Next Article in Journal
Phenolics as GABAA Receptor Ligands: An Updated Review
Previous Article in Journal
Influence of Gamma Irradiation on Different Phytochemical Constituents of Dried Rose Hip (Rosa canina L.) Fruits
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Aziridine Ring Opening as Regio- and Stereoselective Access to C-Glycosyl-Aminoethyl Sulfide Derivatives

Department of Organic and Applied Chemistry, University of Łódź, 91-403 Łódź, Poland
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(6), 1764; https://doi.org/10.3390/molecules27061764
Submission received: 14 February 2022 / Revised: 1 March 2022 / Accepted: 6 March 2022 / Published: 8 March 2022

Abstract

:
A short synthetic route to stereoselective access to C-glycosyl-aminoethyl sulfide derivatives has been developed through the reaction of tributhyltin derivatives of glycals with aziridinecarboaldehyde and the regioselective ring opening of a chiral aziridine with thiophenol. The absolute configurations of the resulting diastereoisomers were determined by 1H NMR spectroscopy.

1. Introduction

Intensively developed in recent years, asymmetric synthesis has proved to be a powerful tool in the synthesis of drugs and natural products as well as in the transformation of readily available simple compounds into chiral building blocks used for the synthesis of more complex connections [1,2,3]. Although sugars are the most readily available raw materials, they were long considered of little use due to the presence of polar functional groups. However, it turns out th–t the structure is in fact a great advantage, enabling wide-ranging possibilities of modification, making them very useful synthetic tools. Research conducted at the borderline of chemistry, biology and medicine indicates an urgent need for the synthesis of natural and non-natural saccharides and glycoconjugates of well-defined structure and composition. Due to their participation in many important biochemical processes, the increasing interest in them is justified. These compounds can serve as probes in research aimed at elucidating complex functions that carbohydrates play at the molecular level in living organisms, and additionally, can be used in the synthesis of new drugs based on carbohydrates [4,5,6,7,8,9,10,11,12,13,14]. An extremely important element of such building blocks is the glycosidic bond. O- and N-glycosidic bonds occur in nature, however, studies confirm their insufficient resistance to chemical and enzymatic hydrolysis. Replacing these bonds with C-C bonds has proved to be a very good solution. This modification increases the resistance required in therapeutic conditions while maintaining biological activity and good tolerance by living organisms [6,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]. An interesting and original idea is the coupling of sugar and aziridine, which leads to a completely new group of C-glycosides with promising biological properties. Chiral aziridines are useful intermediates in the synthesis of biologically significant compounds due to their ability to undergo nucleophilic ring opening reactions [33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50]. Considering our previous results on carbohydrate chemistry [51,52,53,54,55,56,57,58] and based on our experience in synthesis and catalytic activity in the asymmetric synthesis of chiral aziridines [59,60,61,62,63,64,65], we decided to couple aziridines to glycals with D-gluco and D-galacto configurations via C-glycosidic bonding, with the final formation of C-glycosyl-aminoethyl sulfide derivatives.

2. Results and Discussion

Aminoethyl sulfide derivatives and, in particular, phenylaminoethylsulfides (PAES) have numerous applications due to their interesting biological properties. Derivatives of this type are synthetic substrates for dopamine β-hydroxylase (DBH; EC 1.14.17.1) [66,67]. As shown by May, PAES has indirect sympathomimetic activity in vivo and inhibits reflex tachycardia induced by vasodilating antihypertensive drugs. PAES, or structurally similar derivatives, may therefore be useful in the control of hypertension (Figure 1a) [68,69]. In addition, compounds containing an aminoethyl sulfide moiety have the ability to inhibit adenosine deaminase (ADA) (Figure 1b) [70], show β-adrenoreceptor blocking properties (Figure 1c) [71], and are used as inhibitors of DNA methyltransferases (Figure 1d) [72].
The new C-glycosyl-aminoethyl sulfide derivatives 1720 with D-gluco and D-galacto configurations were easily prepared, according to Scheme 1.
Commercially available tri-O-acetyl-D-glucal and tri-O-acetyl-D-galactal were deacetylated in the presence of sodium methanolate in methanol and produced D-glucal (1) and D-galactal (2) in quantitative yields [73]. In the next step, saccharides 1 and 2 were treated with triisopropylsilyl chloride in the presence of imidazole in DMF to obtain the O-silylated derivatives 4 and 5 [74]. However, under these conditions, the 4-hydroxyl group of the D-galacto derivative 2 remained unprotected. Protection of the free hydroxyl group required the use of triisopropylsilyl triflate and 2,6-lutidine (Scheme 1) [75]. Then, the protected derivatives 4 and 5 were subjected to 1-deprotonation with tert-butyllithium and a reaction with tributyltin chloride [74]. The cases of 1-deprotonation of tribenzyl- and tris-(tertbutyldimethyl)-derivatives of D-glucal were also reported by several research groups, but they caused a number of problems and were characterized by low yields (10–30%) [76,77,78,79,80]. For this reason, tin derivatives 6 and 7 were prepared by deprotonation of the tris(triisopropylsilyl) glycals derivatives 4 and 5 with tert-butyllithium and quenching with tributyltin chloride (85% and 82% yields, respectively) [74]. A key step in the synthesis of C-glycosides was the coupling reaction of tin derivatives 6 and 7 with (S)-1-triphenylmethylaziridine-2-carbaldehyde (8), obtained according to literature procedures [81]. On the basis of the available databases, we can indicate only one example of the preparation of C-glycosides in the reaction of a tin derivative of D-glucal with an aldehyde obtained from diacetone-D-glucose. As reported by Whiting, the reaction resulted in isomeric C-disaccharides in a 2.1:1 ratio with a 31% yield from the stannane [74]. The reaction conducted under similar conditions, in the presence of n-BuLi in THF and at −78 °C for 2 h (Procedure A, Scheme 1) gave the desired C-glycoside analogs 912 in satisfactory yields. It seemed interesting to investigate both the stereochemistry of the addition of the organometallic compound to the aldehyde group of optically pure aziridine 8, as well as to determine the influence of the metal cation on the stereochemistry of the reaction. Therefore, subsequent experiments were carried out in the presence of magnesium cation (Procedure B, Scheme 1).
Selectivity, in the addition reaction of organometallic reagents to aziridine 2-carboxyaldehydes, was described ten years ago by Jackson and Borhan [82]. In such addition reactions, new stereogenic centers are formed, leading to the formation of syn- or anti-adducts depending on the kind of metal. Those that are strong coordinators favor syn selectivity, which can be rationalized by a chelatation-based transition state, whereas metals that coordinate poorly, or conditions that suppress chelatation, favor the anti-adducts predicted by the Felkin–Anh model.
Initially, we performed the addition of the tin derivative of D-glucal 6 to aldehyde 8 obtaining a mixture of diastereoisomeric C-glycosides in the ratio erythro-9:threo-10 = 1:3 with a total yield of 40% (Table 1, entry 1).
The introduction of magnesium cation into the reaction medium (Table 1, entry 2) resulted in an increase of yield to 55% and a simultaneous decrease in stereoselectivity (erythro-9:threo-10 = 4:5). Importantly, diastereoisomeric products 9 and 10 were successfully purified and separated by flash chromatography, and all subsequent modifications were carried out on pure stereoisomers. Another experiment was carried out with the tin derivative of D-galactal 7 (Table 1, entry 3). The carbon–carbon bond formation process under these conditions occurred with a slightly higher yield (45%) and excellent stereoselectivity, compared to the analogous D-glucal derivative reaction. The same reaction performed in the presence of a magnesium cation resulted in C-glycosides 11 and 12 in an identical erythro/threo ratio of 1:9 (Table 1, entry 4) but with a much higher yield (65% total yield).
Unfortunately, the separation of diastereoisomeric products 11 and 12 was unsuccessful, so subsequent transformations were performed on their mixture.
The above results indicate that chelation is a minor factor in stereoselectivity. For galactal derivatives, the reactions take place mainly according to the Felkin–Ahn model (1:9). However, for glucal, stronger coordination is manifested by an increase in the share of the chelating model, but it does not exceed 50%. At the present stage of research, we do not have adequate experimental material that would allow for any generalizations.
The absolute configurations of the resulting diastereoisomers 912 were determined by 1H NMR spectroscopy. According to literature data, the assignment of the absolute stereochemistry of the products is made by measuring the coupling constant between two neighboring methine protons, or the chemical shift of the methine proton on the hydroxyl-bearing carbon [83,84,85,86]. As reported by Lee [83], for the derivatives with the structure shown in the figure below (Figure 2), the coupling constant of the methine protons of the threo isomers (anti orientation of both protons) was always larger (J = 4.4–6.0 Hz) than those of erythro isomers (J = 1.9–3.7 Hz) in which protons are syn oriented. Moreover, the chemical shifts of the methine proton on the hydroxyl-bearing carbon of the threo isomers were always in higher field values (for example, 4.23 ppm for R1 = CH(Me)Ph) than those of the erythro isomers (4.66 ppm for R1 = CH(Me)Ph).
Based on the above reports, the absolute configuration of alcohols 912 was determined. Like the structures described by Lee, they possess the (S) configuration of the asymmetric carbon atom in the aziridine ring, which does not change during the reaction with the tin derivative of D-glucal 6 and D-galactal 7. The recorded spectral data for D-glucal derivatives 9 and 10 indicate that diastereoisomer 9 of lower polarity (Rf = 0.45; petroleum ether:diethyl ether = 15:1) has the erythro-(S,S) configuration [4.40 (d, J = 2.2, CHOH)], while the second diastereoisomer 10 of higher polarity (Rf = 0.34; petroleum ether:diethyl ether = 15:1) is a threo-(R,S) isomer [3.92 ppm (d, J = 5.8, CHOH)] (Table 2). The stereochemistry of the products of the reaction of the tin derivative of D-galactal 7 with aldehyde 8 was determined in a similar manner. The methine proton signal at the carbon-containing hydroxyl group for the threo stereoisomer (R,S)-12 occurred at lower ppm values (3.91 ppm) and the coupling constant of 5.8 Hz.
In contrast, the erythro isomer (S,S)-11 gave a proton signal of the CHOH moiety, similar to the erythro isomer of glucal derivative 9, at 4.33 ppm in the form of a doublet with the small coupling constant J = 2.4 Hz (Table 2).
In the next step, deprotection of hydroxyl groups was performed for pure diastereoisomers 9 and 10 with the D-gluko configuration and for the mixture of erythro and threo derivatives of D-galactal (11:12 = 1:9). The reactions carried out in THF at room temperature in the presence of tetrabutylammonium fluoride [87] afforded C-glycosides 1316 with free hydroxyl groups in the saccharide ring with yields of 13:95%, 14:95%, 15 and 16:96%. The absolute configurations of the obtained C-glycosyl derivatives confirm the recorded spectral data. Although the coupling constants between two neighboring methine protons could not be determined (the signals of protons on the hydroxyl-bearing carbon appeared as broadened singlets), the chemical shift of the methine proton of CHOH moiety confirms the assigned configurations. The signals of the threo isomers 14 and 16 were in a higher field than those of the erythro isomers 13 and 15 (Table 2). The last step in the planned sequence of transformations (Scheme 1) was the reaction nucleophilic ring opening of aziridine with thiophenol. As is already known, the thiol group opens the aziridine ring particularly easily, and importantly, this reaction is fully regioselective, and the attack occurs on the less substituted carbon in aziridine [35,36]. All ring opening reactions were carried out in methylene chloride at room temperature, using three times the excess of thiophenol over the starting C-glycosides 1316 (Scheme 1) [88,89]. Final products 1720 were obtained with yields of 67% for 17, 68% for 18, 72% for 19 and 20, respectively.

