Next Article in Journal
Platinum Group Metal-Free Catalysts for Oxygen Reduction Reaction: Applications in Microbial Fuel Cells
Next Article in Special Issue
Recent Approaches to Chiral 1,4-Dihydropyridines and their Fused Analogues
Previous Article in Journal
Electrochemical Reactors for CO2 Conversion
Previous Article in Special Issue
Selective Hydration of Nitriles to Corresponding Amides in Air with Rh(I)-N-Heterocyclic Complex Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Investigation of Pinane-Based Chiral Tridentate Ligands in the Asymmetric Addition of Diethylzinc to Aldehydes

1
Institute of Pharmaceutical Chemistry, University of Szeged, Interdisciplinary excellent center, H-6720 Szeged, Eötvös utca 6, Hungary
2
MTA-SZTE Stereochemistry Research Group, Hungarian Academy of Sciences, H-6720 Szeged, Eötvös utca 6, Hungary
3
Interdisciplinary Centre of Natural Products, University of Szeged, H-6720 Szeged, Eötvös utca 6, Hungary
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(5), 474; https://doi.org/10.3390/catal10050474
Submission received: 7 April 2020 / Revised: 22 April 2020 / Accepted: 24 April 2020 / Published: 26 April 2020
(This article belongs to the Special Issue Catalysis in Heterocyclic and Organometallic Synthesis)

Abstract

:
A library of pinane-based chiral aminodiols, derived from natural (−)-β-pinene, were prepared and applied as chiral catalysts in the addition of diethylzinc to aldehydes. (−)-β-Pinene was reacted to provide 3-methylenenopinone, followed by a reduction of the carbonyl function to give a key allylic alcohol intermediate. Stereoselective epoxidation of the latter and subsequent ring opening of the resulting oxirane with primary and secondary amines afforded aminodiols. The regioselectivity of the ring closure of the N-substituted secondary aminodiols with formaldehyde was examined and exclusive formation of oxazolidines was observed. Treatment of the allylic alcohol with benzyl bromide provided the corresponding O-benzyl derivative, which was transformed into O-benzyl aminodiols by aminolysis. Ring closure of the N-isopropyl aminodiol derivative with formaldehyde resulted in spirooxazolidine. The obtained potential catalysts were applied in the reaction of both aromatic and aliphatic aldehydes to diethylzinc providing moderate to good enantioselectivities (up to 87% ee). Through the use of molecular modeling at an ab initio level, this phenomenon was interpreted in terms of competing reaction pathways. Molecular modeling at the RHF/LANL2DZ level of theory was successfully applied for interpretation of the stereochemical outcome of the reactions leading to display excellent (R) enantioselectivity in the examined transformation.

Graphical Abstract

1. Introduction

Chiral synthons, applied successfully in asymmetric homogenous and heterogeneous catalysis, have achieved increasing importance in organic chemistry in recent years [1,2,3]. The enantioselective addition of dialkylzinc to aldehydes catalyzed by different types of chiral ligands has been investigated intensively [4,5,6], because the preparation of enantiomerically pure or enriched alcohols is of considerable interest for the synthesis of bioactive compounds [7,8,9] and natural products [10,11].
Aromatic and aliphatic aminodiols bearing a 1,2- or a 1,3-aminoalcohol moiety have proven to be highly efficient building blocks [12,13,14,15]. They have been applied as starting materials in the stereoselective synthesis of compounds of pharmacological interest, including 1,3-oxazines [16], 1,3-thiazines [17,18,19] and 2-iminothiazolidines [20]. In addition to their synthetic importance, aminodiols can also be applied as chiral ligands and auxiliaries in enantioselective transformations [21,22,23,24] including intramolecular radical cyclizations [25], intramolecular [2+2] photocycloaddition [26] and Grignard addition [27,28]. Therefore, it is not surprising that the preparation of new chiral aminodiols has been a topic of increased interest. Although numerous enantiopure chelating ligands have been prepared [29,30,31,32,33,34], there is still a need for new types obtainable by concise syntheses from inexpensive starting materials.
Naturally occurring chiral monoterpenes such as (+)- and (–)-α-pinene [21,35,36], (+)-carene [37,38], (+)-camphor [31,39], (–)-menthone [34], (–)-fenchone [29], (+)-sabinol [40], (–)-myternol [21,41], (–)-pulegone [42], neoisopulegol [43] and (S)-perillyl alcohol [44] have been widely used as key intermediates for the synthesis of chiral catalysts and auxiliaries for asymmetric synthesis.
Besides their chemical interests, many compounds containing 1-amino-2,3-diol functionalities display remarkable pharmacological activities. For example, Abbott-aminodiol bears the core of the drug Zankiren® exhibiting renin-inhibitory activity [45,46]. Aminodiols have also been reported to have expressed biological properties such as a gastro-protective effect [47] or HIV protease [48] inhibition and antiviral activity [49,50,51].
In the present paper we report the diastereoselective synthesis of new aminodiol derivatives as potential chiral ligands in the asymmetric addition of Et2Zn to aldehydes starting from commercially available natural (−)-β-pinene. These compounds, the regioisomers of pinane-based 3-amino-1,2-diols, were prepared from α-pinene [36]. In addition, we planned to develop a molecular model through which the interpretation of the catalytic pathway of the reaction and the catalytic activities of the chiral aminodiol derivatives should be possible.

2. Results

2.1. Synthesis of Allylic Alcohol 4a

Key intermediate 3-methylenenopinone 3 was prepared from commercially available (−)-β-pinene 1 by oxidation with the NaIO4/RuCl3 system to afford (−)-nopinone 2. The resulting (−)-nopinone was converted into (−)-3-methylenenopinone 3 by applying formaldehyde in alkaline condition according to literature methods [52,53] (Scheme 1).
Reduction of 3 with NaBH4 in various solvents gave a mixture of 4a and 4b (Scheme 2, Table 1). It is important to note that whereas allylic alcohol 4a was formed in a highly stereoselective manner, 4b exists as a 4:1 mixture of two cis-diastereomers (diexo and diendo, based on 1H-NMR measurement and comparison with data in the literature) [54,55].
When Et2O was applied as solvent, 4a was formed as the main product (4a:4b = 3:1), whereas the ratio of the two products in EtOH changed to 4a:4b = 1:3. In contrast, when MeOH was used as a solvent, the two products formed in a 1:1 ratio. In addition, it is interesting to note that the ratio of 4a and 4b also depended on the temperature. At –20 °C in MeOH, compound 4a was obtained as the major product (Table 1), although 4a and 4b could not be separated by conventional technics. Applying the condition of Luche reaction, in the presence of CeCl3 as additive, 4a was obtained as the single product. This procedure not only allowed highly regioselective reduction, but also enhanced reaction rate. The probable reason is the effect of cerium, a hard Lewis acid. Despites its weak acidity, it certainly contributes to both the regioselectivity and the high reaction rate though the coordination to the oxygen of the carbonyl function [56].

2.2. Synthesis of (–)-β-Pinene-Based Aminodiols

Epoxidation of 4a with t-BuOOH in the presence of VO(acac)2 as catalyst furnished epoxide 5 as a single product in a stereoselective reaction [57,58]. Since purification of epoxide 5 could not be effectively performed without its decomposition, the crude product with a purity of approximately 92% (based on 1H NMR measurement) was treated with various amines to perform the aminolysis of the oxirane ring. Our previous results clearly demonstrated that when aminodiols were applied as catalysts, their N-substituents definitely influenced the enantioselectivity of their catalyzed reaction [22,37,59,60]. Consequently, aminodiol library 6–13 was prepared by aminolysis of 5 with secondary and primary amines in the presence of lithium perchlorate as catalyst (Scheme 3).
Primary aminodiol 14 was obtained in moderate yield by debenzylation of the corresponding N-benzyl aminodiol 6 under standard condition by hydrogenation over Pd/C (Scheme 4).
The ring closure reaction of 6 and 9 aminodiols with formaldehyde was also investigated to study the regioselectivity of the reaction [38,42,60]. When these aminodiols were reacted with formaldehyde under mild conditions, spirooxazolidine 15 and 16 were obtained in highly regioselective ring closure, with similar regioselectivity as observed in the case of pinane-based regioisomers [36]. This regioselectivity, however, is opposite to those of the carene-based analogues reported recently (Scheme 4) [37,38].
On the other hand, to assess the importance of the secondary hydroxyl group in the catalytic application of our aminodiols, allylic alcohol 4a was transformed into O-benzyl derivative 17. The separation of 16 and benzyl bromide was unsuccessful using classical chromatography methods. Therefore, 17 was transformed with mCPBA to epoxide 18 and the latter could be easily purified (in contrast to its epoxyalcohol analogue 5) on a gram scale by simple column chromatography in good yield [61,62,63,64]. The aminolysis of the formed oxirane ring of 18 with different amines afforded O-benzyl aminodiols 19–22 (Scheme 5) [22,59].
The relative configurations of both compounds 6–13 and 19–22 were assigned by means of NOESY experiments: reliable NOE signals were observed between the H-10 and H-2, H-6 as well as H-2 and H-6 protons (Figure 1).
The regioselectivity of the ring closure of 22 with formaldehyde resulting in spirooxazolidine 23 (Scheme 6) was also investigated [38,42,60].

