Next Article in Journal
Photoelectrochemical Oxidation in Ambient Conditions Using Earth-Abundant Hematite Anode: A Green Route for the Synthesis of Biobased Polymer Building Blocks
Previous Article in Journal
Preliminary Study of Extended Ripening Effects on Peptides Evolution and DPPH Radical Scavenging Activity in Mexican Goat Cheese
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Asymmetric Simmons-Smith Cyclopropanation and Diethylzinc Addition to Aldehydes Promoted by Enantiomeric Aziridine-Phosphines

by
Aleksandra Buchcic-Szychowska
1,
Justyna Adamczyk
1,
Lena Marciniak
2,
Adam Marek Pieczonka
1,
Anna Zawisza
1,
Stanisław Leśniak
1 and
Michał Rachwalski
1,*
1
Department of Organic and Applied Chemistry, University of Lodz, Tamka 12, 91-403 Lodz, Poland
2
The Bio-Med-Chem Doctoral School of the University of Lodz and Lodz Institutes of the Polish Academy of Sciences, University of Lodz, 91-403 Lodz, Poland
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(8), 968; https://doi.org/10.3390/catal11080968
Submission received: 31 July 2021 / Revised: 7 August 2021 / Accepted: 12 August 2021 / Published: 13 August 2021

Abstract

:
During an implementation of current research, a set of optically pure chiral aziridines and aziridine imines bearing a phosphine moiety was prepared with high values of chemical yield. The above chiral heteroorganic derivatives were tested for catalytic utility as chiral ligands in asymmetric Simmons-Smith cyclopropanation and asymmetric diethylzinc addition to various aldehydes. Most of the desired products were formed in high chemical yields, with satisfactory values of enantiomeric excess (sometimes more than 90%) and diastereomeric ratios (in case of cyclopropanation reaction).

Graphical Abstract

1. Introduction

The asymmetric construction of chiral non-racemic organic compounds is still an extremely important issue for many research teams around the world. Both the chemical yield and optical purity of the desired products of stereocontrolled synthesis depend on the selection of an appropriate promoter for such a reaction. This role can be played by a chiral ligand [1,2], an organocatalyst [3,4], a chiral auxiliary [5], bifunctional catalysts (molecule possessing two distinct functional groups) [6,7] or another cooperative catalytic system [8].
The asymmetric Simmons-Smith reaction is one of the approaches in the synthesis of cyclopropane derivatives [9]. A cyclopropane ring with unusual electronic and steric properties is present in many natural products [10] and drugs [11,12,13,14,15]. Molecules containing this three-membered motif can be constructed in transformations with or without transition metals [16] in the presence of, e.g., complexes of porphyrin [17], proline derivatives [18,19], chiral phase-transfer catalysts [20] or as demonstrated in our laboratory, using chiral aziridinyl ligands containing sulfinyl moiety [21].
In turn, asymmetric diethylzinc addition to carbonyl compounds (especially aromatic aldehydes) is one of the most exploited test reactions in the chemical literature [22,23]; however, it is still a very simple and effective tool for studying the catalytic activity of chiral ligands. Based on our experience, ligands containing an aziridine fragment show very high efficiency in reactions of this type [23,24,25,26,27,28].
Chiral systems containing a phosphine group and an aziridine ring within one molecule are not the subject of many reports in the chemical literature [29]. Enriching the current state of knowledge about such chiral junctions, we recently published a successful application of aziridine phosphines in the asymmetric Friedel-Crafts reaction [30].
Taking into consideration all of the aforementioned circumstances, we decided to synthesize a series of optically pure aziridines and aziridine imines possessing phosphine motif in order to check their catalytic activities in the asymmetric Simmons-Smith cyclopropanation and asymmetric diethylzinc addition to aldehydes. Thus, we wanted to show that chiral aziridine-phosphines are universal catalysts operating in various types of asymmetric transformations.

2. Results and Discussion

2.1. Preparation of the Aziridine-Phosphines 18

The chiral enantiomeric aziridine-phosphines 18 (Figure 1) were prepared as previously described [30], through triethoxysilane-mediated reduction of the corresponding phosphine oxides [30].

2.2. Synthesis of Imine-Phosphines 912

Chiral imine-phosphines 912 (Scheme 1) were obtained starting from appropriate 1-(2-aminoalkyl)aziridines, which were formed as a result of the zinc bromide-mediated self-ring opening reaction of NH-aziridines, as previously described [31]. The corresponding 1-(2-aminoalkyl)aziridines were then reacted with 2-(diphenylphosphino)benzaldehyde affording 1-(2-iminealkyl)aziridine catalysts functionalized by a phosphine group in excellent yields (Scheme 1).