3. Materials and Methods

Commercially available chemicals used in this work were purchased from Sigma-Aldrich (Darmstadt, Germany) and were used as supplied, without additional purification. NMR spectra were recorded in CDCl3 on a Bruker Avance III (600 MHz for 1H NMR, 150 MHz for 13C NMR) (Billerica, MA, USA); coupling constants are reported in hertz (Hz). The rotations were measured using an Anton Paar MCP 500 polarimeter (Anton Paar GmbH, Graz, Austria). Melting points are uncorrected. Chromatographic purification of compounds was achieved with 230−400 mesh size silica gel. The progress of reactions was monitored by silica gel thin-layer chromatography plates (Merck TLC Silicagel 60 F254) (Merck Millipore, Darmstadt, Germany).
Copies of 1H and 13C NMR spectra of compounds 920 are included in the Supplementary Material.

3.1. General Procedure for the Synthesis of Glycals

A catalytic amount of MeONa (0.03 g, 0.57 mmol) was added to a solution of tri-O-acetyl-D-glycal (2.5 g, 9.18 mmol) in methanol (25 mL) and the resulting reaction mixture was stirred at rt. The progress of reactions was monitored by silica gel thin-layer chromatography plates. After 30 min, the solution was filtered on a Schott funnel over a resin layer (Amberlite® IR120) and celite. Evaporation of the organic solvent afforded a pure product.

3.1.1. D-Glucal (1)

Colorless solid, 1.67 g, 99% yield; Rf = 0.06 (hexane/ethyl acetate, 7:3); [α]D20 = −7.6 (c 0.9, CHCl3), {Lit. [90]: [α]D20 = −8.0 (c 1.19, H2O)}; m.p. = 54–56 °C, {Lit. 90]: m.p. = 58–60 °C}; δH (600 MHz, D2O): 3.61 (dd, 1H, J = 9.0, 7.1, H-4), 3.72–3.86 (m, 3H, H-5, 2H-6), 4.17 (dt, 1H, J = 7.1, 2.0, H-3), 4.73 (dd, 1H, J = 6.0, 2.0, H-2), 6.35 (dd, 1H, J = 6.0, 1.4, H-1).
1H NMR spectral data matched that reported by Crotti [91].

3.1.2. D-Galactal (2)

Yellow oil, 1.70 g, 99% yield; Rf = 0.06 (hexane/ethyl acetate, 7:3); m.p. = 90–52 °C, {Lit. [92]: m.p. = 89–91 °C}; δH (600 MHz, CDCl3): 3.72 (d, 1H, J = 5.9, H-4), 3.77 (dd, 1H, J = 11.6, 5.2, H-6), 3.77–3.80 (m, 1H, H-5), 3.84 (dd, 1H, J = 11.6, 5.9, H-6), 3.90–3.95 (m, 3H, 3OH), 4.34–4.38 (m, 1H, H-3), 4.64 (dd, 1H, J = 6.2, 2.2, H-2), 6.36 (dd, 1H, J = 6.2, 1.6, H-1).
Spectroscopic data are in accordance with Refs. [75,93].

3.1.3. 3,6-Di-O-(triisopropylsilyl)-D-galactal (3)

Colorless oil, 3.95 g, 93% yield; Rf = 0.79 (hexane/ethyl acetate, 20:1); [α]D20 = −33.81 (c 1.0, CHCl3), {Lit. [94]: [α]D20 = −34.0 (c 1.28, CHCl3)}; δH (600 MHz, CDCl3): 1.06–1.10 (m, 42H, TIPS-H), 1.55 (s, 1H, OH), 3.87 (dd, 1H, J = 7.2, 5.7, H-5), 3.92 (dd, 1H, J = 9.7, 5.9, H-6), 3.99–4.02 (m, 1H, H-4), 4.04 (dd, 1H, J = 9.7, 7.4, H-6), 4.56–4.61 (m, 2H, H-2, H-3), 6.34 (d, 1H, J = 4.9, H-1).
The synthesis and spectroscopic data are in accordance with Ref. [93].

3.1.4. 3,4,6-Tris-O-(triisopropylsilyl)-D-glucal (4)

Colorless oil, 6.05 g, 75% yield; Rf = 0.95 (hexane/ethyl acetate, 25:1); [α]D20 = −17.77 (c 0.6, CHCl3), {Lit. [95]: [α]D20 = −21.4 (c 1.0, CHCl3)}; IR (film): 3066 (ν=C-H), 2942, 2867 (νC-H), 1645 (νC=C), 1068 (νC-O); δH (600 MHz, CDCl3): 1.06 (s, 63H, TIPS-H), 3.82 (dd, 1H, J = 11.3, 3.8, H-6), 3.95 (dt, 1H, J = 5.2, 1.9, H-3), 4.04–4.08 (m, 2H, H-4, H-6), 4.22–4.25 (m, 1H, H-5), 4.80 (ddd, 1H, J = 6.6, 5.3, 1.7, H-2), 6.35 (d, 1H, J = 6.4, H-1); δC (150 MHz, CDCl3): 12.2, 12.5, 12.7 (CH), 18.1, 18.2, 18.3 (CH3), 62.3 (C-6), 65.3 (C-3), 70.5 (C-4), 80.9 (C-5), 100.5 (C-2), 143.1 (C-1).
The synthesis and spectroscopic data are in accordance with Refs. [74,95].

3.1.5. 3,4,6-Tri-O-(triisopropylsilyl)-D-galactal (5)

Colorless oil, 1.5 g, 75% yield; Rf = 0.95 (hexane/ethyl acetate, 20:1); [α]D20 = −17.07 (c 0.6, CHCl3); IR (film): 3064, 3018 (ν=C-H), 2944, 2867 (νC-H), 1641 (νC=C), 1087 (νC-O); δH (600 MHz, CDCl3): 1.04–1.17 (m, 63H, TIPS-H), 4.02–4.40 (m, 5H, H-3, H-4, H-5, 2H-6), 4.80 (bs, 1H, H-2), 6.24 (d, 1H, J = 6.1, H-1); δC (150 MHz, CDCl3): 11.9, 12.0, 12.2 (CH), 18.0, 18.1, 18.2, 18.4 (CH3), 62.8 (C-6), 68.1 (C-3), 74.2 (C-4), 86.2 (C-5), 98.1 (C-2), 142.9 (C-1).
The synthesis and spectroscopic data are in accordance with Refs. [93,95].

3.2. General Procedure for the Synthesis of Tributhyltin Derivatives of Glycals

The tributhyltin derivatives were synthesized according to a literature procedure [11]. Where 3,4,6-Tris-O-(triisopropylsilyl)-D-glycal 4 or 5 (1.0 g, 1.62 mmol) was dissolved in dry THF (4 mL) and the solution was stirred under nitrogen, cooled to −78 °C, and treated with t-BuLi (1.7 mol·dm−3, 3.82 mL, 6.5 mmol) in one addition. The solution was warmed to 0 °C and stirred for 1.5 h, then was cooled to −78 °C, and tributyltin chloride (1.1 mL, 4.1 mmol) was added. The solution was again warmed to 0 °C and stirred for 45 min before the reaction was quenched with water (10 mL). The solution was poured into diethyl ether (10 mL) and the organic phase was separated. The aqueous phase was extracted with diethyl ether and the combined extracts were washed successively with water (10 mL) and brine (10 mL), dried over anhydrous MgSO4 and evaporated. The crude product was purified by column chromatography (hexane/ethyl acetate, 25:1).