2.3. Application of Aminodiols as Chiral Ligands for Catalytic Addition of Diethylzinc to Aldehydes

Applying aminodiols 6–15 and 19–23 as chiral catalysts in the addition of diethylzinc to benzaldehyde (24), enantiomeric mixture of (S)- and (R)-1-phenyl-1-propanol 25 was obtained (Scheme 7).
The results are presented in Table 2. The enantiomeric excess of 1-phenyl-1-propanols (S)-25 and/or (R)-25 was determined by chiral GC (CHIRASIL-DEX CB column) according to literature methods [65,66]. Low to good enantioselectivities were observed. The results found clearly show that all aminodiols favored the formation of the (R)-enantiomer of 25. In contrast, the application of 20 led to (S)-enantiomer 25 as the main product. Aminodiol 10 and 21 afforded the best ee value (ee = 80%) with an (R)-selectivity, whereas O-benzyl aminodiol 20 showed the best ee value (ee = 74%) with an (S)-selectivity. Moreover, enantioselectivities were also observed in the addition of diethylzinc to benzaldehyde catalyzed by aminodiols 6–8, whereas lower, but still good selectivities were obtained with the use of O-benzyl aminodiol derivatives 19–22. We suppose that the highly rigid structure of O-benzyl aminodiol derivatives in the transition states leads to better selectivities when compared to flexible moieties. Furthermore, our results clearly indicate that the spirooxazolidine ring (ligand 15 and 23) has weaker catalytic performance compared with fused 1,3-oxazine systems [37,38]. These results show good accordance with those observed with sabinane- or pinane-based spirooxazolidines were reported in our earlier studies [40,60].
The best (R)-selectivity can be explained with the steric effect of O-benzyl and N-(S)-1-phenylethyl substituent as it is given on Figure 2. The carbon of ethyl group of Et2Zn can attack the carbonyl group from the less hindered Re face resulting in (R)-25 as a main product.
With best catalysts 20 and 21, the diethylzinc addition reaction was extended to further aromatic and aliphatic aldehydes (Scheme 8). Our results are presented in Table 3. The enantiomeric purities of the 1-aryl and 1-alkyl-1-propanols obtained were determined by GC on a CHIRASIL-DEX CB column or by chiral HPLC analysis on a Chiralcel OD-H column, according to the literature methods [37].
In order to get insight into the mechanism of chiral control over the ethyl-transfer exerted by the pinane ligand, first we carried out modeling studies at the Hartree–Fock level of theory [67] using LANL2DZ basis set [68] on 27A comprising benzaldehyde coordinated to Zn-centers and on 27C with covalently bonded R-carbinol (Figure 3). Both complexes were identified as local minima on the potential energy surface (PES). The transition state of the ethyl transfer 27B was located by QST2 method [69] as a saddle point on the PES. In accord with the general expectations, the ethyl transfer was found to be a highly exothermic step accompanied by a significant decrease in the Gibbs free energy (‒50.5 kcal/mol), but proceeds via a high activation barrier (+30.5 kcal/mol). It is of pronounced significance that the attempts to find a local minimum representing 26A, the benzaldehyde complex preformed for the ethyl-transfer leading to S-adduct have failed so far, as the optimization of all the tentative initial structures led to 27A, the complex mentioned above, that is preformed for the formation of R-carbinol. These findings suggest that the formation of S-carbinol can be ascribed to a competitive process that takes place without the involvement of the pinane ligand affording racemic product, while the investigated ligand seems to promote the exclusive formation of the R-carbinol product.
All calculations were carried out by using Gaussian 09 software package [70]. The optimized structures are available from the authors.

3. Discussion

Starting from natural (−)-β-pinene, a monoterpene-based 3-amino-1,2-diol library has been created via the epoxide ring opening of epoxyalcohol as key intermediate, whereas the reactions of N-substituted aminodiols with formaldehyde resulted in spirooxazolidines with high regioselectivity. Moreover, O-benzylation of key allylic alcohol intermediate led to O-benzyl aminodiols by aminolysis of its epoxide. The ring closure of O-benzyl, N-isopropyl aminodiol furnished the corresponding spirooxazolidine. Aminodiol derivatives were proven reliable chiral catalysts in the enantioselective addition of diethylzinc to aldehydes. The enantioselective nature of the catalytic activity proved to be N- substituent-dependent, and molecular modeling was applied to explain this phenomenon. As a result of the modeling, the O-benzyl and N-(S)-1-phenylethyl substituent aminodiol 21 provided high enantiomeric excess values (80% ee with (R) selectivity) in the model reactions. This ligand also proved to be excellent catalysts in the additions of diethylzinc to either aromatic or aliphatic aldehydes.

4. Materials and Methods

4.1. Materials and General Methods

1H- and 13C- NMR spectra were obtained on a Bruker Avance DRX 500 (Bruker Biospin, Karlsruhe, Baden Württemberg, Germany) [500 and 125 MHz, respectively, δ = 0 ppm (TMS)]. Chemical shifts (δ) are expressed in ppm and related to TMS as internal reference. J values are given in Hz. GC measurements were made on a Perkin-Elmer Autosystem KL GC consisting of a Flame Ionization Detector (Perkin-Elmer Corp., Norwalk, CT, USA) and a Turbochrom Workstation data system (Perkin-Elmer Corporation Norwalk, USA). O-acetyl derivatives of chiral secondary alcohol enantiomers were separated on a CHIRASIL-DEX CB column (2500 mm × 0.265 mm I.D., Agilent Technologies, Inc., Santa Clara, CA, USA). Microanalyses were achieved on a Perkin–Elmer 2400 elemental analyzer (PerkinElmer Inc., Waltham, MA, USA).
Optical rotations were determined with a Perkin–Elmer 341 polarimeter (PerkinElmer Inc., Shelton, CT, USA). Melting points were measured on a Kofler apparatus (Nagema, Dresden, Germany) and the values are uncorrected. Column chromatography of crude products was performed on Merck Kieselgel 60 (230–400 mesh ASTM, Merck Co., Darmstadt, Germany). Headaway of reactions was followed on Merck Kieselgel 60 F254-precoated TLC plates (0.25 mm thickness, Merck Co., Darmstadt, Germany).
(−)-β-Pinene 1 is commercially available from Merck Co (Cat. No.: 402753, Merck Co., Darmstadt, Germany) and its ee value was defined by Merck Co as 97%. The purity of crude products was examined by 1H NMR in each case and we could not observe the presence of any other diastereoisomer in any case. All chemicals and solvents were used as supplied (Molar Chemicals Ltd., Halásztelek, Hungary; Merck Ltd., Budapest, Hungary and VWR International Ltd., Debrecen, Hungary). THF and toluene were dried over Na wire. Synthesis of (−)-nopinone 2 and (−)-3-methylenenopinone 3 were carried out as given in literature procedures, and all physical and chemical properties of 2 and 3 were similar to those described therein [50,51]. All 1H-, 13C- NMR, HMQC, HMBC and NOESY spectra are found in the Supporting Information.

4.2. (1R,2R,5R)-6,6-Dimethyl-3-Methylenebicyclo [3.1.1]heptan-2-ol (4a)

A suspension of CeCl3.7H2O (2.46 g, 6.6 mmol) in MeOH (50.0 mL) was added to an ice-cooled solution of 3 (1.0 g, 6.6 mmol) in MeOH (50.0 mL). The reaction mixture was stirred in an ice bath for 30 min before NaBH4 (0.5 g, 13.2 mmol) was slowly added to the mixture. Stirring was continued for 30 min at 0 °C. When the reaction was complete, the mixture was evaporated at 20 °C then poured into brine and the product was extracted with Et2O (3 × 150 mL). The combined organic phase was washed with 3.5% HCl aqueous solution (100 mL) and dried (Na2SO4). After evaporation of the solvent in vacuo, the crude product 4a was used without further purification for the next step.
Yield: 87%, white crystals. m.p.: 51–55 °C. [ α ] D 20 =+78 (c 0.255, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.84–0.87 (1H, m), 1.03 (3H, s), 1.25 (3H, s), 1.94–1.95 (1H, m), 2.09–2.12 (1H, m), 2.31–2.36 (1H, m), 2.56–2.77 (2H, m), 4.48 (1H, s), 5.16–5.42 (2H, m). 13C NMR (125 MHz, CDCl3): δ = 21.7, 27.1, 29.6, 34.4, 38.1, 39.6, 46.4, 75.6, 114.2, 148.6. Anal. Calculated for C10H16O (152.24): C 78.90, H 10.59; found: C 78.95, H 10.57.

4.3. (1R,2R,5R)-6,6-Dimethylspiro[bicyclo[3.1.1]heptane-3,2’-oxiran]-2-ol (5)

To the solution of 4a (1.0 g, 6.6 mmol) in dry toluene (50 mL), catalytic amount of VO(acac)2 (7.0 mg) was added. The mixture was stirred for 30 min then t-BuOOH (70% solution in water, 1.7 g, 13.2 mmol), dried briefly (Na2SO4), was added dropwise at 25 °C. Stirring was continued (20 h) whereupon KOH (2.0 g) in brine (80.0 mL) was added. The mixture was extracted with toluene (3 × 100 mL), the organic layer was dried (Na2SO4) and evaporation at 20 °C gave compound 5 (84%) as yellow oil. Crude product 5 was used for the next step.

4.4. General Procedure for Ring Opening of Epoxide 5 with Primary and Secondary Amines

To a solution of the appropriate amine (1.2 mmol) in MeCN (5.0 mL) and LiClO4 (0.06 g, 0.6 mmol), solution of epoxide 5 or 18 (0.6 mmol) in MeCN (5.0 mL) was added. After 6 h reflux the reaction was found to be completed (indicated by TLC), and the mixture was evaporated to dryness, the residue was dissolved in water (15.0 mL) and extracted with CH2Cl2 (3 × 50 mL). The combined organic phase was dried (Na2SO4), filtered and concentrated. The purification of the crude product was accomplished by column chromatography on silica gel with an appropriate solvent mixture resulting in compounds 613 or 1922, respectively.

4.4.1. (1R,2R,3S,5R)-3-((Benzylamino)methyl)-6,6-Dimethylbicyclo [3.1.1]heptane-2,3-diol (6)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 1:2). Yield: 95%, yellow crystals; m.p.: 170–175 °C. [ α ] D 20 = +7 (c = 0.27, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.80 (1H, d, J = 10.8 Hz), 1.09 (3H, s), 1.20 (3H, s), 1.87–1.90 (2H, m), 2.17–2.23 (3H, m), 2.66 (2H, s), 3.81–3.84 (2H, m), 7.26–7.35 (5H, m). 13C NMR (125 MHz, CDCl3): δ = 21.9, 27.0, 27.5, 37.2, 40.4, 40.6, 47.0, 52.3, 59.7, 66.1, 73.8, 129.1, 129.7, 130.2, 130.2. Anal. Calculated for C17H25NO2 (275.39): C 74.14, H 9.15, N 5.09; found: C 74.19, H 5.13, N 5.11.

4.4.2. (1R,2S,3S,5R)-6,6-Dimethyl-3-((((R)-1-phenyl-ethyl)amino)methyl)bicyclo[3.1.1]heptane-2,3- diol (7)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 1:2) to Yield: 43%, yellow crystals, m.p.: 112–116 °C. [ α ] D 20 =+36 (c=0.26, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.76–0.80 (1H, m), 1.07 (3H, s), 1.19 (3H, s), 1.41 (3H, d, J = 6.5 Hz), 1.78–1.86 (2H, m), 2.11–2.22 (3H, m), 2.48 (1H, d, J = 11.4 Hz), 2.60 (1H, d, J = 11.4 Hz), 3.81–3.84 (2H, m), 7.24–7.36 (5H, m). 13C NMR (125 MHz, CDCl3):δ = 21.3, 23.9, 27.3, 27.8, 37.3, 41.0, 41.3, 47.3, 58.1, 62.6, 67.3, 76.5, 126.6, 127.3, 128.7. Anal. Calculated for C18H27NO2 (289.42): C 74.70, H 9.40, N 4.84; found: C 74.73, H 9.44, N 4.80.

4.4.3. (1R,2S,3S,5R)-6,6-Dimethyl-3-((((S)-1-phenyl-ethyl)amino)methyl)bicyclo[3.1.1]heptane-2,3- diol (8)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 1:2) to Yield: 45%, yellow oil. [ α ] D 20 = −3 (c = 0.27, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.68 (1H, d, J = 9.1 Hz), 1.08 (3H, s), 1.18 (3H, s), 1.45 (3H, d, J = 6.5 Hz), 1.81–1.86 (2H, m), 2.14–2.20 (3H, m), 2.46–2.48 (1H, d, J = 11.9 Hz), 2.56–2.58 (1H, d, J = 11.9 Hz), 3.75–3.84 (2H, m), 7.24–7.35 (5H, m). 13C NMR (125 MHz, CDCl3): δ = 21.5, 23.7, 27.3, 27.7, 37.3, 41.0, 41.5, 47.2, 58.6, 62.7, 67.3, 76.1, 126.5, 127.4, 128.7. Anal. Calculated for C18H27NO2 (289.42): C 74.70, H 9.40, N 4.84; found: C 74.73, H 9.44, N 4.80.

4.4.4. (1R,2R,3S,5R)-3-((Disopropylamino)methyl)-6,6-Dimethylbicyclo-[3.1.1]heptane-2,3-diol (9)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 1:2) then recrystallized in Et2O. Yield: 32%, yellow crystals, m.p.: 190–194 °C.   [ α ] D 20 = +8 (c = 0.26, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.88 (1H, d, J = 10.7 Hz), 1.04 (3H, s), 1.22 (3H, s), 1.44 (6H, d, J = 6.2 Hz), 1.95–1.97 (2H, m), 2.24–2.35 (3H, m), 3.00 (1H, d, J = 11.9 Hz), 3.12 (1H, d, J = 11.9 Hz), 3.49–3.55 (1H, m), 4.15 (1H, s). 13C NMR (125 MHz, CDCl3): δ = 19.2, 19.4, 22.0, 27.0, 27.5, 37.3, 40.5, 42.0, 47.0, 52.7, 58.7, 66.1, 73.6. Anal. Calculated for C13H25NO2 (227.35): C 68.68, H 11.08, N 6.16; found: C 68.70, H 11.03, N 6.12.

4.4.5. (1R,2S,3S,5R)-3-((Dibenzylamino)methyl)-6,6-Dimethylbicyclo[3.1.1]heptane-2,3-diol (10)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 9:1). Yield: 40%, yellow oil.   [ α ] D 20 =+16 (c=0.255, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.51 (1H, d, J = 9.2 Hz), 1.06 (3H, s), 1.15 (3H, s), 1.57 (2H, s), 1.80–1.82 (1H, m), 1.90–1.94 (1H, m), 2.08–2.11 (3H, m), 2.63–2.66 (1H, d, J = 13.7 Hz), 2.72–2.75 (1H, d, J = 13.7 Hz), 3.05 (1H, d, J = 6.1 Hz), 3.49 (1H, d, J = 4.0 Hz), 3.74 (4H, s), 3.78–3.80 (1H, m), 4.62 (1H, s), 7.23–7.34(10H, m). 13C NMR (125 MHz, CDCl3): δ = 12.4, 27.3, 27.8, 37.2, 41.2, 43.1, 47.2, 59.6, 67.9, 68.9, 77.4, 127.3, 128.4, 129.2, 139.0. Anal. Calculated for C24H31NO2 (365.52): C 78.86, H 8.55, N 3.83; found: C 78.83, H 8.50, N 3.79.

4.4.6. (1R,2S,3S,5R)-3-((Benzyl((R)-1-phenylethyl)amino)methyl)-6,6-dimethylbicyclo[3.1.1] heptane-2,3- diol (11)

Purified by chromatography on silica gel column (n-hexane/EtOAc = 9:1). Yield: 30%, white crystals, m.p.: 120–125 °C.   [ α ] D 20 = +45 (c = 0.28, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.49 (1H, d, J = 10.1 Hz), 1.05 (3H, s), 1.14 (3H, s), 1.41–1.42 (3H, d, J = 6.8 Hz), 1.57 (1H, s), 1.82–1.83 (1H, m), 1.98–2.03 (2H, m), 2.08–2.12 (1H, m), 2.15-2.18 (1H, m), 2.48–2.50 (1H, d, J = 13.7 Hz), 2.80–2.83 (1H, d, J = 13.7 Hz), 2.95 (1H, d, J = 5.7 Hz), 3.55–3.57 (1H, m), 3.77 (2H, s), 4.08–4.12 (1H, m), 4.85 (1H, s), 7.25–7.36 (10 H, m). 13C NMR (125 MHz, CDCl3): δ = 13.4, 21.3, 27.3, 28.0, 37.1, 41.3, 43.8, 47.1, 56.0, 57.6, 64.4, 67.3, 77.5, 127.2, 128.2, 128.2, 128.5, 128.9, 139.7, 141.9. Anal. Calculated for C25H33NO2 (379.54): C 79.11, H 8.76, N 3.69; found: C 79.08, H 8.81, N 3.65.

4.4.7. (1R,2S,3S,5R)-3-((Benzyl((S)-1-phenylethyl)amino)methyl)-6,6-dimethylbicyclo[3.1.1] heptane-2,3- diol (12)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 9:1). Yield: 60%, yellow oil. [ α ] D 20 = −28 (c = 0.26, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.62–0.65 (1H, m), 1.05 (3H, s), 1.15 (3H, s), 1.42 (3H, d, J = 6.3 Hz), 1.76–1.79 (1H, m), 1.81–1.85 (1H, m), 1.99–2.03 (1H, m), 1.15–2.18 (2H, m), 2.58 (1H, d, J = 14.2 Hz), 2.79 (1H, d, J = 13.8 Hz), 3.66–3.68 (1H, d, J = 12.8 Hz), 3.82–3.85 (1H, d, J = 13.6 Hz), 3.86 (1H, s), 5.29 (1H, s), 7.24–7.35 (10H, m). 13C NMR (125 MHz, CDCl3): δ = 12.7, 21.3, 27.3, 28.2, 37.0, 41.1, 43.8, 47.2, 55.4, 57.3, 65.0, 66.8, 78.3, 127.2, 127.3, 128.2, 128.2, 128.5, 128.9, 139.41, 142.0. Anal. Calculated for C25H33NO2 (379.54): C 79.11, H 8.76, N 3.69; found: C 79.08, H 8.81, N 3.65.

4.4.8. (1R,2S,3S,5R)-3-((4-Benzylpiperidin-1-yl)methyl)-6,6-dimethylbicyclo[3.1.1]heptane-2,3-diol (13)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 1:1). Yield: 55%, yellow oil. [ α ] D 20 = +14 (c = 0.26, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.75 (1H, d, J = 9.6 Hz), 1.10 (3H, s), 1.19 (3H, s), 1.25–1.33 (2H, m), 1.51–1.63 (4H, m), 1.85–1.89 (1H, m), 1.91–1.95 (1H, m), 2.17–2.28 (5H, m), 2.48–2.56 (4H, m), 2.89–3.01 (2H, m), 3.68–3.76 (2H, m), 7.12–7.28 (5H, m). 13C NMR (125 MHz, CDCl3): δ = 21.2, 27.5, 28.4, 32.4, 41.2, 43.1, 44.6, 47.1, 54.3, 55.5, 74.4, 152.8, 128.2, 129.1, 131.7. Anal. Calculated for C22H33NO2 (343.51): C 76.92, H 9.68, N 4.08; found: C 76.88, H 9.71, N 4.13.

4.4.9. (1R,2S,3S,5R)-3-((Benzylamino)methyl)-2-(benzyloxy)-6,6-dimethylbicyclo[3.1.1]heptan-3-ol (19)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 1:2). Yield: 32%, yellow oil.   [ α ] D 20 =+35 (c=0.27, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.95 (3H, s), 1.23 (3H, s), 1.79 (2H, d, J = 10.1 Hz), 1.90–1.92 (3H, m), 2.16–2.24 (2H, m), 2.42–2.45 (1H, m), 3.31 (1H, d, J = 11.8 Hz), 3.69 (1H, d, J = 13.1 Hz), 3.77 (1H, d, J = 13.5 Hz), 3.87 (1H, d, J = 4.6 Hz), 4.21 (1H, d, J = 10.5 Hz), 4.59 (1H, d, J = 11.4 Hz), 7.10–7.12 (2H, m), 7.20–7.32 (5H, m). 13C NMR (125 MHz, CDCl3): δ = 22.2, 24.7, 27.3, 37.5, 38.9, 40.0, 41.2, 54.2, 56.7, 70.0, 71.4, 89.2, 126.9, 127.4, 127.4, 127.9, 128.4, 128.4, 138.8. Anal. Calculated for C24H31NO2 (365.52): C 78.86, H 8.55, N 3.83; found: C 78.90, H 8.51, N 3.80.

4.4.10. (1R,2S,3S,5R)-2-(Benzyloxy)-6,6-dimethyl-3-((((R)-1-phenylethyl)amino)methyl)bicyclo[3.1.1] heptan-3-ol (20)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 1:2). Yield: 36%, white crystals, m.p.: 84–86 °C.   [ α ] D 20 = +4 (c = 0.26, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.89 (3H, s), 1.22 (3H, s), 1.24 (3H, d, J = 6.1 Hz), 1.79 (1H, d, J = 10.0 Hz), 1.84–1.90 (3H, m), 2.01 (1H, d, J = 10.6 Hz), 2.17–2.22 (1H, m), 2.43–2.46 (1H, dd, J = 4.6 Hz, 11.6 Hz), 3.03 (1H, d, J = 11.5 Hz), 3.61 (1H, dd, J = 6.7 Hz, 13.5 Hz), 3.89 (1H, d, J = 4.3 Hz), 4.19 (1H, d, J = 11.6 Hz), 4.60 (1H, d, J = 10.3 Hz), 6.88–6.90 (2H, m), 7.16–7.19 (3H, m), 7.25–7.37 (5H, m). 13C NMR (125 MHz, CDCl3): δ = 22.2, 24.5, 34.8, 27.3, 37.5, 38.8, 40.0, 47.2, 41.1, 55.3, 58.9, 70.2, 71.3, 89.5, 126.0, 126.6, 127.6, 127.9, 128.4, 128.5. Anal. Calculated for C25H33NO2 (379.25): C 79.11, H 8.76, N 3.69; found: C 79.10, H 8.80, N 3.73.

4.4.11. (1R,2S,3S,5R)-2-(Benzyloxy)-6,6-dimethyl-3-((((S)-1-phenylethyl)amino)methyl)bicyclo[3.1.1] heptan-3-ol (21)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 1:2). Yield: 30%, yellow oil.   [ α ] D 20 = −10 (c = 0.28, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.95 (3H, s), 1.20 (3H, d, J = 7.0 Hz), 1.22 (3H, s), 1.75 (1H, d, J = 10.1 Hz), 1.81–1.83 (2H, m), 1.86–1.88 (1H, m), 2.10 (1H, d, J = 11.2 Hz), 2.16–2.20 (1H, m), 2.44 (1H, q, J = 4.9 Hz, 11.0 Hz), 3.21 (1H, d, J = 11.0 Hz), 3.70 (1H, dd, J = 6.3 Hz, 13.4 Hz), 3.90 (1H, d, J = 4.2 Hz), 4.26 (1H, d, J = 11.5 Hz), 4.63 (1H, d, J = 11.5 Hz), 7.19-7.38 (10H, m). 13C NMR (125 MHz, CDCl3): δ = 22.2, 24.2, 24.7, 27.3, 37.5, 38.9, 40.0, 41.3, 54.7, 57.9, 70.1, 71.3, 89.5, 126.7, 126.9, 127.4, 127.5, 128.4, 128.4. Anal. Calculated for C25H33NO2 (379.54): C 79.11, H 8.76, N 3.69; found: C 79.08, H 8.71, N 3.74.

4.4.12. (1R,2S,3S,5R)-2-(Benzyloxy)-3-((isopropylamino)methyl)-6,6-dimethylbicyclo[3.1.1]heptan-3- ol (22)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 1:9) then recrystallized from Et2O. Yield: 36%, yellow crystals, m.p.: 176–179 °C. [ α ] D 20 = +33 (c = 0.26, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.87 (3H, d, J = 6.1 Hz), 1.04 (3H, s), 1.16 (3H, d, J = 7.0 Hz), 1.30 (3H, s), 1.74 (1H, d, J = 10.3 Hz), 2.01–2.03 (1H, m), 2.07–2.16 (2H, m), 2.25–2.30 (1H, m), 2.62 (1H, dd, J = 4.9 Hz, 6.0 Hz), 3.13–3.23 (2H, m), 3.30–3.34 (1H, m), 4.28 (1H, d, J = 4.2 Hz), 4.41 (1H, d, J = 8.8 Hz), 4.56 (1H, d, J = 9.3 Hz), 6.40 (1H, s), 6.83 (1H, s), 7.29–7.43 (5H, m). 13C NMR (125 MHz, CDCl3): δ = 18.4, 19.4, 22.2, 24.4, 27.0, 37.1, 39.6, 39.8, 40.7, 51.9, 52.7, 69.8, 71.1, 85.8, 128.4, 128.7, 129.2, 137.4. Anal. Calculated for C20H31NO2 (317.24): C 75.67, H 9.84, N 4.41; found: C 75.70, H 9.81, N 4.46.

4.5. (1R,2R,3S,5R)-3-(Aminomethyl)-6,6-dimethylbicyclo-[3.1.1]heptane-2,3-diol (14)

To a suspension of 5% Pd/C (87 mg) in n-hexane/EtOAc = 1:1 (20 mL) was added aminodiol 6 (0.29 g, 1.0 mmol) in n-hexane/EtOAc = 1:1 (20 mL). The mixture was stirred under a hydrogen atmosphere at 25 °C. The reaction was monitored by means of TLC and was completed after 24 h stirring at room temperature. The resulting mixture was filtered through a Celite pad and the solution was evaporated to dryness. The obtained crude product was recrystallized in Et2O, resulting in primary aminodiol 14 as the single product.
Yield: 74%, yellow crystals, m.p.: 204–207 °C. [ α ] D 20 = +10 (c = 0.27, MeOH). 1H NMR (500 MHz, DMSO-d6): δ = 0.90 (1H, d, J = 10.5 Hz), 1.03 (3H, s), 1.17 (3H, s), 1.93–2.03 (2H, m), 2.08–2.09 (1H, m), 2.19–2.24 (1H, m), 2.72 (1H, d, J = 12.6 Hz), 2.83 (1H, d, J = 12.5 Hz), 3.74 (1H, s), 4.78 (1H, s), 5.69 (1H, s). 13C NMR (125 MHz, DMSO-d6): δ = 22.4, 27.6, 27.7, 37.3, 41.0, 47.1, 53.4, 66.2, 73.8. Anal. Calculated for C10H19NO2 (185.27): C 64.83, H 10.34, N 7.56; found: C 64.80, H 10.31, N 7.61.

4.6. General Procedure for Ring Closure of Aminodiol Derivatives with Formaldehyde

To the solution of aminodiols 6, 9 or 22 (1.8 mmol) in Et2O (5 mL), 35% aqueous formaldehyde (20 mL) was added in one portion. The mixture was stirred at room temperature for 1 h, than made alkaline with 10% aqueous KOH (20 mL) and extracted with Et2O (3 × 50 mL). After drying (Na2SO4) and solvent evaporation, crude products 15, 16 or 23 were purified by column chromatography with an appropriate solvent mixture.

4.6.1. (1R,2S,3S,5R)-3’-Benzyl-6,6-dimethylspiro [bicyclo[3.1.1]heptane-3,5’-oxazolidin]-2-ol (15)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 2:1). Yield: 98%, yellow oil. [ α ] D 20 = +8 (c = 0.25, MeOH). 1H NMR (500 MHz, DMSO-d6): δ = 0.80 (1H, d, J = 10.3 Hz), 1.01 (3H, s), 1.14 (3H, s), 1.78–1.82 (1H, m), 2.01–2.04 (1H, m), 2.12–2.23 (3H, m), 2.83 (1H, d, J = 10.4 Hz), 2.92 (1H, d, J = 9.8 Hz), 3.74 (2H, s), 3.82 (1H, t, J = 3.6 Hz), 3.89 (1H, d, J = 3.6 Hz), 4.25 (1H, d J = 3.4 Hz), 4.32 (1H, d, J = 3.4 Hz), 7.23–7.35 (5H, m). 13C NMR (125 MHz, DMSO-d6): δ = 22.3, 27.1, 27.5, 37.1, 40.2, 42.5, 46.3, 56.5, 71.9, 78.0, 79.2, 86.3, 127.4, 128.8, 139.4. Anal. Calculated for C18H25NO2 (287.19): C 75.22, H 8.77, N 4.87; found: C 75.21, H 8.80, N 4.82.

4.6.2. (1R,5R)-3’-Isopropyl-6,6-dimethylspiro[bicyclo[3.1.1]heptane-3,5’-oxazolidin]-2-ol (16)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 1:9). Yield: 40%, yellow oil.   [ α ] D 20 =+6 (c=0.255, MeOH). 1H NMR (500 MHz, DMSO-d6): δ = 0.89-0.91. (1H, d, J = 10.4 Hz), 0.98 (3H, s), 1.01-1.04 (6H, dd, J = 6.14 Hz), 1.15 (3H, s), 1.24 (1H, s), 1.78–1.81 (1H, m), 2.00–2.03 (1H, m), 2.09–2.19 (3H, m), 2.47 (1H, s), 2.64-2.68 (1H, m), 2.85-2.86 (1H, m), 1.32-1.37 (1H, m), 4.20 (1H, s), 4.34 (1H, s). 13C NMR (125 MHz, DMSO-d6): δ = 22.2, 22.3, 26.9, 27.5, 37.3, 41.6, 46.3, 51.9, 69.4, 78.6, 78.9, 85.2. Anal. Calculated for C14H25NO2 (239,36): C 70.25, H 10.53, N 5.85; found: C 70.28, H 10.50, N 5.89.

4.6.3. (1R,2S,3S,5R)-2-(Benzyloxy)-3’-isopropyl-6,6-dimethylspiro[bicyclo[3.1.1]heptane-3,5’- oxazolidine] (23)

Purified by column chromatography on silica gel (n-hexane/EtOAc = 4:1). Yield: 50%, yellow oil.   [ α ] D 20 = +32 (c = 0.23, MeOH). 1H NMR (500 MHz, CDCl3): δ = 0.94 (3H, s), 1.05 (6 H, d, J = 7.0 Hz), 1.22 (3H, s), 1.54 (1H, d, J = 10.0 Hz), 1.87–1.91 (1H, m), 2.16–2.23 (3H, m), 2.29–2.32 (1H, m), 2.43–2.48 (2H, m), 3.78–3.81 (2H, m), 4.24 (1H, d, J = 3.4 Hz), 4.46 (1H, d, J = 4.0 Hz), 4.58 (1H, d, J = 11.9 Hz), 4.62 (1H, d, J = 11.7 Hz), 7.23–7.35 (5H, m). 13C NMR (125 MHz, CDCl3): δ = 22.0, 22.2, 22.5, 25.2, 27.1, 37.5, 39.9, 40.6, 42.7, 52.2, 58.6, 71.6, 82.5, 84.0, 87.8, 127.1, 127.2, 128.1, 139.3. Anal. Calculated for C21H31NO2 (329.48): C 76.55, H 9.48, N 4.25; found: C 76.51, H 9.50, N 4.27.

4.7. (1R,2S,5R)-2-(Benzyloxy)-6,6-dimethyl-3-methylenebicyclo[3.1.1]heptane (17)

A suspension of NaH (60% purity, 0.26 g, 6.6 mmol) in dry THF (10.0 mL) was added to a solution of 3 (1.0 g, 6.6 mmol) in dry THF (20.0 mL). The reaction mixture was stirred at 25 °C for 30 min than KI (1.1 g, 6.6 mmol) and benzyl bromide (1.2 mL, 13.2 mmol) were added to the suspension. After stirring for 6 h at 60 °C the reaction was completed (monitored by means of TLC) and the mixture was poured into saturated NH4Cl solution (30 mL) and extracted with EtOAc (3 × 50 mL). The combined organic phase was dried over anhydrous Na2SO4. The solvent was evaporated in vacuo and the crude product 17 was used for the next step.

4.8. (1R,2S,3S,5R)-2-(Benzyloxy)-6,6-dimethylspiro[bicyclo[3.1.1]heptane-3,2’-oxirane] (18)

To a mixture of solution of 17 (0.4 g, 1.65 mmol) in CH2Cl2 (20 mL) and Na2HPO4·2H2O (0.88 g, 4.95 mmol) in water (20 mL), m-chloroperbenzoic acid (70% purity, 0.81 g, 3.3 mmol) was added at 0 °C. The reaction was completed after 2 h stirring at 25 °C (indicated by means of TLC), than the mixture was separated and the aqueous phase was extracted with CH2Cl2 (3 × 50 mL). The organic layer was washed with a 5% KOH solution (3 × 20 mL), then dried (Na2SO4) and evaporated in vacuo. The crude product was purified by column chromatography on silica gel (n-hexane/EtOAc = 19:1) to provide 20 as the single product.
Yield: 47%, colorless oil. [ α ] D 20 = +48 (c = 0.26, MeOH). 1H NMR (500 MHz, CDCl3): δ = 1.10 (3H, s), 1.19–1.22 (1H, m), 1.24–1.26 (1H, m), 1.27 (3H, s), 1.34–1.35 (1H, m), 1.90 (1H, dd, J =3.6 Hz, J = 10.9 Hz), 2.01–2.04 (1H, m), 2.36–2.42 (3H, m), 2.69 (1H, d, J =6.0 Hz), 3.32 (1H, d, J = 5.4 Hz), 3.75 (1H, d, J = 2.5 Hz), 4.34 (1H, d, J = 12.3 Hz), 4.56 (1H, d, J = 11.4 Hz), 7.25–7.34 (5H, m). 13C NMR (125 MHz, CDCl3): δ = 22.5, 26.8, 28.4, 36.0, 37.3, 40.8, 42.9, 53.3, 58.4, 71.0, 85.0, 127.3, 127.4, 128.3, 138.8. Anal. Calculated for C17H22O2 (258.36): C 79.03, H 8.58; found: C 79.07, H 8.61.

4.9. General Procedure for the Reaction of Diethylzinc with Aldehydes in the Presence of Chiral Catalysts

The mixture of appropriate catalyst (0.15 mmol) and 1 M Et2Zn in n-hexane solution (1.5 mL, 1.5 mmol) was stirred for 25 min in argon atmosphere at room temperature and then appropriate aldehyde (1.5 mmol) was added to the mixture in one portion. After 20 h of stirring at room temperature, the reaction was quenched with saturated NH4Cl solution (15 mL) and extracted with EtOAc (2 × 20 mL). The combined organic layer was washed with H2O (10 mL) and dried (Na2SO4) and the solvent was evaporated under vacuum resulting in 25 and 29ae. The ee and absolute configuration of the resulting phenyl-1-propanol (25) were determined by chiral GC on a Chirasil-DEX CB column after O-acetylation in Ac2O/DMPA/pyridine [65,66] and without derivatization for 1-cyclohexyl-1-propanol (29d) and 3-heptanol (29e) [37]. Identification of 29ac was done by chiral HPLC analysis on a Chiralcel OD-H column with V(n-hexane)/V(2-propanol) = 98:2 mixture, 1.0 mL/min, 210 nm and the direction of the optical rotation of products was also checked [37].

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/5/474/s1, 1H-, 13C- NMR, HMQC, HMBC and NOESY spectra of new compounds as Figures S3–S83.

Author Contributions

Z.S. conceived and designed the experiments; M.R. performed the experiments, analyzed the data; all authors (M.R., T.M.L., F.F. and Z.S.) discussed the results and contributed to write the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by EU-Funded Hungarian Grant GINOP, grant number 2.3.2-15-2016-00012 and the APC was funded by Hungarian Scientific Research Found (OTKA), grant number K-115731.

Acknowledgments

We are grateful for financial supports from the EU-funded Hungarian grant GINOP-2.3.2-15-2016-00012 and Hungarian Research Foundation (OTKA No. K 115731).

Conflicts of Interest

The authors declare no conflict of interest. The founding sponsors had no role in the design of the study; in the collection, analyses or interpretation of data; in the writing of the manuscript and in the decision to publish the results. Sources of funding for the study.

References

  1. Caprio, V.; Williams, J.M.J. Catalysis in Asymmetric Synthesis, 2nd ed.; Wiley: Hoboken, NJ, USA, 2009; ISBN 978-1-4051-9091-6. [Google Scholar]
  2. Satyanarayana, T.; Kagan, H.B. The multi-substrate screening of asymmetric catalysts. Adv. Synth. Catal. 2005, 347, 737–748. [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. Noyori, R.; Kitamura, M. Enantioselective addition of organometallic reagents to carbonyl compounds: Chirality transfer, multiplication, and amplification. Angew. Chem. Int. Ed. Engl. 1991, 30, 49–69. [Google Scholar] [CrossRef]
  5. Soai, K.; Niwa, S. Enantioselective addition of organozinc reagents to aldehydes. Chem. Rev. 1992, 92, 833–856. [Google Scholar] [CrossRef]
  6. Avalos, M.; Babiano, R.; Cintas, P.; Jiménez, J.L.; Palacios, J.C. Nonlinear stereochemical effects in asymmetric reactions. Tetrahedron Asymmetry 1997, 8, 2997–3017. [Google Scholar] [CrossRef]
  7. Hanessian, S.; Soma, U.; Dorich, S.; Deschênes-Simard, B. Total synthesis of (+)-ent-Cyclizidine: Absolute configurational confirmation of antibiotic M146791. Org. Lett. 2011, 13, 1048–1051. [Google Scholar] [CrossRef] [PubMed]
  8. Kim, H.C.; Kang, S.H. Total synthesis of Azithromycin. Angew. Chem. Int. Ed. 2009, 48, 1827–1829. [Google Scholar] [CrossRef]
  9. Pastó, M.; Moyano, A.; Pericàs, M.A.; Riera, A. An enantioselective, stereodivergent approach to anti- and syn-α-hydroxy-β-amino acids from anti-3-amino-1,2-diols. Synthesis of the ready for coupling taxotere® side chain. Tetrahedron Asymmetry 1996, 7, 243–262. [Google Scholar] [CrossRef]
  10. Ohsaki, A.; Ishiyama, H.; Yoneda, K.; Kobayashi, J. Secu’amamine A, a novel indolizidine alkaloid from Securinega suffruticosa var. amamiensis. Tetrahedron Lett. 2003, 44, 3097–3099. [Google Scholar] [CrossRef]
  11. Azumi, M.; Ogawa, K.; Fujita, T.; Takeshita, M.; Yoshida, R.; Furumai, T.; Igarashi, Y. Bacilosarcins A and B, novel bioactive isocoumarins with unusual heterocyclic cores from the marine-derived bacterium Bacillus subtilis. Tetrahedron 2008, 64, 6420–6425. [Google Scholar] [CrossRef]
  12. Faigl, F.; Erdélyi, Z.; Deák, S.; Nyerges, M.; Mátravölgyi, B. A new pyrrolidine-derived atropisomeric amino alcohol as a highly efficient chiral ligand for the asymmetric addition of diethylzinc to aldehydes. Tetrahedron Lett. 2014, 55, 6891–6894. [Google Scholar] [CrossRef]
  13. Pu, L.; Yu, H.-B. Catalytic asymmetric organozinc additions to carbonyl compounds. Chem. Rev. 2001, 101, 757–824. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, M.-C.; Li, G.-W.; Hu, W.-B.; Hua, Y.-Z.; Song, X.; Lu, H.-J. A mathematical expression for the enantioselectivity and thermodynamic factors in the conformational equilibrium of catalysts in the addition of diethylzinc to benzaldehyde. Tetrahedron Asymmetry 2014, 25, 1360–1365. [Google Scholar] [CrossRef]
  15. Wiseman, J.M.; McDonald, F.E.; Liotta, D.C. 1-Deoxy-5-hydroxysphingolipids as new anticancer Principles: An efficient procedure for stereoselective syntheses of 2-amino-3,5-diols. Org. Lett. 2005, 7, 3155–3157. [Google Scholar] [CrossRef] [PubMed]
  16. Woltering, T.J.; Wostl, W.; Hilpert, H.; Rogers-Evans, M.; Pinard, E.; Mayweg, A.; Göbel, M.; Banner, D.W.; Benz, J.; Travagli, M.; et al. BACE1 inhibitors: A head group scan on a series of amides. Bioorg. Med. Chem. Lett. 2013, 23, 4239–4243. [Google Scholar] [CrossRef] [PubMed]
  17. Kai, H.; Morioka, Y.; Tomida, M.; Takahashi, T.; Hattori, M.; Hanasaki, K.; Koike, K.; Chiba, H.; Shinohara, S.; Kanemasa, T.; et al. 2-Arylimino-5,6-dihydro-4H-1,3-thiazines as a new class of cannabinoid receptor agonists. Part 2: Orally bioavailable compounds. Bioorg. Med. Chem. Lett. 2007, 17, 3925–3929. [Google Scholar] [CrossRef]
  18. Kai, H.; Morioka, Y.; Murashi, T.; Morita, K.; Shinonome, S.; Nakazato, H.; Kawamoto, K.; Hanasaki, K.; Takahashi, F.; Mihara, S.; et al. 2-Arylimino-5,6-dihydro-4H-1,3-thiazines as a new class of cannabinoid receptor agonists. Part 1: Discovery of CB2 receptor selective compounds. Bioorg. Med. Chem. Lett. 2007, 17, 4030–4034. [Google Scholar] [CrossRef]
  19. Kai, H.; Morioka, Y.; Koriyama, Y.; Okamoto, K.; Hasegawa, Y.; Hattori, M.; Koike, K.; Chiba, H.; Shinohara, S.; Iwamoto, Y.; et al. 2-Arylimino-5,6-dihydro-4H-1,3-thiazines as a new class of cannabinoid receptor agonists. Part 3: Synthesis and activity of isosteric analogs. Bioorg. Med. Chem. Lett. 2008, 18, 6444–6447. [Google Scholar] [CrossRef]
  20. Xu, X.; Qian, X.; Li, Z.; Song, G.; Chen, W. Synthesis and fungicidal activity of fluorine-containing phenylimino-thiazolidines derivatives. J. Fluor. Chem. 2005, 126, 297–300. [Google Scholar] [CrossRef]
  21. Cherng, Y.-J.; Fang, J.-M.; Lu, T.-J. Pinane-type tridentate reagents for enantioselective reactions: Reduction of ketones and addition of diethylzinc to aldehydes. J. Org. Chem. 1999, 64, 3207–3212. [Google Scholar] [CrossRef]
  22. Bergmeier, S.C. The synthesis of vicinal amino alcohols. Tetrahedron 2000, 56, 2561–2576. [Google Scholar] [CrossRef]
  23. Braga, A.L.; Rubim, R.M.; Schrekker, H.S.; Wessjohann, L.A.; De Bolster, M.W.G.; Zeni, G.; Sehnem, J.A. The facile synthesis of chiral oxazoline catalysts for the diethylzinc addition to aldehydes. Tetrahedron Asymmetry 2003, 14, 3291–3295. [Google Scholar] [CrossRef]
  24. Matsuda, K.; Yamamoto, S.; Irie, M. Diastereoselective cyclization of a diarylethene having a chiral N-phenylethylamide substituent in crystals. Tetrahedron Lett. 2001, 42, 7291–7293. [Google Scholar] [CrossRef]
  25. Pedrosa, R.; Andrés, C.; Duque-Soladana, J.P.; Rosón, C.D. Regio- and stereoselective 6-exo-trig radical cyclisations onto chiral perhydro-1,3-benzoxazines: Synthesis of enantiopure 3-alkylpiperidines. Tetrahedron Asymmetry 2000, 11, 2809–2821. [Google Scholar] [CrossRef]
  26. Pedrosa, R.; Andrés, C.; Nieto, J.; Del Pozo, S. Synthesis of enantiopure 3-azabicyclo[3.2.0]heptanes by diastereoselective intramolecular [2 + 2] photocycloaddition reactions on chiral perhydro-1,3-benzoxazines. J. Org. Chem. 2003, 68, 4923–4931. [Google Scholar] [CrossRef]
  27. Alberola, A.; Andrés, C.; Pedrosa, R. Diastereoselective ring opening of 2-substituted N-Benzyl-4,4, 7α-trimethyl-trans-octahydro-1,3-benzoxazines by Grignard reagents. Highly enantioselective synthesis of primary amines. Synlett 1990, 1990, 763–765. [Google Scholar] [CrossRef]
  28. Andrés, C.; Nieto, J.; Pedrosa, R.; Villamañán, N. Synthesis of enantiopure primary amines by stereoselective ring opening of chiral octahydro-1,3-benzoxazines by Grignard and Organoaluminum reagents. J. Org. Chem. 1996, 61, 4130–4135. [Google Scholar] [CrossRef]
  29. Philipova, I.; Dimitrov, V.; Simova, S. Synthesis of new enantiopure aminodiols and their use as ligands for the addition of diethylzinc to benzaldehyde. Tetrahedron Asymmetry 1999, 10, 1381–1391. [Google Scholar] [CrossRef]
  30. Zhang, A.-L.; Yang, L.-W.; Yang, N.-F.; Liu, D.-C. Investigation of catalytic activity and catalytic mechanism of chiral amino diol tridentate ligands in the asymmetric addition of aldehydes in the present of methyllithium reagent. J. Organomet. Chem. 2014, 768, 50–55. [Google Scholar] [CrossRef]
