polina A. Petrova
1,
Denis V. Sudarikov
1,2,*,
Larisa L. Frolova
1,
Roman V. Rumyantcev
3,
Svetlana A. Rubtsova
1 and
Aleksandr V. Kutchin
1
1
Institute of Chemistry, FRC “Komi Scientific Centre”, Ural Branch of the Russian Academy of Sciences Pervomayskaya St. 48, 167000 Syktyvkar, Russia
2
Biologically Active Terpenoids Laboratory, Kazan Federal University, 18 Kremlevskaya Street, 420008 Kazan, Russia
3
G.A. Razuvaev Institute of Organometallic Chemistry of Russian Academy of Sciences, 49 Tropinina St., 603950 Nizhny Novgorod, Russia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(20), 7068; https://doi.org/10.3390/molecules27207068
Submission received: 11 October 2022 / Revised: 14 October 2022 / Accepted: 18 October 2022 / Published: 20 October 2022
(This article belongs to the Special Issue Insights for Organofluorine Chemistry)

Abstract

:
For the first time, monoterpene trifluoromethylated β-hydroxy-benzyl-O-oximes were synthesized in 81–95% yields by nucleophilic addition of the Ruppert–Prakash reagent (TMSCF3) to the corresponding β-keto-benzyl-O-oximes based on (+)-nopinone, (−)-verbanone and (+)-camphoroquinone. Trifluoromethylation has been determined to entirely proceed chemo- and stereoselective at the C=O rather than C=N bond. Trifluoromethylated benzyl-O-oximes were reduced to the corresponding α-trifluoromethyl-β-amino alcohols in 82–88% yields. The structure and configuration of the compounds obtained have been established.

1. Introduction

Derivatives of monoterpenoids have a wide spectrum of antimicrobial activity against certain pathogenic species of bacteria and fungi [1]. The binding site of cyclic terpene hydrocarbons is located in the cell membrane of pathogenic microorganisms [2]. Many monoterpenoids, such as α- and β-pinenes, γ-terpinene, limonene, are capable of inhibiting respiration and other energy-dependent processes localized in fungal cell membranes [3].
It is known that the introduction of fluorine-containing groups into the molecule of a substance leads to an increase in membrane permeability, as well as an increase in resistance to biodegradation in comparison with their non-fluorinated analogues [4,5]. For this reason, about 25% of all modern pharmaceuticals contain fluorine atoms [6,7,8]. These transformations can lead to a change in the biological activity of the resulting compounds, as well as a new way of substrate–receptor interactions in comparison with hydrocarbon analogues [9,10,11,12].
Natural asymmetric molecules are a good starting point for the synthesis of chiral compounds because they are usually enantiomerically pure, obtained from renewable sources, and in most cases inexpensive. Terpenes are excellent natural asymmetric building blocks: they are mainly produced by various plants, some of them can be converted into more complex compounds used, for example, as ligands or catalysts for asymmetric reactions [13].
It is known that chiral β-amino alcohols [14,15], including terpene ones [16,17,18,19], are organocatalysts for a wide range of reactions of asymmetric synthesis, such as Diels–Alder cycloaddition, 1,3-dipolar cycloaddition, aldol condensation, Michael addition, cascade cyclization, Morita–Baylis–Hillman reaction, Friedel–Kraftz alkylation of indoles, allylation of isatins, and epoxidation of olefins.
It has been shown that the introduction of fluoroalkyl groups, including the trifluoromethyl group, into many chiral ligands, chiral auxiliaries, and chiral substrates improved their ability to induce asymmetry in stereoselective reactions [20,21,22,23]. Chiral α-trifluoromethyl-β-amino alcohols improved the stereoselectivity of addition reactions of diethylzinc and the Reformatsky reagent to carbonyl compounds and imines compared to their non-fluorinated analogs [22,24]. The authors attribute the effect of increasing stereoselectivity and reaction rate to strong electron-withdrawing properties, a large steric effect, as well as electrostatic repulsion between the local negative charge of the trifluoromethyl group and the charge of attacking nucleophiles [22].
In addition, fluorine-containing compounds are easily identified by 19F NMR spectroscopy and related homo- and heterocorrelation techniques due to the fact that the nucleus of the 19F fluorine atom has a spin ½, with an unprecedented natural abundance (100%) and a relatively high gyromagnetic ratio (83% of γ1H), which results in a strong signal. The large range of chemical shifts observed for fluorine nuclei means that 19F NMR spectroscopy is a very sensitive source of changes in the electronic environment and changes in the local dielectric medium [25,26]. These advantages, as well as the absence of background noise and the considerable simplicity of 19F NMR spectra compared to 13C, 1H, 15N nuclei, make it possible to study fluorine-containing compounds in biological media [25,27,28], to study the mechanisms and kinetics of reactions [29,30,31], including catalytic reactions [32,33]. Chiral fluorine-containing derivatizing agents make it easy to evaluate the enantiomeric purity of amines and amino alcohols by 19F NMR [34,35,36,37,38].
Based on the foregoing, the synthesis of chiral trifluoromethylated amino alcohols based on natural monoterpenoids is of undoubted interest. In this work, based on verbanone, nopinone, and camphorquinone, we synthesized the corresponding ketooximes, benzyl-O-oximes, trifluoromethylated benzyl-O-oxymoalcohols, and trifluoromethylated amino alcohols.

