Abstract
New racemic vicinal amino alcohol derivatives with 4‐benzylidenecyclohexane skeleton and axial chirality have been prepared. A preparatively easy and efficient protocol for resolution of the N‐benzoylamino alcohol is described. Using a 250 × 20 mm (L × ID) Chiralpak® IA column, and the appropriate mixture of n‐hexane/ethanol/chloroform as eluent, both enantiomers of N‐benzoylamino alcohol 3 are obtained with >99% enantiomeric excess (ee) by successive injections of a solution of the racemic sample in chloroform. The obtained axially chiral vicinal amino alcohol is used to synthesize structurally novel bisoxazoline ligands in high yields.
Keywords: axial chirality, bisoxazolines, chiral stationary phases, enantiomeric separation, vicinal amino alcohols
A new axially chiral vicinal amino alcohol is used to synthesize structurally novel bisoxazoline ligands in high yields.

1. INTRODUCTION
Chiral 1,2‐amino alcohols are essential structural motifs widely found in natural and synthetic biologically active compounds displaying very diverse activities, 1 , 2 with a tremendous versatility in asymmetric synthesis as building blocks, 3 auxiliaries, 4 or ligands for metal catalyzed reactions 5 and organocatalysis. 6 In most cases, chirality of the 1,2‐amino alcohol is due to the presence of a chiral center, but molecular dissymmetry is not restricted to the presence of a chiral center. Compounds featuring a chirality axis or plane are in fact of tremendous importance 7 and are present in many natural compounds, 8 , 9 , 10 ligands or catalysts for asymmetric synthesis, or synthetic intermediates. 11 , 12 , 13 Synthesis, properties, and applications of axially chiral atropoisomers (biphenyls and derivatives) and allenes have been widely studied. However, the preparation and use of chiral alkylidene cycloalkanes has been much less studied, in spite of the fact that alkylidene cycloalkanes are chiral compounds, fully stable at the stereogenic axis.
Considering the value of both chiral 1,2‐amino alcohols and axially chiral compounds in asymmetric synthesis, we wondered about the potential application of new axially chiral 1,2‐amino alcohols belonging to the alkylidene cycloalkane family.
Focusing on the synthesis and isolation of new α‐amino acids and derivatives with unusual structural features, 14 , 15 , 16 we developed a high‐performance liquid chromatographic (HPLC) resolution protocol 17 for the enantioseparation of unusual amino acid derivatives containing a cyclohexylidene moiety on an analytical and semipreparative scale. 18 , 19 , 20 Now we wish to describe the synthesis and resolution of axially chiral 1,2‐amino alcohols containing a 4‐benzylidenecyclohexane moiety.
2. MATERIALS AND METHODS
2.1. General information
Unless otherwise specified, all reagents were obtained from commercial suppliers and used without purification. For anhydrous conditions, reactions were carried out under Ar in solvents dried using a solvent purification system (SPS). n‐Hexane, ethanol, acetone, and chloroform used for HLPC separations were chromoscan grade from LabScan. Heating was performed using an oil bath mounted on a hot plate magnetic stirrer. Whenever possible, the reactions were monitored by TLC. TLC analysis was performed on precoated silica gel polyester plates with an F254 indicator, and products were visualized using ultraviolet (UV) light (254 nm) and ethanolic phosphomolybdic acid solutions followed by heating. Column chromatography was performed on silica gel (Kiesegel 60, 230–400 Mesh) with air pressure or alumina (Aluminum Oxide, 90 Neutral, 50‐200 μm). Methyl 1‐benzamido‐4‐oxocyclohexane‐1‐carboxylate 1 was prepared as previously described. 21
2.2. Instrumentation
HPLC separations were carried out on a Waters HPLC system consisting of an M‐600 low‐pressure gradient pump, an M‐2996 photodiode array detector, and an M‐2487 dual wavelength absorbance detector, to monitor analytical and preparative separations, respectively. The chromatographic data were acquired and processed with Millennium® chromatography manager software (Waters). A Rheodyne 7125 syringe‐loading sample injector was equipped with 20‐ and 500‐μl loops, respectively, for analytical or semipreparative chromatography. Commercially available polysaccharide chiral stationary phases based on immobilized amylose tris(3,5‐dimethylphenylcarbamate), Daicel Chiralpak® IA column, and immobilized cellulose tris(3,5‐dimethylphenylcarbamate), Daicel Chiralpak® IB column, were used. Melting points were determined in open capillaries using a Gallenkamp capillary melting point apparatus and are not corrected. The FT‐IR spectra of oils were recorded as thin films on NaCl plates; the FT‐IR spectra of solids were recorded on pressed KBr pellets using a Thermo Nicolet Avatar 360 FT‐IR spectrophotometer. Optical rotations were measured on a Jasco 1020 digital polarimeter at λ 589 nm and 25°C in cells with 1 or 10 cm path length. 1H NMR and 13C NMR spectra were acquired in deuterated solvents on a Bruker AV‐400 spectrometer operating at 400 MHz for 1H NMR and 100 MHz for 13C NMR using the solvent residual resonance as the internal standard. 22 Spectra were acquired at room temperature unless otherwise stated using a 5‐mm probe. High‐resolution mass spectra were recorded from methanolic solutions on a Bruker Dalton MICROTOF‐Q (quadrupole time‐of‐flight) microinstrument using the positive electrospray ionization mode (ESI+).
2.3. HPLC analytical assays
The HPLC analytical assays were carried out operating under isocratic conditions at room temperature on Chiralpak® IA and Chiralpak® IB 250 × 4.6 mm (L × ID) columns. Different binary and ternary mixtures of solvents were used as eluents. Samples were manually injected. The flow rate was 1 ml/min. The analyte concentration in injected solutions was 5 mg/ml, and the injection volume was 5 μl. Detection was performed at 250 nm. The capacity (k′), selectivity (α), and resolution (R s ) factors were calculated according to the equations k′ = (t r − t 0)/t 0, a = k′ 2/k′ 1, R s = 1.18(t r2 − t r1)/(W 0.5h1 + W 0.5h2). Subscripts 1 and 2 refer to the first and second eluted enantiomer, respectively; t r are their retention times and W 0.5h denote their full width to the half maximum of each peak; t 0 is the dead time.