  31. Stoyanova, M.P.; Shivachev, B.L.; Nikolova, R.P.; Dimitrov, V. Highly efficient synthesis of chiral aminoalcohols and aminodiols with camphane skeleton. Tetrahedron Asymmetry 2013, 24, 1426–1434. [Google Scholar] [CrossRef]
  32. Schwindt, M.A.; Belmont, D.T.; Carlson, M.; Franklin, L.C.; Hendrickson, V.S.; Karrick, G.L.; Poe, R.W.; Sobieray, D.M.; Van De Vusse, J. Unique and efficient synthesis of [2S-(2 R*,3S*,4R*)]-2-amino-1-cyclohexyl-6-methyl-3,4-heptanediol, a popular C-terminal component of many renin inhibitors. J. Org. Chem. 1996, 61, 9564–9568. [Google Scholar] [CrossRef]
  33. Rigoli, J.W.; Guzei, I.A.; Schomaker, J.M. Aminodiols via stereocontrolled oxidation of methyleneaziridines. Org. Lett. 2014, 16, 1696–1699. [Google Scholar] [CrossRef] [PubMed]
  34. Panev, S.; Linden, A.; Dimitrov, V. Chiral aminoalcohols with a menthane skeleton as catalysts for the enantioselective addition of diethylzinc to benzaldehyde. Tetrahedron Asymmetry 2001, 12, 1313–1321. [Google Scholar] [CrossRef]
  35. Cherng, Y.J.; Fang, J.M.; Lu, T.J. A new pinane-type tridentate modifier for asymmetric reduction of ketones with lithium aluminum hydride. Tetrahedron Asymmetry 1995, 6, 89–92. [Google Scholar] [CrossRef]
  36. Szakonyi, Z.; Hetényi, A.; Fulop, F. Synthesis of enantiomeric spirooxazolines and spirooxazolidines by the regioselective ring closure of (–)-α-pinene-based aminodiols. Arkivoc 2007, 2008, 33. [Google Scholar]
  37. Szakonyi, Z.; Csillag, K.; Fülöp, F. Stereoselective synthesis of carane-based aminodiols as chiral ligands for the catalytic addition of diethylzinc to aldehydes. Tetrahedron Asymmetry 2011, 22, 1021–1027. [Google Scholar] [CrossRef]
  38. Szakonyi, Z.; Csőr, Á.; Csámpai, A.; Fülöp, F. Stereoselective synthesis and modelling-driven optimisation of Carane-based aminodiols and 1,3-oxazines as catalysts for the enantioselective addition of diethylzinc to benzaldehyde. Chem. Eur. J. 2016, 22, 7163–7173. [Google Scholar] [CrossRef]
  39. Lee, D.-S.; Chang, S.-M.; Ho, C.-Y.; Lu, T.-J. Enantioselective addition of diethylzinc to aldehydes catalyzed by chiral O, N, O-tridentate phenol ligands derived from Camphor. Chirality 2016, 28, 65–71. [Google Scholar] [CrossRef]
  40. Tashenov, Y.; Daniels, M.; Robeyns, K.; Van Meervelt, L.; Dehaen, W.; Suleimen, Y.; Szakonyi, Z. Stereoselective syntheses and application of chiral bi- and tridentate ligands derived from (+)-Sabinol. Molecules 2018, 23, 771. [Google Scholar] [CrossRef] [Green Version]
  41. Csillag, K.; Németh, L.; Martinek, T.A.; Szakonyi, Z.; Fülöp, F. Stereoselective synthesis of pinane-type tridentate aminodiols and their application in the enantioselective addition of diethylzinc to benzaldehyde. Tetrahedron Asymmetry 2012, 23, 144–150. [Google Scholar] [CrossRef]
  42. Gonda, T.; Szakonyi, Z.; Csámpai, A.; Haukka, M.; Fülöp, F. Stereoselective synthesis and application of tridentate aminodiols derived from (+)-pulegone. Tetrahedron Asymmetry 2016, 27, 480–486. [Google Scholar] [CrossRef]
  43. Le, T.M.; Szilasi, T.; Volford, B.; Szekeres, A.; Fülöp, F.; Szakonyi, Z. Stereoselective synthesis and investigation of Isopulegol-based chiral ligands. Int. J. Mol. Sci. 2019, 20, 4050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Outouch, R.; Oubaassine, S.; Ait Ali, M.; El Firdoussi, L.; Spannenberg, A. Crystal structure of (1S,2R,4S)-1-[(morpholin-4-yl)methyl]-4-(prop-1-en-2-yl)cyclohexane-1,2-diol. Acta Crystallogr. Sect. E Crystallogr. Commun. 2015, 71, 79–81. [Google Scholar] [CrossRef] [PubMed]
  45. Rosenberg, S.H.; Spina, K.P.; Woods, K.W.; Polakowski, J.; Martin, D.L.; Yao, Z.; Stein, H.H.; Cohen, J.; Barlow, J.L. Studies directed toward the design of orally active renin inhibitors. 1. Some factors influencing the absorption of small peptides. J. Med. Chem. 1993, 36, 449–459. [Google Scholar] [CrossRef] [PubMed]
  46. Beaulieu, P.L.; Gillard, J.; Bailey, M.; Beaulieu, C.; Duceppe, J.-S.; Lavallée, P.; Wernic, D. Practical synthesis of BILA 2157 BS, a potent and orally active renin inhibitor: Use of an enzyme-catalyzed hydrolysis for the preparation of homochiral succinic acid derivatives. J. Org. Chem. 1999, 64, 6622–6634. [Google Scholar] [CrossRef]
  47. Kawai, A.; Hara, O.; Hamada, Y.; Shioiri, T. Stereoselective synthesis of the hydroxy amino acid moiety of Al-77-B, a gastroprotective substance from Bacillus pumilus Al-77. Tetrahedron Lett. 1988, 29, 6331–6334. [Google Scholar] [CrossRef]
  48. Wang, G.T.; Li, S.; Wideburg, N.; Krafft, G.A.; Kempf, D.J. Synthetic chemical diversity: Solid phase synthesis of libraries of C2 symmetric inhibitors of HIV protease containing diamino diol and diamino alcohol cores. J. Med. Chem. 1995, 38, 2995–3002. [Google Scholar] [CrossRef]
  49. Grajewska, A.; Rozwadowska, M.D. Stereoselective synthesis of cytoxazone and its analogues. Tetrahedron Asymmetry 2007, 18, 803–813. [Google Scholar] [CrossRef]
  50. Mishra, R.K.; Coates, C.M.; Revell, K.D.; Turos, E. Synthesis of 2-Oxazolidinones from β-Lactams: Stereospecific Total Synthesis of (−)-Cytoxazone and All of Its Stereoisomers. Org. Lett. 2007, 9, 575–578. [Google Scholar] [CrossRef]
  51. Allepuz, A.C.; Badorrey, R.; Díaz-de-Villegas, M.D.; Gálvez, J.A. Diastereoselective reduction of ketimines derived from (R)-3,4-dihydroxybutan-2-one: An alternative route to key intermediates for the synthesis of anticancer agent ES-285. Tetrahedron Asymmetry 2010, 21, 503–506. [Google Scholar] [CrossRef]
  52. Malkov, A.V.; Baxendale, I.R.; Bella, M.; Langer, V.; Fawcett, J.; Russell, D.R.; Mansfield, D.J.; Valko, M.; Kočovský, P. Synthesis of new chiral 2,2′-bipyridyl-type ligands, their coordination to Molybdenum(0), Copper(II), and Palladium(II), and application in asymmetric allylic Substitution, allylic oxidation, and cyclopropanation. Organometallics 2001, 20, 673–690. [Google Scholar] [CrossRef]
  53. Chelucci, G.; Loriga, G.; Murineddu, G.; Pinna, G.A. Synthesis of chiral C2-symmetric 1,10-Phenanthrolines from naturally occurring monoterpenes. Synthesis 2003, 0073–0078. [Google Scholar] [CrossRef]
  54. Badjah-Hadj-Ahmed, A.Y.; Meklati, B.Y.; Waton, H.; Pham, Q.T. Structural studies in the bicyclo[3.1.1]heptane series by 1H and 13C NMR. Magn. Reson. Chem. 1992, 30, 807–816. [Google Scholar] [CrossRef]
  55. Molander, G.A.; Cavalcanti, L.N. Oxidation of Organotrifluoroborates via Oxone. J. Org. Chem. 2011, 76, 623–630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Gemal, A.L.; Luche, J.L. Lanthanoids in organic synthesis. 6. Reduction of alpha.-enones by sodium borohydride in the presence of lanthanoid chlorides: Synthetic and mechanistic aspects. J. Am. Chem. Soc. 1981, 103, 5454–5459. [Google Scholar] [CrossRef]
  57. Friedrich, D.; Bohlmann, F. Total synthesis of various elemanolides. Tetrahedron 1988, 44, 1369–1392. [Google Scholar] [CrossRef]
  58. Rodríguez-Berríos, R.R.; Torres, G.; Prieto, J.A. Stereoselective VO(acac)2 catalyzed epoxidation of acyclic homoallylic diols. Complementary preparation of C2-syn-3,4-epoxy alcohols. Tetrahedron 2011, 67, 830–836. [Google Scholar] [CrossRef] [Green Version]
  59. Shivani Pujala, B.; Chakraborti, A.K. Zinc(II) perchlorate hexahydrate Catalyzed opening of epoxide ring by amines: Applications to synthesis of (RS)/(R)-Propranolols and (RS)/(R)/(S)-Naftopidils. J. Org. Chem. 2007, 72, 3713–3722. [Google Scholar] [CrossRef]