2. Results

For this study, we used ketooximes 13 based on (+)-nopinone [39], (−)-verbanone [18], and (+)-camphoroquinone [40], which were synthesized according to already known methods. The corresponding benzyl-O-oximes 46 were obtained from the resulting oximes 13 in 68, 55, 45% yields, respectively (Scheme 1). The benzyl group was previously introduced to protect the OH group of the oxime before trifluoromethylation of the obtained compounds.
The IR spectra of compounds 46 contain absorption bands characteristic of the carbonyl group in the region of 1715–1717 cm−1, characteristic of the C=N–O group in the region of 1584–1684 cm−1. The 1H NMR spectra contain the signals of the protons of the methylene group C-1′ at 5.37 ppm for compound 4, at 5.4 ppm for 5; at 5.28 ppm for 6 and a multiplet of the phenyl fragment in the range 7.31–7.41 ppm for 46. In the 13C NMR spectra, the signals of these functional groups are present at 77.9 ppm for 4, at 78 ppm for 5, at 77.3 ppm for 6 and the multiplet of the phenyl fragment in the regions 128.2–128.4 ppm for 46, respectively.
Significantly, oximes 46 have two reaction centers C=O and C=N bonds, and both of them can be subjected to trifluoromethylation, for example, as demonstrated in [41,42,43,44] for imines and sulfinimines, which react at the C=N bond, and for monoterpene ketooximes 46 undergoing trifluoromethylation to yield the products solely at the C=O bond.
Nucleophilic addition of the Ruppert–Prakash reagent–trifluoromethyltrimethylsilane (TMSCF3) [45] to β-keto-benzyl-O-oximes 46 at the double C=O bond is carried out in THF at 4 °C in an argon atmosphere in the presence of an initiator—cesium fluoride (CsF). At the first stage, trimethylsilyl ethers 79 are formed, which, after addition of tetrabutylammonium fluoride hydrate (TBAF·3H2O), form new trifluoromethyl alcohols 1012 in 81, 89 and 95% yields, respectively (Scheme 2).
The IR spectra of compounds 1012 contain absorption bands characteristic of the hydroxyl group in the region 3437–3560 cm−1, characteristic of the C=N–O group in the region 1618–1688 cm−1, absorption bands corresponding to the CF3 group at 1271, 1171, and 1096 cm−1 for 10; at 1283, 1169, 1105 cm−1 for 11; at 1265, 1180, and 1103 cm−1 for 12.
The 1H NMR spectra of compounds 1012 contain singlets of the proton of the hydroxyl group at 3.27 ppm for 10; 3.20 ppm for 11; 2.76 for 12 and a multiplet of the phenyl fragment in the range of 7.31–7.41 ppm for 1012. The 13C NMR spectra show quartets of the C-2 carbon atom at 77.9 ppm (JF 27.6 Hz) for 10; at 78.9 ppm (JF 27.6 Hz) for 12. There is a quartet of C-4 carbon at 78.6 ppm (JF 26.5 Hz) in the 13C NMR spectrum of compound 11. The quartet of the C-10 carbon of compound 10 is present at 125.0 ppm (JF 288.6 Hz). The quartets of the C-11 carbon atom of compounds 11, 12 are present at 125.2 ppm (JF 288.6 Hz) for 11; at 125.2 ppm (JF 287.5 Hz) for 12. Singlets of the trifluoromethyl group of compounds 1012 appear in the 19F NMR spectra in the range from −71.8 to −74.6 ppm.
For each of the benzyl-O-oximes 46 in the trifluoromethylation reaction, only (2S)-10, (4R)-11, (2R)-12 diastereomers are formed from two theoretically possible diastereomers in 81, 89 and 95% yields, respectively (Scheme 2).
The configuration of C-2, C-4 and C-2 atoms of compounds 1012, respectively, was established by 1H NMR NOESY spectroscopy by the presence of NOE interactions between the protons of the hydroxyl group and the C-8 methyl group in compounds 10 and 11, between the protons of the hydroxyl group and the C-9 methyl group of compound 12 (Figure 1).
Trifluoromethylated benzyl-O-oximes 1012 based on (+)-nopinone, (−)-verbanone and (+)-camphoroquinone were reduced with LiAlH4 to the corresponding amines isolated as hydrochlorides 1315 in 81–95% yields (Scheme 3).
The IR spectra of compounds 1315 contain absorption bands characteristic of the hydroxyl group in the region of 3298–3558 cm−1, characteristic of the NH3+ group in the region of 2922–3080 cm−1, absorption bands corresponding to the CF3 group in the regions of 1128–1198 cm−1 for 13, 1126–1194 cm−1 for 14, and 1121–1186 cm−1 for 15.
The 1H NMR spectra of compounds 1315 contain singlets of the protons of the OH and NH3+ groups at 4.75 ppm and there are no signals of the phenyl fragment compared to the original substrates. The 13C NMR spectra of compounds 13 and 15 contain quartets of the C-2 carbon atom at 78.2 ppm (JF 25.4 Hz) for 13; at 80.6 ppm (JF 26.5 Hz) for 15. There is a quartet of the C-4 carbon atom at 78.1 ppm (JF 26.5 Hz) in the 13C NMR spectrum of compound 14. The quartet of the trifluoromethyl group C-10 of compound 13 is at 125.3 ppm (JF 288.6 Hz). Quartets of the trifluoromethyl groups C-11 of compounds 14, 15 are present at 125.3 ppm (JF 287.5 Hz) for 14; at 125.4 ppm (JF 288.6 Hz) for 15, respectively. There are singlets in the range from −72.6 to −68.2 ppm in the 19F NMR spectra of 1315.
The configuration of the C-3 atom of compounds 1315 was established by 1H NOESY NMR spectroscopy by the presence of NOE interactions between the H-3 protons and the methyl group C-8 in 13 and 14, between the H-3 protons and the methyl group C-9 in the compound 15 (Figure 2).
A single crystal of free amine 16 was obtained after alkaline extraction of hydrochloride 14 with Et2O. The configuration of free amine 16 was confirmed by X-ray diffraction analysis (Figure 3). This compound crystallizes in the chiral space group P212121 of the orthorhombic system. There are two independent molecules (A and B) of 16 in the asymmetric unit cell. They have the same molecular structure. The root–mean–square deviation of atomic positions of A and B molecules is 0.056 Å. The carbon atoms C(1), C(2), C(3), C(4), and C(5) lie almost in the same plane. The average deviation of atoms from the plane is 0.085 Å. The trifluoromethyl group, the amino group, and the methylene group are on the same side of this plane. The main geometric characteristics in 16 are in good agreement with related carbocyclic compounds [18,46].
In a crystal, neighboring molecules are oriented in such a way that intermolecular hydrogen bonds are realized O-H...O (2.03 Å), O-H...N (2.07 Å), N-H...N (2.47 Å), and N-H...O (2.48 Å). As a result, endless molecular chains A-B-A-B are formed.