2.3. Asymmetric Simmons-Smith Cyclopropanation Promoted by Aziridine-Phosphines 18 and Imine-Phosphines 912

Equipped with all optically pure phosphorus-functionalized aziridines 112, we decided to check their utility as chiral ligands catalyzing an asymmetric cyclopropanation reaction of cinnamyl alcohol (13) as a model substrate. The transformations were carried out in DCM with 2 equiv. of diethylzinc and 3 equiv. of diiodomethane in the presence of a ligand (10 mol%) (Scheme 2). All the results are presented in Table 1.
The analysis of the results collected in Table 1 leads to the following conclusions. Ligands 14 containing a methylene linker connecting a phenyl ring with an aziridine nitrogen atom were characterized by moderate catalytic efficiency leading to cyclopropyl derivative 14 in chemical yields varying from 63 to 72%, with enantiomeric excess in the range of 53 to 65% and with a moderate diastereomeric ratio of 5:1 (Table 1, entries 1–4). Much better catalytic activity was shown by aziridine-phosphines 58 (without methylene subunit). Their use in the asymmetric cyclopropanation allowed us to obtain the desired chiral product 14 in high yields of around 90%, 80–98% of ee, and with reasonable diastereoselectivity (10:1–15:1) (Table 1, entries 5–8). In turn, the last group of chiral ligands, namely, imine-phosphines 912, exhibited indirect catalytic activity to systems with and without a methylene linker. In these cases, cyclopropane 14 was formed in 79–85% yields, 68–70% of ee, and a 10:1 diastereomeric ratio (Table 1, entries 9–12).
Finally, there is one more interesting aspect to note. The change in the absolute configuration on the carbon atom of the aziridine moiety of the ligand leads to the cyclopropanation product 14 with opposite absolute configuration (Table 1, entries 1, 5 and 10). We observed a similar relationship while conducting research on asymmetric Michael [33], Mannich [34] and Friedel-Crafts [30] reactions. Thus, the application of enantiomeric catalysts provides access to both enantiomers of the cyclopropanation product 14.

2.4. Asymmetric Simmons-Smith Cyclopropanation Promoted by Aziridine-Phosphine 6

Considering the highest catalytic activity of ligand 6, we decided to study the scope of the asymmetric cyclopropanation in the presence of 6 by the application of another allylic alcohol under the same conditions (Scheme 3). All the results are shown in Table 2.
The data collected in Table 2 clearly indicate a very high catalytic activity of aziridine-phosphine 6 in the asymmetric Simmons-Smith cyclopropanation of (E)-allylic alcohols 1517 using diethylzinc and diiodomethane in the presence of diethylaluminum chloride leading to the appropriate cyclopropyl systems in high chemical yields (up to 90%) with a high degree of enantioselectivity (up to 90% ee) and diastereoselectivity (15:1 of dr). Lower enantio- and diastereoselectivity in the case of the reaction of (Z)-cinnamyl alcohol 18 (Table 2, entry 4) is in accordance with previous studies on the asymmetric cyclopropanation of this system [21].

2.5. Asymmetric Diethylzinc Addition to Benzaldehyde Catalyzed by Aziridine-Phosphines 18 and Imine-Phosphines 912

Based on our experience in the field of asymmetric transformations in the presence of zinc ions catalyzed by aziridine derivatives [35], in the next stage of research, we decided to check the catalytic activity of the chiral systems 112 in one of the most basic model asymmetric reactions involving zinc ions, namely, asymmetric diethylzinc addition to benzaldehyde (Scheme 4). The results of these tests are shown in Table 3.
The results summarized in Table 3 show that all chiral systems 112 are capable of catalyzing an asymmetric addition reaction of diethylzinc to benzaldehyde. The best results in terms of chemical yields and enantioselectivities were achieved using aziridine-phosphines 58 without a methylene linker (Table 3, entries 5–8) with the most active system 6 leading to the desired chiral alcohol 19a in 95% yield and with 96% ee. Ligands 912 (imines) exhibited lower activity (yield and ee around 80%) (Table 3, entries 9–12). In turn, aziridine-phosphines 14 containing a methylene linker showed only moderate efficiency as chiral ligands (Table 3, entries 1–4), which is in line with our previous observations.

2.6. Asymmetric Addition of Et2Zn to Aldehydes Promoted by Aziridine-Phosphine 6

Equipped with the most effective catalyst 6, we decided to expand the scope of its application using other aldehydes for this purpose (Scheme 5). The results are shown in Table 4.
As shown in Table 4, aziridine-phosphine 6 is prone to efficiently catalyzing an asymmetric addition of diethylzinc to aldehydes in most cases. The use of n-butyraldehyde (Table 4, entry 1), 2- and 4-methyl, methoxy-, bromo- and 4-trifluoromethylbenzaldehydes (Table 4, entries 2–8) led to the corresponding chiral alcohols in chemical yields with enantioselectivities up to 90%. Slightly worse results in terms of yield and ee were obtained for cinnamaldehyde (Table 4, entry 9) and thiophene-2-carbaldehyde (Table 4, entry 11). Disappointingly, no additional product was observed in the case of 4-nitrobenzaldehyde (Table 4, entry 10) and furan-2-carbaldehyde (Table 4, entry 12).

2.7. Asymmetric Cyclopropanation of Cinnamyl Alcohol and Diethylzinc Addition to Benzaldehyde Catalyzed by Aziridine-Phosphine Oxides

Finally, we decided to check whether the presence of the phosphine group is important for the course of both title asymmetric reactions. To this end, we carried out the cyclopropanation of cinnamyl alcohol and the addition of diethylzinc to benzaldehyde under the same conditions as before using two enantiomerically pure phosphinoyl-aziridines 20 and 21, which were prepared as previously described [33,34] (Scheme 6). The results are summarized in Table 5 and Table 6.
The inspection of Table 5 and Table 6 clearly indicates that the presence of the phosphine group in the structure of the chiral ligand has a very strong influence on the efficient course of both asymmetric transformations. The application of two chiral ligands bearing phosphinoyl (P=O) moiety instead of phosphine produced cyclopropyl derivative 14 and the chiral alcohol 19a only in moderate chemical yields and with moderate enantiomeric excess and diastereoselectivity.