3.2.1. 1-(Tributylstannyl)-3,4,6-tris-O-(triisopropylsilyl)-D-glucal (6)

Colorless oil, 0.98 g, 85% yield, Rf = 0.94 (hexane/ethyl acetate, 25:1); [α]D20 = −23.87 (c 0.5, CHCl3); IR (film): 2942, 2867 (νC-H), 1654 (νC=C), 1068 (νC-O); δH (600 MHz, CDCl3): 0.86–0.94 (m, 15H, n-Bu), 1.06 (s, 63H, TIPS-H), 1.27–1.34 (m, 6H, n-Bu), 1.48–1.55 (m, 6H, n-Bu), 3.85 (dt, 1H, J = 5.0, 2.6, H-3), 3.91 (dd, 1H, J = 11.3, 5.0, H-6), 3.96 (dd, 1H, J = 11.3, 7.0, H-6), 4.06–4.08 (m, 1H, H-4), 4.08–4.12 (m, 1H, H-5), 4.83 (dd, 1H, J = 5.2, 1.6, H-2); δC (150 MHz, CDCl3): 9.7 (CH2CH2CH2CH3), 12.3, 12.7, 12.8 (CH(CH3)2), 13.8 (CH3CH2CH2CH2), 18.2, 18.3, 18.4 (CH(CH3)2), 27.5 (CH3CH2CH2CH2), 29.1 (CH3CH2CH2CH2), 62.6 (C-6), 65.3 (C-3), 70.5 (C-4), 80.8 (C-5), 111.5 (C-2), 162.6 (C-1).
1H and 13C NMR spectral data matched that reported by Whiting [74].

3.2.2. 1-(Tributylstannyl)-3,4,6-tris-O-(triisopropylsilyl)-D-galactal (7)

Colorless oil, 0.95 g, 82% yield; Rf = 0.95 (hexane/ethyl acetate, 25:1); [α]D20 = −27.82 (c 0.5, CHCl3); IR (film): 2942, 2867 (νC-H), 1641 (νC=C), 1072 (νC-O); δH (600 MHz, CDCl3): 0.89–0.93 (m, 15H, n-Bu), 1.05–1.13 (m, 63H, TIPS-H), 1.30–1.36 (m, 6H, n-Bu), 1.50–1.56 (m, 6H, n-Bu), 4.06–4.26 (m, 5H, H-3, H-4, H-5, 2H-6), 4.84 (bs, 1H, H-2); δC (150 MHz, CDCl3): 9.7 (CH2CH2CH2CH3), 12.2, 12.8 (CH(CH3)2), 13.8 (CH3CH2CH2CH2), 18.2, 18.5 (CH(CH3)2), 27.5 (CH3CH2CH2CH2), 29.1 (CH3CH2CH2CH2), 61.4 (C-6), 64.7 (C-3), 70.6 (C-4), 81.1 (C-5), 112.9 (C-2), 162.6 (C-1); elementar analysis: C45H96O4Si3Sn (904.56 g/mol) calculated: C% 59.77, H% 10.70; found: C% 59.94, H% 10.86.

3.3. General Procedure for the Reaction of Derivatives of Glycals with Aziridine Aldehyde 8

Procedure A: 0.25 g (0.28 mmol) of tributhyltin derivative 6 or 7 was dissolved in 1.3 mL of dry THF and the solution was stirred under argon, cooled to −78 °C, and treated with n-BuLi (0.25 mL, 0.33 mmol) added dropwise. The solution was stirred at this temperature for 15 min, then 0.10 g (0.33 mmol) of aldehyde 8, previously dissolved in 1 mL of dry THF, was added, and stirred for 1.5 h before the reaction was quenched with 5 mL of water. The solution was poured into methylene chloride (10 mL), the organic phase was separated and washed successively with water (3 × 5 mL) and brine (5 mL) and finally dried over anhydrous MgSO4, filtered and evaporated. The crude product was purified by column chromatography.
Procedure B: Preparing of MgBr2: 0.049 g (12 mmol) of magnesium turnings was placed in 6 mL of dry THF under an argon atmosphere, and then 0.17 mL (12 mmol) of 1,2-dibromoethane was added. The reaction was gently heated until the magnesium was completely dissolved. Then, 0.25 g (0.28 mmol) of tributhyltin derivative 6 or 7 was dissolved in 1.3 mL of dry THF and the solution was stirred under argon, cooled to −78 °C, and treated with n-BuLi (0.25 mL, 0.33 mmol) added dropwise. The solution was stirred at this temperature for 15 min, then 1 mL (0.33 mmol) previously prepared MgBr2 solution was added, continuing stirring for the next 15 min. In the next step, 0.10 g (0.33 mmol) of aldehyde 8 dissolved in 1 mL of dry THF was added and stirred for 1.5 h before the reaction was quenched with 5 mL of water. The solution was poured into methylene chloride (10 mL), the organic phase was separated and washed successively with water (3 × 5 mL) and brine (5 mL) and finally dried over anhydrous MgSO4, filtered and evaporated. The crude product was purified by column chromatography.
The reaction of the tin derivative of D-glucal 4 and D-galactal 5 with the aldehyde 8 resulted in a mixture of erythro-(S,S) and threo-(R,S) diastereoisomers. Pure stereoisomers of D-glucal 9 and 10 were isolated by flash chromatography on the apparatus Reveleris ®X2.

3.3.1. erythro-(S)-[3,4,6-Tris-O-(triisopropylsilyl)-D-glucal-1-yl][(S)-1-triphenylmethylaziridin-2-yl]methanol (9)

Colorless solid, 55% yield; Rf = 0.45 (petroleum ether/diethyl ether, 15:1); [α]D20 = −13.04 (c 0.5, CHCl3) IR (KBr): 3479 (νO-H), 3058, 3018 (ν=C-H), 2925, 2865 (νC-H), 1596, 1469 (νCAr-CAr), 1099 (νC-O); δH (600 MHz, CDCl3): 0.87–0.93 (m, 21H, TIPS-H), 0.97 (d, 1H, J = 6.4, CH2N), 0.98–1.07 (m, 42H, TIPS-H), 1.84 (d, 1H, J = 3.1, CH2N), 1.92 (ddd, 1H, J = 6.4, 3.1, 2.2, CHN), 3.50 (s, 1H, OH), 3.81 (dd, 1H, J = 11.2, 4.1, H-6), 3.92–3.95 (m, 1H, H-3), 3.96 (dd, 1H, J = 11.2, 7.6, H-6), 3.98–4.01 (m, 1H, H-4), 4.12–4.16 (m, 1H, H-5), 4.40 (d, 1H, J = 2.2, CHOH), 4.96 (d, 1H, J = 5.3, H-2), 7.18–7.23 (m, 3H, C6H5), 7.23–7.30 (m, 6H, C6H5), 7.40 (d, 6H, J = 7.6, C6H5); δC (150MHz, CDCl3): 12.2, 12.5, 12.5 (TIPS-C), 18.1, 18.2, 18.3, 18.4 (TIPS-C), 22.5 (CH2N), 34.8 (CHN), 62.3 (C-6), 65.9 (C-3), 66.6 (CHOH), 70.2 (C-4), 74.1 (C(C6H5)3), 81.2 (C-5), 94.5 (C-2), 127.0, 127.8, 129.4 (C6H5), 144.3 (C6H5), 152.5 (C-1); HRMS (EI): calculated for C55H89NO5Si3 M.+ 928.6127; found 928.6112.

3.3.2. threo-(R)-[3,4,6-Tris-O-(triisopropylsilyl)-D-glucal-1-yl][(S)-1-triphenylmethylaziridin-2-yl]methanol (10)

Colorless solid, 55% yield; Rf = 0.34 (petroleum ether/diethyl ether, 15:1); [α]D20 = −17.29 (c 0.6, CHCl3); IR (KBr): 3457 (νO-H), 3058, 3020 (ν=C-H), 2943, 2866 (νC-H), 1675 (νC=C), 1520,1464 (νCAr-CAr), 1062 (νC-O); δH (600 MHz, CDCl3): 0.95–1.04 (m, 63H, TIPS-H), 1.10 (d, 1H, J = 6.4, CH2N), 1.58 (ddd, 1H, J = 6.4, 5.8, 3.0, CHN), 1.83 (d, 1H, J = 3.0, CH2N), 2.41 (s, 1H, OH), 3.80 (d, 2H, J = 6.2, 2H-6), 3.92 (d, 1H, J = 5.8, CHOH), 3.95–3.98 (m, 1H, H-3), 4.01–4.04 (m, 1H, H-4), 4.15–4.19 (m, 1H, H-5), 4.88 (d, 1H, J = 4.4, H-2), 7.16–7.20 (m, 3H, C6H5), 7.22–7.27 (m, 6H, C6H5), 7.50 (d, 6H, J = 7.6, C6H5); δC (150 MHz, CDCl3): 12.1, 12.5, 12.6 (TIPS-C), 18.1, 18.2, 18.3, 18.4 (TIPS-C), 25.2 (CH2N), 37.3 (CHN), 61.7 (C-6), 66.2 (C-3), 70.2 (C-4), 73.9 (C(C6H5)3), 75.0 (CHOH), 81.0 (C-5), 96.6 (C-2), 126.8, 127.6, 129.7 (C6H5), 144.6 (C6H5), 151.9 (C-1); HRMS (EI): calculated for C55H89NO5Si3 M+ 928.6127; found 928.6112.