3. Materials and Methods

3.1. General Information

FT-IR spectra were recorded on a Shimadzu IR Prestige 21 on thin films or KBr pellets; ν in cm−1. 1H and 13C NMR spectra were registered on a Bruker Avance 300 spectrometer (300.17 MHz for 1H, 75.48 MHz for 13C and 282.44 MHz for 19F) in CDCl3, J in Hz (See Supplementary Materials). The signals were assigned using COSY, NOESY, HSQC, HMBC techniques, and 13C NMR spectra in J-modulation mode. Automatic analyzer EA 1110 CHNS-O was employed for elemental analysis. The melting points were measured on a Sanyo Gallenkamp MPD350.BM3.5 and were not corrected. Optical rotations were performed with automatized digital polarimeter Optical Activity PolAAr 3001. Thin layer chromatography (TLC) was performed on Sorbfil plates; spots were visualized by treatment with 10% phosphomolybdic acid in ethanol, 5% vanillin and 0.005% H2SO4 in ethanol, 5% KMnO4, and 0.005% H2SO4 in H2O. Silica gel 60 (70–230 mesh, Alfa Aesar, Lancashire, UK) was used for column chromatography (CC). For both TLC and CC the same eluent systems were used.
X-ray Data Collection and Structure Refinement. The diffraction data for compound 16 were collected on a Bruker D8 Quest diffractometer (Mo-Kα radiation, ω-scan technique, λ = 0.71073 Å) at 298(2) K. The intensity data were integrated by the SAINT [47] program. The structure was solved by dual methods [48] and was refined on F hkl 2 using the SHELXTL package [49]. The SADABS program [50] was used to perform absorption corrections. All non-hydrogen atoms were refined anisotropically. All H-atoms, with the exception of hydrogens of the hydroxyl and amino groups, were placed in calculated positions and were refined using a riding model (Uiso(H) = 1.5Ueq(C) for CH3 groups and Uiso(H) = 1.2Ueq(C) for other groups). The H(1)-H(6) atoms in 16 were located from the differential Fourier map and were refined isotropically. CCDC 2191818 contains the supplementary crystallographic data accessed on 19 October 2022. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via https://www.ccdc.cam.ac.uk/structures.
Commercially available reagents such as (trifluoromethyl)trimethylsilane TMSCF3 (purity 98%, Alfa Aesar, Lancashire, UK), caesium fluoride CsF (purity 98%, Alfa Aesar, Lancashire, UK), tetra-N-butylammonium fluoride trihydrate TBAF·3H2O (purity 98%, Alfa Aesar, Lancashire, UK), lithium aluminum hydride LiAlH4 (purity 95%, Sigma-Aldrich, St. Louis, USA) were used directly as supplied without further purification. All solvents used for the reactions were distilled. (1R,5R,E)-3-(Hydroxyimino)-6,6-dimethylbicyclo[3.1.1]heptan-2-one (1), mp 180 °C, [39], (1S,4S,5S,Z)-3-(hydroxyimino)-4,6,6-trimethylbicyclo[3.1.1]heptan-2-one (2), mp 135 °C, [18], and (1S,4R,E)-3-(hydroxyimino)-1,7,7-trimethylbicyclo[2.2.1]heptan-2-one (3), mp 118 °C, [40] were synthesized in accordance with known methods.

3.2. General Procedure for the Synthesis of Benzyl-O-Oximes 4 and 5

In a two-necked flask equipped with a stirrer and a reflux condenser, oxime 1 or 2 (2.63 mmol) in 15 mL of acetonitrile was placed under argon. Cs2CO3 (5.27 mmol) was then added, and benzyl chloride (5.27 mmol) was added dropwise after 5 min of stirring. The resulting mixture was stirred for 3 h at room temperature. The reaction progress was monitored by TLC (eluent, chloroform). The solvent was distilled off under vacuum, H2O (30 mL) was added to the residue, extracted with Et2O, the organic layer was washed with brine and dried over Na2SO4. The solvent was distilled off under reduced pressure. The reaction products were isolated by silica gel column chromatography.
(1R,5R,E)-3-((Benzyloxy)imino)-6,6-dimethylbicyclo[3.1.1]heptan-2-one (4). Yield: 68%; light brown oil; [ α ] D 25 = +19.70 (c = 0.99 in CHCl3); Rf 0.38 (petr. ether/EtOAc, 3:1); 1H NMR (CDCl3, δ, ppm, J/Hz): 0.91 (s, 3H, H8), 1.38 (s, 3H, C9H3), 1.53 (d, 1H, J = 11.0, H7α), 2.25–2.32 (m, 1H, H5), 2.65–2.88 (m, 4H, H1, H7β, H4α, H4β), 5.37 (s, 2H, H1′α, H1′β), 7.32–7.41 (m, 5H, H3′, H4′, H5′, H6′, H7′); 13C NMR (CDCl3, δ, ppm): 21.5 (C8), 26.2 (C9), 28.0 (C7), 28.6 (C4), 37.6 (C5), 41.7 (C6), 56.6 (C1), 77.9 (C1′), 128.2 (C5′), 128.3 (C4′,6′), 128.4 (C3′,7′), 136.6 (C2′), 152.7 (C3), 198.2 (C2); IR spectrum (KBr, ν, cm−1): 1715 (C=O), 1684 (C=N−O); elemental analysis calcd (%) for C16H19NO2: C 74.68, H 7.44, N 5.44; found: C 74.08, H 7.26, N 5.14.
(1S,4S,5S,E)-3-((Benzyloxy)imino)-4,6,6-trimethylbicyclo[3.1.1]heptan-2-one (5). Yield: 55%; light brown oil; [ α ] D 25 = −1.2 (c = 1.0 in CHCl3); Rf 0.29 (CHCl3/petr. ether, 2:1); 1H NMR (CDCl3, δ, ppm, J/Hz): 1.05 (s, 3H, H8), 1.37 (d, J = 7.2, 3H, H10), 1.39 (s, 3H, C9H3), 1.45 (d, 1H, J = 11.0, H7α), 2.17 (td, 1H, J = 5.8, 3.0, H1), 2.66 (ddd, 1H, J = 11.0, 6.1, 5.8, H7β), 2.73 (t, 1H, J = 5.8, H5), 3.18 (qd, 1H, J = 7.1, 3.0, H2), 5.32–5.42 (m, 2H, H1′α, H1′β), 7.31–7.41 (m, 5H, H3′, H4′, H5′, H6′, H7′); 13C NMR (CDCl3, δ, ppm): 16.5 (C10), 23.9 (C8), 27.3 (C9), 28.2 (C7), 36.9 (C2), 42.3 (C6), 45.7 (C1), 56.6 (C5), 78.0 (C1′), 128.1 (C5′), 128.2 (C4′,6′), 128.4 (C3′,7′), 136.6 (C2′), 156.3 (C3), 199.0 (C4); IR spectrum (KBr, ν, cm−1): 1717 (C=O), 1584 (C=N−O); elemental analysis calcd (%) for C17H21NO2: C 75.25, H 7.80, N 5.16; found: C 75.05, H 7.52, N 5.01.
(1S,4R,E)-3-((Benzyloxy)imino)-1,7,7-trimethylbicyclo[2.2.1]heptan-2-one (6). In a two-necked flask equipped with a stirrer and a reflux condenser, (+)-camphorquinone oxime 3 (0.59 mmol) in 3 mL of THF (dry) was placed under argon. After ice-bath cooling the mixture, t-BuOK (0.65 mmol) was added and the flask was purged with argon. Benzyl chloride (2.06 mmol) was added after 20 min of stirring. The resulting mixture was stirred overnight at room temperature. The progress of the reaction was monitored by TLC (eluent, petr.ether:EtOAc, 10:1). At the end of the reaction, H2O (20 mL) was added, the reaction mixture was extracted with diethyl ether, the organic layer was washed with brine and dried over Na2SO4. The solvent was distilled off under reduced pressure. The reaction product was isolated by silica gel column chromatography. Yield: 45%; light brown oil; [ α ] D 25 = −134.4 (c = 1.15 in CHCl3); Rf 0.23 (petr. ether/Et2O, 5:1); 1H NMR (CDCl3, δ, ppm, J/Hz): 0.88 (s, 3H, H8), 0.98 (s, 3H, C9H3), 1.03 (s, 3H, C10H3), 1.47–1.62 (m, 2H, H5α, H6α), 1.71–1.83 (m, 1H, H6β), 1.96–2.06 (m, 1H, H5β), 3.21 (d, 1H, J = 4.4, H4), 5.28 (s, 2H, H1′α, H1′β), 7.33–7.37 (m, 5H, H3′, H4′, H5′, H6′, H7′); 13C NMR (CDCl3, δ, ppm): 9.0 (C10), 17.6 (C9), 20.6 (C8), 23.9 (C5), 30.7 (C6), 44.8 (C7), 47.4 (C4), 58.5 (C1), 77.3 (C1′), 128.0 (C5′), 128.1 (C4′,6′), 128.4 (C3′,7′), 137.0 (C2′), 159.3 (C3), 203.9 (C2); IR spectrum (KBr, ν, cm−1): 1715 (C=O), 1634 (C=N−O); elemental analysis calcd (%) for C17H21NO2: C 75.25, H 7.80, N 5.16; found: C 75.10, H 7.56, N 5.09.