3. Materials and Methods

3.1. Materials

Dichloromethane was dried over calcium hydride. Toluene and diethyl ether were distilled from sodium benzophenone ketyl radical. All the NMR spectra were accomplished using a Bruker(Bruker, Billerica, MA, USA) instrument at 600, 150 and 243 MHz, respectively, with CDCl3 as a solvent and TMS as an internal standard. Data are presented as singlet (s), doublet (d), triplet (t), quartet (q), multiplet (m) and broad singlet (br. s). Column chromatography was performed using Merck 60 silica gel. TLC was accomplished on Merck 60 F254 silica gel plates (Merck Group (Merck KGaA), Darmstadt, Germany). The plates were visualized using UV light (254 nm). The enantiomeric excess (ee) values were determined by HPLC with chiral support (Chiralcel OD-H or OD columns). The corresponding chiral phosphine-aziridines 18 were synthesized according to general procedures reported in our group [30].

3.2. Methods

3.2.1. Synthesis of Imine-Phosphines 912—General Procedure

The appropriate 1-(2-aminoalkyl)aziridines were prepared as previously described [31]. A solution of the 1-(2-aminoalkyl)aziridine (1 mmol) and 2-(diphenylphosphino)benzaldehyde (1 mmol) in methanol (10 mL) was refluxed for 16 h. After this time, the solvent was evaporated in vacuo, and the crude product was purified via flash chromatography (silica gel and hexane/ethyl acetate in gradient) to afford the appropriate chiral imine-phosphines as colorless viscous oils. Copies of NMR spectra of compounds 912 are included in the Supplementary Materials.
(E)-1-(2-(diphenylphosphaneyl)phenyl)-N-((S)-1-((S)-2-isopropylaziridin-1-yl)-3-methylbutan-2-yl)methanimine 9
Colorless oil, 97% yield; [α]D20 = +4.83 (c 0.5, CHCl3);
1H NMR (CDCl3, 600 MHz): δ = 0.58 (d, J = 6.9 Hz, 1H); 0.75–0.81 (m, 9H, 3xCH3); 0.82–0.86 (m, 3H, CH3); 0.87–0.89 (m, 1H); 0.95–1.01 (m, 2H); 1.04–1.08 (m, 2H); 1.12–1.19 (m, 1H); 1.44 (d, J = 3.4 Hz, 1H); 1.73 (dd, J1 = 7.0 Hz, J2 = 12.0 Hz, 1H); 1.89–1.95 (m, 1H); 2.84 (dd, J1 = 5.4 Hz, J2 = 11.9 Hz, 1H); 3.08 (m, 1H); 6.87–6.91 (m, 1H); 7.25–7.31 (m, 9H); 7.32–7.37 (m, 6H); 7.38–7.43 (m, 1H); 7.95–7.98 (m, 1H); 8.81 (d, J = 4.62 Hz, 1H) ppm;
13C NMR (150 MHz, CDCl3): δ = 17.70 (CH3), 19.13 (CH3), 20.00 (CH3), 20.69 (CH3), 30.98 (CH), 31.53 (CH2), 31.89 (CH), 46.63 (CH), 64.46 (CH2), 77.60 (CH), 128.37 (Carom.), 128.59 (Carom.), 128.72 (Carom.), 129.90 (Carom.), 133.38 (Carom.), 134.11 (Carom.), 137.03 (Carom.), 137.33 (Carom.), 139.57 (Carom.), 159.13 (d, J = 72.0 Hz, Cimine) ppm;
31P NMR (243 MHz, CDCl3): δ = −13.2 ppm;
Anal. Calcd. for C29H35N2P; C, 78.70; H, 7.97; N, 6.33; P, 7.00; Found: C, 78.72; H, 7.95; N, 6.32; P, 7.01.
(E)-1-(2-(diphenylphosphaneyl)phenyl)-N-((R)-1-((R)-2-isopropylaziridin-1-yl)-3-methylbutan-2-yl)methanamine 10
Colorless oil, 97% yield; [α]D20 = −4.80 (c 0.5, CHCl3);
1H NMR (CDCl3, 600 MHz): δ = 0.58 (d, J = 6.9 Hz, 1H); 0.75–0.81 (m, 9H, 3xCH3); 0.82–0.86 (m, 3H, CH3); 0.87–0.89 (m, 1H); 0.95–1.01 (m, 2H); 1.04–1.08 (m, 2H); 1.12–1.19 (m, 1H); 1.44 (d, J = 3.4 Hz, 1H); 1.73 (dd, J1 = 7.0 Hz, J2 = 12.0 Hz, 1H); 1.89–1.95 (m, 1H); 2.84 (dd, J1 = 5.4 Hz, J2 = 11.9 Hz, 1H); 3.08 (m, 1H); 6.87–6.91 (m, 1H); 7.25–7.31 (m, 9H); 7.32–7.37 (m, 6H); 7.38–7.43 (m, 1H); 7.95–7.98 (m, 1H); 8.81 (d, J = 4.62Hz, 1H) ppm;