3.3.3. erythro-(S)-[3,4,6-Tris-O-(triisopropylsilyl)-D-galactal-1-yl][(S)-1-triphenylmethylaziridin-2-yl]methanol (11) and threo-(R)-[3,4,6-Tris-O-(triisopropylsilyl)-D-galactal-1-yl][(S)1-triphenylmethylaziridin-2-yl]methanol (12)

Colorless solid, 65% yield; Rf = 0.47 (hexane/ethyl acetate, 25:1); [α]D20 = −22.42 (c 0.3, CHCl3); IR (KBr): 3457 (νO-H), 3058, 3018 (ν=C-H), 2927, 2865 (νC-H), 1672 (νC=C), 1596, 1436 (νCAr-CAr), 1097 (νC-O); δH (600 MHz, CDCl3): 0.91–1.09 (m, 63H, TIPS-H), 1.15 (d, 1H, J = 6.4, CH2N), 1.47 (ddd, 1H, J = 6.4, 5.8, 2.9, CHN), 1.97 (bs, 1H, CH2N), 2.02 (d, 1H, J = 3.1, CH2N, erythro), 2.38 (s, 1H, OH), 3.58 (s, 1H, OH, erythro), 3.91 (d, 1H, J = 4.8, CHOH), 3.94–4.30 (m, 5H, H-3, H-4, H-5, 2H-6), 4.33 (d, 1H, J = 2.4, CHOH, eryhtro), 4.79 (bs, 1H, H-2), 4.90 (bs, 1H, H-2, erythro), 7.17–7.22 (m, 3H, C6H5), 7.23–7.30 (m, 6H, C6H5), 7.41 (d, J = 7.7, C6H5, erythro) 7.50 (d, 6H, J = 7.8, C6H5); δC (150 MHz, CDCl3): 12.1, 12.7 (TIPS-C), 18.1, 18.2, 18.3, 18.4, 18.4 (TIPS-C), 29.9 (CH2N), 37.4 (CHN), 61.0 (C-6), 64.3 (C-3), 70.2 (C-4), 73.6 (C(C6H5)3), 73.8 (CHOH), 80.8 (C-5), 97.6 (C-2), 99.1 (C-2, erythro), 126.9, 127.0, 127.7, 127.8, 128.9, 129.4, 129.6, 131.0 (C6H5), 144.3 (C6H5, erythro), 144.5 (C6H5), 152.7 (C-1); elementar analysis: C55H89NO5Si3 (928.56 g/mol) calculated: C% 71.14, H% 9.66, N% 1.51; found: C% 71.08, H% 9.66, N% 1.48.

3.4. General Procedure for Deprotection of Hydroxyl Groups

In the round bottom flask, there are 4 equivalents of tetrabutylammonium fluoride in 2.5 mL of dry THF, then the flask was secured with a septum and CaCl2 tube. 1 Equivalent of the compound 912 was dissolved in 2.5 mL of dry THF and added slowly to the tetrabutylammonium fluoride solution. Stirring was continued for 18 h at room temperature. After this time, the solvent was evaporated, and the residue was dissolved in ethyl acetate (15 mL), washed with brine (15 mL) and then dried over anhydrous MgSO4. After filtration and evaporation of the solvent, the product was purified by column chromatography using ethyl acetate and methanol (25:1).

3.4.1. erythro-(S)-[D-Glucal-1-yl][(S)-1-triphenylmethylaziridin-2-yl]methanol (13)

Colorless solid, 95% yield; Rf = 0.56 (ethyl acetate/methanol, 25:1); [α]D20 = −10.13 (c 0.6, CHCl3); IR (KBr): 3450 (νO-H), 3052 (ν=C-H), 2970, 2855 (νC-H), 1627 (νC=C), 1592, 1466 (νCAr-CAr), 1067 (νC-O); δH (600 MHz, CDCl3): 1.12 (d, 1H, J = 5.2, CH2N), 1.69 (m, 1H, CHN), 1.84 (d, 1H, J = 2.4, CH2N), 3.65–3.71 (m, 2H, H-4, H-5), 3.77 (d, 1H, J = 11.7, H-6), 3.82 (d, 1H, J = 11.7, H-6), 4.08–4.17 (m, 1H, H-3), 4.30 (d, 1H, J = 3.1, CHOH), 4.74 (bs, 1H, H-2), 7.19 (t, 3H, J = 7.2, C6H5), 7.22–7.27 (m, 6H, C6H5), 7.40 (d, 6H, J = 7.7, C6H5); δC (150 MHz, CDCl3): 29.8 (CHN), 34.4 (CHN), 61.2 (C-6), 67.9 (CHOH), 69.4 (C-4), 70.0 (C-3), 74.1 (C(C6H5)3), 78.5 (C-5), 99.6 (C-2), 127.1, 127.8, 129.4 (C6H5), 144.1 (C6H5), 154.3 (C-1); MS-EI m/z: 482.1 [M + Na]+; TOF MS ES+ calculated for C28H29NO5Na [M].+ 482.1943; found 482.1950.

3.4.2. threo-(R)-[D-Glucal-1-yl][(S)-1-triphenylmethylaziridin-2-yl]methanol (14)

Colorless solid, 95% yield; Rf = 0.53 (ethyl acetate/methanol, 25:1); [α]D20 = −7.62 (c 0.4, CHCl3); IR (KBr): 3453 (νO-H), 3048 (ν=C-H), 2955, 2868 (νC-H), 1634 (νC=C), 1575, 1472 (νCAr-CAr), 1059 (νC-O); δH (600 MHz, CDCl3): 1.12 (d, 1H, J = 5.3, CH2N), 1.62 (m, 1H, CHN), 1.79 (d, 1H, J = 2.3, CH2N), 3.48 (dd, 1H, J = 8.7, 7.2, H-4), 3.58 (d, 1H, J = 11.6, H-6), 3.63 (d, 4H, J = 9.6, H-5), 3.74 (d, 1H, J = 11.6, H-6), 3.94 (bs, 1H, CHOH), 4.11 (d, 1H, J = 7.2, H-3), 4.80 (bs, 1H, H-2), 7.13 (t, 3H, J = 7.1, C6H5), 7.17–7.23 (m, 6H, C6H5), 7.38 (d, 6H, J = 7.3, C6H5); δC (150 MHz, CDCl3): 32.1 (CHN), 35.8 (CHN), 60.5 (C-6), 70.0 (C-3, C-4), 70.7 (CHOH), 74.0 (C(C6H5)3), 78.2 (C-5), 99.9 (C-2), 127.0, 127.7, 128.3, 128.9, 129.7, (C6H5), 144.2 (C6H5), 154.8 (C-1); MS-EI m/z: 482.1 [M + Na]+; TOF MS ES+ calculated for C28H29NO5Na [M].+ 482.1943; found 482.1942.

3.4.3. erythro-(S)-[D-Glalactal-1-yl][(S)-1-triphenylmethylaziridin-2-yl]methanol (15) and threo-(R)-[D-Galactal-1-yl][(S)-1-triphenylmethylaziridin-2-yl]methanol (16)

Colorless solid, 96% yield, Rf = 0.54 (ethyl acetate/methanol, 25:1); [α]D20 = −3.85 (c 0.4, CHCl3); IR (KBr): 3477 (νO-H), 3063 (ν=C-H), 2975 (νC-H), 1653 (νC=C), 1543, 1491 (νCAr-CAr), 1068 (νC-O); δH (600 MHz, CDCl3): 1.27 (d, 1H, J = 5.2, CH2N), 1.66 (bs, 1H, OH), 1.89–1.94 (m, 1H, CHN), 1.97 (d, 1H, J = 3.4, CHN), 2.20 (bs, 1H, OH), 2.37 (bs, 1H, OH), 3.38 (bs, 1H, OH), 3.77 (dd, 1H, J = 12.5, 5.7, H-6), 3.82–3.86 (m, 2H, H-4, H-5), 3.87 (dd, 1H, J = 12.5, 5.3, H-6), 4.04 (bs, 1H, CHOH), 4.28 (bs, 1H, CHOH, erythro), 4.29–4.34 (m, 1H, H-3), 4.79 (bs, 1H, J = 4.5, H-2, erythro), 4.89 (dd, 1H, J = 5.7, 1.6, H-2), 7.23 (t, 3H, J = 7.2, C6H5), 7.29 (t, 6H, J = 7.1, C6H5), 7.42 (d, J = 7.8, C6H5, erythro), 7.46 (d, 6H, J = 7.5, C6H5); δC (150 MHz, CDCl3): 29.5 (CHN, erythro), 29.8 (CHN), 34.8 (CHN, erythro), 35.4 (CHN), 62.8 (C-6, erythro), 62.9 (C-6), 64.6 (C-3, erythro), 64.7 (C-3), 66.4 (C-4), 67.7 (CHOH), 68.4 (CHOH, erythro), 73.9 (C(C6H5)3), 74.2 (C(C6H5)3, erythro), 76.6 (C-5, erythro), 76.7 (C-5), 97.6 (C-2), 99.2 (C-2, erythro), 127.2, 127.7, 129.8 (C6H5), 127.8, 129.5 (C6H5, erythro), 144.0 (C6H5), 144.1 (C6H5, erythro), 154.2 (C-1, erythro), 155.7 (C-1); MS-EI m/z: 482.1 [M + Na]+; TOF MS ES+ calculated for C28H29NO5Na [M].+ 482.1943; found 482.1952.

3.5. General Procedure for Aziridine Ring Opening Reaction

In a round bottom flask, 1 equivalent of compound 1316 was dissolved in 1 mL of methylene chloride and then 3 equivalents of thiophenol were added. The mixture was stirred at room temperature for 2–6 h (controlled by TLC tests). The crude product was dissolved in methylene chloride and purified on a preparative plate using ethyl acetate and methanol (25:1).

3.5.1. erytro-(1S,2R)-1-[(1-Hydroxy-3-(phenylthio)-2-(triphenylmethylamino)propyl)]-D-glucal (17)

White solid, 67% yield; Rf = 0.58 (ethyl acetate/methanol, 25:1); [α]D20 = −7.68 (c 0.6, CHCl3); IR (KBr): 3385 (νO-H), 3083, 3057, 3031 (ν=C-H), 2923, 2852 (νC-H), 1636 (νC=C), 1594, 1576, 1521, 1447 (νCAr-CAr), 644 (νC-S); δH (600 MHz, CDCl3): 2.06 (bs, 4H, 4OH), 2.86–2.97 (m, 3H, CH2S, CHN), 3.49–3.59 (m, 2H, H-4, H-5) 3.63–3.58 (m, 1H, H-3), 3.69 (d, 1H, J = 12.5, H-6), 3.77 (d, 1H, J = 12.5, H-6), 4.12 (d, 1H, J = 5.1, CHOH), 4.70 (bs, 1H, H-2), 7.08 (d, 2H, J = 7.1, C6H5), 7.11–7.23 (m, 13H, C6H5), 7.46 (d, 6H, J = 7.6, C6H5); δC (150 MHz, CDCl3): 36.0 (CH2S), 54.1 (CHN), 61.3 (C-6), 69.8 (CHOH), 70.0 (C-3), 70.6 (C-4), 71.1 (C(C6H5)3), 77.8 (C-5), 98.9 (C-2), 126.2, 126.5, 1279, 128.8, 128.9, 129.9, 136.4 (C6H5), 146.3 (C6H5), 154.4 (C-1); MS-EI m/z: 592.7 [M + Na]+; TOF MS ES+ calculated for C34H35NO5NaS [M].+ 592.2134; found 592.2150.

3.5.2. threo-(1R,2R)-1-[(1-Hydroxy-3-(phenylthio)-2-(triphenylmethylamino)propyl)]-D-glucal (18)

Colorless solid, 68% yield; Rf = 0.51 (ethyl acetate/methanol, 25:1); [α]D20 = + 13.8 (c 0.4, CHCl3); IR (KBr): 3382 (νO-H), 3079, 3043, 3031 (ν=C-H), 2918, 2843 (νC-H), 1642 (νC=C), 1589, 1571, 1520, 1437 (νCAr-CAr), 637 (νC-S); δH (600 MHz, CDCl3): 1.66 (bs, 4H, OH), 2.12 (dd, 1H, J = 12.8, 6.2, CH2S), 2.65 (dd, 1H, J = 12.8, 1.7, CH2S), 3.21 (t, 1H, J = 5.9, CHN), 3.58–3.69 (m, 2H, H-5, H-6), 3.81 (t, 1H, J = 9.4, H-4), 3.86–3.93 (m, 2H, H-6, CHOH), 4.12 (d, 1H, J = 7.3, H-3), 4.65 (s, 1H, H-2), 7.07 (d, 2H, J = 8.4, C6H5), 7.15–7.28 (m, 13H, C6H5), 7.48 (d, 6H, J = 7.4, C6H5); δC (150 MHz, CDCl3): 36.5 (CH2S), 54.3 (CHN), 60.6 (C-6), 68.3 (C-4), 69.7 (C-3), 71.2 (C(C6H5)3), 73.0 (CHOH), 79.0 (C-5), 104.0 (C-2), 126.5, 126.9, 128.2, 129.0, 130.1, 136.7 (C6H5), 146.2 (C6H5), 151.9 (C-1); MS-EI m/z: 529.2 [M + Na]+; TOF MS ES+ calculated for C34H35NO5NaS [M].+ 592.2134; found 592.2156.

3.5.3. erytro-(1S,2R)-1-[(1-Hydroxy-3-(phenylthio)-2-(triphenylmethylamino)propyl)]-D-galactal (19) and threo-((1R,2R)-1-[(1-hydroxy-3-(phenylthio)-2-(triphenylmethylamino)propyl)]-D-galactal (20)

Colorless solid, 72% yield, Rf = 0.58 (ethyl acetate/methanol, 25:1); [α]D20 = −3.67 (c 0.6, CHCl3); IR (KBr): 3390 (νO-H), 3093, 3066, 3042 (ν=C-H), 2939, 2866 (νC-H), 1655 (νC=C), 1592, 1569, 1526, 1437 (νCAr-CAr), 651 (νC-S); δH (600 MHz, CDCl3): 1.97 (bs, 1H, OH, threo), 2.74 (s, 3H, 3OH, threo), 2.90 (dd, 1H, J = 15.3, 10.3, CH2S, threo), 3.14–3.19 (m, 2H, CH2S, CHN, threo), 3.73 (dd, 1H, J = 13.4, 3.4, H-6, threo), 3.86 (dd, 1H, J = 13.4, 1.1, H-6, threo), 3.94–3.97 (m, 1H, H-5, threo), 4.02–4.05 (m, 1H, H-3, threo), 4.14 (t, 1H, J = 4.1, H-4, threo), 4.20 (bs, 1H, CHOH, threo), 4.87 (d, 1H, J = 4.7, H-2, erythro), 5.05 (dd, 1H, J = 4.6, 1.0, H-2, threo), 7.17 (t, 3H, J = 7.4, C6H5, threo), 7.19–7.27 (m, 17H, C6H5, threo); δC (150 MHz, CDCl3): 39.7 (CH2S), 52.1 (CHN), 61.0 (C-6), 61.5 (C-3), 67.1 (C-4), 70.9 (CHOH), 76.2 (C-5), 82.2 (C(C6H5)3), 98.0 (C-2), 127.0, 127.4, 129.0, 129.4, 130.3 (C6H5), 147.0 (C6H5), 155.0 (C-1); MS-EI m/z: 592.1 [M + Na]+; TOF MS ES+ calculated for C34H35NO5NaS [M]+ 592.2134; found 592.2141.

4. Conclusions

In conclusion, we have developed a simple and stereoselective methodology for the synthesis of C-glycosyl-aminoethyl sulfide derivatives of potential biological interest by a reaction of tributhyltin derivatives 6 and 7 of glycals with aziridinecarboaldehyde 8 and the regioselective ring opening of a chiral aziridine with thiophenol. The absolute configurations of the resulting diastereoisomers were determined via 1H NMR spectroscopy. The obtained results indicate that chelation is a less important factor influencing stereoselectivity. For galactal derivatives, the reactions proceed mainly according to the Felkin–Anh model, leading predominantly to the threo product. However, for glucal derivative, a higher contribution of “chelation-controlled” carbon–carbon bond formation was observed, which results in an increase of the erythro isomer. At the present stage of the study, we do not have adequate experimental material to explain the higher diastereoselectivity of the D-galacto derivative 7 in comparison to D-gluco 6.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27061764/s1, Figure S1: 1H NMR (600 MHz, CD3Cl) spectrum of 9, Figure S2: 13C NMR (150 MHz, CD3Cl) spectrum of 9, Figure S3: 1H-1H COSY spectrum of 9, Figure S4: 1H-13C HMQC spectrum of 9, Figure S5: 1H NMR (600 MHz, CD3Cl) spectrum of 10, Figure S6: 13C NMR (150 MHz, CD3Cl) spectrum of 10, Figure S7: 1H-1H COSY spectrum of 10, Figure S8: 1H-13C HMQC spectrum of 10, Figure S9: 1H NMR (600 MHz, CD3Cl) spectrum of 11 and 12, Figure S10: 13C NMR (150 MHz, CD3Cl) spectrum of 11 and 12, Figure S11: 1H NMR (600 MHz, CD3Cl) spectrum of 13, Figure S12: 13C NMR (150 MHz, CD3Cl) spectrum of 13, Figure S13: 1H NMR (600 MHz, CD3Cl) spectrum of 14, Figure S14: 13C NMR (150 MHz, CD3Cl) spectrum of 14, Figure S15: 1H NMR (600 MHz, CD3Cl) spectrum of 15 and 16, Figure S16: 13C NMR (150 MHz, CD3Cl) spectrum of 15 and 16, Figure S17: 1H NMR (600 MHz, CD3Cl) spectrum of 17, Figure S18: 13C NMR (150 MHz, CD3Cl) spectrum of 17, Figure S19: 1H NMR (600 MHz, CD3Cl) spectrum of 18, Figure S20: 13C NMR (150 MHz, CD3Cl) spectrum of 18, Figure S21: 1H NMR (600 MHz, CD3Cl) spectrum of 19 and 20, Figure S22: 13C NMR (150 MHz, CD3Cl) spectrum of 19 and 20.