  60. Szakonyi, Z.; Hetényi, A.; Fülöp, F. Synthesis and application of monoterpene-based chiral aminodiols. Tetrahedron 2008, 64, 1034–1039. [Google Scholar] [CrossRef]
  61. Jia, Y.X.; Wu, B.; Li, X.; Ren, S.K.; Tu, Y.Q.; Chan, A.S.C.; Kitching, W. Synthetic studies of the HIV-1 protease inhibitive didemnaketals: Stereocontrolled synthetic approach to the key mother spiroketals. Org. Lett. 2001, 3, 847–849. [Google Scholar] [CrossRef]
  62. Kim, J.H.; Lim, H.J.; Cheon, S.H. Synthesis of (+)-hernandulcin and (+)-epihernandulcin. Tetrahedron Lett. 2002, 43, 4721–4722. [Google Scholar] [CrossRef]
  63. Waddell, T.G.; Ross, P.A. Chemistry of 3,4-epoxy alcohols. Fragmentation reactions. J. Org. Chem. 1987, 52, 4802–4804. [Google Scholar] [CrossRef]
  64. Kim, J.H.; Lim, H.J.; Cheon, S.H. A facile synthesis of (6S,1′S)-(+)-hernandulcin and (6S,1′R)-(+)-epihernandulcin. Tetrahedron 2003, 59, 7501–7507. [Google Scholar] [CrossRef]
  65. Tanaka, T.; Yasuda, Y.; Hayashi, M. New chiral Schiff base as a tridentate ligand for catalytic enantioselective addition of diethylzinc to aldehydes. J. Org. Chem. 2006, 71, 7091–7093. [Google Scholar] [CrossRef] [PubMed]
  66. Jimeno, C.; Pastó, M.; Riera, A.; Pericàs, M.A. Modular amino alcohol ligands containing bulky alkyl groups as chiral controllers for Et2Zn addition to aldehydes: Illustration of a design principle. J. Org. Chem. 2003, 68, 3130–3138. [Google Scholar] [CrossRef] [PubMed]
  67. Roothaan, C.C.J. Self-consistent field theory for open shells of electronic systems. Rev. Mod. Phys. 1960, 32, 179–185. [Google Scholar] [CrossRef]
  68. Hay, P.J.; Wadt, W.R. Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270–283. [Google Scholar] [CrossRef]
  69. Peng, C.; Ayala, P.Y.; Schlegel, H.B.; Frisch, M.J. Using redundant internal coordinates to optimize equilibrium geometries and transition states. J. Comput. Chem. 1996, 17, 49–56. [Google Scholar] [CrossRef]
  70. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09 (Revision A.02); Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
Scheme 1. Preparation of (−)-3-methylenenopinone.
Scheme 1. Preparation of (−)-3-methylenenopinone.
Catalysts 10 00474 sch001
Scheme 2. (i) NaBH4 (2 equ.) and CeCl3.7H2O (1 equ.), 0 °C, 0.5 h, 87%.
Scheme 2. (i) NaBH4 (2 equ.) and CeCl3.7H2O (1 equ.), 0 °C, 0.5 h, 87%.
Catalysts 10 00474 sch002
Scheme 3. (i) VO(acac)2, 70% t-BuOOH (2 equ.), dry toluene, 25 °C, 12 h, 76% and (ii) R1R2NH (2 equ.), LiClO4 (1 equ.), MeCN, 70–80 °C, 6 h, 32%–94%.
Scheme 3. (i) VO(acac)2, 70% t-BuOOH (2 equ.), dry toluene, 25 °C, 12 h, 76% and (ii) R1R2NH (2 equ.), LiClO4 (1 equ.), MeCN, 70–80 °C, 6 h, 32%–94%.
Catalysts 10 00474 sch003
Scheme 4. (i) From 6 (R = Bn), 5% Pd/C, H2 (1 atm), MeOH, 25 °C, 24 h, 74% and (ii) 35% HCHO, Et2O, 25 °C, 1 h, 98% (15, R = Bn), 40% (16, R = CH(Me)2).
Scheme 4. (i) From 6 (R = Bn), 5% Pd/C, H2 (1 atm), MeOH, 25 °C, 24 h, 74% and (ii) 35% HCHO, Et2O, 25 °C, 1 h, 98% (15, R = Bn), 40% (16, R = CH(Me)2).
Catalysts 10 00474 sch004
Scheme 5. (i) 60% NaH (1 equ.), dry THF, 25 °C, 1 h; then BnBr (1.5 equ.), KI (1 equ.), 60–70 °C, 12 h, 62%; (ii) mCPBA (2 equ.), Na2HPO4. 2H2O (3 equ.), 25 °C, 2 h, 46% and (iii) RNH2 (2 equ.), LiClO4 (1 equ.), MeCN, 70–80 °C, 6 h, 30%–36%.
Scheme 5. (i) 60% NaH (1 equ.), dry THF, 25 °C, 1 h; then BnBr (1.5 equ.), KI (1 equ.), 60–70 °C, 12 h, 62%; (ii) mCPBA (2 equ.), Na2HPO4. 2H2O (3 equ.), 25 °C, 2 h, 46% and (iii) RNH2 (2 equ.), LiClO4 (1 equ.), MeCN, 70–80 °C, 6 h, 30%–36%.
Catalysts 10 00474 sch005
Figure 1. Determination of the structure of aminodiols by NOESY.
Figure 1. Determination of the structure of aminodiols by NOESY.
Catalysts 10 00474 g001
Scheme 6. (i) 35% HCHO, Et2O, 25 °C, 1 h, 50%.
Scheme 6. (i) 35% HCHO, Et2O, 25 °C, 1 h, 50%.
Catalysts 10 00474 sch006
Scheme 7. Model reaction for enantioselective catalysis.
Scheme 7. Model reaction for enantioselective catalysis.
Catalysts 10 00474 sch007
Figure 2. Proposed transition state for the asymmetric addition with 21.
Figure 2. Proposed transition state for the asymmetric addition with 21.
Catalysts 10 00474 g002
Scheme 8. Model reaction for enantioselective catalysis.
Scheme 8. Model reaction for enantioselective catalysis.
Catalysts 10 00474 sch008
Figure 3. Reaction coordinate with activation barrier and thermodynamics of the ethyl transfer connecting complexes 27A and 27C via transition state 27B optimized by the HF/LANL2DZ method.
Figure 3. Reaction coordinate with activation barrier and thermodynamics of the ethyl transfer connecting complexes 27A and 27C via transition state 27B optimized by the HF/LANL2DZ method.
Catalysts 10 00474 g003
Table 1. Stereoselective reduction of 3 according to Scheme 1.
Table 1. Stereoselective reduction of 3 according to Scheme 1.
EntryReductantAdditiveSolventT (°C)t (h)Ratio 4a/4b [a]Yield [b] (%)
1NaBH4-MeOH−2063:186
2NaBH4-MeOH011:186
3NaBH4CeCl3MeOH00.5100:087
5NaBH4-Et2O033:176
4NaBH4-EtOH0481:384
[a] Based on 1H-NMR measurements of the crude product. [b] Isolated, combined yield of 4a and 4b.
Table 2. Addition of diethylzinc to benzaldehyde, catalyzed by aminodiol derivatives.
Table 2. Addition of diethylzinc to benzaldehyde, catalyzed by aminodiol derivatives.
EntryLigand aYield b (%)eec (%)Configuration d
16835(R)
279223(R)
388031(R)
49854(R)
5108580(R)
611903(R)
7129516(R)
8138060(R)
91483--
10158326(R)
11168535(R)
12199573(R)
13208774(S)
14218380(R)
15228040(R)
16239310(R)
[a] 10 mol%. [b] Are given after silica column chromatography. [c] Determined by measuring the ee of the crude product by GC (Chirasil-DEX CB column). [d] Determined by comparing optical rotations and the tR of GC analysis with the literature data [65,66].
Table 3. Addition of diethylzinc to aldehydes, catalyzed by 10 mol % 20 or 21.
Table 3. Addition of diethylzinc to aldehydes, catalyzed by 10 mol % 20 or 21.
EntryCatalystProductsRYield a (%)eeb (%)Configuration c
12029a(4-MeO)C6H48092(S)
22029b(3-MeO)C6H47884(S)
32029c(3-Me)C6H47578(S)
42129a(4-MeO)C6H48385(R)
52129b(3-MeO)C6H49287(R)
62129c(3-Me)C6H48084(R)
72129dcyclohexyl8548(R)
82129en-butyl8045(R)
[a] Are given after silica column chromatography. [b] Determined on the crude product by HPLC (Chiracel OD-H) or GC (Chirasil-DEX CB column). [c] Determined by comparing the tR of the HPLC analysis and the optical rotation with the literature data [37].

Share and Cite

MDPI and ACS Style

Raji, M.; Le, T.M.; Fülöp, F.; Szakonyi, Z. Synthesis and Investigation of Pinane-Based Chiral Tridentate Ligands in the Asymmetric Addition of Diethylzinc to Aldehydes. Catalysts 2020, 10, 474. https://doi.org/10.3390/catal10050474

AMA Style

Raji M, Le TM, Fülöp F, Szakonyi Z. Synthesis and Investigation of Pinane-Based Chiral Tridentate Ligands in the Asymmetric Addition of Diethylzinc to Aldehydes. Catalysts. 2020; 10(5):474. https://doi.org/10.3390/catal10050474

Chicago/Turabian Style

Raji, Mounir, Tam Minh Le, Ferenc Fülöp, and Zsolt Szakonyi. 2020. "Synthesis and Investigation of Pinane-Based Chiral Tridentate Ligands in the Asymmetric Addition of Diethylzinc to Aldehydes" Catalysts 10, no. 5: 474. https://doi.org/10.3390/catal10050474

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