13C NMR (150 MHz, CDCl3): δ = 17.70 (CH3), 19.13 (CH3), 20.00 (CH3), 20.69 (CH3), 30.98 (CH), 31.53 (CH2), 31.89 (CH), 46.63 (CH), 64.46 (CH2), 77.60 (CH), 128.37 (Carom.), 128.59 (Carom.), 128.72 (Carom.), 129.90 (Carom.), 133.38 (Carom.), 134.11 (Carom.), 137.03 (Carom.), 137.33 (Carom.), 139.57 (Carom.), 159.13 (d, J = 72.0 Hz, Cimine) ppm;
31P NMR (243 MHz, CDCl3): δ = -13.2 ppm;
Anal. Calcd. for C29H35N2P; C, 78.70; H, 7.97; N, 6.33; P, 7.00; Found: C, 78.72; H, 7.95; N, 6.32; P, 7.01.
(E)-1-(2-(diphenylphosphaneyl)phenyl)-N-((S)-1-((S)-2-isobutylaziridin-1-yl)-4-methylpentan-2-yl)methanimine 11
Colorless oil, 94% yield; [α]D20 = +4.76 (c 0.5, CHCl3);
1H NMR (CDCl3, 600 MHz): δ = 0.65 (d, J = 6.3 Hz, 3H, CH3); 0.69 (d, J = 6.8 Hz, 1H); 0.73 (d, J = 6.5 Hz, 3H, CH3); 0.79 (d, J = 9.2 Hz, 3H, CH3); 0.84 (d, J = 1.0 Hz, 3H, CH3); 0.96–1.0 (m, 4H); 1.21 (d, J = 6.0 Hz, 1H); 1.39–1.42 (m, 1H); 1.43 (d, J = 3.8 Hz, 1H); 1.81 (dd, J1 = 7.5 Hz, J2 = 11.6 Hz, 1H); 2.78 (dd, J1 = 4.8 Hz, J2 = 11.7 Hz, 1H); 3.43–3.48 (m, 1H); 6.86–6.89 (m, 1H); 7.22–7.27 (m, 3H); 7.29–7.31 (m, 1H); 7.32–7.37 (m, 8H); 7.38–7.43 (m, 3H); 7.92–8.0 (m, 1H); 8.92 (d, J = 4.86 Hz, 1H) ppm;
13C NMR (150 MHz, CDCl3): δ = 21.06 (CH3), 21.32 (CH), 22.53 (CH3), 22.86 (CH3), 23.57 (CH3), 27.00 (CH), 33.84 (CH2), 39.09 (CH), 42.08 (CH2), 42.81 (CH), 67.01 (CH2), 70.02 (CH2), 127.99 (Carom.), 128.65 (Carom.), 129.11 (Carom.), 130.07 (Carom.), 130.70 (Carom.), 133.22 (Carom.), 134.07 (Carom.), 136.13 (Carom.), 136.69 (Carom.), 137.38 (Carom.), 139.48 (Carom.), 158.91 (d, J = 78.0 Hz, Cimine) ppm;
31P NMR (243 MHz, CDCl3): δ = −13.41 ppm;
Anal. Calcd. for C31H39N2P C, 79.11; H, 8.35; N, 5.95; P, 6.58; Found: C, 79.21; H, 8.30; N, 5.90; P, 6.58.
(E)-N-((S)-1-((S)-2-benzylaziridin-1-yl)-3-phenylpropan-2-yl)-1-(2-(diphenylphosphaneyl)phenyl)methanimine 12
Colorless oil, 98% yield; [α]D20 = +3.90 (c 0.5, CHCl3);
1H NMR (CDCl3, 600 MHz): δ = 1.23 (d, J = 5.8 Hz, 1H); 1.50 (d, J = 3.2 Hz, 1H); 1.57–1.59 (m, 1H); 1.82 (dd, J1 = 7.8 Hz, J2 = 11.5 Hz, 1H); 2.21–2.27 (m, 1H); 2.58–2.64 (m, 1H); 2.70 (dd, J1 = 4.2 Hz, J2 = 13.4 Hz, 2H); 2.86–2.91 (m, 1H); 3.52–3.56 (m, 1H); 6.99–7.01 (m, 3H); 7.21–7.24 (m, 4H); 7.28–7.35 (m, 17H); 8.62 (d, J = 4.4 Hz, 1H) ppm;
13C NMR (150 MHz, CDCl3): δ = 33.33 (CH2), 39.40 (CH2), 40.83 (CH2), 42.08 (CH2), 41.23 (CH), 65.50 (CH2), 73.80 (CH), 125.95 (Carom.), 126.13 (Carom.), 126.43 (Carom.), 128.06 (Carom.), 128.29 (Carom.), 128.60 (Carom.), 128.77 (Carom.), 129.14 (Carom.), 129.53 (Carom.), 130.01 (Carom.), 133.75 (Carom.), 133.93 (Carom.), 134.04 (Carom.), 134.16 (Carom.), 137.14 (Carom.), 137.54 (Carom.), 137.68 (Carom.), 138.85 (Carom.), 139.37 (Carom.), 159.52 (d, J = 69.60 Hz, Cimine) ppm;
31P NMR (CDCl3, 243 MHz): δ = −13.10 ppm;
Anal. Calcd. for C37H35N2P C, 82.50; H, 6.55; N, 5.20; P, 5.75; Found: C, 82.42; H, 6.58; N, 5.26; P, 5.74.