3.3. General Procedure for Trifluoromethylation of β-Keto-Benzyl-O-Oximes 46

In a two-necked flask equipped with a stirrer and reflux condenser, cooled in an ice bath under argon, benzyl-O-oxime 4 (or 5, 6, 1.36 mmol) was placed in 6 mL of THF (dry). After cooling the mixture, CsF (0.68 mmol) and TMSCF3 (4.08 mmol) were added with stirring. The resulting mixture was stirred for 4 h (control by TLC until the disappearance of the substrate). After that, the ice-bath was removed and TBAF·3H2O (1.36 mmol) was added. The progress of the reaction was monitored by TLC (eluent, pet.ether:EtOAc, 3:1). A saturated solution of NH4Cl (20 mL) was added, the reaction products were extracted with diethyl ether, the organic layer was washed with brine and dried over Na2SO4. The solvent was distilled off under vacuum. The reaction products were isolated by column chromatography.
((1R,2S,5R,E)-2-Hydroxy-6,6-dimethyl-2-(trifluoromethyl)bicyclo[3.1.1]heptan-3-one O-benzyl oxime (10). Yield: 81%; light brown oil; [ α ] D 25 = −9.69 (c = 0.98 in CHCl3); Rf 0.39 (petr. ether/EtOAc, 10:1); 1H NMR (CDCl3, δ, ppm, J/Hz): 0.95 (s, 3H, H8), 1.35 (s, 3H, C9H3), 1.53 (d, 1H, J = 11.3, H7α), 1.99–2.04 (m, 1H, H5), 2.34 (t, 1H, J = 5.9, H1), 2.40–2.54 (m, 2H, H7β, H4α), 3.02 (dt, 1H, J = 18.7, 3.0, H4β), 3.27 (s, 1H, OH), 5.16–5.25 (m, 2H, H1′α, H1′β), 7.34–7.39 (m, 5H, H3′, H4′, H5′, H6′, H7′); 13C NMR (CDCl3, δ, ppm): 21.9 (C8), 26.7 (C7), 26.8 (C9), 30.2 (C4), 37.4 (C5), 40.0 (C6), 45.2 (C1), 76.7 (C1′), 77.9 (q, JF = 27.6, C2), 125.0 (q, JF = 288.6, C10), 128.0 (C5′,4′,6′), 128.4 (C3′,7′), 137.4 (C2′), 156.7 (C3); 19F NMR (CDCl3, δ, ppm, J/Hz): –74.6 (s, 3F, C11F3); IR spectrum (KBr, ν, cm−1): 3560 (OH), 1627 (C=N−O), 1271, 1171, 1096 (CF3); elemental analysis calcd (%) for C17H20F3NO2: C 62.38, H 6.16, N 4.28; found: C 62.01, H 6.12, N 4.17.
(1S,2R,4S,5S,E)-2-Hydroxy-4,6,6-trimethyl-2-(trifluoromethyl)bicyclo[3.1.1]heptan-3-one O-benzyl oxime (11). Yield: 89%; light brown oil; [ α ] D 26 = +17.9 (c = 1.0 in CHCl3); Rf 0.20 (petr. ether/CH2Cl2, 2:1); 1H NMR (CDCl3, δ, ppm, J/Hz): 1.11 (s, 3H, H8), 1.35 (s, 3H, C9H3), 1.41 (d, 1H, J = 11.6, H7α), 1.47 (d, J = 7.2, 3H, H10), 1.85 (td, 1H, J = 5.8, 2.0, H1), 2.33 (t, 1H, J = 5.8, H5), 2.42 (dt, 1H, J = 11.6, 6.1, H7β), 2.92 (qd, 1H, J = 7.1, 2.0, H2), 3.20 (s, 1H, OH), 5.13–5.22 (m, 2H, H1′α, H1′β), 7.32–7.40 (m, 5H, H3′, H4′, H5′, H6′, H7′); 13C NMR (CDCl3, δ, ppm): 19.1 (C10), 24.2 (C8), 27.0 (C7), 27.9 (C9), 39.2 (C2), 39.7 (C6), 44.9 (C5), 46.7 (C1), 77.0 (C1′), 78.6 (q, JF = 26.5, C4), 125.2 (q, JF = 288.6, C11), 127.9 (C5′), 128.0 (C4′,6′), 128.4 (C3′,7′), 137.4 (C2′), 158.5 (C3); 19F NMR (CDCl3, δ, ppm, J/Hz): –74.2 (s, 3F, C11F3); IR spectrum (KBr, ν, cm−1): 3557 (OH), 1618 (C=N−O), 1265, 1180, 1103 (CF3); elemental analysis calcd (%) for C18H22F3NO2: C 63.33, H 6.50, N 4.10; found: C 63.13, H 6.34, N 4.32.
(1R,3R,4S,E)-3-Hydroxy-4,7,7-trimethyl-3-(trifluoromethyl)bicyclo[2.2.1]heptan-2-one O-benzyl oxime (12). Yield: 95%; light brown oil; [ α ] D 25 = −37.9 (c = 0.82 in CHCl3); Rf 0.37 (petr. ether/Et2O, 10:1); 1H NMR (CDCl3, δ, ppm, J/Hz): 0.94 (s, 3H, H8), 1.04 (s, 3H, C9H3), 1.07 (s, 3H, C10H3), 1.36–1.45 (m, 1H, H5α), 1.58–1.89 (m, 3H, H6α, H6β, H5β), 2.76 (s, 1H, OH), 3.12 (d, 1H, J = 4.4, H4), 5.09–5.18 (m, 2H, H1′α, H1′β), 7.30–7.39 (m, 5H, H3′, H4′, H5′, H6′, H7′); 13C NMR (CDCl3, δ, ppm): 9.8 (C10), 18.7 (C9), 21.9 (C8), 22.3 (C5), 28.9 (C6), 48.0 (C7), 48.3 (C4), 52.6 (C1), 76.2 (C1′), 78.9 (q, JF = 27.6, C2), 125.2 (q, JF = 287.5, C11), 127.8 (C5′), 127.9 (C4′,6′), 128.3 (C3′,7′), 137.8 (C2′), 164.7 (C3); 19F NMR (CDCl3, δ, ppm, J/Hz): –71.8 (s, 3F, C11F3); IR spectrum (KBr, ν, cm−1): 3437 (OH), 1688 (C=N−O), 1283, 1169, 1105 (CF3); elemental analysis calcd (%) for C18H22F3NO2: C 63.33, H 6.50, N 4.10; found: C 63.71, H 6.62, N 4.38.