2.4. HPLC semipreparative assays
The HPLC semipreparative resolution of compound rac‐3 was carried out operating under isocratic conditions at room temperature on a 250 × 20 mm (L × ID) Chiralpak® IA column. A ternary mixture of n‐hexane/ethanol/chloroform was used as the eluent. Injections and collections were made manually. The flow rate was 0.8 ml/min. The wavelength for UV detection was 270 nm. The column loading capacity, W s (defined as the maximum sample mass that the column can hold), was experimentally calculated for the analytical 250 × 4.6 mm (L × ID) Chiralpak® IA column by injecting increasing amounts of sample with a concentration of 200 mg/ml.
2.4.1. Methyl 1‐benzamido‐4‐benzylidenecyclohexane‐1‐carboxylate ( rac‐2)
To a stirred suspension of benzyltriphenylphosphonium chloride (7.77 g, 20 mmol) in dry THF (75 ml) under argon at room temperature, potassium tert‐butoxide (2.24 g, 20 mmol) was added and stirring was continued for 45 min. A solution of methyl 1‐benzamido‐4‐oxocyclohexane‐1‐carboxylate 1 (2.75 g, 10 mmol) in dry THF (30 ml) was then added and the reaction mixture was stirred under argon at room temperature for 24 h. On completion, the reaction mixture was acidified to about pH 2 by the addition of 1 N hydrochloric acid solution. The organic solvent was evaporated in vacuo, and the aqueous layer was extracted with diethyl ether (3 × 50 ml). The combined organic layers were dried over anhydrous MgSO4, filtered, and evaporated under reduced pressure. The residue was purified by column chromatography on silica gel (eluent: Et2O/hexanes 4:1) to give 3.11 g (90% yield) of compound rac‐2 as a colorless oil that solidified by stirring with a small amount of diethyl ether. White solid; m.p. 160°C; 1H NMR (400 MHz, CDCl3, δ): 7.79–7.82 (m, 2H), 7.50–7.55 (m, 1H), 7.42–7.47 (m, 2H), 7.29–7.35 (m, 2H), 7.18–7.24 (m, 3H), 6.35 (s, 2H), 3.75 (s, 3H), 2.81 (dt, J = 13.9, 4.5 Hz, 1H), 2.42–2.49 (m, 2H), 2.16–2.38 (m, 4H), 2.08 (ddd, J = 13.6, 11.5, 4.4 Hz, 1H); 13C{1H} APT NMR (100 MHz, CDCl3, δ): 173.9 (C), 167.2 (C), 139.0 (C), 137.5 (C), 134.1 (C), 131.7 (CH), 128.8 (CH), 128.5 (CH), 128.1 (CH), 127.0 (CH), 126.2 (CH), 123.8 (CH), 59.1 (C), 52.4 (CH3), 33.8 (CH2), 33.2 (CH2), 31.9 (CH2), 24.2 (CH2); IR (nujol): ν = 3308 (NH), 1742 (COO), 1638 (CONH), 1579, 1557 (C=C), 1527 (NH δ) cm−1; HRMS (ESI, m/z): [M + Na]+ calcd. for C22H23NNaO3, 372.1560; found, 372.1570.
2.4.2. N‐(4‐Benzylidene‐1‐(hydroxymethyl)cyclohexyl)benzamide ( rac‐3)
To a stirred solution of compound rac‐2 (1.51 g, 4.32 mmol) in dry diethyl ether (50 ml) under argon at 0°C, a 2 M solution of lithium borohydride in dry THF (4.34 ml, 8.68 mmol) was added and the resulting mixture was stirred for 1 h at room temperature. After completion, the reaction was quenched at 0°C by the slow addition of saturated aqueous ammonium chloride solution (35 ml). The solution was extracted with dichloromethane (3 × 25 ml). The combined organic layers were dried over anhydrous MgSO4, filtered, and evaporated under reduced pressure to give 1.30 g (94% yield) of compound rac‐3. White solid; m.p. 126°C; 1H NMR (400 MHz, CDCl3, δ): 7.74–7.77 (m, 2H), 7.50–7.54 (m, 1H), 7.41–7.46 (m, 2H), 7.30–7.35 (m, 2H), 7.19–7.24 (m, 3H), 6.35 (s, 1H), 6.28 (s, 1H), 4.85 (bs, 1H), 3.81 (s, 2H), 2.68 (dt, J = 14.6, J = 5.2, 1H), 2.28–2.42 (m, 3H), 2.16–2.24 (m, 1H), 2.04–2.13 (m, 1H), 1.82 (ddd, J = 13.5, J = 8.6, J = 6.4, 1H), 1.69 (ddd, J = 13.5, J = 10.6, J = 4.5, 1H); 13C{1H} APT NMR (100 MHz, CDCl3, δ): 168.8 (C), 139.7 (C), 137.5 (C), 134.7 (C), 131.7 (CH), 128.8 (CH), 128.6 (CH), 128.1 (CH), 126.9 (CH), 126.2 (CH), 123.7 (CH), 69.2 (CH2), 58.5 (C), 33.3 (CH2), 32.7 (CH2), 32.0 (CH2), 24.2 (CH2); IR (nujol): ν = 3304 (NH), 3218 (OH), 1632 (CONH), 1600, 1577 (C=C), 1535 (NH δ) cm−1; HRMS (ESI, m/z): [M + Na]+ calcd. for C21H23NNaO2, 344.1661; found, 344.1633.
