Next Article in Journal
Assessment of the In Vivo Genotoxicity of New Lead Compounds to Treat Sickle Cell Disease
Next Article in Special Issue
Proline-Catalysed Amination Reactions in Cyclic Carbonate Solvents
Previous Article in Journal
Synthesis, Characterization and Spectral Properties of Substituted Tetraphenylporphyrin Iron Chloride Complexes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Enantioselective Aryl Transfer to Aldehydes Using Chiral 2,2’-Bispyrrolidine-Based Salan Ligands

1
School of Chemistry and Chemical Engineering, State Key Laboratory of Coordination Chemistry, Nanjing University, Nanjing, 210093, China
2
School of Chemistry and Materials Science, Shanxi Normal University, Linfen, 041004, China
3
State Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, Chinese Academy of Sciences, Shanghai, 200032, China
4
School of Chemistry and Chemical Engineering, Key Laboratory of Mesoscopic Chemistry of MOE, Nanjing University, Nanjing, 210093, China
*
Authors to whom correspondence should be addressed.
Molecules 2011, 16(4), 2971-2981; https://doi.org/10.3390/molecules16042971
Submission received: 24 December 2010 / Revised: 15 March 2011 / Accepted: 16 March 2011 / Published: 6 April 2011
(This article belongs to the Special Issue Catalytic Asymmetric Synthesis)

Abstract

:
Chiral C2-symmetric diamines have emerged as versatile auxiliaries or ligands in numerous asymmetric transformations. Chiral 2,2’-bispyrrolidine-based salan ligands were prepared and applied to the asymmetric aryl transfer to aldehydes with arylboronic acids as the source of transferable aryl groups. The corresponding diarylmethanols were obtained in high yields with moderate to good enantioselectivitives of up to 83% ee.

1. Introduction

Chiral diarylmethanols are important intermediates and precursors for the synthesis of pharmacologically and biologically active compounds [1,2,3,4,5,6,7,8]. Therefore, the development of effective catalyst systems for the synthesis of these compounds is of significant importance for organic chemists. The scientifically important protocols for the synthesis of chiral diarylmethanols commonly involve two strategies: (1) the asymmetric reduction of prochiral diaryl ketones [9,10,11,12,13], (2) the enantioselective aryl transfer to aromatic aldehydes [14,15,16]. The reduction method requires an ortho substituent on one of the aryls or electronic different aryl groups for optimum results. The second method seems easy to realize chiral induction due to the large steric and electronic differences between an aryl group and a hydrogen atom on the aldehyde substrates with diphenylzinc. As reported previously, many functionalized diarylzincs used as the transferring nucleophiles are unstable and difficult to synthesize, so the method of the aryl transfer to aldehyde is greatly limited. Recently, an elegant method that the arylzinc species prepared in situ by transmetalation between organoboron [17,18,19] or organoboronic derivatives [20,21,22,23,24,25,26] and diethylzinc has been proposed as an alternative for the synthesis of salt-free organozinc reagents. We have also successfully developed an efficient and practical method for the synthesis of diarylmethanols by transmetalation using the arylboronic acid in the presence of trimethylgallium [27]. These methods allow the exploitation of a broad range of substituted aryl transfer reagents since numerous arylboronic acids are commercially available, and a lot of excellent ligands were developed and applied to the asymmetric aryl transfer reaction with good results [28,29,30,31,32,33,34,35,36,37,38,39,40,41]. For the future, the introduction of the new, effective and more easily available catalysts is also a field of continuous interest for the catalytic aryl transfer reaction.
Chiral C2-symmetric diamines have emerged as versatile auxiliaries or ligands in numerous asymmetric transformations [42,43,44]. (R,R)-2,2’-bispyrrolidine, initially developed by Hirama, was synthesized by various routes [45,46,47,48,49,50], and its derivatives had been successfully employed as chiral ligands or organocatalysts in many asymmetric reactions [51,52,53,54,55,56,57,58,59]. So far, the application of 2,2’-bispyrrolidine-based salan ligands [60,61] in asymmetric catalysis has not been reported. We describe herein our efforts toward the synthesis of optically active diarylmethanols through the asymmetric aryl transfer to aldehydes under the catalysis of (R,R)-2,2’-bispyrrolidine-based salan ligands.