3.4. General Procedure for the Preparation of Trifluoromethylated Amino Alcohols 1315

Trifluoromethylated β-trifluoromethyl-β-hydroxy-benzyl-O-oxime 10 (or 11, 12, 0.54 mmol) in 15 mL of dry Et2O was placed into a two-necked flask equipped with a stirrer, cooled to 4 °C in an argon atmosphere. LiAlH4 (1.65 mmol) was added in portions with stirring. The reaction mixture was stirred at room temperature for a day, the progress of the reaction was monitored by TLC (eluent, pet.ether: EtOAc, 10:1). The mixture was cooled again in an ice bath, then Et2O (20 mL) was added and 5% KOH solution was carefully poured until phase separation and the aqueous layer was extracted with Et2O. The combined organic phases were washed with brine and dried over Na2SO4. The solvent was distilled off in vacuo. The resulting amines were isolated in the hydrochlorides form in Et2O solution by blowing dry HCl into the flask until the precipitation ceased. The hydrochlorides were purified by washing with a mixture of hexane- Et2O (1:1).
(1R,2S,3S,5R)-2-Hydroxy-6,6-dimethyl-2-(trifluoromethyl)bicyclo[3.1.1]heptan-3-ammonium chloride (13). Yield: 88%; white powder; mp = 218 °C (decomposition); [ α ] D 27 = +33.3 (c = 0.74 in MeOH); 1H NMR (D2O, δ, ppm, J/Hz): 1.04 (s, 3H, H8), 1.30 (s, 3H, C9H3), 1.36 (d, 1H, J = 11.6, H7α), 1.78 (ddd, 1H, J = 14.1, 6.4, 1.5, H4β), 2.05–2.12 (m, 1H, H5), 2.37 (t, 1H, J = 5.8, H1), 2.40–2.49 (m, 1H, H7β), 2.68 (ddt, 1H, J = 13.8, 11.0, 3.2, H4β); 4.04 (dd, 1H, J = 10.3, 6.2, H3), 4.75 (s, 4H, OH, NH3+); 13C NMR (D2O, δ, ppm): 22.6 (C8), 26.0 (C7), 26.5 (C9), 31.9 (C4), 38.6 (C6), 39.0 (C5), 47.2 (C1), 51.3 (C3), 78.2 (q, JF = 25.4, C2), 125.3 (q, JF = 288.6, C10); 19F NMR (D2O, δ, ppm, J/Hz): –72.6 (s, 3F, C10F3); IR spectrum (KBr, ν, cm−1): 3558 (OH), 2959 (NH3+), 1601 (N), 1198, 1148, 1128 (CF3); elemental analysis calcd (%) for C10H17ClF3NO: C 46.25, H 6.60, N 5.39; found: C 46.61, H 6.82, N 5.51.
(1S,2R,3R,4S,5S)-2-Hydroxy-4,6,6-trimethyl-2-(trifluoromethyl)bicyclo[3.1.1]heptan-3-ammonium chloride (14). Yield: 89%; white powder; mp = 220 °C (decomposition); [ α ] D 25 = −17.7 (c = 0.5 in MeOH); 1H NMR (D2O, δ, ppm, J/Hz): 1.15 (s, 3H, H8), 1.25 (d, J = 6.9, 3H, H10), 1.34 (s, 3H, C9H3), 1.34 (d, 1H, J = 11.3, H7α), 2.00 (t, 1H, J = 5.5, H1), 2.92 (quin, 1H, J = 7.5, H2), 2.39 (t, 1H, J = 5.8, H5), 2.49 (dt, 1H, J = 11.8, 6.1, H7β), 3.91 (d, 1H, J = 8.8, H3), 4.75 (s, 4H, OH, NH3+); 13C NMR (D2O, δ, ppm): 18.4 (C10), 23.9 (C8), 27.9 (C7), 27.9 (C9), 39.0 (C6), 40.1 (C2), 46.3 (C1), 48.5 (C5), 59.7 (C3), 78.1 (q, JF = 26.5, C4), 125.3 (q, JF = 287.5, C11); 19F NMR (D2O, δ, ppm, J/Hz): –71.1 (s, 3F, C11F3); IR spectrum (KBr, ν, cm−1): 3298 (OH), 2955, 2922 (NH3+), 1580 (N-H), 1194, 1150, 1126 (CF3); elemental analysis calcd (%) for C11H19ClF3NO: C 48.27, H 7.00, N 5.12; found: C 48.11, H 6.82, N 5.12. A single crystal of free amine 16 was obtained after alkaline extraction of hydrochloride 14 with Et2O. A colorless prismatic crystal of the orthorhombic system had size 0.56 × 0.47 × 0.42 mm, space group P212121, a = 7.8532(5), b = 15.7765(9), c = 19.0439(12) Å, α = β = γ = 90°, V = 2359.5(3) Å3, Z = 8, μ = 0.117 mm−1, dcalc = 1.336 g/cm3, F(000) = 1008. A dataset of 34,841 reflections was collected at scattering angles 2.139° < θ < 26.361°, of which 4820 were independent (Rint = 0.0287), including 4036 reflections with I > 2σ(I). The final refinement parameters were R1 = 0.0496, wR2 = 0.1013 (all data), R1 = 0.0376, wR2 = 0.0944 [I > 2σ(I)] with GooF = 1.062. Δρe = 0.190/−0.169 e Å–3; Flack parameter = −0.16(15).
(1R,2R,3R,4S)-3-Hydroxy-4,7,7-trimethyl-3-(trifluoromethyl)bicyclo[2.2.1]heptan-2-ammonium chloride (15). Yield: 82%; white powder; mp = 222 °C (decomposition); [ α ] D 26 = −11.9 (c = 0.8 in MeOH); 1H NMR (D2O δ, ppm, J/Hz): 0.95 (s, 3H, H8), 0.99 (s, 3H, C10H3), 1.14 (s, 3H, C9H3), 1.35–1.50 (m, 1H, H5α), 1.64–1.69 (m, 2H, H6α, H6β, H5β), 1.76–1.87 (m, 1H, H5β), 2.14 (t, 1H, J = 4.0, H4), 3.95–3.96 (m, 1H, H3), 4.75 (br.s, 4H, OH, NH3+); 13C NMR (D2O, δ, ppm): 10.3 (C10), 18.0 (C5), 19.3 (C9), 19.6 (C8), 27.9 (C6), 47.6 (C4), 48.0 (C7), 53.4 (C1), 60.7 (C3), 80.6 (q, JF = 26.5, C2), 125.4 (q, JF = 288.6, C11); 19F NMR (CDCl3, δ, ppm, J/Hz): –68.2 (s, 3F, C11F3); IR spectrum (KBr, ν, cm−1): 3339 (OH), 3080, 2964 (NH3+), 1585 (N-H), 1186, 1143, 1121 (CF3); elemental analysis calcd (%) for C11H19ClF3NO: C 48.27, H 7.00, N 5.12; found: C 48.61, H 7.12, N 5.34.