3.2.2. Asymmetric Simmons–Smith Cyclopropanation—General Procedure

The corresponding allylic alcohols (substrates) were prepared using the literature protocols [40,41]. The solution of chiral ligand in anhydrous CH2Cl2 (5 mL) was cooled to 0 °C and alcohol (0.5 mmol), diethylzinc (1.0 M in hexanes, 1 mmol) and CH2I2 (0.6 mmol) were added. The resulting mixture was stirred for 4 h at room temperature. The reaction was quenched by adding a 2 M aqueous solution of NaOH, the phases were separated and the aqueous layer was extracted with dichloromethane (3 × 10 mL). Combined layers were dried with MgSO4 and evaporated in vacuo. The pure product was obtained by purification via column chromatography on silica gel (hexane:ethyl acetate 2:1). The spectroscopic data of products were consistent with the literature data [21]. Selected HPLC chromatograms of the cyclopropanation products are included in the Supplementary Materials.

3.2.3. Asymmetric Addition of Diethylzinc to Aldehydes—General Procedure

Diethylzinc (1.0 M in hexanes, 3 mmol) was added to the solution of the chiral ligand (0.1 mmol) in toluene (5 mL) at 0 °C. The mixture was stirred for 30 min at this temperature, and an aldehyde (1 mmol) was added. The resulting solution was stirred for 2 h at 0 °C, then overnight at room temperature. The reaction was quenched by adding 5% aqueous solution of HCl. The layers were separated, and the aqueous phase was extracted with diethyl ether (3 × 10 mL). The combined organic layers were washed with brine, dried over MgSO4 and evaporated in vacuo. The crude product was purified by column chromatography on silica gel (hexane:ethyl acetate 5:1) [42]. The NMR spectra of the products were consistent with the literature data [37,38,39]. Selected HPLC chromatograms of the addition products are included in the Supplementary Materials.

4. Conclusions

The chiral aziridine-phosphines exhibited a very high catalytic activity in the asymmetric Simmons-Smith cyclopropanation of allylic alcohols and in the asymmetric addition of diethylzinc to aldehydes. The corresponding cyclopropanation and addition products were achieved in satisfactory chemical yields and with a high level of enantioselectivity and diastereoselectivity. The absolute configuration of a carbon atom of aziridine has a crucial influence on the stereochemical outcome of both title reactions. Moreover, the presence of the phosphine moiety in the structure of the chiral ligand has a very strong influence on the efficient course of both asymmetric transformations.
By carrying out these asymmetric reactions, we showed that aziridine-phosphines are versatile catalysts; thus, it is certainly necessary to continue research on their effectiveness in asymmetric synthesis.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11080968/s1: copies of NMR spectra of compounds 912 and selected HPLC chromatograms for Simmons-Smith cyclopropanation and diethylzinc addition products.