2. Results and Discussion

A preliminary study was performed to test the catalytic property of the ligands L1-L6 (Figure 1) in the asymmetric phenyl transfer reaction to 4-nitrobenzaldehyde at 0 °C. As is evident from Table 1, the resulting products could be obtained in moderate yield, but low enantioselectivity when (1R,2R)-cyclohexane-1,2-diamine-based ligands L1-L4 were tested (Table 1, entries 1-4). Gratifyingly, we found that the ligands L5 and L6 were more effective in this reaction (Table 1, entries 5-6). The ee value of the product could be increased to 63% when the reaction was carried out at −25 °C (Table 1, entry 7). Increasing catalyst loading had a positive impact on both the yield and enantioselectivity. The best result was obtained in 88% yield with 83% ee while using 20 mol% of L6 (Table 1, entry 9).
After having established the optimal protocol for the asymmetric phenyl transfer reaction, we further extended the reaction to a series of aldehyde substrates (Table 2). The electronic properties of the aromatic rings of the aldehydes have a significant influence on the enantioselectivity in this reaction. The aldehydes with electron-withdrawing substituents provided better results than those with electron-donating substituents in terms of ee values. 4-Nitrobenzaldehyde gave the corresponding diarylmethanol with 83% ee, but 4-methoxybenzaldehyde only with 11% ee (Table 2, entries 1, 2 and 10). Similar results were obtained when 3-substituted-benzaldehydes (Table 2, entries 3 and 9) or 2-substituted-benzaldehydes (Table 2, entries 5, 6 and 8) were tested. However, an exception was observed for 2-nitrobenzaldehyde (Table 2, entry 4), presumably caused by the chelating effect of the NO2 group with the lewis acids [62,63]. The enantioselectivity was also found to be influenced by the steric effect with the same exception of 2-nitrobenzaldehyde. ortho-Substituted (-Cl or -Me) benzaldehydes gave higher ee values (Table 2, entries 6 vs 2 or 8 vs 9). It should be noted that the reaction of 2-naphthaldehyde proceeded well, giving 70% ee and good yield (Table 2, entry 11), and α,β-unsaturated cinnamaldehyde gave the corresponding product with only moderate enantioselectivity (Table 2, entry 12).
We also further investigated the asymmetric aryl transfer to aromatic aldehydes with substituted phenylboronic acids. As shown in Table 3, when 4-chlorophenylboronic acid was chosen as the aryl source and 4-nitrobenzaldehyde as the substrate, 71% ee was obtained (Table 3, entry 1). And 54% ee was obtained when 4-methoxylphenylboronic acid was tested (Table 3, entry 3).

3. Experimental

3.1. General

All reactions were carried out under an argon atmosphere using standard Schlenk techniques. Solvents were dried and distilled prior to use according to standard methods. Unless otherwise indicated, all materials were obtained from commercial sources and liquid aldehydes were freshly distilled before use. For thin-layer chromatography (TLC), compounds were visualized by irradiation with UV light on GF 254 silica gel plates. 1H-NMR and 13C-NMR spectra were recorded in CDCl3 on a Bruker ARX-300 spectrometer with chemical shifts being referenced to SiMe4 used as internal standard. The coupling constants J are given in Hz. HPLC analysis were performed on a chiral column (Daicel Chiralcel OB-H, OD-H or AD-H column) on a Chromatography Interface 600 Series Link instrument and Series 200 pump), with Series 200 UV/VIS detection at 254 nm. The solvent system used has hexane (A)-2-propanol (B) in the indicated proportions. Optical rotations were measured on Rudolph Research Analytical Autopol III Automatic Polarimeter equipped with a 100 mm cell. Mass spectra (EI-MS) were taken using a Shimadzu GCMS-QP2010 mass spectrometer. High Resolution Mass Spectra (HRMS) were taken using a LTQ Orbitrap XL ThermoFisher unit.

3.2. Typical Procedure for the Asymmetric Aryl Transfer Reaction

In a 20 mL flame-dried Schlenk reaction tube, diethylzinc (0.9 mmol, 6 equiv, 1.5 M in toluene solution) was added dropwise to a solution of phenylboronic acid (0.3 mmol, 2 equiv) in toluene (3 mL) under an argon atmosphere. After stirring for 12 h at 60 °C, a toluene solution of L6 (20 mol%) was introduced. The reaction was stirred for an additional 30 minutes and cooled to −25 °C followed by the addition of aldehydes (0.15 mmol). After completion of the reaction (monitored by TLC), the reaction solution was quenched with saturated aqueous NH4Cl (3 mL) and further extracted with ethyl acetate (3 × 5 mL). The combined organic layer was dried over Na2SO4. Evaporation of the solvent gave the crude product, which was further purified by preparative TLC to afford the corresponding chiral diarylmethanols.
(S)-4-Nitrophenyl(phenyl)methanol (3a). 1H-NMR: δ 8.19 (d, J = 7.2 Hz, 2H), 7.58 (d, J = 7.2 Hz, 2H), 7.37–7.33 (m, 5H), 5.92 (s, 1H), 2.25 (brs, 1H). 83% ee determined by HPLC with a Chiralcel OB-H column (A/B = 70:30, 0.8 mL/min, uv 230 nm): tR = 21.05 min (minor), tR = 35.74 min (major). [α]D23 = +31.6 (c = 0.50, EtOH).
(S)-4-Chlorophenyl(phenyl)methanol (3b). 1H-NMR: δ 7.38–7.33 (m, 4H), 7.31–7.27 (m, 5H), 5.80 (s, 1H), 2.20 (brs, 1H). 41% ee determined by HPLC with a Chiralcel OB-H column (A/B = 90:10, 1.0 mL/min, uv 230 nm): tR = 10.21 min (minor), tR = 18.33 min (major). [α]D23 = +5.9 (c = 0.64, EtOH).
(S)-3-Nitrophenyl(phenyl)methanol (3c). 1H-NMR: δ 8.30 (s, 1H), 8.11 (d, J = 8.1 Hz, 1H), 7.72 (d, J = 7.8 Hz, 1H), 7.50 (t, J = 7.8 Hz, 1H), 7.40–7.29 (m, 5H), 5.92 (s, 1H), 2.13 (brs, 1H). 75% ee determined by HPLC with a Chiralcel OB-H column (A/B = 80:20, 0.8 mL/min, uv 230 nm): tR = 34.19 min (minor), tR = 47.40 min (major). [α]D23 = +42.5 (c = 0.40, EtOH).
(R)-2-Nitrophenyl(phenyl)methanol (3d). 1H-NMR: δ 7.94 (dd, J = 7.8, 1.5 Hz, 1H), 7.75 (dd, J = 7.8, 1.5 Hz, 1H), 7.64 (dt, J = 7.5, 1.2 Hz, 1H), 7.46 (t, J = 7.8, 1.5 Hz, 1H), 7.36–7.29 (m, 5H), 6.44 (s, 1H), 2.02 (brs, 1H). 41% ee determined by HPLC with a Chiralpark AD-H column (A/B = 90:10, 0.8 mL/min, uv 254 nm): tR = 13.57 min (major), tR = 14.62 min (minor); [α]D23 = 11.2 (c = 0.32, EtOH).
(R)-2-Trifluoromethylphenyl(phenyl)methanol (3e). 1H-NMR: δ 7.66 (t, J = 7.8 Hz, 2H), 7.55 (t, J = 7.8 Hz, 1H), 7.42–7.32 (m, 5H), 7.30–7.27 (m, 1H), 6.32 (s, 1H), 1.99 (brs, 1H). 80% ee determined by HPLC with a Chiralcel OD-H column (A/B = 90:10, 0.5 mL/min, uv 254 nm): tR = 9.33 min (major), tR = 11.79 min (minor). [α]D23 = −37.2 (c = 0.5, EtOH).
(S)-2-Chlorophenyl(phenyl)methanol (3f). 1H-NMR: δ 7.60 (d, J = 7.8 Hz, 1H), 7.42–7.39 (m, 2H), 7.36–7.28 (m, 5H), 7.25–7.22 (m, 1H), 6.24 (s, 1H), 2.05 (brs, 1H). 73% ee determined by HPLC with a Chiralcel OB-H column (A/B = 90:10, 1.0 mL/min, uv 230 nm): tR = 8.97 min (minor), tR = 10.00 min (major). [α]D23 = −20.6 (c = 0.64, EtOH).
(S)-3-Bromophenyl(phenyl)methanol (3g). 1H-NMR: δ 7.57 (s, 1H), 7.42–7.35 (m, 5H), 7.33–7.27 (m, 2H), 7.20 (t, J = 7.8 Hz, 1H), 5.78 (s, 1H), 2.33 (brs, 1H). 26% ee determined by HPLC with a Chiralcel OB-H column (A/B = 90:10, 1.0 mL/min, uv 230 nm): tR = 15.17 min (minor), tR = 27.81 min (major). [α]D23 = +11.4 (c = 0.76, EtOH).
(S)-2-Methylphenyl(phenyl)methanol (3h). 1H-NMR: δ 7.53 (d, J = 9.0 Hz, 1H), 7.34–7.21(m, 7H), 7.16 (t, J = 8.1 Hz, 1H), 6.02 (s, 1H), 2.26 (s, 3H), 1.95 (s, 1H). 60% ee determined by HPLC with a Chiralcel OB-H column (A/B = 90:10, 1.0 mL/min, uv 230 nm): tR = 9.47 min (minor), tR = 10.6 min (major). [α]D23 = −19.3 (c = 0.30, EtOH).
(S)-3-Methylphenyl(phenyl)methanol (3i). 1H-NMR: δ 7.41–7.34 (m, 4H), 7.30–7.27 (m, 2H), 7.24–7.16 (m, 2H), 7.09 (d, J = 7.5 Hz, 1H), 5.81 (s, 1H), 2.35 (s, 3H), 2.02 (brs, 1H). 52% ee determined by HPLC with a Chiralcel OB-H column (A/B = 90:10, 1.0 mL/min, uv 230 nm): tR = 12.39 min (minor), tR = 21.34 min (major). [α]D23 = −15.8 (c = 0.34, EtOH).
(S)-4-Methoxylphenyl(phenyl)methanol (3j). 1H-NMR: δ 7.37–7.34 (m, 3H), 7.30–7.26 (m, 4H), 6.88 (d, J = 9.0 Hz, 2H), 5.81 (s, 1H), 3.78 (s, 3H), 2.23 (brs, 1H). 11% ee determined by HPLC with a Chiralcel OB-H column (A/B = 90:10, 1.0 mL/min, uv 230 nm): tR = 21.79 min (minor), tR = 24.03 min (major). [α]D23 = 8.1 (c = 0.42, EtOH).
(S)-2-Naphathyl(phenyl)methanol (3k). 1H-NMR: δ 7.90 (s, 1H), 7.86–7.79 (m, 3H), 7.50–7.42 (m, 5H), 7.38–7.28 (m, 3H), 6.01 (s, 1H), 2.06 (brs, 1H). 70% ee determined by HPLC with a Chiralcel OD-H column (A/B = 85:15, 0.8 mL/min, uv 230 nm): tR = 12.71 min (major), tR = 15.08 min (minor). [α]D23 = −18.4 (c = 0.46, EtOH).
(R)-1, 3-Diphenylprop-2-en-1-ol (3l). 1H-NMR: δ 7.47–7.33 (m, 5H), 7.32–7.27 (m, 3H), 7.29–7.24 (m, 2H), 6.70 (d, J = 15.6 Hz, 1H), 6.40 (dd, J = 6.3, 15.6 Hz, 1H), 5.40 (d, J = 6.6 Hz, 1H), 2.15 (brs, 1H). 47%ee determined by HPLC with a Chiralcel OD-H column (A/B = 80:20, 0.8 mL/min, uv 254 nm): tR = 9.31 min (minor), tR = 11.14 min (major). [α]D20 = +13.5 (c = 0.40, EtOH).
4-Chlorophenyl(4-nitrophenyl)methanol (4a). 1H-NMR: δ 8.20 (d, J = 8.7 Hz, 2H), 7.55 (d, J = 8.7 Hz, 2H), 7.36–7.27 (m, 4H), 5.90 (s, 1H), 2.04 (brs, H). 71% ee determined by HPLC with a Chiralcel OB-H column (A/B = 80:20, 0.8 mL/min, uv 230 nm): tR = 22.92 min (minor), tR = 25.07 min (major). [α]D23 = −19.5 (c = 0.64, EtOH).
4-Chlorophenyl(3-bromophenyl)methanol (4b). 1H-NMR: δ 7.53 (s, 1H), 7.41 (d, J = 7.5 Hz, 1H), 7.39–7.28 (m, 4H), 7.25–7.24 (m, 2H), 7.20 (t, J = 7.5 Hz, 1H), 5.77 (s, 1H), 2.04 (brs, 2H). 55% ee determined by HPLC with a Chiralcel OD-H column (A/B = 85:15, 0.8 mL/min, uv 230 nm): tR = 7.80 min (major), tR = 8.58 min (minor). [α]D23 = +22.8 (c = 0.60, EtOH).
(S)-4-Methoxylphenyl(4-nitrophenyl)methanol (4c). 1H-NMR: δ 8.19 (d, J = 8.7 Hz, 2H), 7.57 (d, J = 8.4 Hz, 2H), 7.25 (d, J = 8.4 Hz, 2H), 6.89 (d, J = 8.7 Hz, 2H), 5.88 (s, 1H), 3.80 (s, 3H), 2.20 (s, 1H). 54%ee determined by HPLC with a Chiralpark AD-H column (A/B = 85:15, 0.8 mL/min, uv 254 nm): tR = 15.60 min (minor), tR = 19.25 min (major). [α]D23 = +27.9 (c = 0.44, EtOH).
4-Methoxylphenyl(3-nitrophenyl)methanol (4d). 1H-NMR: δ 8.28 (s, 1H), 8.10 (d, J = 8.1 Hz, 1H), 7.71 (d, J = 8.1 Hz, 1H), 7.49 (t, J = 8.1 Hz, 1H), 7.27 (d, J = 6.6 Hz, 2H), 6.89 (d, J = 6.9 Hz, 2H), 5.88 (s, 1H), 3.80 (s, 3H), 2.28 (brs, 1H). 24% ee determined by HPLC with a Chiralcel OD-H column (A/B = 85:15, 0.8 mL/min, uv 230 nm): tR =15.08 min (major), tR =16.23 min (minor). [α]D23 = +23.8 (c = 0.50, EtOH).
3, 5-Dimethylphenyl(phenyl)methanol (4e). 1H-NMR: δ 7.42–7.28 (m, 5H), 7.01 (s, 2H), 6.93 (s, 1H), 5.77 (s, 1H), 2.31 (s, 6H), 2.18 (brs, 1H). 48% ee determined by HPLC with a Chiralcel OD-H column (A/B = 90:10, 0.8 mL/min, uv 254 nm): tR = 8.67 min (minor), tR = 9.58 min (major). [α]D23 = +20.4 (c = 0.65, EtOH).

4. Conclusions

In summary, we haved reported the asymmetric aryl transfer to aldehydes with arylboronic acids as aryl sources in the presence of the chiral 2,2’-bispyrrolidine-based ligand L6. The corresponding diarylmethanols could be obtained in high yields with moderate to good enantioselectivities. Further work on the asymmetric addition mechanism and the broad application of chiral 2,2’-bispyrrolidine-based ligands in other asymmetric catalytic reactions are now in progress in our laboratory.

Acknowledgements

We gratefully acknowledge the National Natural Science Foundation of China (20832001, 20972065, 21074054) and the National Basic Research Program of China (2007CB925103, 2010CB923303) for their financial support. The Major Scientific and Technological Special Project (2009ZX09103-081) is also acknowledged.

References and Notes

  1. Shafi’ee, A.; Hite, G. Absolute configurations of the pheniramines, methyl phenidates, and pipradrols. J. Med. Chem. 1969, 12, 266–270. [Google Scholar] [CrossRef] [PubMed]
  2. Ebnǒther, A.; Weber, H.P. Synthesis and absolute configuration of clemastine and its isomers. Helv. Chim. Acta 1976, 59, 2462–2468. [Google Scholar] [CrossRef] [PubMed]
  3. Meguro, K.; Aizawa, M.; Sohda, T.; Kawamatsu, A.; Nagaoka, A. New 1,4-dihydropyridine derivatives with potent and long-lasting hypotensive effect. Chem. Pharm. Bull. 1985, 33, 3787–3797. [Google Scholar] [CrossRef] [PubMed]
  4. Toda, F.; Tanaka, K.; Koshiro, K. A new preparative method for optically active diarylcarbinols. Tetrahedron Asymmetry 1991, 2, 873–874. [Google Scholar] [CrossRef]
  5. Casy, A.F.; Drake, A.F.; Ganellin, C.R.; Mercer, A.D.; Upton, C. Stereochemical studies of chiral H-1 antagonists of histamine: The resolution, chiral analysis, and biological evaluation of four antipodal pairs. Chirality 1992, 4, 356–366. [Google Scholar] [CrossRef] [PubMed]
  6. Stanev, S.; Rakovska, R.; Berova, N.; Snatzke, G. Synthesis, absolute configuration and circular dichroism of some diarylmethane derivatives. Tetrahedron Asymmetry 1995, 6, 183–198. [Google Scholar] [CrossRef]
  7. Botta, M.; Summa, V.; Corelli, F.; Di Pietro, G.; Lombardi, P. Synthesis of aryl 2-benzofuranyl and 2-indolyl carbinols of high enantiomeric purity via palladium-catalyzed heteroannulation of chiral arylpropargylic alcohols. Tetrahedron Asymmetry 1996, 7, 1263–1266. [Google Scholar] [CrossRef]
  8. Torrens, A.; Castrillo, J.A.; Claparols, A.; Redondo, J. Enantioselective synthesis of (R)- and (S)-cizolirtine. Application of oxazaborolidine-catalyzed asymmetric borane reduction to azolyl phenyl ketones. Synlett 1999, 765–767. [Google Scholar] [CrossRef]
  9. Corey, E.J.; Helal, C.J. Reduction of carbonyl compounds with chiral oxazaborolidine catalysts: A new paradigm for enantioselective catalysis and a powerful new synthetic method. Angew. Chem. Int. Ed. 1998, 37, 1986–2012. [Google Scholar] [CrossRef]
  10. Corey, E.J.; Bakshi, R.K.; Shibata, S. Highly enantioselective borane reduction of ketones catalyzed by chiral oxazaborolidines. Mechanism and synthetic implications. J. Am. Chem. Soc. 1987, 109, 5551–5553. [Google Scholar] [CrossRef]
  11. Corey, E.J. New enantioselective routes to biologically interesting compounds. Pure Appl. Chem. 1990, 62, 1209–1216. [Google Scholar] [CrossRef]
  12. Ohkuma, T.; Koizumi, M.; Ikehira, H.; Yokozawa, T.; Noyori, R. Selective hydrogenation of benzophenones to benzhydrols. Asymmetric synthesis of unsymmetrical diarylmethanols. Org. Lett. 2000, 2, 659–662. [Google Scholar] [CrossRef] [PubMed]
  13. Noyori, R.; Ohkuma, T. Rapid, productive and stereoselective hydrogenation of ketones. Pure Appl. Chem. 1999, 71, 1493–1501. [Google Scholar] [CrossRef]
  14. Pu, L.; Yu, H.-B. Catalytic asymmetric organozinc additions to carbonyl compounds. Chem. Rev. 2001, 101, 757–824. [Google Scholar] [CrossRef] [PubMed]
  15. Bolm, C.; Hildebrand, J.P.; Muńiz, K.; Hermanns, N. Catalyzed asymmetric arylation reacti­ons. Angew. Chem. Int. Ed. 2001, 40, 3284–3308. [Google Scholar] [CrossRef]
  16. Schmidt, F.; Stemmler, R.T.; Rudolph, J.; Bolm, C. Catalytic asymmetric approaches towards enantiomerically enriched diarylmethanols and diarylmethylamines. Chem. Soc. Rev. 2006, 35, 454–470. [Google Scholar] [PubMed]
  17. Langer, F.; Schwink, L.; Devasagayaraj, A.; Chavant, P.-Y.; Knochel, P. Preparation of functiona­lized dialkylzincs via a boron-zinc exchange. Reactivity and catalytic asymmetric addition to aldehydes. J. Org. Chem. 1996, 61, 8229–8243. [Google Scholar] [CrossRef] [PubMed]
  18. Oppolzer, W.; Radinov, R.N.; Sayed, E.E. Catalytic asymmetric synthesis of macrocyclic (E)-allylic alcohols from ω-alkynals via intramolecular 1-alkenylzinc/aldehyde additions. J. Org. Chem. 2001, 66, 4766–4770. [Google Scholar] [CrossRef] [PubMed]
  19. Dahmen, S.; Brase, S. [2, 2] Paracyclophane-based N,O-ligands in alkenylzinc additions to Alde­hydes. Org. Lett. 2001, 3, 4119–4122. [Google Scholar] [CrossRef] [PubMed]
  20. Bolm, C.; Rudolph, J. Catalyzed asymmetric aryl transfer reactions to aldehydes with boronic acids as aryl source. J. Am. Chem. Soc. 2002, 124, 14850–14851. [Google Scholar] [CrossRef] [PubMed]
  21. Rudolph, J.; Schmidt, F.; Bolm, C. Highly enantioselective synthesis of secondary alcohols using triphenylborane. Adv. Synth. Catal. 2004, 346, 867–872. [Google Scholar] [CrossRef]
  22. Braga, A.L.; Luedtke, D.S.; Vargas, F.; Paixao, M.W. Catalytic enantioselective arylation of aldehydes: Boronic acids as a suitable source of transferable aryl groups. Chem. Commun. 2005, 2512–2514. [Google Scholar] [CrossRef] [PubMed]
  23. Wu, X.-Y.; Liu, X.-Y.; Zhao, G. Catalyzed asymmetric aryl transfer reactions to aldehydes with boroxines as aryl source. Tetrahedron Asymmetry 2005, 16, 2299–2305. [Google Scholar] [CrossRef]
  24. Liu, X.-Y.; Wu, X.-Y.; Chai, Z.; Wu, Y.-Y.; Zhao, G.; Zhu, S.-Z. Highly effective and recyclable dendritic ligands for the enantioselective aryl transfer reactions to aldehydes. J. Org. Chem. 2005, 70, 7432–7435. [Google Scholar] [CrossRef] [PubMed]
  25. Bolm, C.; Schmidt, F.; Zani, L. Synthesis of new chiral hydroxy oxazolines and their use in the catalytic asymmetric phenyl transfer to aldehydes. Tetrahedron Asymmetry 2005, 16, 1367–1376. [Google Scholar] [CrossRef]
  26. Dahmen, S.; Lormann, M. Triarylborane ammonia complexes as ideal precursors for arylzinc reagents in asymmetric catalysis. Org. Lett. 2005, 7, 4597–4600. [Google Scholar] [CrossRef] [PubMed]
  27. Jia, X.-F.; Fang, L.; Lin, A.-J.; Pan, Y.; Zhu, C.-J. Highly efficient and facile aryl transfer to aldehydes using ArB(OH)2-GaMe3. Synlett 2009, 3, 495–499. [Google Scholar]
  28. Ji, J.-X.; Wu, J.; Au-Yeung, T.-T.-L.; Yip, C.W.; Haynes, R.K.; Chan, A.S.C. Highly enantioselective phenyl transfer to aryl aldehydes catalyzed by easily accessible chiral tertiary aminonaphthol. J. Org. Chem. 2005, 70, 1093–1095. [Google Scholar] [CrossRef] [PubMed]
  29. Ito, K.; Tomita, Y.; Katsuki, T. Enantioselective phenyl transfer to aldehydes using 1,1'-bi-2-naphthol-3,3'-dicarboxamide as chiral auxiliary. Tetrahedron Lett. 2005, 46, 6083–6086. [Google Scholar] [CrossRef]
  30. Braga, A.L.; Milani, P.; Vargas, F.; Paixāo, M.W.; Sehnem, J.A. Modular chiral thiazolidine catalysts in asymmetric aryl transfer reactions. Tetrahedron Asymmetry 2006, 17, 2793–2797. [Google Scholar] [CrossRef]
  31. Wu, P.-Y.; Wu, H.-L.; Uang, B.-J. Asymmetric synthesis of functionalized diarylmethanols catalyzed by a new γ-amino thiol. J. Org. Chem. 2006, 71, 833–835. [Google Scholar] [CrossRef] [PubMed]
  32. Lu, G.; Kwong, F.-Y.; Ruan, J.W.; Li, Y.-M.; Chan, A.S.C. Highly enantioselective addition of in situ prepared arylzinc to aldehydes catalyzed by a series of atropisomeric binaphthyl-derived amino alcohols. Chem. Eur. J. 2006, 12, 4115–4120. [Google Scholar] [CrossRef] [PubMed]
  33. Jin, M.J.; Sarkar, S.M.; Lee, D.H.; Qiu, H.L. Highly enantioselective aryl transfer to aldehydes: A remarkable effect of sulfur substitution in amino thioacetate ligands. Org. Lett. 2008, 10, 1235–1237. [Google Scholar] [CrossRef] [PubMed]
  34. Huang, X.-B.; Wu, L.-L.; Xu, J.-Q.; Zong, L.-L.; Hu, H.-W.; Cheng, Y.-X. Enantioselective arylation of aldehydes catalyzed by a soluble optically active polybinaphthols ligand. Tetrahedron Lett. 2008, 49, 6823–6826. [Google Scholar] [CrossRef]
  35. DeBerardinis, A.M.; Turlington, M.; Pu, L. Activation of functional arylzincs prepared from aryl iodides and highly enantioselective addition to aldehydes. Org. Lett. 2008, 10, 2709–2712. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, M.-C.; Wang, X.-D.; Ding, X.; Liu, Z.-K. Catalytic asymmetric aryl transfer: Highly enantioselective preparation of (R)- and (S)-diarylmethanols catalyzed by the same chiral ferrocenyl aziridino alcohol. Tetrahedron 2008, 64, 2559–2564. [Google Scholar] [CrossRef]
  37. Ruan, J.-W.; Lu, G.; Xu, L.-J.; Li, Y.-M.; Chan, A.S.C. Catalytic asymmetric alkynylation and arylation of aldehydes by an H8-binaphthyl-based amino alcohol ligand. Adv. Synth. Catal. 2008, 350, 76–84. [Google Scholar] [CrossRef]
  38. Wang, M.-C.; Zhang, Q.-J.; Zhao, W.-X.; Wang, X.-D.; Ding, X.; Jing, T.-T.; Song, M.-P. N-(Ferrocenylmethyl)azetidin-2-yl(diphenyl)methanol for catalytic asymmetric addition of organozinc reagents to aldehydes. J. Org. Chem. 2008, 73, 168–176. [Google Scholar] [CrossRef] [PubMed]
  39. Braga, A.L.; Paixa, M.W.; Westermann, B.; Schneider, P.H.; Wessjohann, L.A. Acceleration of arylzinc formation and its enantioselective addition to aldehydes by microwave irradiation and aziridine-2-methanol catalysts. J. Org. Chem. 2008, 73, 2879–2882. [Google Scholar] [CrossRef] [PubMed]
  40. Salvi, L.; Kim, J.G.; Walsh, P.J. Practical catalytic asymmetric synthesis of diaryl-, aryl heteroaryl-, and diheteroarylmethanols. J. Am. Chem. Soc. 2009, 131, 12483–12493. [Google Scholar] [CrossRef] [PubMed]
  41. Wouters, A.D.; Trossini, G.H.G.; Stefani, H.A.; Lüdtke, D.S. Enantioselective arylations catalyzed by carbohydrate-based chiral amino alcohols. Eur. J. Org. Chem. 2010, 2351–2356. [Google Scholar] [CrossRef]
  42. Alexakis, A.; Mangeney, P. Advanced Asymmetric Synthesis; Stephenson, G.R., Ed.; Chapman & Hall: London, UK, 1996; Chapter 5; p. 93. [Google Scholar]
  43. Bennani, Y.L.; Hannessian, S. Trans-1,2-diaminocyclohexane derivatives as chiral reagents, scaffolds, and ligands for catalysis: Applications in asymmetric synthesis and molecular recognition. Chem. Rev. 1997, 97, 3161–3195. [Google Scholar] [CrossRef] [PubMed]
  44. Lucet, D.; Le Gall, T.; Mioskowski, C. The chemistry of vicinal diamines. Angew. Chem. Int. Ed. 1998, 37, 2580–2627. [Google Scholar] [CrossRef]
  45. Hirama, M.; Oishi, T.; Ito, S. Asymmetric dihydroxylation of alkenes with osmium tetroxide: Chiral N, N'-dialkyl-2,2'-bipyrrolidine complex. Chem. Commun. 1989, 10, 665–666. [Google Scholar] [CrossRef]
  46. Oishi, T.; Hirama, M. Highly enantioselective dihydroxylation of trans-disubstituted and monosubstituted olefins. J. Org. Chem. 1989, 54, 5834–5835. [Google Scholar] [CrossRef]
  47. Oishi, T.; Hirama, M.; Sita, L.R.; Masamune, S. Synthesis of chiral 2, 2'-bipyrrolidine derivatives. Synthesis 1991, 789–792. [Google Scholar] [CrossRef]
  48. Kotsuki, H.; Kuzume, H.; Gohda, T.; Fukuhara, M.; Ochi, M.; Oishi, T.; Hirama, M.; Shirot, M. New convenient, enantiospecific synthesis of (S, S)- and (R, R)-2,2'-bipyrrolidine derivatives. Tetrahedron Asymmetry 1995, 6, 2227–2236. [Google Scholar] [CrossRef]
  49. Alexakis, A.; Tomassini, A.; Chouillet, C.; Roland, S.; Mangeney, P.; Bernardinelli, G. A new efficient synthesis of (R, R)-2,2'-bipyrrolidine: An interesting chiral 1,2-diamine with C2 symmetry. Angew. Chem. Int. Ed. 2000, 39, 4093–4095. [Google Scholar] [CrossRef]
  50. Denmark, S.E.; Fu, J.-P.; Lawler, M.J. (R, R)-2,2'- bipyrrolidine and (S, S)-2,2'-bipyrrolidine: Useful ligands for asymmetric synthesis. Org. Synth. 2006, 83, 121–130. [Google Scholar]
  51. Denmark, S.E.; Fu, J.-P. Catalytic, Enantioselective addition of substituted allylic trichlorosilanes using a rationally-designed 2,2'-bispyrrolidine-based bisphosphoramide. J. Am. Chem. Soc. 2001, 123, 9488–9489. [Google Scholar] [CrossRef] [PubMed]
  52. Alexakis, A.; Andrey, O. Diamine-catalyzed asymmetric michael additions of aldehydes and ketones to nitrostyrene. Org. Lett. 2002, 4, 3611–3614. [Google Scholar] [CrossRef] [PubMed]
  53. Andrey, O.; Alexakis, A.; Bernardinelli, G. Asymmetric michael addition of α-Hydroxyketones to nitroolefins catalyzed by chiral diamine. Org. Lett. 2003, 5, 2559–2561. [Google Scholar] [CrossRef] [PubMed]
  54. Andrey, O.; Alexakis, A.; Tomassini, A.; Bernardinelli, G. The use of N-alkyl-2,2’-bipyrrolidine derivatives as organocatalysts for the asymmetric michael addition of ketones and aldehydes to nitroolefins. Adv. Synth. Catal. 2004, 346, 1147–1168. [Google Scholar] [CrossRef]
  55. Mossé, S.; Alexakis, A. First organocatalyzed asymmetric michael addition of aldehydes to vinyl Sulfones. Org. Lett. 2005, 7, 4361–4364. [Google Scholar] [CrossRef] [PubMed]
  56. Alexakis, A.; Tomassini, A.; Andrey, O.; Bernardinelli, G. Diastereoselective alkylation of (arene)tricarbonylchromium and ferrocene complexes using a chiral, C2-symmetrical 1,2-diamine as auxiliary. Eur. J. Org. Chem. 2005, 1332–1339. [Google Scholar] [CrossRef]
  57. Denmark, S.E.; Fu, J.-P.; Lawler, M.J. Chiral phosphoramide-catalyzed enantioselective addition of allylic trichlorosilanes to aldehydes. Preparative studies with bidentate phosphorus-based amides. J. Org. Chem. 2006, 71, 1523–1536. [Google Scholar] [CrossRef] [PubMed]
  58. Chen, M.S.; White, M.C. A predictably selective aliphatic C-H oxidation reaction for complex molecule synthesis. Science 2007, 318, 783–787. [Google Scholar] [CrossRef] [PubMed]
  59. Suzuki, K.; Oldenburg, P.D.; Que, L., Jr. Iron-catalyzed asymmetric olefin cis-dihydroxylation with 97% enantiomeric excess. Angew. Chem. Int. Ed. 2008, 47, 1887–1889. [Google Scholar] [CrossRef] [PubMed]
  60. Sergeeva, E.; Kopilov, J.; Goldberg, I.; Kol, M. Salan ligands assembled around chiral bipyrrolidine: Predetermination of chirality around octahedral Ti and Zr centres. Chem. Commun. 2009, 21, 3053–3055. [Google Scholar] [CrossRef] [PubMed]
  61. Sergeeva, E.; Kopilov, J.; Goldberg, I.; Kol, M. 2,2'-Bipyrrolidine versus 1,2-diaminocyclohexane as chiral cores for helically wrapping diamine-diolate ligands. Inorg. Chem. 2009, 48, 8075–8077. [Google Scholar] [CrossRef] [PubMed]
  62. Jia, Y.-X.; Zhu, S.-F.; Yang, Y.; Zhou, Q.-L. Asymmetric friedel-crafts alkylations of indoles with nitroalkenes catalyzed by Zn(II)-bisoxazoline complexes. J. Org. Chem. 2006, 71, 75–80. [Google Scholar] [CrossRef] [PubMed]
  63. Arai, T.; Yokoyama, N. Tandem Catalytic Asymmetric friedel-crafts/henry reaction: Control of three contiguous acyclic stereocenters. Angew. Chem. Int. Ed. 2008, 47, 4989–4992. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Samples of the compounds are available from the authors.
Figure 1. Structures of Ligands L1-L6.
Figure 1. Structures of Ligands L1-L6.
Molecules 16 02971 g001
Table 1. Asymmetric Phenyl Transfer to 4-nitrobenzaldehyde. a
Table 1. Asymmetric Phenyl Transfer to 4-nitrobenzaldehyde. a
Molecules 16 02971 i001
EntryLigandMol%T(°C)Yield(%)bEe(%)c
1L1100666
2L21007311
3L31006916
4L4100803
5L51007431
6L61008443
7L610-257063
8L615-258071
9L620-258883(S) d
a All the reactions were carried out on 0.2 mmol scale of substrates with 2 equiv of arylboronic acid and 6 equiv of Et2Zn in toluene for 24 h. b Isolated yields. c Determined by HPLC with a Chiralcel OB-H column. d The absolute configuration of the products were determined by comparison with literature values.
Table 2. Asymmetric Phenyl Transfer to Aromatic Aldehydes. a
Table 2. Asymmetric Phenyl Transfer to Aromatic Aldehydes. a
Molecules 16 02971 i002
EntryArProductYield (%)bEe (%)c,d
14-NO2C6H43a8883(S)
24-ClC6H43b8041(S)
33-NO2C6H43c9175(S)
42-NO2C6H43d8541(R)
52-CF3C6H43e7280(R)
62-ClC6H43f8473(S)
73-BrC6H43g8526(S)
82-MeC6H43h7660(S)
93-MeC6H43i7852(S)
104-MeOC6H43j8011(S)
112-C10H73k8070(S)
12PhCH=CH3l7647(R)
Table 3. Asymmetric Aryl Transfer to Aldehydes. a
Table 3. Asymmetric Aryl Transfer to Aldehydes. a
Molecules 16 02971 i003
EntryAr1Ar2ProductYield (%)bEe (%)c
14-Cl C6H44-NO2C6H44a7571
24-Cl C6H43-BrC6H44b8255
34-MeOC6H44-NO2C6H44c7854(S)d
44-MeOC6H43-NO2C6H44d7024
53,5-diMeC6H3C6H54e7048
a All reactions were carried out on 0.15 mmol scale of substrates with 2 equiv of arylboronic acid and 6 equiv of Et2Zn in toluene at −25 °C for 24 h in the presence of 20 mol% ligand; b Isolated yields; c Enantiomeric excess was determined by HPLC with a Chiralcel OB-H, OD-H or AD-H column; d The absolute configuration of the products were determined by comparison with literature values.

Share and Cite

MDPI and ACS Style

Jia, X.; Lin, A.; Mao, Z.; Zhu, C.; Cheng, Y. Catalytic Enantioselective Aryl Transfer to Aldehydes Using Chiral 2,2’-Bispyrrolidine-Based Salan Ligands. Molecules 2011, 16, 2971-2981. https://doi.org/10.3390/molecules16042971

AMA Style

Jia X, Lin A, Mao Z, Zhu C, Cheng Y. Catalytic Enantioselective Aryl Transfer to Aldehydes Using Chiral 2,2’-Bispyrrolidine-Based Salan Ligands. Molecules. 2011; 16(4):2971-2981. https://doi.org/10.3390/molecules16042971

Chicago/Turabian Style

Jia, Xuefeng, Aijun Lin, Zhijie Mao, Chengjian Zhu, and Yixiang Cheng. 2011. "Catalytic Enantioselective Aryl Transfer to Aldehydes Using Chiral 2,2’-Bispyrrolidine-Based Salan Ligands" Molecules 16, no. 4: 2971-2981. https://doi.org/10.3390/molecules16042971

Article Metrics

Back to TopTop