4. Conclusions

Thus, trifluoromethylated amino alcohols based on pinane and bornane monoterpenoids have been synthesized for the first time. The addition of the Ruppert–Prakash reagent to β-keto-benzyl-O-oximes, as well as the reduction of β-hydroxy-benzyl-O-oximes to the corresponding amino alcohols proceed stereoselectively with the formation of one of the diastereomers. Trifluoromethylation has been determined to entirely proceed at the C=O rather than C=N bond.
All compounds are isolated individually; the structure and configuration are proven by NMR and IR spectroscopy, elemental, and X-ray diffraction analysis. The obtained compounds may be of interest as biologically active substances and/or their precursors, as well as new chiral fluorine-containing auxiliaries, ligands or organocatalysts containing a trifluoromethyl group.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/molecules27207068/s1. The 1H NMR, 13C NMR and IR spectra of novel compounds.

Author Contributions

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

Funding

The study was financially supported by the Ministry of Science and Higher Education of the Russian Federation (state assignment, reg. № 122040600073-3).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was supported by Kazan Federal University Strategic Academic Leadership Program (PRIORITY-2030). This work was performed using equipment from the Center for Collective Use “Khimia” of the Institute of Chemistry of the Komi Scientific Center, Ural Branch of the Russian Academy of Sciences. The X-ray investigation of compound 16 was performed using the equipment of the center for collective use “Analytical Center of the IOMC RAS” with the financial support of the grant “Ensuring the development of the material and technical infrastructure of the centers for collective use of scientific equipment” (Unique identifier RF-2296.61321X0017, Agreement Number 075-15-2021-670).

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors, but can be specially prepared upon prior request.