2.4.3. (R a )‐N‐(4‐Benzylidene‐1‐(hydroxymethyl)cyclohexyl)benzamide [(R a )‐3] and (S a )‐N‐(4‐benzylidene‐1‐(hydroxymethyl)cyclohexyl)benzamide [(S a )‐3]
Six hundred milligrams of rac‐3 dissolved in CHCl3 (3 ml) were resolved by successive injections of 200 μl of solution (W s = 40 mg) on a 250 × 20 mm (L × ID) Chiralpak® IA column and using a ternary mixture n‐Hex/EtOH/CHCl3 (88/6/6) as the eluent (flow rate: 0.8 ml/min). A total of 15 injections were performed, with one injection performed every 8 min. Four separate fractions were collected. The first, second, third, and fourth fractions contained, respectively, 100/0 (235 mg), 75/25 (80 mg), 3/97 (135 mg), and 0.5/99.5 (150 mg) mixtures of (R a )‐3 and (S a )‐3. Recrystallization of the fourth fraction from ethanol/ether provided 45 mg of enantiomerically pure (S a )‐3. (R a )‐3: White solid, m.p. = 138°C; [α]25 D = −136.6 (c = 0.49 in CHCl3). (S a )‐3: White solid, m.p. = 137°C; [α]25 D = 133.2 (c = 0.48 in CHCl3). Spectroscopic data for (R a )‐3 and ( S a )‐3 were identical to those given above for the racemic compound.
2.4.4. (1‐Amino‐4‐benzylidenecyclohexyl)methanol ( rac‐4)
To a stirred solution of compound rac‐3 (0.90 g, 2.80 mmol) in ethanol (20 ml), a solution of potassium hydroxide (2.01 g, 35.82 mmol) in water (25 ml) was added and the resulting mixture was stirred under reflux conditions for 3 h. Ethanol was evaporated in vacuo, and the aqueous layer was extracted with diethyl ether (2 × 25 ml). The combined organic layers were then extracted with 1 N hydrochloric acid solution (3 × 25 ml). The combined extracts were basified by the addition of 6 N sodium hydroxide aqueous solution. The basic solution was extracted again with dichloromethane (3 × 25 ml). The combined organic layers were dried over anhydrous MgSO4, filtered, and evaporated under reduced pressure to give 0.51 g (84% yield) of compound rac‐4. White solid; m.p. 117°C; 1H NMR (400 MHz, CDCl3, δ): 7.29–7.34 (m, 2H), 7.18–7.21 (m, 3H), 6.29 (s, 1H), 3.41 (s, 2H), 2.54 (dt, J = 14.5, J = 5.7, 1H), 2.32–2.46 (m, 2H), 2.28 (dt, J = 14.1, J = 5.7, 1H), 2.10 (bs, 3H), 1.54–1.67 (m, 2H), 1.44–1.52 (m, 2H); 13C{1H} APT NMR (100 MHz, CDCl3, δ): 141.2 (C), 137.9 (C), 128.8 (CH), 128.0 (CH), 126.0 (CH), 123.0 (CH), 69.9 (CH2), 52.2 (C), 36.5 (CH2), 36.0 (CH2), 32.3 (CH2), 24.3 (CH2); IR (nujol): ν = 3335, 3281 (NH2), 3155 (OH), 1649, 1591 (C=C) cm−1; HRMS (ESI, m/z): [M + H]+ calcd. for C14H20NO, 218.1539; found, 218.1535.
2.4.5. (R a )‐(1‐Amino‐4‐benzylidenecyclohexyl)methanol [(R a )‐4] and (S a )‐(1‐amino‐4‐benzylidenecyclohexyl)methanol [(S a )‐4]
Hydrolysis of enantiomerically pure (R a )‐3 and (S a )‐3 as described above provided (R a )‐4 and (S a )‐4, respectively. (Ra)‐4: White solid, m.p. = 68°C; [α]25 D = −6.9 (c = 0.54 in EtOH). (S a )‐3: White solid, m.p. = 68°C; [α]25 D = 7.0 (c = 0.31 in EtOH). Spectroscopic data for (R a )‐3 and (S a )‐3 were identical to those given above for the racemic compound.
2.4.6. (5R,5′R)‐2,2′‐(Propane‐2,2‐diyl)bis(8‐((Z)‐benzylidene)‐3‐oxa‐1‐azaspiro[4.5]dec‐1‐ene) [(R a ,R a )‐5]
To a stirred solution of 2,2‐dimethyl‐malononitrile (51.7 mg, 0.55 mmol) in dry toluene (7 ml) under argon at room temperature, dry 23 zinc triflate (200 mg, 0.55 mmol) was added and the resulting mixture was stirred for 5 min. Then a solution of β‐amino alcohol (R a )‐4 (217 mg, 1 mmol) in dry toluene (5 ml) was added and the solution was heated under reflux for 48 h. The system was allowed to cool to room temperature and the reaction solution was then washed with brine (2 × 20 ml) and 5% aqueous solution of NaHCO3 (3 × 15 ml). The organic layer was dried over anhydrous MgSO4, filtered, and evaporated under reduced pressure. The residue was purified by column chromatography on alumina previously oven‐dried for 1 h at 100°C (eluent: hexanes/Et2O 4:1) to give 135 mg (55% yield) of compound (R a ,R a )‐5. Colorless oil; [α]25 D = −36.1 (c = 0.39 in CHCl3); 1H NMR (400 MHz, CDCl3, δ): 7.29–7.35 (m, 2H), 7.17–7.22 (m, 3H), 6.32 (s, 1H), 4.05 (s, 2H), 2.76 (ddd, J = 14.0, J = 6.7, J = 4.5, 1H), 2.56 (ddd, J = 13.8, J = 6.8, J = 4.9, 1H), 2.14–2.24 (m, 2H), 1.89 (ddd, J = 12.6, J = 9.8, J = 4.5, 1H), 1.78 (ddd, J = 13.9, J = 9.9, J = 4.5, 1H), 1.65–1.73 (m, 1H), 1.55–1.62 (m, 1H), 1.51 (s, 3H); 13C{1H} APT NMR (100 MHz, CDCl3, δ): 167.6 (C), 140.5 (C), 138.0 (C), 128.8 (CH), 128.0 (CH), 126.0 (CH), 123.1 (CH), 77.2 (C), 70.4 (C), 38.4 (CH2), 37.7(CH2), 33.3(CH2), 25.2(CH2), 24.4(CH3); IR (nujol): ν = 1658 (C=N) cm−1; HRMS (ESI, m/z): [M + H]+ calcd. for C33H39N2O2, 495.3006; found, 495.3008.
2.4.7. 2,6‐Bis((R)‐8‐((Z)‐benzylidene)‐3‐oxa‐1‐azaspiro[4.5]dec‐1‐en‐2‐yl)pyridine [(R a ,R a )‐6]
To a stirred solution of pyridine‐2,6‐dicarbonitrile (71 mg, 0.55 mmol) in dry toluene (7 ml) under argon at room temperature, dry 23 zinc triflate (20 mg, 0.055 mmol) was added and the resulting mixture was stirred for 5 min. Then a solution of β‐amino alcohol (R a )‐4 (217 mg, 1 mmol) in dry toluene (5 ml) was added and the solution was heated under reflux for 24 h. The system was allowed to cool to room temperature, and the reaction solution was diluted by the addition of ethyl acetate (15 ml). The obtained solution was then washed with brine (2 × 20 ml) and aqueous saturated solution of NaHCO3 (3 × 15 ml). The organic layer was dried over anhydrous MgSO4, filtered, and evaporated under reduced pressure. The residue was purified by column chromatography on alumina previously oven‐dried for 1 h at 100°C (eluent: gradually form Et2O/hexanes/4:1 to Et2O) to give 185 mg (70% yield) of compound (R a ,R a )‐6. Colorless oil; [α]25 D = −23.6 (c = 0.30 in CHCl3) [[α]25 D = 24.5 (c = 0.10 in CHCl3) for (S a ,S a )‐6]; 1H NMR (400 MHz, CDCl3, δ): 8.20 (d, J = 7.9, 2H), 7.86 (t, J = 7.9, 1H), 7.29–7.34 (m, 4H), 7.17–7.23 (m, 6H), 6.35 (s, 2H), 4.31 (s, 4H), 2.80 (ddd, J = 13.9, J = 7.5, J = 4.8, 2H), 2.68 (ddd, J = 12.7, J = 7.5, J = 4.4, 2H), 2.36 (ddd, J = 12.8, J = 8.1, J = 4.0, 2H), 2.27 (ddd, J = 13.4, J = 8.8, J = 4.1, 2H), 2.01 (ddd, J = 12.9, J = 8.7, J = 4.7, 2H), 1.90 (ddd, J = 13.1, J = 8.6, J = 4.5, 2H), 1.80 (ddd, J = 12.5, J = 7.7, J = 4.7, 2H), 1.65–1.73 (m, 2H); 13C{1H} APT NMR (100 MHz, CDCl3, δ): 161.0 (C), 147.0 (C), 140.4 (C), 138.0 (C), 137.3 (CH), 128.9 (CH), 128.1 (CH), 126.1 (CH), 125.8 (CH), 123.4 (CH), 78.1 (C), 71.5 (C), 39.0 (CH2), 38.3 (CH2), 33.3 (CH2), 29.7 (CH2), 25.2 (CH2); IR (nujol): ν = 1589 (C=CPy) cm−1; HRMS (ESI, m/z): [M + Na]+ calcd. for C35H35N3NaO2, 552.2621; found, 552.2592.
3. RESULTS AND DISCUSSION
We started the synthesis from the known 4‐substituted cyclohexanone 1, prepared through Diels–Alder reaction between methyl 2‐benzamidoacrylate and 2‐trimethylsilyloxy‐1,3‐butadiene in the presence of ZnI2 as a catalyst. 21 Wittig olefination of this ketone featuring a symmetry plane with benzyltriphenylphosphonium bromide led to axially chiral alkene rac‐2 in 90% yield, as a racemic mixture. Racemic N‐benzoyl amino alcohol rac‐3 was obtained in 94% yield by reduction of the ester moiety of compound rac‐2 with lithium borohydride. Final basic hydrolysis led to the desired axially chiral 1,2‐amino alcohol rac‐4 in 84% yield, as a racemic mixture (Scheme 1).
SCHEME 1.

Synthesis of axially chiral 1,2‐amino alcohol rac‐4
With axially chiral compound rac‐2, rac‐3, and rac‐4 in hand, we studied their enantiomeric separation by high‐performance liquid chromatography using chiral stationary phases based on immobilized 3,5‐dimethylphenylcarbamate derivatives of amylose or cellulose at an analytical level using 250 × 4.6 mm (L × ID) columns, namely Chiralpak® IA and Chiralpak® IB. The capacity (k′), selectivity (α), and resolution (R s ) factors for each column in the enantioseparation of all compounds using different eluents were determined (Table 1).
TABLE 1.
Chromatographic data for the analytical HPLC resolution of rac‐2, rac‐3, and rac‐4 on Chiralpak® IA and Chiralpak® IB using different mobile phases
| Compound | Column | Eluent | k′ | α | Rs |
|---|---|---|---|---|---|
| rac‐2 | IA | n‐Hex/EtOH (90/10) | 1.54 | 1.00 | ‐ |
| rac‐2 | IB | n‐Hex/EtOH (90/10) | 3.32 | 1.05 | 0.5 |
| rac‐3 | IA | n‐Hex/EtOH (90/10) | 1.74 | 1.21 | 2.65 |
| rac‐3 | IB | n‐Hex/EtOH (90/10) | 2.08 | 1.00 | ‐ |
| rac‐4 | IA | n‐Hex/EtOH (90/10) | 0.97 | 1.55 | 1.90 |
| rac‐4 | IA | n‐Hex/EtOH (95/5) | 2.07 | 1.33 | 2.12 |
| rac‐4 | IB | n‐Hex/EtOH (95/5) | 2.60 | 1.00 | ‐ |
| rac‐3 | IA | n‐Hex/EtOH/Acetone (92/5/3) | 2.57 | 1.12 | 1.24 |
| rac‐3 | IA | n‐Hex/EtOH/CHCl3 (90/8/2) | 2.08 | 1.19 | 1.80 |
| rac‐4 | IA | n‐Hex/EtOH/Acetone (95/3/2) | 2.10 | 1.18 | 1.28 |
| rac‐4 | IA | n‐Hex/EtOH/CHCl3 (90/8/2) | 1.17 | 1.27 | 1.30 |
| rac‐4 | IA | n‐Hex/EtOH/CHCl3 (92/6/2) | 1.78 | 1.28 | 1.95 |
| rac‐3 | IA | n‐Hex/EtOH/CHCl3 (90/4/6) | 3.11 | 1.12 | 1.80 |
| rac‐3 | IA | n‐Hex/EtOH/CHCl3 (90/6/4) | 2.16 | 1.13 | 1.94 |
| rac‐3 | IA | n‐Hex/EtOH/CHCl3 (85/10/5) | 1.22 | 1.15 | 1.17 |
| rac‐3 | IA | n‐Hex/EtOH/CHCl3 (88/6/6) | 2.12 | 1.12 | 1.86 |
Note: Chromatographic conditions: Room temperature, 250 × 4.6 mm (L × ID) columns, injection volume: 5 μl, flow rate 1 ml/min; UV detection 250 nm.
A selectivity value of about 1.15, which allows resolution values higher than 1.5, is required for an easy separation at analytical scale. From results using binary mixtures of n‐hexane/ethanol as mobile phase, it was concluded that both rac‐3 and rac‐4 are suitable analytes for enantioseparation with these chiral stationary phases. In both cases, Chiralpak® IA column is the only one that provides selectivity. Nevertheless, the low solubility of both racemates in the binary mixture of solvents, that causes the precipitation of the compounds into the column, hampered its use on a semipreparative scale.
In order to enhance the solubility of racemates and then be able to move from analytical conditions to the semipreparative scale, the addition of acetone or chloroform as a third component to the eluting mixture was evaluated for compounds rac‐3 and rac‐4. In all cases, the presence of acetone or chloroform in the eluent led to a decrease in selectivity and resolution, but chloroform mixtures led to better values, some times in the same range as binary mixtures. Unfortunately, even in this ternary mixture, solubility of compound rac‐4 remained too low to allow a good loading capacity in semipreparative assays. Analyte rac‐3 was then selected for further optimization to extend the study to the semipreparative‐scale enantioseparation. Taking into account selectivity, resolution and solubility (improved with increasing percentage of chloroform in the mixture), ternary mixture n‐hexane/ethanol/chloroform 88/6/6 was chosen to perform the semipreparative resolution. Figure 1 shows the chromatographic resolution of rac‐3, analytical HPLC with the optimized ternary mixture.
FIGURE 1.

HPLC analytical resolution of rac‐3 at room temperature on a 250 × 4.6 mm (L × ID) Chiralpak® IA column. Mobile phase composition: n‐Hex/EtOH/CHCl3 88/6/6 (v/v/v); flow rate: 1 ml/min; UV detection: 250 nm
At this point, the column saturation capacity (defined as the maximum sample mass that the column can hold in mg) and the optimum sample volume of the column were determined in an experimental approach. Starting from the previously selected elution conditions on the analytical column and using a solution of 200 mg/ml of compound 3 in chloroform (the most concentrated solution that can be obtained), gradually increased volumes of the sample were injected. The chromatographic data obtained on working in the overload mode on the analytical column are shown in Table 2.
TABLE 2.
Chromatographic data for the resolution of amino acid derivatives rac‐3 on Chiralpak® IA data working in an overload mode in the analytical column at room temperature
| V (μl) | k′ | α | R s |
|---|---|---|---|
| 5 | 1.90 | 1.12 | 1.16 |
| 10 | 1.90 | 1.14 | 0.90 |
| 15 | 1.86 | 1.13 | 0.80 |
| 20 | 1.86 | 1.14 | 0.70 |
Note: Overload mode, eluent n‐Hex/EtOH/CHCl3 (88/6/6), c = 200 mg/ml, flow rate 0.8 ml/min; UV detection 270 nm.
Increasing the injected volume, selectivity remained in the same range and resolution was worse even with an injected volume of 5 μl. With this analyte, a significant peak tailing was observed working in overload mode, probably due to the strong interaction of the benzamido group with the carbamate moiety on the selector of the chiral stationary phase. This behavior points out the difficulty in obtaining fractions containing the second eluted enantiomer as single component. That being the case, we decided to work with a large injected volume in a first round to obtain mixtures enriched in either the first or the second eluted enantiomer and perform an additional purification later. We opted to inject 10 μl of a 200 mg/ml solution to work in an overloaded mode on the analytical column (250 × 4.6 mm (L × ID)). That means working with 200‐μl injections of the same concentrated solution (40 mg of analyte per injection) on a 250 × 20 mm (L × ID) column and a flow rate of 16 ml/min to keep the same chromatographic parameters (Figure 2).
FIGURE 2.

Chromatogram for the enantioseparation of rac‐3 operating in an overload mode at room temperature on a 250 × 20 mm (L × ID) Chiralpak® IA column. Injection volume: 200 μl, c = 200 mg/ml. Mobile phase composition: n‐Hex/EtOH/CHCl3 88/6/6 (v/v/v); flow rate: 16 ml/min; UV detection: 270 nm
The semipreparative resolution of compound rac‐3 on a 250 × 20 mm (L × ID) Chiralpak® IA column was achieved by successive injections of a solution of the sample in chloroform (Figure 3). In order to enhance throughput, injections were partially overlapped. For each run, four separate fractions were collected and combined with equivalent fractions. The four combined fractions were concentrated and reinjected onto the analytical chiral column to determine their enantiomeric purity; 15 injections of 200 μl of a 200 mg/ml solution (40 mg of rac‐3 per injection) provided the following: First fraction, 235 mg of the first eluted enantiomer in enantiomerically pure form; second fraction, 80 mg of a 75/25 mixture enriched in the first eluted enantiomer; third fraction, 135 mg of a 97/3 mixture enriched in the second eluted enantiomer; and fourth fraction, 150 mg of a 99.5/0.5 mixture enriched in the second eluted enantiomer.
FIGURE 3.

Semipreparative chromatogram for the enantioseparation of rac‐3 at room temperature on a 250 × 20 mm (L × ID) Chiralpak® IA column. Injection volume: 200 μl, c = 200 mg/ml. Mobile phase composition: n‐Hex/EtOH/CHCl3 88/6/6 (v/v/v); flow rate: 16 ml/min; UV detection: 270 nm. Repetitive injection every 8 min
Recrystallization of the last fraction from ethanol/diethyl ether provided 45 mg of the second eluted enantiomer in enantiomerically pure form. A 99/1 mixture of enantiomers was recovered from mother liquor and combined with the third fraction containing a 97/3 mixture enriched in the second eluted enantiomer. To these combined fractions, a similar protocol of semipreparative resolution was applied, and in this case, three separate fractions were collected. In this way, six injections of 200 μl of a 200 mg/ml solution (40 mg of analyte per injection) provided the following: First fraction, 29 mg of an 87/13 mixture enriched in the second eluted enantiomer; second fraction, 180 mg of 99.5/0.5 mixture enriched in the second eluted enantiomer; and third fraction, 30 mg of the second eluted enantiomer in enantiomerically pure form.
In summary, from 600 mg of amido alcohol rac‐3, 235 mg of the enantiomerically pure first eluted enantiomer, 180 mg the second eluted enantiomer with a 99% enantiomeric purity, and 75 mg of the second eluted enantiomer in enantiomerically pure form, enough to characterization purposes, were obtained. The corresponding analytical check of the several collected fractions in the resolution of rac‐3 is shown in Figure 4.
FIGURE 4.

Analytical check of fractions collected in the enantioseparation of rac‐3 at room temperature on a 250 × 4.6 mm (L × ID) Chiralpak® IA column. Mobile phase composition: n‐Hex/EtOH/CHCl3 88/6/6 (v/v/v); flow rate: 1 ml/min; UV detection: 250 nm. (A) First eluted enantiomer in enantiomerically pure form. (B) Second eluted enantiomer in enantiomerically pure form. (C) 99.5/0.5 mixture enriched in the second eluted enantiomer. (D) 87/13 mixture enriched in the second eluted enantiomer
After resolution by semipreparative HPLC, each enantiomer of compound 3 was hydrolyzed as described above and the assignment of the absolute configuration of β‐amino alcohol 4 2 nd derived from the of second eluted enantiomer of 3 was determined. Absolute configuration determination was performed by 1H using both enantiomers of Boc‐β‐phenylglycine (BPG) as chiral solvating agent (CSA), as described by Pazos et al. 24 for others β‐amino alcohols. Previously, the unambiguous assignment of each signal to the corresponding hydrogens of the cyclohexyl moiety of amino alcohol 4 was performed using bidimensional NMR experiments (COSY and 1H‐13C HSQC).
The 1H NMR spectrum of an equimolecular mixture of the enantiopure amino alcohol 4 2 nd and (R)‐BPG performed in CDCl3 was compared with that obtained in a parallel way with a mixture of 4 2 nd and (S)‐BPG. Comparison of (R)‐BPG/4 2 nd and (S)‐BPG/4 2 nd 1H‐NMR spectra showed that the signals due to the C(γ) protons were shifted downfield in the spectrum of (S)‐BPG/4 2 nd [Δδ RS < 0]. On the other hand, signals due to the C(γ′) protons were shifted upfield in (S)‐BPG/4 2 nd [Δδ RS > 0] (Figure 5).
FIGURE 5.

Partial 1H NMR spectra (400 MHz) of β‐amino alcohol 4 2 nd after the addition of 1 equiv. of (R)‐BPG (bottom line) and (S)‐BPG (top line) in CDCl3. The upfield and downfield shifts are highlighted
This different behavior of the shifts of C(γ) and C(γ′) protons in both complexes and the resulting signs of Δδ RS are fully consistent with an association between the enantiopure amino alcohol 4 2 nd and the chiral solvating agent according to the model previously proposed by Pazos et al. 24 Noncovalent interactions between the CSA and the β‐amino alcohol led to an arrangement where a side of the amino alcohol is located under the shielding cone of the BPG phenyl group. This side is different depending on the configuration of BPG (Figure 6). In such way, for the enantiomer studied, C(γ) hydrogens resonate at a higher field in the presence of (R)‐BPG than in the presence of (S)‐BPG [Δδ RS < 0], and the opposite occurs for C(γ′) hydrogens. According to this model, the absolute configuration of 1,2‐amino alcohol 4 2 nd derived from basic hydrolysis of second eluted enantiomer of 3 should be S a .
FIGURE 6.

Association of (R)‐BPG and (S)‐BPG with the S a enantiomer of β‐amino alcohol 4 and sign distribution of Δδ RS
One prominent use of chiral amino alcohols is the synthesis of compounds containing a chiral oxazoline ring, which have become one of the most successful, versatile, and commonly used classes of ligands for asymmetric catalysis. As a consequence of their ready accessibility, modular nature, and applicability in metal‐catalyzed transformations these ligands have been used with great success in a wide range of asymmetric reactions. 25 , 26 , 27 , 28 , 29 With both enantiomers of axially chiral amino alcohol 4 in enantiomerically pure form in hand, a new class of bis(oxazoline) and pyridine bis(oxazoline) ligands with a chiral axis at the C4 position have been prepared from (R a )‐4 according to Cornejo et al. 30 one‐pot procedure. Condensation of the chiral β‐amino alcohol (R a )‐4 (2 mmol) with 2,2‐dimethylmalononitrile (1 mmol) using stoichiometric amounts of zinc triflate (1 mmol) under refluxing toluene gave bis(oxazoline) (R a ,R a )‐5 in 55% yield after 48 h (Scheme 2).
SCHEME 2.

Synthesis of bis(oxazoline) (R a ,R a )‐5
The same reaction using pyridine‐2,6‐dicarbonitrile required only 24 h to give pyridine bis(oxazoline) (R a ,R a )‐6 in 70% yield. (Scheme 3). In this reaction, only catalytic amounts of Zn (OTf)2 were needed.
SCHEME 3.

Synthesis of pyridine bis(oxazoline) (R a ,R a )‐6
4. CONCLUSION
In summary, we have described here an easy and efficient protocol for the synthesis and semipreparative HPLC resolution of new axially chiral β‐amino alcohols. Box and pybox ligands with new structural features have been obtained in just one step using Zn (OTf)2 to promote condensation of the β‐amino alcohol with the corresponding dicarbonitrile.
Supporting information
Data S1. Supporting Information
ACKNOWLEDGMENTS
Financial support from the Government of Aragón (E45_20R) is acknowledged. Authors wish to thank A. I. Pallarés Pallarés, and M. Nuño Tutor for some HPLC experiments and J. Alegre Fernández de Heredia for taking part in pybox synthesis, as part of their Project‐oriented Studies during the course of their studies of Chemistry Degree (University of Zaragoza).
López‐Ram‐de‐Víu P, Gálvez JA, Díaz‐de‐Villegas MD. Synthesis, resolution, and absolute configuration determination of a vicinal amino alcohol with axial chirality. Application to the synthesis of new box and pybox ligands. Chirality. 2022;34(8):1140‐1150. doi: 10.1002/chir.23475
Funding information Government of Aragón, Grant/Award Number: E45_20R
Contributor Information
Pilar López‐Ram‐de‐Víu, Email: pilopez@unizar.es.
José A. Gálvez, Email: jagl@unizar.es.
DATA AVAILABILITY STATEMENT
Additional supporting information may be found online in the Supporting Information section at the end of this article.
REFERENCES
- 1. Bergmeier SC. The synthesis of vicinal amino alcohols. Tetrahedron. 2000;56:2561‐2576. doi: 10.1016/S0040-4020(00)00149-6 [DOI] [Google Scholar]
- 2. Gupta P, Mahajanb N. Biocatalytic approaches towards the stereoselective synthesis of vicinal amino alcohols. New J Chem. 2018;42:12296‐12327. doi: 10.1039/c8nj00485d [DOI] [Google Scholar]
- 3. Heravi MM, Lashaki TB, Fattahi B, Zadsirjan V. Application of asymmetric sharpless aminohydroxylation in total synthesis of natural products and some synthetic complex bio‐active molecules. RSC Adv. 2018;8(12):6634‐6659. doi: 10.1039/c7ra12625e [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. Ager DJ, Prakash I, Schaad DR. 1,2‐Amino alcohols and their heterocyclic derivatives as chiral auxiliaries in asymmetric synthesis. Chem Rev. 1996;96(2):835‐875. doi: 10.1021/cr9500038 [DOI] [PubMed] [Google Scholar]
- 5. Fache F, Schulz E, Tommasino ML, Lemaire M. Nitrogen‐containing ligands for asymmetric homogeneous and heterogeneous catalysis. Chem Rev. 2000;100(6):2159‐2232. doi: 10.1021/cr9902897 [DOI] [PubMed] [Google Scholar]
- 6. Reddy UVS, 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(24):4124‐4143. doi: 10.1002/ejoc.201600164 [DOI] [Google Scholar]
- 7. Mancinelli M, Bencivenni G, Pecorari D, Mazzanti A. Stereochemistry and recent applications of axially chiral organic molecules. Eur J Org Chem. 2020;2020(27):4070‐4086. doi: 10.1002/ejoc.201901918 [DOI] [Google Scholar]
- 8. Tajuddeen N, Bringmann G. N,C‐coupled naphthylisoquinoline alkaloids: a versatile new class of axially chiral natural products. Nat Prod Rep. 2021;38:2154‐2186. doi: 10.1039/d1np00020a [DOI] [PubMed] [Google Scholar]
- 9. Zask A, Murphy J, Ellestad GA. Biological stereoselectivity of atropisomeric natural products and drugs. Chirality. 2013;25:265‐274. doi: 10.1002/chir.22145 [DOI] [PubMed] [Google Scholar]
- 10. Hoffmann‐Röder A, Krause N. Synthesis and properties of allenic natural products and pharmaceuticals. Angew Chem Int Ed. 2004;43:1196‐1216. doi: 10.1002/anie.200300628 [DOI] [PubMed] [Google Scholar]
- 11. Kitagawa O. Chiral Pd‐catalyzed enantioselective syntheses of various N–C axially chiral compounds and their synthetic applications. Acc Chem Res. 2021;54:719‐730. doi: 10.1021/acs.accounts.0c00767 [DOI] [PubMed] [Google Scholar]
- 12. Yu S, Ma S. Allenes in catalytic asymmetric synthesis and natural product syntheses. AngewChemInt Ed. 2012;51:3074‐3112. doi: 10.1002/anie.201101460 [DOI] [PubMed] [Google Scholar]
- 13. Bringmann G, Tasler S, Pfeifer RM, Breuning M. The directed synthesis of axially chiral ligands, reagents, catalysts, and natural products through the ‘lactone methodology’. J Organometallic Chem. 2002;661:49‐65. doi: 10.1016/S0022-328X(02)01819-3 [DOI] [Google Scholar]
- 14. Cativiela C, Díaz‐de‐Villegas MD, Gálvez JA. Synthesis and chemical resolution of unique α,β‐didehydroamino acids with a chiral axis. Tetrahedron Lett. 1999;40:1027‐1030. doi: 10.1016/S0040-4039(98)02517-9 [DOI] [Google Scholar]
- 15. Cativiela C, Díaz‐de‐Villegas MD, Gálvez JA, Su G. Synthesis and conformational properties of model dipeptides containing novel axially chiral α,β‐didehydroamino acids at the (i + 1) position of a β‐turn conformation. Tetrahedron. 2004;60:11923‐11932. doi: 10.1016/j.tet.2004.09.066 [DOI] [Google Scholar]
- 16. Cativiela C, Díaz‐de‐Villegas MD, Gálvez JA, Sub G. Horner‐Wadsworth‐Emmons reaction for the synthesis of unusual α,β‐didehydroamino acids with a chiral axis. Arkivoc. 2004;iv:59‐66. doi: 10.3998/ark.5550190.0005.408 [DOI] [Google Scholar]
- 17. López‐Ram‐de‐Víu P, Gálvez JA, Díaz‐de‐Villegas MD. High‐performance liquid chromatographic enantioseparation of unusual amino acid derivatives with axial chirality on polysaccharide‐based chiral stationary phases. J Chromatogr a. 2005;1390:78‐85. doi: 10.1016/j.chroma.2015.02.055 [DOI] [PubMed] [Google Scholar]
- 18. Cox GB. Preparative Enantioselective Chromatography. Blackwell Publishing Ltd; 2005. [Google Scholar]
- 19. Sardella R, Ianni F, Marinozzi M, Macchiarulo A, Natalini B. Laboratory‐scale preparativeenantioseparations of pharmaceutically relevant compounds on commercially available chiral stationary phases for HPLC. Curr. Med. Chem. 2017;24:796‐817. doi: 10.2174/0929867323666160907111107 [DOI] [PubMed] [Google Scholar]
- 20. Pinto MMM, Fernandes C, Tiritan ME. Chiral separations in preparative scale: a medicinal chemistry point of view. Molecules. 2020;25:1931 doi: 10.3390/molecules25081931 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21. Avenoza A, Cativiela C, Busto JH, Fernández‐Recio MA, Peregrina JM, Rodríguez F. New synthesis of 7‐azabicyclo[2.2.1]heptane‐1‐carboxylic Acid. Tetrahedron. 2001;57(3):545‐548. doi: 10.1016/S0040-4020(00)01023-1 [DOI] [Google Scholar]
- 22.Residual solvent signals were set according toFulmer GR, Miller AJM, Sherden NH, et al. NMR chemical shifts of trace impurities: common laboratory solvents, organics, and gases in deuterated solvents relevant to the organometallic chemist. Organometallics. 2010;29(9):2176‐2179. doi: 10.1021/om100106e [DOI] [Google Scholar]
- 23.It is very important to use dry zinc triflate to ensure that the reaction takes place until completion. Otherwise, reaction does not progress properly from mono (oxazoline) to bis(oxazoline), and, furthermore, the presence of that intermediate makes it difficult to isolate the desired bis(oxazoline). Thus, commercial zinc triflate is kept dry in a desiccator and, before use, it is dried under vacuum at 125 for 2h according toCorey EJ, Shimoji K. Magnesium and zinc‐catalyzed thioketalization. Tetrahedron Lett. 1983;24:169‐172. doi: 10.1016/S0040-4039(00)81357-X [DOI] [Google Scholar]
- 24. Pazos Y, Leiro V, Seco JM, Quiñoa E, Riguera R. Boc–phenylglycine: a chiral solvating agent for the assignment of the absolute configuration of amino alcohols and their ethers by NMR. Tetrahedron: Asymmetry. 2004;15(12):1825‐1829. doi: 10.1016/j.tetasy.2004.04.032 [DOI] [Google Scholar]
- 25. Connon R, Roche B, Rokade BV, Guiry PJ. Further developments and applications of oxazoline‐containing ligands in asymmetric catalysis. Chem Rev. 2021;121:6373‐6521. doi: 10.1021/acs.chemrev.0c00844 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Yang G, Zhang W. Renaissance of pyridine‐oxazolines as chiral ligands for asymmetric catalysis. ChemSoc Rev. 2018;47:1783‐1810. doi: 10.1039/c7cs00615b [DOI] [PubMed] [Google Scholar]
- 27. Babu SA, Krishnan KK, Ujwaldev SM, Anilkumar G. Applications of pybox complexes in asymmetric catalysis. Asian J Org Chem. 2018;7:1033‐1053. doi: 10.1002/ajoc.201800094 [DOI] [Google Scholar]
- 28. Liao A, Sun XL, Tang Y. Side arm strategy for catalyst design: modifying bisoxazolines for remote control of enantioselection and related. Acc Chem Res. 2014;47:2260‐2272. doi: 10.1021/ar800104y [DOI] [PubMed] [Google Scholar]
- 29. Desimoni G, Faita G, Jørgensen KA. Update 1 of: C2‐symmetric chiral bis(oxazoline) ligands in asymmetric catalysis. Chem Rev. 2011;11:PR284‐PR437. doi: 10.1021/cr100339a [DOI] [PubMed] [Google Scholar]
- 30. Cornejo A, Fraile JM, García JI, et al. An efficient and general one‐pot method for the synthesis of chiral bis(oxazoline) and pyridine bis(oxazoline) ligands. Synlett. 2005;(15):2321‐2324. doi: 10.1055/s-2005-872672 [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data S1. Supporting Information
Data Availability Statement
Additional supporting information may be found online in the Supporting Information section at the end of this article.