Author Contributions

Conceptualization and methodology, S.L. and M.R.; software, A.M.P., A.Z. and M.R.; investigation, A.B.-S., A.Z., J.A., L.M., A.M.P. and M.R.; writing—original draft preparation, M.R.; writing—review and editing, S.L. and M.R.; supervision, M.R. and S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Centre (NCN) (grant no. 2016/21/B/ST5/00421 for M.R.). The scientific grant InterChemMed No. POWR.03.02.00-00-00-I029/16 to A.B. is also acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Carroll, M.P.; Guiry, P.J. P,N ligands in asymmetric catalysis. Chem. Soc. Rev. 2014, 43, 819–833. [Google Scholar] [CrossRef]
  2. Li, Y.-M.; Kwong, F.-Y.; Yu, W.-Y.; Chan, A.S.C. Recent advances in developing new axially chiral phosphine ligands for asymmetric catalysis. Coord. Chem. Rev. 2007, 251, 2119–2144. [Google Scholar] [CrossRef]
  3. Gaunt, M.J.; Johansson, C.C.C.; McNally, A.; Vo, N.T. Enantioselective organocatalysis. Drug Discov. Today 2007, 12, 8–27. [Google Scholar] [CrossRef] [PubMed]
  4. Han, B.; He, X.-H.; Liu, Y.-Q.; He, G.; Peng, C.; Li, J.-L. Asymmetric organocatalysis: An enabling technology for medicinal chemistry. Chem. Soc. Rev. 2021, 50, 1522–1586. [Google Scholar] [CrossRef] [PubMed]
  5. Diaz-Muñoz, G.; Miranda, I.L.; Sartori, S.K.; De Rezende, D.C.; Nogueira Diaz, M.A. Use of chiral auxiliaries in the asymmetric synthesis of biologically active compounds: A review. Chirality 2019, 31, 776–812. [Google Scholar] [CrossRef] [PubMed]
  6. Dixon, D.J. Bifunctional catalysis. Beilstein J. Org. Chem. 2016, 12, 1079–1080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Yuan, Y.-H.; Han, X.; Zhu, F.-P.; Tian, J.-M.; Zhang, F.-M.; Zhang, X.-M.; Tu, Y.-Q.; Wang, S.-H.; Guo, X. Development of bifunctional organocatalysts and application to asymmetric total synthesis of naucleofficine I and II. Nat. Commun. 2019, 10, 3394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Peters, R. (Ed.) Cooperative Catalysis: Designing Efficient Catalysts for Synthesis; Wiley-VCH Verlag GmbH & Co. KgaA: Weinheim, Germany, 2015; pp. 1–456. [Google Scholar]
  9. Werth, J.; Uyeda, C. Regioselective Simmons-Smith-type cyclopropanations of polyalkenes enabled by transition metal catalysis. Chem. Sci. 2018, 9, 1604–1609. [Google Scholar] [CrossRef] [Green Version]
  10. Časar, Z. Synthetic Approaches to Contemporary Drugs that Contain the Cyclopropyl Moiety. Synthesis 2020, 52, 1315–1345. [Google Scholar] [CrossRef]
  11. Ezawa, T.; Kawashima, Y.; Noguchi, T.; Jung, S.; Imai, N. Convenient synthesis of memantine analogues containing a chiral cyclopropane skeleton as a sigma-1 receptor agonist. Tetrahedron Asymmetry 2017, 28, 266–281. [Google Scholar] [CrossRef]
  12. Sun, M.; Li, W.-D.Z.; Qiu, F.G. Asymmetric Synthesis of the DEFG Rings of Solanoeclepin A. Org. Lett. 2019, 21, 644–647. [Google Scholar] [CrossRef]
  13. Talele, T.T. The “Cyclopropyl Fragment” is a Versatile Player that Frequently Appears in Preclinical/Clinical Drug Molecules. J. Med. Chem. 2016, 59, 8712–8756. [Google Scholar] [CrossRef] [PubMed]
  14. Chawner, S.J.; Cases-Thomas, M.J.; Bull, J.A. Divergent Synthesis of Cyclopropane-Containing Lead-Like Compounds, Fragments and Building Blocks through a Cobalt Catalyzed Cyclopropanation of Phenyl Vinyl Sulfide. Eur. J. Org. Chem. 2017, 5015–5024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Novakov, I.A.; Babushkin, A.S.; Yablokov, A.S.; Nawrozkij, M.B.; Vostrikova, O.V.; Shejkin, D.S.; Mkrtchyan, A.S.; Balakin, K.V. Synthesis and structure—activity relationships of cyclopropane-containing analogs of pharmacologically active compounds. Russ. Chem. Bull. Int. Ed. 2018, 67, 395–418. [Google Scholar] [CrossRef]
  16. Wu, W.; Lin, Z.; Jiang, H. Recent advances in the synthesis of cyclopropanes. Org. Biomol. Chem. 2018, 16, 7315–7329. [Google Scholar] [CrossRef] [PubMed]
  17. Berkessel, A.; Ertürk, E.; Neudörfl, J.M. Asymmetric cyclopropanation of olefins catalyzed by a chiral cobalt(II) porphyrin. Org. Commun. 2017, 10, 79–89. [Google Scholar] [CrossRef]
  18. Meazza, M.; Ashe, M.; Shin, H.Y.; Yang, H.S.; Mazzanti, A.; Yang, J.W.; Rios, R. Enantioselective Organocatalytic Cyclopropanation of Enals Using Benzyl Chlorides. J. Org. Chem. 2016, 81, 3488–3500. [Google Scholar] [CrossRef] [Green Version]
  19. Dočekal, V.; Petrželová, S.; Císařová, I.; Veselý, J. Enantioselective Cyclopropanation of 4-Nitroisoxazole Derivatives. Adv. Synth. Catal. 2020, 362, 2597–2603. [Google Scholar] [CrossRef]
  20. Herchl, R.; Waser, M. Asymmetric cyclopropanation of chalcones using chiral phase-transfer catalysts. Tetrahedron Lett. 2013, 54, 2472–2475. [Google Scholar] [CrossRef] [Green Version]
  21. Rachwalski, M.; Kaczmarczyk, S.; Leśniak, S.; Kiełbasiński, P. Highly Efficient Asymmetric Simmons-Smith Cyclopropanation Promoted by Chiral Heteroorganic Aziridinyl Ligands. ChemCatChem 2014, 6, 873–875. [Google Scholar] [CrossRef]
  22. Pu, L.; Yu, H.-B. Catalytic Asymmetric Organozinc Additions to Carbonyl Compounds. Chem. Rev. 2001, 101, 757–824. [Google Scholar] [CrossRef]
  23. Rachwalski, M.; Jarzyński, S.; Leśniak, S. Aziridine ring-containing chiral ligands as highly efficient catalysts in asymmetric synthesis. Tetrahedron Asymmetry 2013, 24, 421–425. [Google Scholar] [CrossRef]