References

  1. Nikitina, L.E.; Artemova, N.P.; Startseva, V.A. Natural and Thiomodified Monoterpenoids (Russian Edition); LAP Lambert: Saarbrücken, Germany, 2011; ISBN 978-3-8484-3023-9. [Google Scholar]
  2. Sikkema, J.; de Bont, J.A.; Poolman, B. Mechanisms of membrane toxicity of hydrocarbons. Microbiol. Rev. 1995, 59, 201–222. [Google Scholar] [CrossRef] [PubMed]
  3. Uribe, S.; Pena, A. Toxicity of allelopathic monoterpene suspensions on yeast dependence on droplet size. J. Chem. Ecol. 1990, 16, 1399–1408. [Google Scholar] [CrossRef] [PubMed]
  4. Meanwell, N.A. Synopsis of Some Recent Tactical Application of Bioisosteres in Drug Design. J. Med. Chem. 2011, 54, 2529–2591. [Google Scholar] [CrossRef] [PubMed]
  5. Gupta, S.P. Roles of Fluorine in Drug Design and Drug Action. Lett. Drug Des. Discov. 2019, 16, 1089–1109. [Google Scholar] [CrossRef]
  6. Wang, J.; Sánchez-Roselló, M.; Aceña, J.L.; del Pozo, C.; Sorochinsky, A.E.; Fustero, S.; Soloshonok, V.A.; Liu, H. Fluorine in Pharmaceutical Industry: Fluorine-Containing Drugs Introduced to the Market in the Last Decade (2001–2011). Chem. Rev. 2014, 114, 2432–2506. [Google Scholar] [CrossRef]
  7. Liu, X.; Xu, C.; Wang, M.; Liu, Q. Trifluoromethyltrimethylsilane: Nucleophilic Trifluoromethylation and Beyond. Chem. Rev. 2015, 115, 683–730. [Google Scholar] [CrossRef]
  8. Johnson, B.M.; Shu, Y.-Z.; Zhuo, X.; Meanwell, N.A. Metabolic and Pharmaceutical Aspects of Fluorinated Compounds. J. Med. Chem. 2020, 63, 6315–6386. [Google Scholar] [CrossRef]
  9. Yoder, N.C.; Kumar, K. Fluorinated amino acids in protein design and engineering. Chem. Soc. Rev. 2002, 31, 335–341. [Google Scholar] [CrossRef]
  10. Isanbor, C.; O’Hagan, D. Fluorine in medicinal chemistry: A review of anti-cancer agents. J. Fluor. Chem. 2006, 127, 303–319. [Google Scholar] [CrossRef]
  11. Hagmann, W.K. The Many Roles for Fluorine in Medicinal Chemistry. J. Med. Chem. 2008, 51, 4359–4369. [Google Scholar] [CrossRef]
  12. Qiu, X.; Qing, F. Recent Advances in the Synthesis of Fluorinated Amino Acids. Eur. J. Org. Chem. 2011, 2011, 3261–3278. [Google Scholar] [CrossRef]
  13. Ibn El Alami, M.S.; El Amrani, M.A.; Agbossou-Niedercorn, F.; Suisse, I.; Mortreux, A. Chiral Ligands Derived from Monoterpenes: Application in the Synthesis of Optically Pure Secondary Alcohols via Asymmetric Catalysis. Chem. -A Eur. J. 2015, 21, 1398–1413. [Google Scholar] [CrossRef] [PubMed]
  14. Reddy, U.V.S.; Chennapuram, M.; Seki, C.; Kwon, E.; Okuyama, Y.; Nakano, H. Catalytic Efficiency of Primary β-Amino Alcohols and Their Derivatives in Organocatalysis. Eur. J. Org. Chem. 2016, 2016, 4124–4143. [Google Scholar] [CrossRef]
  15. Nakano, H.; Owolabi, I.A.; Chennapuram, M.; Okuyama, Y.; Kwon, E.; Seki, C.; Tokiwa, M.; Takeshita, M. β-Amino Alcohol Organocatalysts for Asymmetric Additions. Heterocycles 2018, 97, 647–667. [Google Scholar] [CrossRef]
  16. Siyutkin, D.E.; Kucherenko, A.S.; Frolova, L.L.; Kuchin, A.V.; Zlotin, S.G. N-Pyrrolidine-2-ylmethyl)-2-hydroxy-3-aminopinanes as novel organocatalysts for asymmetric conjugate additions of ketones to α-nitroalkenes. Tetrahedron Asymmetry 2013, 24, 776–779. [Google Scholar] [CrossRef]
  17. Banina, O.A.; Sudarikov, D.V.; Nigmatov, A.G.; Frolova, L.L.; Slepukhin, P.A.; Zlotin, S.G.; Kutchin, A.V. Carane amino alcohols as organocatalysts in asymmetric aldol reaction of isatin with acetone. Russ. Chem. Bull. 2017, 66, 293–296. [Google Scholar] [CrossRef]
  18. Frolova, L.L.; Sudarikov, D.V.; Alekseev, I.N.; Banina, O.A.; Slepukhin, P.A.; Kutchin, A.V. Synthesis of new enantiomerically pure β-amino alcohols of the pinane series. Russ. J. Org. Chem. 2017, 53, 335–343. [Google Scholar] [CrossRef]
  19. Zlotin, S.G.; Banina, O.A.; Sudarikov, D.V.; Nigmatov, A.G.; Frolova, L.L.; Kutchin, A.V. Asymmetric aldol reaction of isatins with acetone in the presence of terpene amino alcohols. Mendeleev Commun. 2020, 30, 147–149. [Google Scholar] [CrossRef]
  20. Fache, F. Asymmetric fluorous catalysis: The particular case of nitrogen-containing chiral auxiliaries. New J. Chem. 2004, 28, 1277–1283. [Google Scholar] [CrossRef]
  21. Uneyama, K. Organofluorine Chemistry; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2008; ISBN 1-4051-7293-2. [Google Scholar]
  22. Harada, A.; Fujiwara, Y.; Katagiri, T. Improvement of the asymmetry-inducing ability of a trifluoromethylated amino alcohol by electron donation to a CF3 group. Tetrahedron Asymmetry 2008, 19, 1210–1214. [Google Scholar] [CrossRef]
  23. Mlostoń, G.; Obijalska, E.; Heimgartner, H. Synthesis of β-amino-α-trifluoromethyl alcohols and their applications in organic synthesis. J. Fluor. Chem. 2010, 131, 829–843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Xu, X.-H.; Qiu, X.-L.; Qing, F.-L. Synthesis and utilization of trifluoromethylated amino alcohol ligands for the enantioselective Reformatsky reaction and addition of diethylzinc to N-(diphenylphosphinoyl)imine. Tetrahedron 2008, 64, 7353–7361. [Google Scholar] [CrossRef]
  25. Dahanayake, J.N.; Kasireddy, C.; Karnes, J.P.; Verma, R.; Steinert, R.M.; Hildebrandt, D.; Hull, O.A.; Ellis, J.M.; Mitchell-Koch, K.R. Progress in Our Understanding of 19F Chemical Shifts. In Annual Reports on NMR Spectroscopy; Elsevier: Amsterdam, The Netherlands, 2018; Volume 93, pp. 281–365. ISBN 978-0-12-814913-3. [Google Scholar]
  26. Howe, P.W. Recent developments in the use of fluorine NMR in synthesis and characterisation. Prog. Nucl. Magn. Reson. Spectrosc. 2020, 118–119, 1–9. [Google Scholar] [CrossRef] [PubMed]
  27. Dalvit, C.; Vulpetti, A. Fluorine-Protein Interactions and 19F NMR Isotropic Chemical Shifts: An Empirical Correlation with Implications for Drug Design. ChemMedChem 2011, 6, 104–114. [Google Scholar] [CrossRef] [PubMed]
  28. Gee, C.T.; Arntson, K.E.; Urick, A.K.; Mishra, N.K.; Hawk, L.M.L.; Wisniewski, A.J.; Pomerantz, W.C.K. Protein-observed 19F-NMR for fragment screening, affinity quantification and druggability assessment. Nat. Protoc. 2016, 11, 1414–1427. [Google Scholar] [CrossRef]
  29. Reich, R.M.; Kaposi, M.; Pöthig, A.; Kühn, F.E. Kinetic studies of fluorinated aryl molybdenum(ii) tricarbonyl precursors in epoxidation catalysis. Catal. Sci. Technol. 2016, 6, 4970–4977. [Google Scholar] [CrossRef] [Green Version]
  30. Zientek, N.; Laurain, C.; Meyer, K.; Paul, A.; Engel, D.; Guthausen, G.; Kraume, M.; Maiwald, M. Automated data evaluation and modelling of simultaneous 19 F-1 H medium-resolution NMR spectra for online reaction monitoring. Org. Magn. Reson. 2016, 54, 513–520. [Google Scholar] [CrossRef]
  31. Weidener, D.; Singh, K.; Blümich, B. Synthesis of α-fluoro-α,β-unsaturated esters monitored by 1D and 2D benchtop NMR spectroscopy. Org. Magn. Reson. 2019, 57, 852–860. [Google Scholar] [CrossRef]
  32. Zhang, S.; Li, L.; Hu, Y.; Li, Y.; Yang, Y.; Zha, Z.; Wang, Z. Highly Enantioselective Construction of Fluoroalkylated Quaternary Stereocenters via Organocatalytic Dehydrated Mannich Reaction of Unprotected Hemiaminals with Ketones. Org. Lett. 2015, 17, 5036–5039. [Google Scholar] [CrossRef]
  33. Ibba, F.; Pupo, G.; Thompson, A.L.; Brown, J.M.; Claridge, T.D.W.; Gouverneur, V. Impact of Multiple Hydrogen Bonds with Fluoride on Catalysis: Insight from NMR Spectroscopy. J. Am. Chem. Soc. 2020, 142, 19731–19744. [Google Scholar] [CrossRef]
  34. Zhao, Y.; Swager, T.M. Simultaneous Chirality Sensing of Multiple Amines by 19F NMR. J. Am. Chem. Soc. 2015, 137, 3221–3224. [Google Scholar] [CrossRef]
  35. Jang, S.; Park, H.; Duong, Q.H.; Kwahk, E.-J.; Kim, H. Determining the Enantiomeric Excess and Absolute Configuration of In Situ Fluorine-Labeled Amines and Alcohols by 19F NMR Spectroscopy. Anal. Chem. 2021, 94, 1441–1446. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, W.; Xia, X.; Bian, G.; Song, L. A chiral sensor for recognition of varied amines based on 19F NMR signals of newly designed rhodium complexes. Chem. Commun. 2019, 55, 6098–6101. [Google Scholar] [CrossRef] [PubMed]
  37. Apparu, M.; Ben Tiba, Y.; Léo, P.-M.; Hamman, S.; Coulombeau, C. Determination of the enantiomeric purity and the configuration of β-aminoalcohols using (R)-2-fluorophenylacetic acid (AFPA) and fluorine-19 NMR: Application to β-blockers. Tetrahedron Asymmetry 2000, 11, 2885–2898. [Google Scholar] [CrossRef]
  38. Sabot, C.; Mosser, M.; Antheaume, C.; Mioskowski, C.; Baati, R.; Wagner, A. Novel chiral derivatizing isothiocyanate-based agent for the enantiomeric excess determination of amines. Chem. Commun. 2009, 23, 3410–3412. [Google Scholar] [CrossRef]
  39. Binder, C.M.; Bautista, A.; Zaidlewicz, M.; Krzemiński, M.P.; Oliver, A.; Singaram, B. Dual Stereoselectivity in the Dialkylzinc Reaction Using (−)-β-Pinene Derived Amino Alcohol Chiral Auxiliaries. J. Org. Chem. 2009, 74, 2337–2343. [Google Scholar] [CrossRef]
  40. Chuo, T.H.; Boobalan, R.; Chen, C. Camphor-Based Schiff Base Of 3-Endo-Aminoborneol (SBAB): Novel Ligand for Vanadium-Catalyzed Asymmetric Sulfoxidation and Subsequent Kinetic Resolution. ChemistrySelect 2016, 1, 2174–2180. [Google Scholar] [CrossRef]
  41. Prakash, G.K.S.; Mandal, M.; Olah, G.A. Asymmetric Synthesis of Trifluoromethylated Allylic Amines Using α,β-Unsaturated N-tert-Butanesulfinimines. Org. Lett. 2001, 3, 2847–2850. [Google Scholar] [CrossRef]
  42. Petrov, V.A.; Davidson, F.; Marshall, W. Reactions of quadricyclane with fluorinated nitrogen-containing compounds. Synthesis of 3-aza-4-perfluoroalkyl-tricyclo [4.2.1.02,5]non-3,7-dienes. J. Fluor. Chem. 2004, 125, 1621–1628. [Google Scholar] [CrossRef]
  43. Radchenko, D.S.; Michurin, O.M.; Chernykh, A.V.; Lukin, O.; Mykhailiuk, P.K. An easy synthesis of α-trifluoromethyl-amines from aldehydes or ketones using the Ruppert-Prakash reagent. Tetrahedron Lett. 2013, 54, 1897–1898. [Google Scholar] [CrossRef]
  44. Sudarikov, D.V.; Krymskaya, Y.V.; Il’Chenko, N.O.; Slepukhin, P.A.; Rubtsova, S.A.; Kutchin, A.V. Synthesis and biological activity of fluorine-containing amino derivatives based on 4-caranethiol. Russ. Chem. Bull. 2018, 67, 731–742. [Google Scholar] [CrossRef]
  45. Prakash, G.K.S.; Yudin, A.K. Perfluoroalkylation with Organosilicon Reagents. Chem. Rev. 1997, 97, 757–786. [Google Scholar] [CrossRef]
  46. Jaworska, M.; Pawlak, T.; Kruszyński, R.; Ćwiklińska, M.; Krzemiński, M. NMR Crystallography Comparative Studies of Chiral (1R,2S,3R,5R)-3-Amino-6,6-dimethylbicyclo [3.1.1]heptan-2-ol and Its p-Toluenesulfonamide Derivative. Cryst. Growth Des. 2012, 12, 5956–5965. [Google Scholar] [CrossRef]
  47. SAINT. Data Reduction and Correction Program; Bruker AXS: Fitchburg, WI, USA, 2014. [Google Scholar]
  48. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  50. Krause, L.; Herbst-Irmer, R.; Sheldrick, G.M.; Stalke, D. Comparison of silver and molybdenum microfocus X-ray sources for single-crystal structure determination. J. Appl. Crystallogr. 2015, 48, 3–10. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of benzyl-O-oximes 46 from oximes 13. Reagents and conditions: (a) Cs2CO3, BnCl, MeCN, rt, 48 h. (b) t-BuOK, BnCl, THF, 4 °C → rt, 24 h.
Scheme 1. Synthesis of benzyl-O-oximes 46 from oximes 13. Reagents and conditions: (a) Cs2CO3, BnCl, MeCN, rt, 48 h. (b) t-BuOK, BnCl, THF, 4 °C → rt, 24 h.
Molecules 27 07068 sch001
Scheme 2. Synthesis of β-trifluoromethyl-β-hydroxy-benzyl-O-oximes 1012 from β-keto-benzyl-O-oximes 46. Reagents and conditions: (a) TMSCF3, CsF, THF, 4 °C → rt, 4 h. (b) TBAF·3H2O, THF, rt, 30 min.
Scheme 2. Synthesis of β-trifluoromethyl-β-hydroxy-benzyl-O-oximes 1012 from β-keto-benzyl-O-oximes 46. Reagents and conditions: (a) TMSCF3, CsF, THF, 4 °C → rt, 4 h. (b) TBAF·3H2O, THF, rt, 30 min.
Molecules 27 07068 sch002
Figure 1. Structure and NOE interactions of compounds (2S)-10, (4R)-11, (2R)-12.
Figure 1. Structure and NOE interactions of compounds (2S)-10, (4R)-11, (2R)-12.
Molecules 27 07068 g001
Scheme 3. Synthesis of hydrochlorides of trifluoromethylated amino alcohols 1315 from trifluoromethylated benzyl-O-oximes 1012. Reagents and conditions: (a) LiAlH4, Et2O, 4 °C → rt, 12 h. (b) HCl (gas), Et2O, 4 °C.
Scheme 3. Synthesis of hydrochlorides of trifluoromethylated amino alcohols 1315 from trifluoromethylated benzyl-O-oximes 1012. Reagents and conditions: (a) LiAlH4, Et2O, 4 °C → rt, 12 h. (b) HCl (gas), Et2O, 4 °C.
Molecules 27 07068 sch003
Figure 2. Structure and NOE interactions of compounds (2S,3S)-13, (3R,4R)-14, (2R,3R)-15.
Figure 2. Structure and NOE interactions of compounds (2S,3S)-13, (3R,4R)-14, (2R,3R)-15.
Molecules 27 07068 g002
Figure 3. Molecular structure of two independent molecules (A,B) of compound 16 with thermal ellipsoids drawn at the 30% probability level.
Figure 3. Molecular structure of two independent molecules (A,B) of compound 16 with thermal ellipsoids drawn at the 30% probability level.
Molecules 27 07068 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Petrova, P.A.; Sudarikov, D.V.; Frolova, L.L.; Rumyantcev, R.V.; Rubtsova, S.A.; Kutchin, A.V. Synthesis of Trifluoromethylated Monoterpene Amino Alcohols. Molecules 2022, 27, 7068. https://doi.org/10.3390/molecules27207068

AMA Style

Petrova PA, Sudarikov DV, Frolova LL, Rumyantcev RV, Rubtsova SA, Kutchin AV. Synthesis of Trifluoromethylated Monoterpene Amino Alcohols. Molecules. 2022; 27(20):7068. https://doi.org/10.3390/molecules27207068

Chicago/Turabian Style

Petrova, Polina A., Denis V. Sudarikov, Larisa L. Frolova, Roman V. Rumyantcev, Svetlana A. Rubtsova, and Aleksandr V. Kutchin. 2022. "Synthesis of Trifluoromethylated Monoterpene Amino Alcohols" Molecules 27, no. 20: 7068. https://doi.org/10.3390/molecules27207068

Article Metrics

Back to TopTop