Author Contributions

Conceptualization, methodology, A.Z. and S.L.; writing—original draft preparation, A.Z.; writing—review and editing, A.Z., S.L., A.T. and M.M.; experimental part, A.Z., A.T. and M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Higher Education conferred on the basis of the decision number 201536/E-345/M/2018 and by Student Research Grants at the University of Lodz.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Farina, V.; Reeves, J.T.; Senanayake, C.H.; Song, J.J. Asymmetric synthesis of active pharmaceutical ingredients. Chem. Rev. 2006, 106, 2734–2793. [Google Scholar] [CrossRef] [PubMed]
  2. Caner, H.; Groner, E.; Levy, L.; Agranat, I. Trends in the development of chiral drugs. Drug Discov. Today 2004, 9, 105–110. [Google Scholar] [CrossRef]
  3. Nag, A. Asymmetric Synthesis of Drugs and Natural Products; CRC Press: Boca Raton, FL, USA, 2018. [Google Scholar]
  4. Varki, A.; Cummings, R.; Esko, J.; Freeze, H.; Hart, G.; Marth, J. Essentials of Glycobiology; Cold Spring Harbor Laboratory Press: New York, NY, USA, 1999. [Google Scholar]
  5. Wong, C.-H. Carbohydrate-Based Drug Discovery; Wiley-VCH: Weinheim, Germany, 2003; Volumes 1 and 5. [Google Scholar]
  6. Nuzzi, A.; Massi, A.; Dondoni, A. General Synthesis of C-Glycosyl Amino Acids via Proline-Catalyzed Direct Electrophilic α-Amination of C-Glycosylalkyl Aldehyde. Org. Lett. 2008, 10, 4485–4488. [Google Scholar] [CrossRef] [PubMed]
  7. Schäfer, A.; Henkensmeier, D.; Kröger, L.; Thiem, J. Aziridine ring opening as regio- and stereoselective access to O-glycosyl amino acids and their transformation into O-glycopeptide mimetics. Tetrahedron Asymmetry 2009, 20, 902–909. [Google Scholar] [CrossRef]
  8. Seitz, O. Glycopeptide Synthesis and the Effects of Glycosylation on Protein Structure and Activity. ChemBioChem 2000, 1, 214–246. [Google Scholar] [CrossRef]
  9. Sears, P.; Wong, C.-H. Carbohydrate Mimetics: A New Strategy for Tackling the Problem of Carbohydrate-Mediated Biological Recognition. Angew. Chem. Int. Ed. 1999, 38, 2301–2324. [Google Scholar] [CrossRef]
  10. Dondoni, A.; Marra, A.; Massi, A. Stereoselective synthesis of the C-linked analogue of β-D-galactopyranosyl-L-serine. Tetrahedron 1998, 54, 2827–2832. [Google Scholar] [CrossRef]
  11. Lay, L.; Meldal, M.; Nicotra, F.; Panza, L.; Russo, G. Stereoselective synthesis of the C-analogue of β-D-glucopyranosyl serin. J. Chem. Soc. Chem. Commun. 1997, 1469–1470. [Google Scholar] [CrossRef]
  12. Debenham, S.D.; Debenham, J.S.; Burk, M.J.; Toone, E.J. Synthesis of Carbon-Linked Glycopeptides through Catalytic Asymmetric Hydrogenation. J. Am. Chem. Soc. 1997, 119, 9897–9898. [Google Scholar] [CrossRef]
  13. Fisher, J.F.; Harrison, A.W.; Bundy, G.L.; Wilkinson, K.F.; Bush, B.D.; Ruwart, M.J. Peptide to glycopeptide: Glycosylated oligopeptide renin inhibitors with attenuated in vivo clearance properties. J. Med. Chem. 1991, 34, 3140–3143. [Google Scholar] [CrossRef]
  14. Kröger, L.; Henkensmeier, D.; Schäfe, A.; Thiem, J. Novel O-glycosyl amino acid mimetics as building blocks for O-glycopeptides act as inhibitors of galactosidases. Bioorgan. Med. Chem. Lett. 2004, 14, 73–75. [Google Scholar] [CrossRef] [PubMed]
  15. Ernst, B.; Magnani, J. From carbohydrate leads to glycomimetic drugs. Nat. Rev. Drug. Disc. 2009, 8, 661–677. [Google Scholar] [CrossRef] [PubMed]
  16. Levy, D.E.; Tang, C. The Chemistry of C-Glycosides; Pergamon: Oxford, UK, 1995. [Google Scholar]
  17. Dondoni, A.; Mariotti, D.; Marra, A. Synthesis of α- and β-Glycosyl Asparagine Ethylene Isosteres (C-Glycosyl Asparagines) via Sugar Acetylenes and Garner Aldehyde Coupling. J. Org. Chem. 2002, 67, 4475–4486. [Google Scholar] [CrossRef] [PubMed]
  18. Marcaurelle, L.A.; Bertozzi, C.R. New Directions in the Synthesis of Glycopeptide Mimetics. Chem. Eur. J. 1999, 5, 1384–1390. [Google Scholar] [CrossRef]
  19. Schäfer, A.; Thiem, J. Synthesis of Novel Donor Mimetics of UDP-Gal, UDP-GlcNAc, and UDP-GalNAc as Potential Transferase Inhibitors. J. Org. Chem. 2000, 65, 24–29. [Google Scholar] [CrossRef]
  20. Wittmann, V.; Kessler, H. Stereoselective Synthesis of C-Glycoside with a Glycosyl Dianio. Angew. Chem. Int. Ed. Engl. 1993, 105, 1091–1093. [Google Scholar] [CrossRef] [Green Version]
  21. Westermann, B.; Walter, A.; Florke, U.; Altenbach, H.-J. Chiral Auxiliary Based Approach Toward the Synthesis of C-Glycosylated Amino Acids. Org. Lett. 2001, 3, 1375–1378. [Google Scholar] [CrossRef]
  22. Togo, H.; He, W.; Waki, Y.; Yokoyama, M. C-Glycosidation Technology with Free Radical Reactions. Synlett 1998, 7, 700–717. [Google Scholar] [CrossRef]
  23. Paterson, D.E.; Griffin, F.K.; Alcaraz, M.-L.; Taylor, R.J.K. A Ramberg−Bäcklund Approach to the Synthesis of C-Glycosides, C-Linked Disaccharides, and C-Glycosyl Amino Acids. Eur. J. Org. Chem. 2002, 7, 1323–1336. [Google Scholar] [CrossRef]
  24. Dondoni, A.; Marra, A. Methods for anomeric carbon-linked and fused sugar amino acid synthesis: The gateway to artificial glycopeptides. Chem. Rev. 2000, 100, 4395–4422. [Google Scholar] [CrossRef]
  25. Xu, X.; Fakha, G.; Sinou, D. Stereoselective synthesis of C-glycosyl analoguesphenylalanine. Tetrahedron 2002, 58, 7539–7544. [Google Scholar] [CrossRef]
  26. Li, X.; Takahashi, H.; Ohtake, H.; Ikegami, S. Synthesis of ketosyl spiro-isoxazolidine by 1,3-dipolar cycloaddition of 1-methylenesugars with nitrones—A new access to C-glycosyl amino acids. Heterocycles 2003, 59, 547–571. [Google Scholar] [CrossRef]
  27. Gustafsson, T.; Saxin, M.; Kihlberg, J. Synthesis of a C-Glycoside Analogue of β-D-Galactosylthreonine. J. Org. Chem. 2003, 68, 2506–2509. [Google Scholar] [CrossRef] [PubMed]
  28. Dondoni, A.; Giovannini, P.P.; Massi, A. Assembling heterocycle-tethered C-glycosyl and alpha-amino acid residues via 1,3-dipolar cycloaddition reactions. Org. Lett. 2004, 6, 2929–2932. [Google Scholar] [CrossRef] [PubMed]
  29. Postema, M.H.D.; Piper, J.L. Synthesis of Some Biologically Relevant β-C-Glycoconjugates. Org. Lett. 2003, 5, 1721–1723. [Google Scholar] [CrossRef] [PubMed]
  30. Chambers, D.J.; Evans, G.R.; Fairbanks, A.J. Synthesis of C-glycosyl amino acids: Scope and limitations of the tandem Tebbe/Claisen approach. Tetrahedron Asymmetry 2005, 16, 45–55. [Google Scholar] [CrossRef]
  31. Dondoni, A.; Massi, A.; Aldhoun, M. Hantzsch-Type Three-Component Approach to a New Family of Carbon-Linked Glycosyl Amino Acids. Synthesis of C-Glycosylmethyl Pyridylalanines. J. Org. Chem. 2007, 72, 7677–7687. [Google Scholar] [CrossRef]
  32. McGarvey, G.; Benedum, T.; Schmidtmann, F. Development of Co- and Post-Translational Synthetic Strategies to C-Neoglycopeptides. Org. Lett. 2002, 4, 3591–3594. [Google Scholar] [CrossRef]
  33. Müller, P.; Nury, P. Copper-Catalyzed Desymmetrization of N-Sulfonylaziridines with Methylmagnesium Halides. Org. Lett. 1999, 1, 439–442. [Google Scholar] [CrossRef]
  34. Hu, X.E.; Kim, N.K.; Ledoussal, B.; Colson, A.-O. Regio- and stereo-controlled copper organometallic addition to a piperidinyl aziridine: Synthesis of trans 3-amino-4-alkyl-piperidines. Tetrahedron Lett. 2002, 43, 4289–4293. [Google Scholar] [CrossRef]
  35. Crousse, B.; Narizuka, S.; Bonnet-Delpon, D.; Begue, J.-P. First Stereoselective Synthesis of cis 3-CF3-Aziridine-2-carboxylates. A Route to New (Trifluoromethyl) α-Functionalised β-Amino Acids. Synlett 2001, 5, 679–681. [Google Scholar] [CrossRef]
  36. Lugiņina, J.; Uzuleņ, J.; Posevins, D.; Turks, M. Ring-Opening of Carbamate-Protected Aziridines and Azetidines in Liquid Sulfur Dioxide. Eur. J. Org. Chem. 2016, 9, 1760–1771. [Google Scholar] [CrossRef]
  37. Chakraborty, T.K.; Ghosh, A.; Raju, T.V. Efficient Ring Opening Reactions of N-Tosyl Aziridines with Amines and Water in Presence of Catalytic Amount of Cerium(IV) Ammonium Nitrate. Chem. Lett. 2003, 32, 82–83. [Google Scholar] [CrossRef] [Green Version]
  38. O’Neil, I.A.; Woolley, J.C.; Southern, J.M.; Hobbs, H. The synthesis of β-N-tosylamino hydroxylamines via the ring opening of N-tosylaziridines and their use in reverse Cope cyclisation. Tetrahedron Lett. 2001, 42, 8243–8245. [Google Scholar] [CrossRef]
  39. Nishikawa, T.; Ishikawa, M.; Wade, K.; Isobe, M. Total Synthesis of α-C-Mannosyltryptophan, a Naturally Occurring C-Glycosyl Amino Acid. Synlett 2001, 945–947. [Google Scholar] [CrossRef] [Green Version]
  40. Wu, J.W.; Hou, X.-L.; Dai, L.-X. Effective Ring-Opening Reaction of Aziridines with Trimethylsilyl Compounds:  A Facile Access to β-Amino Acids and 1,2-Diamine Derivatives. J. Org. Chem. 2000, 65, 1344–1348. [Google Scholar] [CrossRef]
  41. Matsubara, S.; Kodama, T.; Utimoto, K. Yb(CN)3-catalyzed reaction of aziridines with cyanotrimethylsilane. A facile synthesis of optically pure β-amino nitriles. Tetrahedron Lett. 1990, 31, 6379–6380. [Google Scholar] [CrossRef]
  42. Wu, B.; Gallucci, J.C.; Parquett, J.R.; RajanBabu, T.V. Regiodivergent Ring Opening of Chiral Aziridine. Angew. Chem. Int. Ed. 2009, 48, 1126–1129. [Google Scholar] [CrossRef]
  43. Sugihara, Y.; Limura, S.; Nakayama, J. Aza-pinacol rearrangement: Acid-catalyzed rearrangement of aziridines to imines. Chem. Commun. 2002, 134–135. [Google Scholar] [CrossRef]
  44. Prasad, B.A.B.; Sekar, G.; Singh, V.K. An efficient method for the cleavage of aziridines using hydroxyl compound. Tetrahedron Lett. 2000, 41, 4677–4679. [Google Scholar] [CrossRef]
  45. Mao, H.; Joly, G.J.; Peeters, K.; Hoornaert, G.J.; Compernolle, F. Synthesis of 1-deoxymannojirimycin analogues using N-tosyl and N-nosyl activated aziridines derived from 1-amino-1-deoxyglucitol. Tetrahedron 2001, 57, 6955–6967. [Google Scholar] [CrossRef]
  46. Satoh, T.; Matsue, R.; Fujii, T.; Morikawa, S. Cross-coupling of nonstabilized aziridinylmagnesiums with alkylhalides catalyzed by Cu(I) iodide: A new synthesis of amines bearing a quaternary chiral center and an asymmetric synthesis of both enantiomers of the amines from one chiral starting material. Tetrahedron 2001, 57, 3891–3898. [Google Scholar] [CrossRef]
  47. Davis, F.A.; Liu, H.; Reddy, G.V. 2-Methyl N-(p-toluenesulfinyl)aziridine-2-carboxylic acid: Asymmetric synthesis of α-methylphenylalanine and α-methyl-β-phenylserine. Tetrahedron Lett. 1996, 37, 5473–5476. [Google Scholar] [CrossRef]
  48. Jarzyński, S.; Leśniak, S.; Rachwalski, M. Synthesis of enantiomerically pure 2-(N-aryl, N-alkyl-aminomethyl)aziridines: A new class of ligands for highly enantioselective asymmetric synthesis. Tetrahedron Asymmetry 2017, 28, 1808–1816. [Google Scholar] [CrossRef]
  49. Olofsson, B.; Somfai, P. Divergent Synthesis of D-erythro-Sphingosine, L-threo-Sphingosine, and Their Regioisomers. J. Org. Chem. 2003, 68, 2514–2517. [Google Scholar] [CrossRef]
  50. Fukuta, Y.; Mita, T.; Fukuda, N.; Kanai, M.; Shibasaki, M. De Novo Synthesis of Tamiflu via a Catalytic Asymmetric Ring-Opening of meso-Aziridines with TMSN3. J. Am. Chem. Soc. 2006, 128, 6312–6313. [Google Scholar] [CrossRef]
  51. Frappa, I.; Kryczka, B.; Lhoste, P.; Porwański, S.; Sinou, D.; Zawisza, A. Palladium(0)-Mediated Synthesis of Acetylated Unsaturated 1,4-Disaccharides. J. Carbohydr. Chem. 1998, 17, 1117–1130. [Google Scholar] [CrossRef]
  52. Zawisza, A.; Kryczka, B.; Lhoste, P.; Porwański, S.; Sinou, D. Efficient Palladium(0)-Catalyzed Synthesis of Alkenyl 1-Thioglycosides and Thiodisaccharides. J. Carbohydr. Chem. 2000, 19, 795–804. [Google Scholar] [CrossRef]
  53. Jarosz, S.; Szewczyk, K.; Zawisza, A. Synthesis and thermal stability of secondary sugar allyltin derivatives. Tetrahedron: Asymmetry 2003, 14, 1715–1723. [Google Scholar] [CrossRef]
  54. Szulc, I.; Kołodziuk, R.; Kryczka, B.; Zawisza, A. New phosphine–imine ligands derived from D-gluco- and D-galactosamine in Pd-catalysed asymmetric allylic alkylation. Tetrahedron Lett. 2015, 56, 4740–4743. [Google Scholar] [CrossRef]
  55. Kubiak, A.; Kołodziuk, R.; Porwański, S.; Zawisza, A. Palladium(0)-catalysed synthesis of 2,3- and 3,4-unsaturated aryl β-O-glycosides. Carbohydr. Res. 2015, 417, 34–40. [Google Scholar] [CrossRef] [PubMed]
  56. Janasik, B.; Zawisza, A.; Malachowska, B.; Fendler, W.; Stanislawska, M.; Kuras, R.; Wąsowicz, W. Relationship between arsenic and selenium in workers occupationally exposed to inorganic arsenic. J. Trace Elem. Med. Biol. 2017, 42, 76–80. [Google Scholar] [CrossRef] [PubMed]
  57. Szulc, I.; Kołodziuk, R.; Zawisza, A. New phosphine-imine and phosphine-amine ligands derived from D-gluco-, D-galacto- and D-allosamine in Pd-catalysed asymmetric allylic alkylation. Tetrahedron 2018, 74, 1476–1485. [Google Scholar] [CrossRef]
  58. Robak, J.; Koselak, K.; Zawisza, A.; Porwański, S. Studies on the influence of saccharide fragment of urea organocatalysts on the yield and enantioselectivity of aza-Henry reaction. Arkivoc 2020, 8, 150–160. [Google Scholar] [CrossRef]
  59. 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]
  60. 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]
  61. Rachwalski, M.; Jarzyński, S.; Jasiński, M.; Leśniak, S. Mandelic Acid Derived α-Aziridinyl Alcohols as Highly Efficient Ligands for Asymmetric Additions of Zinc Organyls to Aldehydes. Tetrahedron Asymmetry 2013, 24, 689–693. [Google Scholar] [CrossRef]
  62. Buchcic, A.; Zawisza, A.; Leśniak, S.; Adamczyk, J.; Pieczonka, A.M.; Rachwalski, M. Enantioselective Mannich Reaction Promoted by Chiral Phos-phinoyl-Aziridines. Catalysts 2019, 9, 837. [Google Scholar] [CrossRef] [Green Version]
  63. 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]
  64. Buchcic, A.; Zawisza, A.; Leśniak, S.; Rachwalski, M. Asymmetric Friedel–Crafts Alkylation of Indoles Catalyzed by Chiral Aziridine-Phosphines. Catalysts 2020, 10, 971–980. [Google Scholar] [CrossRef]
  65. 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, 968–978. [Google Scholar] [CrossRef]
  66. May, S.W.; Phillips, R.S. Asymmetric sulfoxidation by dopamine beta-hydroxylase, an oxygenase heretofore considered specific for methylene hydroxylation. J. Am. Chem. Soc. 1980, 102, 5981–5983. [Google Scholar] [CrossRef]
  67. May, S.W.; Phillips, R.S.; Mueller, P.W.; Herman, H.H. Dopamine beta-hydroxylase. Comperative specificities and mechanisms of the oxygenation reaction. J. Biol. Chem. 1981, 256, 8470–8475. [Google Scholar] [CrossRef]
  68. Padgette, S.R.; Herman, H.H.; Han, J.H.; Pollock, S.H.; May, S.W. Antihypertensive Activities of Phenyl Aminoethyl Sulfides, a Class of Synthetic Substrates for Dopamine β-Hydroxylase. J. Med. Chem. 1984, 27, 1354–1357. [Google Scholar] [CrossRef] [PubMed]
  69. Herman, H.H.; Husain, P.A.; Colbert, J.E.; Schweri, M.M.; Pollock, S.H.; Fowler, L.C.; May, S.W. The Enantiomeric Specificity of the Antihypertensive Activity of l-(Phenylthio)-2-aminopropane, a Synthetic Substrate Analogue for Dopamine β-Monooxygenase. J. Med. Chem. 1991, 34, 1082–1085. [Google Scholar] [CrossRef]
  70. Kandalkar, S.R.; Ramaiah, P.A.; Joshi, M.; Wavhal, A.; Waman, Y.; Raje, A.A.; Tambe, A.; Ansari, S.; De, S.; Palle, V.P.; et al. Modifications of flexible nonyl chain and nucleobase head group of (+)-erythro-9-(2′s-hydroxy-3′s-nonyl)adenine [(+)-EHNA] as adenosine deaminase inhibitors. Bioorgan. Med. Chem. 2017, 25, 5799–5819. [Google Scholar] [CrossRef] [PubMed]
  71. Tucker, H.; Coope, J.F. β-Adrenergic blocking agents. 18. 1-(Aryloxy)-3-(arylthioalkylamino)propan-2-ols and 1-substituted alkylthioamino-3-(aryloxy)propan-2-ols. J. Med. Chem. 1978, 21, 769–773. [Google Scholar] [CrossRef]
  72. Erdmann, A.; Menon, Y.; Gros, C.; Molinier, N.; Novosad, N.; Samson, A.; Gregoire, J.-M.; Long, C.; Ausseil, F.; Halby, L.; et al. Design and synthesis of new non nucleoside inhibitors of DNMT3A. Bioorgan. Med. Chem. 2015, 23, 5946–5953. [Google Scholar] [CrossRef]
  73. Bartlett, M.J.; Turner, C.A.; Harvey, J.E. Pd-Catalyzed Allylic Alkylation Cascade with Dihydropyrans: Regioselective Synthesis of Furo[3,2-c]pyrans. Org. Lett. 2013, 15, 2430–2433. [Google Scholar] [CrossRef]
  74. Bearder, J.R.; Dewis, M.L.; Whiting, D.A. Short synthetic route to congeners of the undecose antibiotic herbicidin. J. Chem. Soc. Perkin Trans. 1995, 227–233. [Google Scholar] [CrossRef]
  75. Zhang, S.; Niu, Y.-H.; Ye, X.-S. General Approach to Five-Membered Nitrogen Heteroaryl C-Glycosides Using a Palladium/Copper Cocatalyzed C–H Functionalization Strategy. Org. Lett. 2017, 19, 3608–3611. [Google Scholar] [CrossRef] [PubMed]
  76. Hnnesian, S.; Martin, M.; Desai, R. Formation of C-Glycosides by Polarity Inversion at the Anomeric Centre. J. Chem. Soc. Chem. Commun. 1986, 926–927. [Google Scholar] [CrossRef]
  77. Nicolaou, K.C.; Hwang, C.-K.; Duggan, M.E. Stereospecific Synthesis of 1,1-Dialkylglycosides. J. Chem. Soc. Chem. Commun. 1986, 12, 925–926. [Google Scholar] [CrossRef]
  78. Lesimple, P.; Beau, J.-M.; Jaurand, G.; Sinaÿ, P. Preparation and use of lithiated glycals: Vinylic deprotonation versus tin-lithium exchange from 1-tributylstannyl glycals. Tetrahedron Lett. 1986, 27, 6201–6204. [Google Scholar] [CrossRef]
  79. Crich, D.; Ritchie, T.J. Preparation and reactions of some cyclic orthoester derivatives. Tetrahedron 1988, 44, 2319–2328. [Google Scholar] [CrossRef]
  80. Friesen, R.W.; Sturino, C.F.; Daljeet, A.K.; Kolaczewska, A. Observation of .alpha.-silyl carbanions in the metalation of 3,4,6-tri-O-(tert-butyldimethylsilyl)-D-glucal. J. Org. Chem. 1991, 56, 1944–1947. [Google Scholar] [CrossRef]
  81. Utsunomiya, I.; Fuji, M.; Sato, T.; Natsume, M. Preparation of Alkyl-Substituted Indoles in the Benzene Portion. Part 9. Synthesis of (1aS,8bS)-1-tert-Butyloxycarbonyl-8-formyl-1,1a,2,8b-tetrahydroazirino[2’,3’:3,4]pyrrolo[1,2-a]indole. Model Study for the Enantiospecific Synthesis of Aziridinomitosenes. Chem. Pharm. Bull. 1993, 41, 854–860. [Google Scholar] [CrossRef] [Green Version]
  82. Kulshrestha, A.; Schomaker, J.M.; Holmes, D.; Staples, R.J.; Jackson, J.E.; Borhan, B. Selectivity in the Addition Reactions of Organometallic Reagents to Aziridine-2-carboxaldehydes: The Effects of Protecting Groups and Substitution Patterns. Chem. Eur. J. 2011, 17, 12326–12339. [Google Scholar] [CrossRef]
  83. Hwang, G.-I.; Chung, J.-H.; Lee, W.K. Efficient Synthesis of Ephedra Alkaloid Analogues Using an Enantiomerically Pure N-[(R)-(+)-α-Methylbenzyl]aziridine-2-carboxaldehyde. J. Org. Chem. 1996, 61, 6183–6188. [Google Scholar] [CrossRef]
  84. Goument, B.; Duhamel, L.; Maugé, R. Synthesis of (S)-fenfluramine from (R) or (S) 1-[3-(trifluoromethyl)phenyl]propan-2-ol. Bull. Soc. Chim. Fr. 1993, 130, 450–458. [Google Scholar]
  85. Van der Zeijden, A.A.H. A novel chiral cyclopentadienyl ligand based on ephedrine. Tetrahedron Asymmetry 1995, 6, 913–918. [Google Scholar] [CrossRef]
  86. Hyne, J.B. Preferred residence conformations of diastereoisomeric α-β amino alcohols: An, N.M.R. study of the ephedrines. Can. J. Chem. 1961, 39, 2536–2542. [Google Scholar] [CrossRef] [Green Version]
  87. Cheng, J.C.Y.; Hacksell, U.; Daves, G.D., Jr. Differentially Protected Ribofuranoid Glycals. J. Org. Chem. 1985, 50, 2778–2780. [Google Scholar] [CrossRef]
  88. Bae, J.H.; Shin, S.-H.; Park, C.S.; Lee, W.K. Preparation of cysteinol derivatives by highly regioselective ring opening of nonactivated chiral aziridines by thiols. Tetrahedron 1999, 55, 10041–10046. [Google Scholar] [CrossRef]
  89. Powers, J.C.; Asgian, J.L.; Ekici, Ö.D.; James, K.E. Irreversible inhibitors of serine, cysteine, and threonine proteases. Chem. Rev. 2002, 102, 4639–4750. [Google Scholar] [CrossRef] [PubMed]
  90. González, M.; Gándara, Z.; Pazos, G.; Gómez, G.; Fall, Y. Synthesis of (–)-Muricatacin from Tri-O-acetyl-D-glucal. Synthesis 2013, 45, 625–632. [Google Scholar] [CrossRef]
  91. Di Bussolo, V.; Caselli, M.; Pineschi, M.; Crotti, P. New Stereoselective β-Glycosylation via a Vinyl Oxirane Derived from D-Glucal. Org. Lett. 2002, 4, 3695–3698. [Google Scholar] [CrossRef]
  92. Kozikowski, A.P.; Lee, J. A synthetic approach to the cis-fused marine pyranopyrans,(3E)-and (3Z)-dactomelyne. X-ray structure of a rare organomercurial. J. Org. Chem. 1990, 55, 863–870. [Google Scholar] [CrossRef]
  93. Moore, P.W.; Schuster, J.K.; Stone, R.J.M.; Rhia, L.; Teesdale-Spittle, P.H.; Harvey, J.E. Divergent synthesis of 2-C-branched pyranosides and oxepines from 1, 2-gem-dibromocyclopropyl carbohydrates. Tetrahedron 2014, 70, 7032–7043. [Google Scholar] [CrossRef]
  94. Steunenberg, P.; Jeanneret, V.; Zhu, Y.-H.; Vogel, P. C (1→4)-linked disaccharides through carbonylative Stille cross-coupling. Tetrahedron Asymmetry 2005, 16, 337–346. [Google Scholar] [CrossRef]
  95. Linker, T.; Schanzenbach, D.; Elamparuthi, E. Remarkable Oxidation Stability of Glycals: Excellent Substrates for Cerium(IV)-Mediated Radical Reactions. J. Am. Chem. Soc. 2008, 130, 16003–16010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Literature examples of aminoethyl sulfides with biological properties. (a) PAES, or structurally similar derivatives; (bd) compounds containing an aminoethyl sul-fide moiety.
Figure 1. Literature examples of aminoethyl sulfides with biological properties. (a) PAES, or structurally similar derivatives; (bd) compounds containing an aminoethyl sul-fide moiety.
Molecules 27 01764 g001
Scheme 1. Synthesis of C-glycosyl-aminoethyl sulfide derivatives.
Scheme 1. Synthesis of C-glycosyl-aminoethyl sulfide derivatives.
Molecules 27 01764 sch001
Figure 2. (a) Coupling constants of erythro and threo isomers according to literature data [83,84,85,86]; (b) configuration of erythro and threo obtained isomers 912.
Figure 2. (a) Coupling constants of erythro and threo isomers according to literature data [83,84,85,86]; (b) configuration of erythro and threo obtained isomers 912.
Molecules 27 01764 g002
Table 1. Reaction of tributhyltin derivatives of glycals 6 and 7 with aziridinecarboaldehyde 8, according to Scheme 1.
Table 1. Reaction of tributhyltin derivatives of glycals 6 and 7 with aziridinecarboaldehyde 8, according to Scheme 1.
EntryGlycalProcedureYield (%) 1erythro:threo2
16A401:3
26B554:5
37A451:9
47B651:9
1 Isolated product. 2 Determined by 1H NMR analysis.
Table 2. Characteristic coupling constants and chemical shift values of diastereoisomers 916.
Table 2. Characteristic coupling constants and chemical shift values of diastereoisomers 916.
CompoundCHOH δ (ppm)CHN-CHOH J (Hz)CompoundCHOH δ (ppm)
erythro (S,S)-94.402.2(S,S)-134.30
threo (R,S)-103.925.8(R,S)-143.94
erythro (S,S)-114.332.4(S,S)-154.28
threo (R,S)-123.915.8(R,S)-164.04
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tracz, A.; Malinowska, M.; Leśniak, S.; Zawisza, A. Aziridine Ring Opening as Regio- and Stereoselective Access to C-Glycosyl-Aminoethyl Sulfide Derivatives. Molecules 2022, 27, 1764. https://doi.org/10.3390/molecules27061764

AMA Style

Tracz A, Malinowska M, Leśniak S, Zawisza A. Aziridine Ring Opening as Regio- and Stereoselective Access to C-Glycosyl-Aminoethyl Sulfide Derivatives. Molecules. 2022; 27(6):1764. https://doi.org/10.3390/molecules27061764

Chicago/Turabian Style

Tracz, Aleksandra, Martyna Malinowska, Stanisław Leśniak, and Anna Zawisza. 2022. "Aziridine Ring Opening as Regio- and Stereoselective Access to C-Glycosyl-Aminoethyl Sulfide Derivatives" Molecules 27, no. 6: 1764. https://doi.org/10.3390/molecules27061764

Article Metrics

Back to TopTop