  24. Pieczonka, A.M.; Leśniak, S.; Jarzyński, S.; Rachwalski, M. Aziridinylethers as highly enantioselective ligands for the asymmetric addition of organozinc species to carbonyl compounds. Tetrahedron Asymmetry 2015, 26, 148–151. [Google Scholar] [CrossRef]
  25. Leśniak, S.; Rachwalski, M.; Sznajder, E.; Kiełbasiński, P. New highly efficient aziridine-functionalized tridentate sulfinyl catalysts for enantioselective diethylzinc addition to carbonyl compounds. Tetrahedron Asymmetry 2009, 20, 2311–2314. [Google Scholar] [CrossRef]
  26. Rachwalski, M.; Jarzyński, S.; Jasiński, M.; Leśniak, S. Mandelic acid derived a-aziridinyl alcohols as highly efficient ligands for asymmetric additions of zinc organyls to aldehydes. Tetrahedron Asymmetry 2013, 24, 689–693. [Google Scholar] [CrossRef]
  27. Leśniak, S.; Rachwalski, M.; Jarzyński, S.; Obijalska, E. Lactic acid derived aziridinyl alcohols as highly effective catalysts for asymmetric additions of an organozinc species to aldehydes. Tetrahedron Asymmetry 2013, 24, 1336–1340. [Google Scholar] [CrossRef]
  28. Rachwalski, M. Limonene oxide derived aziridinyl alcohols as highly efficient catalysts for asymmetric additions of organozinc species to aldehydes. Tetrahedron Asymmetry 2014, 25, 219–223. [Google Scholar] [CrossRef]
  29. Doğan, Ö.; Çağli, E. PFAM catalyzed enantioselective diethylzinc addition to imines. Turk. J. Chem. 2015, 39, 290–296. [Google Scholar] [CrossRef]
  30. Buchcic, A.; Zawisza, A.; Leśniak, S.; Rachwalski, M. Asymmetric Friedel-Crafts Alkylation of Indoles Catalyzed by Chiral Aziridine-Phosphines. Catalysts 2020, 10, 971. [Google Scholar] [CrossRef]
  31. Pieczonka, A.M.; Misztal, E.; Rachwalski, M.; Leśniak, S. Synthesis of chiral 1-(2-aminoalkyl)aziridines via a self-opening reaction of aziridine. Arkivoc 2017, 2017, 223–234. [Google Scholar] [CrossRef] [Green Version]
  32. Shitama, H.; Katsuki, T. Asymmetric Simmons-Smith Reaction of Allylic Alcohols with Al Lewis Acid/N Lewis Base Bifunctional Al(Salalen) Catalyst. Angew. Chem. Int. Ed. 2008, 47, 2450–2453. [Google Scholar] [CrossRef]
  33. Wujkowska, Z.; Zawisza, A.; Leśniak, S.; Rachwalski, M. Phosphinoyl-aziridines as a new class of chiral catalysts for enantioselective Michael addition. Tetrahedron 2019, 75, 230–235. [Google Scholar] [CrossRef]
  34. Buchcic, A.; Zawisza, A.; Leśniak, S.; Adamczyk, J.; Pieczonka, A.M.; Rachwalski, M. Enantioselective Mannich reaction promoted by chiral phosphinoyl-aziridines. Catalysts 2019, 9, 837. [Google Scholar] [CrossRef] [Green Version]
  35. Leśniak, S.; Rachwalski, M.; Pieczonka, A.M. Optically Pure Aziridinyl Ligands as Useful Catalysts in the Stereocontrolled Synthesis. Curr. Org. Chem. 2014, 18, 3045–3065. [Google Scholar] [CrossRef]
  36. Zhang, C.-H.; Yan, S.-J.; Pan, S.-Q.; Huang, R.; Lin, J. Highly Enantioselective Addition of Diethylzinc to aldehydes Catalyzed by Novel Chiral tert-Amino Alcohols. Bull. Korean Chem. Soc. 2010, 31, 869–873. [Google Scholar] [CrossRef] [Green Version]
  37. Cere, V.; Peri, F.; Pollicino, S. Regio- and Stereospecific Me3SiI-Promoted Intramolecular Displacement of Hydroxy Group by Sulfide. 2. Polyhydroxylated Thiepane Ring Contraction to Thiolane or Thiane Derivatives. Synthesis of Enantiopure Polyalcohols. J. Org. Chem. 1997, 62, 8572–8574. [Google Scholar] [CrossRef] [PubMed]
  38. Dilek, Ö.; Tezeren, M.A.; Tilki, T.; Ertürk, E. Chiral 2-(hydroxyaryl)alcohols (HAROLs) with a 1,4-diol scaffold as a new family of ligands and organocatalysts. Tetrahedron 2018, 74, 268–286. [Google Scholar] [CrossRef]
  39. Lombardo, M.; Chiarucci, M.; Trombini, C. The first enantioselective addition of diethylzinc to aldehydes in ionic liquids catalysed by a recyclable ion-tagged diphenylprolinol. Chem. Eur. J. 2008, 14, 11288–11291. [Google Scholar] [CrossRef] [PubMed]
  40. Dieckmann, M.; Jang, Y.-S.; Cramer, N. Chiral cyclopentadienyl iridium(III) complexes promote enantioselective cycloisomerizations giving fused cyclopropanes. Angew. Chem. Int. Ed. 2015, 54, 12149–12152. [Google Scholar] [CrossRef] [PubMed]
  41. Ando, K. Highly selective synthesis of (Z)-unsaturated esters by using new Horner-Emmons reagents, ethyl (diarylphosphono)acetates. J. Org. Chem. 1997, 62, 1934–1939. [Google Scholar] [CrossRef] [PubMed]
  42. Rachwalski, M.; Kwiatkowska, M.; Drabowicz, J.; Kłos, M.; Wieczorek, W.M.; Szyrej, M.; Sieroń, L.; Kiełbasiński, P. Enzyme-promoted desymmetrization of bis(2-hydroxymethylphenyl)sulfoxide as a route to tridentate chiral catalysts. Tetrahedron Asymmetry 2008, 19, 2096–2101. [Google Scholar] [CrossRef]
Figure 1. The structures of chiral aziridine-phosphines 18.
Figure 1. The structures of chiral aziridine-phosphines 18.
Catalysts 11 00968 g001
Scheme 1. Synthesis of chiral imine-phosphines 912.
Scheme 1. Synthesis of chiral imine-phosphines 912.
Catalysts 11 00968 sch001
Scheme 2. Asymmetric cyclopropanation of alcohol 13 in the presence of catalysts 112 leading to cyclopropane derivative 14.
Scheme 2. Asymmetric cyclopropanation of alcohol 13 in the presence of catalysts 112 leading to cyclopropane derivative 14.
Catalysts 11 00968 sch002
Scheme 3. Asymmetric cyclopropanation reaction of alcohols 1518 catalyzed by aziridine-phosphine 6 (*—stereogenic center).
Scheme 3. Asymmetric cyclopropanation reaction of alcohols 1518 catalyzed by aziridine-phosphine 6 (*—stereogenic center).
Catalysts 11 00968 sch003
Scheme 4. Asymmetric diethylzinc addition to benzaldehyde promoted by ligands 112.
Scheme 4. Asymmetric diethylzinc addition to benzaldehyde promoted by ligands 112.
Catalysts 11 00968 sch004
Scheme 5. Asymmetric diethylzinc addition to aldehydes catalyzed by ligand 6.
Scheme 5. Asymmetric diethylzinc addition to aldehydes catalyzed by ligand 6.
Catalysts 11 00968 sch005
Scheme 6. Cyclopropanation of 13 and addition of Et2Zn to benzaldehyde in the presence of phosphinoyl-aziridines.
Scheme 6. Cyclopropanation of 13 and addition of Et2Zn to benzaldehyde in the presence of phosphinoyl-aziridines.
Catalysts 11 00968 sch006
Table 1. Asymmetric Simmons-Smith cyclopropanation of cinnamyl alcohol catalyzed by aziridines 112.
Table 1. Asymmetric Simmons-Smith cyclopropanation of cinnamyl alcohol catalyzed by aziridines 112.
EntryLigandProduct 14
Yield (%)ee (%) aDrbAbs. Conf. c
1170625:11R,2R
2272655:11S,2S
3363545:11S,2S
4465585:11S,2S
55909212:11R,2R
66939815:11S,2S
77908210:11S,2S
88918510:11S,2S
99857210:11S,2S
1010837010:11R,2R
1111796810:11S,2S
1212807010:11S,2S
a Determined by chiral HPLC using Chiralcel OD-H column (for major product). b Determined by 1H NMR of the crude mixture. c According to the literature data [21,32]. Conditions: 10 mol% of the ligand, cinnamyl alcohol (0.5 mmol), diethylzinc (1.0 M in hexanes, 1 mmol), diiodomethane (0.6 mmol), DCM (5 mL), 0 °C, 4 h.
Table 2. Asymmetric cyclopropanation promoted by aziridine-phosphine 6.
Table 2. Asymmetric cyclopropanation promoted by aziridine-phosphine 6.
EntrySubstrateYield (%)ee (%) aDrbAbs. Conf. c
115929515:1(1S,2S)
216949315:1(1S,2S)
317919115:1(1S,2S)
41856404:1(1R,2S)
a Determined by chiral HPLC using Chiralcel OD-H column (for major product). b Determined by 1H NMR of the crude mixture. c According to the literature data [21,32]. Conditions: 10 mol% of the catalyst 6, allylic alcohol (0.5 mmol), diethylzinc (1.0 M in hexanes, 1 mmol), diiodomethane (0.6 mmol), DCM (5 mL), 0 °C, 4 h.
Table 3. Asymmetric addition of Et2Zn to benzaldehyde in the presence of 112.
Table 3. Asymmetric addition of Et2Zn to benzaldehyde in the presence of 112.
EntryLigandProduct 19a
Yield (%)ee (%) aAbs. Conf. b
117566(R)
227868(S)
337162(S)
447263(S)
559296(R)
669596(S)
779186(S)
889085(S)
998580(S)
10108380(R)
11118173(S)
12128276(S)
a Determined by chiral HPLC using Chiralcel OD column. b According to the literature data [25,36]. Conditions: 10 mol% of the ligand, benzaldehyde (1 mmol), diethylzinc (1.0 M in hexanes, 3 mmol), toluene (5 mL), 0 °C, 2 h.
Table 4. Asymmetric addition of Et2Zn to aldehydes in the presence of 6.
Table 4. Asymmetric addition of Et2Zn to aldehydes in the presence of 6.
EntryRProducts 19b–m
Yield (%)ee (%) aAbs. Conf. b
1n-Pr9089(S)
22-MeC6H49190(S)
34-MeC6H49093(S)
42-MeOC6H49393(S)
54-MeOC6H49395(S)
62-BrC6H49190(S)
74-BrC6H49392(S)
84-CF3C6H49193(S)
9Vi8785(S)
104-O2NC6H40Nd cNd c
11C4H3S7068(S)
12C4H3O0Nd cNd c
a Determined by chiral HPLC using Chiralcel OD column. b According to the literature data [37,38,39]. c Not determined. Conditions: 10 mol% of the ligand, aldehyde (1 mmol), diethylzinc (1.0 M in hexanes, 3 mmol), toluene (5 mL), 0 °C, 2 h.
Table 5. Cyclopropanation of cinnamyl alcohol catalyzed by phosphine oxides 20 and 21.
Table 5. Cyclopropanation of cinnamyl alcohol catalyzed by phosphine oxides 20 and 21.
EntryLigandProduct 14
Yield (%)ee (%) aDrbAbs. Conf. c
12050415:1(1S,2S)
22155525:1(1S,2S)
a Determined by chiral HPLC using Chiralcel OD-H column (for major product). b Determined by 1H NMR of the crude mixture. c According to the literature data [21,32]. Conditions: 10 mol% of the ligand, cinnamyl alcohol (0.5 mmol), diethylzinc (1.0 M in hexanes, 1 mmol), diiodomethane (0.6 mmol), DCM (5 mL), 0 °C, 4 h.
Table 6. Diethylzinc addition to benzaldehyde catalyzed by phosphine oxides 20 and 21.
Table 6. Diethylzinc addition to benzaldehyde catalyzed by phosphine oxides 20 and 21.
EntryLigandProduct 19a
Yield (%)ee (%) aAbs. Conf. b
1206056(S)
2216862(S)
a Determined using chiral HPLC using Chiralcel OD column. b According to the literature data [25,26,27,28,29,30,31,32,33,34,35,36]. Conditions: 10 mol% of the ligand, benzaldehyde (1 mmol), diethylzinc (1.0 M in hexanes, 3 mmol), toluene (5 mL), 0 °C, 2 h.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Buchcic-Szychowska, A.; Adamczyk, J.; Marciniak, L.; Pieczonka, A.M.; Zawisza, A.; Leśniak, S.; Rachwalski, M. Efficient Asymmetric Simmons-Smith Cyclopropanation and Diethylzinc Addition to Aldehydes Promoted by Enantiomeric Aziridine-Phosphines. Catalysts 2021, 11, 968. https://doi.org/10.3390/catal11080968

AMA Style

Buchcic-Szychowska A, Adamczyk J, Marciniak L, Pieczonka AM, Zawisza A, Leśniak S, Rachwalski M. Efficient Asymmetric Simmons-Smith Cyclopropanation and Diethylzinc Addition to Aldehydes Promoted by Enantiomeric Aziridine-Phosphines. Catalysts. 2021; 11(8):968. https://doi.org/10.3390/catal11080968

Chicago/Turabian Style

Buchcic-Szychowska, Aleksandra, Justyna Adamczyk, Lena Marciniak, Adam Marek Pieczonka, Anna Zawisza, Stanisław Leśniak, and Michał Rachwalski. 2021. "Efficient Asymmetric Simmons-Smith Cyclopropanation and Diethylzinc Addition to Aldehydes Promoted by Enantiomeric Aziridine-Phosphines" Catalysts 11, no. 8: 968. https://doi.org/10.3390/catal11080968

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop