Next Article in Journal
Dimethyl 2-{[2-(2-Methoxy-1-methoxycarbonyl-2-oxoethyl)-4,5,7-trimethoxy-3-(2,4,5-trimethoxyphenyl)-2,3-dihydro-1H-inden-1-yl]methyl}malonate
Next Article in Special Issue
Synthesis and Isolation of Diastereomeric Anomeric Sulfoxides from a d-Mannuronate Thioglycoside Building Block
Previous Article in Journal
2-(Bicyclo[4.2.0]octa-1,3,5-trien-3-yl)-adamantan-2-ol
Previous Article in Special Issue
N′-Acetyl-3-methyl-1,6-diphenyl-1H-pyrazolo[3,4-b]pyridine-4-carbohydrazide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Unexpected Formation of Oxetanes during the Synthesis of Dodeco-6,7-diuloses

by
Marius Bayer
,
Cäcilia Maichle-Mössmer
and
Thomas Ziegler
*
Institute of Organic Chemistry, University of Tuebingen, Auf der Morgenstelle 18, 72076 Tuebingen, Germany
*
Author to whom correspondence should be addressed.
Molbank 2020, 2020(1), M1108; https://doi.org/10.3390/M1108
Submission received: 11 December 2019 / Revised: 8 January 2020 / Accepted: 9 January 2020 / Published: 14 January 2020
(This article belongs to the Collection Molecules from Side Reactions)

Abstract

:
During the synthesis of symmetrical dodeco-6,7-diuloses that are potential candidates for inhibition of glycosidases, an unanticipated epoxide-oxetane rearrangement was observed. A bicyclic sugar consisting of a glycal moiety and an anomeric esterified furanose was oxidized under epoxidation conditions (mCPBA/KF). The isolation of the pure epoxide was not possible since a rapid reversible conversion accompanied by the migration of the ester group took place and resulted in the formation of an unusual oxetane-bridged disaccharide scaffold. X-ray diffractometric structure elucidation and the suggested mechanism of the rearrangement are provided.

Graphical Abstract

1. Introduction

Oxetane rings attached directly to a carbohydrate unit rarely occur in literature, due to steric reasons [1]. The most common representatives are 3,5-anhydrofuranose derivatives, which are formed by nucleophilic [2,3] or additive [4] intramolecular ring closure reactions. In addition to the furanoid 3,5-linked oxetanes, there also exists a limited number of 1,2-fused or anomeric spiro-oxetanes having a pyranoid constitution. These highly strained scaffolds are prepared photochemically [5,6,7] or via template-directing C-glycosylations [8,9]. In synthesis, sugar oxetanes are used as versatile precursors since the rigid conformation and steric repulsion of the bridge favor high stereoselectivities [10]. Furthermore, the detection of antimicrobial properties has brought these structures closer to the focus of current research [4]. In this communication, we introduce a new method for the generation of anomerically bridged oxetanes whose formation was initiated by an unexpected rearrangement of an epoxide assisted by an adjacent ester group.

2. Results and Discussion

Starting from the known compounds 1 [11] and 2 [12] a base-induced coupling of two anomeric centers analogous to Shiozaki’s protocol [13] was performed (Scheme 1). Contrary to Shiozaki’s procedure, which involved the use of toxic stannylated glycals, we could employ 1-phenylsulfinyl glycals since it is known that they are easily lithiated in a similar fashion with one equivalent of phenyllithium [14]. Sulfoxide 1 was treated with phenyllithium at –78 °C followed by quenching with manno-lactone 2 after 5 minutes. The observance of this reaction time is mandatory since a longer reaction time results in the irreversible formation of the protonated glycal species in increasing amounts leading to by-products. The coupled disaccharide 3 was isolated as an anomeric mixture (α/β ratio 1:5) in 80% yield. Next, the anomeric hydroxyl group was masked in order to avoid its coordinating effect on the envisaged stereoselective epoxidation. For this purpose, we decided to choose the ester protecting groups acetyl and benzoyl. Thus, the fully protected disaccharides 4a and 4b were obtained after the acylation step in 96% and 68% yield, respectively. In both cases solely the β-anomers were obtained which was proven by H,H-NOESY NMR spectroscopy (see Supplementary Materials).
Furthermore, the configuration could be confirmed by growing crystals of 5a suitable for X-ray diffraction by recrystallization of the latter from n-hexane and ethyl acetate. Compound 5a was obtained through desilylation of 4a with tetra-n-butylammonium fluoride (Scheme 1). The moderate yield of 71% can be explained by detectable partial cleavage of the acetyl group in alkaline milieu during desilylation. The benzoyl protecting group was significantly more stable under the same conditions resulting in a yield of 89% for the conversion of 4b to 5b. In our efforts to synthesize symmetrical dodeco-6,7-diuloses both sugar moieties need to be manno-configured. Therefore, the next step was to stereoselectively convert the glycal entity into a mannose derivative. Conventional methods for the implementation of stereoselective dihydroxylation on similar allyl alcohol systems use substrate-directing reagents such as molybdenum catalysts [15], vanadyl acetylacetonate (VO(acac)2) [16,17] or meta-chloroperbenzoic acid (mCPBA) [17,18] for epoxidation. Subsequent epoxide opening leads to the desired diol derivatives. When these methodologies were applied to disaccharides 5a and 5b the best results were achieved using the Camps reagent (mCPBA/KF) [18,19]. The addition of potassium fluoride reduces the solubility of mCPBA and mCBA and thus, inhibiting nucleophilic epoxide opening by the acids. However, we could not isolate the pure epoxides 6a and 6b for a rapid rearrangement of the latter to oxetane species 7a and 7b occurred. Upon early quenching of the epoxidation reaction, only 8:1 mixtures of epoxide and oxetane could be isolated in yields of 94% for 6a and 90% for 6b, respectively. In order to enable the complete conversion of the epoxide to the oxetane derivatives the reaction times were extended and 7a and 7b could be obtained neatly in 70% and 63% yields, respectively. Crystals of compound 7a suitable for X-ray crystallography could be obtained by overlaying a saturated solution of 7a in methylene chloride with n-heptane and slowly evaporating the methylene chloride. Thus, the oxetane-bridged disaccharide structure of 7a with two anomeric α-configurations could be unambiguously verified (Scheme 1). For the formation of oxetanes 7 we propose a mechanism deduced from the X-ray data in which ester group migration from position 7 to 6 occurs simultaneously with O-heterocyclic ring extension of the oxirane ring to the less strained oxetane ring (Figure 1). The rearrangement process can also be visualized by 1H-NMR spectroscopy since the H-5 signal of oxetane 7a (4.96 ppm, 4.6 Hz) is significantly shifted to lower field compared to the characteristic doublet of H-5 in epoxide 6a (3.66 ppm, 2.4 Hz) (Figure 1). To some extent, this rearrangement resembles the Ferrier rearrangement of acetylated glycals. Alternatively, a stepwise mechanism via the intermediate formation of carbenium ions can also be envisaged.
It must also be mentioned that the rearrangement appeared to be reversible for in attempts to recrystallize 7a from boiling n-hexane the precipitation of an amorphous substance which turned out to be epoxide 6a was observed. Studies are ongoing to evaluate both the effect of the temperature and of the solvent on the equilibrium ratio between compounds 6a and 7a. After removal of the ester protecting groups of 7a and 7b with ammonia in methanol, oxetane 8 was obtained from both compounds in 99% and 80% yield respectively (Scheme 1). Oxetanes 8 did not recede to the corresponding epoxides according to TLC monitoring. Consequently, it was proven that 7b is also present as an oxetane-fused disaccharide.

3. Materials and Methods

All reactions were performed under an atmosphere of nitrogen using solvents dried by standard procedures. Reaction progress was monitored by TLC on Polygram SIL G/UV254 silica gel plates from Macherey and Nagel, Germany. Detection of spots was affected by carbonizing with sulfuric acid (5% in EtOH), staining by spraying the plates with an alkaline aqueous solution of potassium permanganate or by inspection of the TLC plates under UV light (254 nm). Preparative flash chromatography was performed on silica gel (0.032–0.063 mm) from Macherey-Nagel, using plastic cartridges from Götec. The flowrate was regulated by a Sykam S1122 solvent delivery system. Nuclear magnetic resonance (NMR) spectra were recorded with a Bruker Avance 400 spectrometer and calibrated for the solvent signal (1H: CDCl3: δ = 7.26 ppm; acetone-d6: δ = 2.05 ppm; DMSO-d6: δ = 2.50 ppm; 13C: CDCl3: δ =77.16 ppm; acetone-d6: δ = 29.92 ppm; DMSO-d6: δ = 39.52 ppm). All NMR-assignments were proven by 2D-experiments to be correct. ESI-TOF-HRM spectrometry was performed on a Bruker MAXIS 4G spectrometer. Elemental analyses were obtained from a HEKA tech Euro EA 3000 apparatus. Optical rotations were determined with a Perkin-Elmer Polarimeter 341 in a 10 cm cuvette at 20 °C with a wavelength of 589 nm (Na-lamp). Melting points were measured with a Büchi Melting Point M-560 apparatus.

2,6-Anhydro-5-deoxy-4-O-(tert-butyldimethylsilyl)-1,3:8,9:11,12-tri-O-isopropylidene-d-arabino-l-gulo-dodeco-7-ulo-6-enitol (3)

To a suspension of 1 (2.62 g, 6.18 mmol) and grounded molecular sieve (3 Å, 30 mg) in THF (40 mL) was added phenyllithium (1.9 M in Bu2O, 3.57 mL, 6.79 mmol) at −78 °C over a period of 15 min. After 5 min a further solution containing 2 (1.75 g, 6.79 mmol) in THF (20 mL) was added dropwise over a period of 30 min. Subsequently, the reaction mixture was stirred for 2 h at the same temperature before the cooling bath was removed. The reaction was quenched by addition of sat. aq. NH4Cl-solution (60 mL) at ambient temperature and the molecular sieve was filtered off. The organic layer was separated, the aqueous layer was extracted with CH2Cl2 (3 × 20 mL) and the combined organic layers were dried with Na2SO4. Removing of the solvent followed by column chromatography (methylene chloride/ethyl acetate 5:1) furnished 3 (2.77 mg, 4.97 mmol, 80%, anomeric mixture α:β, 1:5) as a colorless crystalline solid. Rf = 0.22–0.46 (methylene chloride/ethyl acetate 5:1).
3β*(major anomer): 1H-NMR (400 MHz, CDCl3): δ(ppm) = 5.04 (d, J5,4 = 2.2 Hz, 1H, H-5), 4.82 (dd, J9,8 = 5.7 Hz, J9,10 = 3.5 Hz, 1H, H-9), 4.53 (d, J8,9 = 5.9 Hz, 1H, H-8), 4.38–4.44 (m, 1H, H-11), 4.33–4.35 (m, 1H, H-4), 4.15 (dd, J10,11 = 7.2 Hz, J10,9 = 3.5 Hz, 1H, H-10), 4.07–4.11 (m, 1H, H-12a), 4.03 (dd, J = 8.7 Hz, J = 4.9 Hz, 1H, H-12b), 3.86–3.95 (m, 2H, H-1a, H-1b), 3.71–3.76 (m, 2H, H-2, H-3), 2.77 (s, 1H, OH), 1.50, 1.44, 1.42, 1.40, 1.38, 1.31 (6s, 18H, C(CH3)2), 0.87 (s, 9H, SiC(CH3)3), 0.07, 0.06 (2s, 6H,SiCH3). 13C-NMR (101 MHz, CDCl3): δ(ppm) = 150.4 (C-6), 113.3, 109.1 (C(CH3)2), 104.3 (C-5), 103.4 (C-7), 99.4 (C(CH3)2), 86.7 (C-8), 80.0 (C-9), 79.4 (C-10), 73.2 (C-11), 72.9 (C-3), 70.2 (C-2), 68.2 (C-4), 66.7 (C-12), 61.7 (C-1), 28.9, 26.8, 25.8 (C(CH3)2), 25.7 (SiC(CH3)3), 25.4, 25.0, 19.1 (C(CH3)2), 18.1 (SiC(CH3)3), –4.4, –4.8 (SiCH3).
3α*(minor anomer): 1H-NMR (400 MHz, CDCl3): δ(ppm) = 5.07 (d, J5,4 = 2.1 Hz, 1H, H-5), 4.79 (dd, J9,8 = 6.0 Hz, J9,10 = 3.7 Hz, 1H, H-9), 4.62 (d, J8,9 = 6.0 Hz, 1H, H-8), 4.34–4.35 (m, 2H, H-4, H-11), 4.10–4.11 (m, 1H, H-12a), 4.01–4.02 (m, 1H, H-12b), 3.85–3.88 (m, 2H, H-1a, H-1b, 3.76–3.80 (m, 3H, H-2, H-3, H-10), 2.77 (s, 1H, OH), 1.57, 1.49, 1.42, 1.40, 1.36 (5s, 18H, C(CH3)2), 0.88 (s, 9H, SiC(CH3)3), 0.08, 0.07 (2s, 6H, SiCH3). 13C-NMR (101 MHz, CDCl3): δ(ppm) = 150.8 (C-6), 113.6, 109.3 (C(CH3)2), 103.1 (C-5), 100.6 (C-7), 99.6 (C(CH3)2), 80.7 (C-8), 80.0 (C-9), 78.6 (C-10), 73.2 (C-11), 72.8 (C-3), 70.3 (C-2), 67.8 (C-4), 67.2 (C-12), 61.7 (C-1), 28.9, 26.9, 25.9 (C(CH3)2), 25.8 (SiC(CH3)3), 25.2, 24.6, 19.0 (C(CH3)2), 18.2 (SiC(CH3)3), –4.4, –4.8 (SiCH3). HRESIMS m/z 581.27570 (calcd for C27H46O10SiNa, 581.27524); anal. C 58.37, H 8.21, calcd for C27H46O10Si, C 58.04, H 8.39. *Anomers 3β and 3α could be interchanged

2,6-Anhydro-5-deoxy-7-O-acetyl-4-O-(tert-butyldimethylsilyl)-1,3:8,9:11,12-tri-O-isopropylidene-β-d-manno-d-lyxo-dodeco-6-enitol (4a)

A solution of 3 (300 mg, 0.537 mmol) and 4-dimethylaminopyridine (13 mg, 0.107 mmol) in triethylamine (7 mL) was treated with acetic anhydride (508 μL, 5.37 mmol) at 0 °C. After stirring for 12 h at ambient temperature until TLC indicated complete transformation of the starting material, the reaction mixture was diluted with ethyl acetate (20 mL). The organic phase was washed with 1 N hydrochloric acid (3 × 10 mL), sat. aq. NH4Cl-solution (1 × 10 mL) and brine (1 × 10 mL). The solution was dried with Na2SO4, the solvent was removed and the crude product was purified by column chromatography (petroleum ether/ethyl acetate 5:1) to afford 4a (309 mg, 0.515 mmol, 96%) as a colorless crystalline solid. Rf = 0.82 (methylene chloride/ethyl acetate 5:1). [ α ] D 20 = +64.3 (c = 1.0, CHCl3). M.p. 52 °C (n-hexane). 1H-NMR (400 MHz, CDCl3): δ (ppm) = 4.99 (d, J5,4 = 2.2 Hz, 1H, H-5), 4.86 (dd, J9,8 = 5.6 Hz, J9,10 = 3.4 Hz, 1H, H-9), 4.69 (d, J8,9 = 5.6 Hz, 1H, H-8), 4.42 (ddd, J = 8.3 Hz, J = 5.9 Hz, J = 3.9 Hz, 1H, H-11), 4.35 (dd, J4,3 = 6.7 Hz, J4,5 = 2.1 Hz, 1H, H-4), 4.09–4.15 (m, 1H, H-12a), 3.99–4.07 (m, 2H, H-10, H-12b), 3.80–3.89 (m, 2H, H-1aH-1b), 3.61–3.73 (m, 2H, H-2, H-3), 2.01 (s, 3H, COCH3), 1.48, 1.47, 1.45, 1.37, 1.37, 1.33 (6s, 18H, C(CH3)2), 0.83 (s, 9H, SiC(CH3)3), 0.03, 0.03 (2s, 6H, SiCH3).13C-NMR (101 MHz, CDCl3): δ(ppm) = 168.3 (CO), 146.9 (C-6), 113.9, 109.4 (C(CH3)2), 107.4 (C-7), 104.8 (C-5), 99.3 (C(CH3)2), 86.4 (C-8), 81.2 (C-10), 79.6 (C-9), 73.0 (C-3), 72.8 (C-11), 70.2 (C-2), 68.2 (C-4), 67.1 (C-12), 61.7 (C-1), 28.9, 27.0, 25.8 (C(CH3)2), 25.6 (SiC(CH3)3), 25.1, 25.1 (C(CH3)2), 21.5 (COCH3), 19.1 (C(CH3)2), 18.0 (SiC(CH3)3), –4.4, –4.9 (SiCH3). HRESIMS m/z 623.28604 (calcd for C29H48O11SiNa, 623.28581); anal. C 58.10, H 8.21, calcd for C29H48O11Si, C 57.98, H 8.05.

2,6-Anhydro-5-deoxy-7-O-benzoyl-4-O-(tert-butyldimethylsilyl)-1,3:8,9:11,12-tri-O-isopropylidene-β-d-manno-d-lyxo-dodeco-6-enitol (4b)

A solution of 3 (930 mg, 1.67 mmol) and DMAP (41 mg, 0.333 mmol) in pyridine (15 mL) was treated with benzoyl chloride (1.92 mL, 16.7 mmol) at 0 °C. After stirring for 48 h at ambient temperature TLC indicated complete transformation of the starting material and the reaction mixture was diluted with ethyl acetate (60 mL). The organic layer was washed with 1 N hydrochloric acid (3 × 30 mL), sat. aq. NaHCO3-solution (3 × 20 mL) and brine (1 × 20 mL). The solution was dried with Na2SO4, the solvent was removed and the crude product was purified by column chromatography (toluene/ethyl acetate, gradient: 100:1 – 20:1) to afford 4b (749 mg, 1.13 mmol, 68%) as a colorless crystalline solid. Rf = 0.66 (toluene/ethyl acetate 4:1). [ α ] D 20 = +57.1(c = 1.0, CHCl3). M.p. 58 °C. 1H-NMR (400 MHz, CDCl3): δ(ppm) = 7.90–8.02 (m, 2H, Ph), 7.52–7.62 (m, 1H, Ph), 7.39–7.48 (m, 2H, Ph), 5.11 (d, J5,4 = 2.3 Hz, 1H, H-5), 4.90–5.01 (m, 2H, H-8, H-9), 4.47 (ddd, J = 8.2 Hz, J = 6.0 Hz, J11,12b = 4.0 Hz, 1H, H-11), 4.41 (dd, J4,3 = 6.7 Hz, J4,5 = 2.2 Hz, 1H, H-4), 4.08–4.17 (m, 2H, H-10, H-12a), 4.05 (dd, J12b,12a = 9.1 Hz, J12b,11 = 4.0 Hz, 1H, H-12b), 3.66–3.84 (m, 4H, H-1a, H-1b, H-2, H-3), 1.52, 1.48, 1.42, 1.38, 1.37 (5s, 18H, C(CH3)2), 0.86 (s, 9H, SiC(CH3)3), 0.06, 0.06 (2s, 6H, SiCH3).13C-NMR (101 MHz, CDCl3): δ(ppm) = 164.0 (CO), 147.3 (C-6), 133.4, 130.2, 129.9, 128.5 (Ph), 114.2, 109.6 (C(CH3)2), 108.2 (C-7), 105.1 (C-5), 99.5 (C(CH3)2), 86.7 (C-8), 81.8 (C-10), 80.0 (C-9), 73.3 (C-3), 73.0 (C-11), 70.4 (C-2), 68.5 (C-4), 67.3 (C-12), 61.8 (C-1), 29.1, 27.1, 26.0 (C(CH3)2), 25.8 (SiC(CH3)3), 25.4, 25.3, 19.3 (C(CH3)2), 18.2 (SiC(CH3)3), –4.2, –4.7 (SiCH3). HRESIMS m/z 685.30154 (calcd for C34H50O11SiNa, 685.30146); anal. C 61.26, H 8.02, calcd for C34H50O11Si, C 61.61, H 7.60.

General Procedure A for Desilylation of 4a and 4b

Tetra-n-butylammonium fluoride (1.0 M in THF, 1.5 eq.) was added to a solution of 4a or 4b (1.0 mmol) in THF (10 mL) at 0 °C. The reaction mixture was warmed to room temperature and stirring was continued until complete conversion of the starting material was detected by TLC. Subsequently, the solution was diluted with CH2Cl2 (20 mL) and washed with brine (3 × 10 mL). The aqueous phase was extracted with CH2Cl2 (2 × 20 mL) and the combined organic phases were dried with Na2SO4. The solvent was evaporated and the residue was purified by column chromatography (petroleum ether/ethyl acetate 2:1 containing 0.5% Et3N).

2,6-Anhydro-5-deoxy-7-O-acetyl-1,3:8,9:11,12-tri-O-isopropylidene-β-d-manno-d-lyxo-dodeco-6-enitol (5a)

Following general procedure A 5a was obtained as a white amorphous solid and was prepared from 4a in 71% yield stirring the reaction mixture for 20 h. Rf = 0.28 (petroleum ether/ethyl acetate 1:1). [ α ] D 20 = +98.0 (c = 1.0, CHCl3). M.p. 179 °C (n-hexane/ethyl acetate). 1H-NMR (400 MHz, CDCl3): δ(ppm) = 5.14 (d, J5,6 = 2.2 Hz, 1H, H-5), 4.88 (dd, J9,8 = 5.7 Hz, J9,10 = 3.5 Hz, 1H, H-9), 4.71 (d, J8,9 = 5.7 Hz, 1H, H-8), 4.37–4.45 (m, 2H, H-4, H-11), 4.01–4.12 (m, 3H, H-10, H-12a, H-12b), 3.85–3.94 (m, 2H, H-1a, H-1b), 3.68–3.80 (m, 2H, H-2, H-3), 2.02 (s, 3H, COCH3), 1.53, 1.47, 1.45, 1.42, 1.37, 1.33 (6s, 18H, C(CH3)2). 13C-NMR (101 MHz, CDCl3): δ (ppm) = 168.2 (CO), 148.0 (C-6), 113.7, 109.4 (C(CH3)2), 107.1 (C-7), 102.9 (C-5), 99.7 (C(CH3)2), 86.3 (C-8), 81.5 (C-10), 79.6 (C-9), 73.2 (C-3), 72.8 (C-11), 69.9 (C-2), 67.8 (C-4), 67.0 (C-12), 61.5 (C-1), 28.9, 27.0, 25.8, 25.1, 24.9 (C(CH3)2), 21.6 (COCH3), 19.2 (C(CH3)2). HRESIMS m/z 509.19928 (calcd for C23H34O11Na, 509.19933); anal. C 56.57, H 7.17, calcd for C23H34O11, C 56.78, H 7.04.

2,6-Anhydro-5-deoxy-7-O-benzoyl-1,3:8,9:11,12-tri-O-isopropylidene-β-d-manno-d-lyxo-dodeco-6-enitol (5b)

Following general procedure A 5b was obtained as a colorless amorphous solid and was prepared from 4b in 89% yield stirring the reaction mixture for 24 h. Rf = 0.37 (petroleum ether/ethyl acetate 1:1). [ α ] D 20 = +89.4 (c = 1.0, CHCl3). 1H-NMR (400 MHz, CDCl3): δ(ppm) = 7.92–7.99 (m, 2H, Ph), 7.54–7.60 (m, 1H, Ph), 7.41–7.47 (m, 2H, Ph), 5.26 (d, J5,4 = 2.2 Hz, 1H, H-5), 4.99 (dd, J9,8 = 5.9 Hz, J9,10 = 3.6 Hz, 1H, H-9), 4.95 (d, J8,9 = 5.9 Hz, 1H, H-8), 4.41–4.50 (m, 2H, H-4, H-11), 4.09–4.14 (m, 2H, H-10, H-12a), 4.04 (dd, J = 9.1 Hz, J = 4.2 Hz, 1H, H-12b), 3.72–3.88 (m, 4H, H-1a, H-1b, H-2, H-3), 2.11 (br. s., 1H, OH), 1.53, 1.52, 1.42, 1.41, 1.38, 1.36 (6 s, 18H, C(CH3)2).13C-NMR (101 MHz, CDCl3): δ(ppm) = 163.9 (CO), 148.4 (C-6), 133.5, 130.1, 129.9, 128.6 (Ph), 114.0, 109.6 (C(CH3)2), 107.9 (C-7), 103.2 (C-5), 99.9 (C(CH3)2), 86.6 (C-8), 82.0 (C-10), 79.9 (C-9), 73.4 (C-3), 73.0 (C-11), 70.2 (C-2), 68.1 (C-4), 67.2 (C-12), 61.6 (C-1), 29.1, 27.1, 26.0, 25.2, 25.1, 19.4 (C(CH3)2). HRESIMS m/z 571.21543 (calcd for C28H36O11Na, 571.21498); anal. C 61.18, H 6.77, calcd for C28H36O11, C 61.30, H 6.61.

General Procedure B for Epoxidation of 5a and 5b and Further Rearrangement to the Oxetanes

To a solution of mCPBA (2.5 eq.) in CH2Cl2 (10 mL) was added potassium fluoride (5 eq.) and the resulting suspension was stirred for 30 min at ambient temperature. A solution containing 5a or 5b (0.5 mmol) in CH2Cl2 (10 mL) was added in one portion and stirring was continued for 2 h (epoxidation). To enforce the rearrangement to the oxetanes the reaction mixture was stirred for 20 h, the white precipitate was filtered off and the solution was stirred for further 20 h. The reaction was quenched by addition of sat. aq. NaHCO3-solution (10 mL) and sat. aq. Na2S2O3-solution (10 mL). Subsequently, the aqueous layer was extracted by CH2Cl2 (3 × 10 mL), the combined organic layers were dried with Na2SO4, the solvent was removed and the crude product was purified by column chromatography.

5,6-Anhydro-7-O-acetyl-1,3:8,9:11,12-tri-O-isopropylidene-β-d-manno-β-d-manno-dodeco-6,7-diulo-2,6-pyranose-7,10-furanose (6a)

Following general procedure B a mixture of 6a and 7a (8:1) was obtained after quenching the reaction after 2 h as a white amorphous solid and was prepared from 5a in 94% yield. A mixture of petroleum ether/ethyl acetate 3:1 containing 0.5% Et3N was used as the eluent for column chromatography. Rf = 0.58 (petroleum ether/ethyl acetate 1:2). 1H-NMR (400 MHz, CDCl3): δ(ppm) = 4.96 (d, J8,9 = 6.0 Hz, 1H, H-8), 4.91 (dd, J9,8 = 6.0 Hz, J9,10 = 3.9 Hz, 1H, H-9), 4.31–4.38 (m, 1H, H-11), 4.20 (dd, J10,11 = 6.6 Hz, J10,9 = 3.9 Hz, 1H, H-10), 3.85–4.06 (m, 5H, H-1a, H-3, H-4, H-12a, H-12b), 3.77 (dd, J = 10.7 Hz, J = 10.7 Hz, 1H, H-1b), 3.66 (d, J5,4 = 2.4 Hz, 1H, H-5), 3.49 (ddd, J = 10.2 Hz, J = 10.2 Hz, J = 5.7 Hz, 1H, H-2), 2.38 (br. s., 1H, OH), 2.06 (s, 3H, COCH3), 1.57, 1.48, 1.42, 1.39, 1.35, 1.33 (6s, 18H, C(CH3)2). 13C-NMR (101 MHz, CDCl3): δ(ppm) = 169.2 (CO), 114.2, 109.2 (C(CH3)2), 107.9 (C-7), 99.7 (C(CH3)2), 86.2 (C-8), 85.0 (C-6), 82.9 (C-10), 79.9 (C-9), 73.2 (C-11), 73.0 (C-3), 70.6 (C-2), 69.9 (C-4), 66.4, 61.7 (C-1), 57.3 (C-5), 29.1, 27.0, 25.6, 25.3, 24.6 (C(CH3)2), 21.9 (COCH3), 19.3 (C(CH3)2). HRESIMS m/z 525.19448 (calcd for C23H34O12Na, 525.19425).

5,6-Anhydro-7-O-benzoyl-1,3:8,9:11,12-tri-O-isopropylidene-β-d-manno-β-d-manno-dodeco-6,7-diulo-2,6-pyranose-7,10-furanose (6b)

Following general procedure B, a mixture of 6b and 7b (8:1) was obtained after quenching the reaction after 2 h as a white amorphous solid and was prepared from 5b in 90% yield. A mixture of petroleum ether/ethyl acetate 3:1 containing 0.5% Et3N was used as the eluent for column chromatography. Rf = 0.64 (petroleum ether/ethyl acetate 1:2). 1H-NMR (400 MHz, CDCl3): δ(ppm) = 7.92–7.99 (m, 2H, Ph), 7.59–7.63 (m, 1H, Ph), 7.45–7.50 (m, 2H, Ph), 5.18 (d, J8,9 = 6.0 Hz, 1H, H-8), 5.03 (dd, J9,8 = 6.1 Hz, J9,10 = 4.0 Hz, 1H, H-9), 4.34–4.41 (m, 1H, H-11), 4.31 (dd, J10,11 = 6.6 Hz, J10,9 = 3.9 Hz, 1H, H-10), 4.13–4.18 (m, 1H, H-4), 4.03 (dd, J = 8.9 Hz, J = 6.2 Hz, 1H, H-12a), 3.91–3.99 (m, 2H, H-3, 12b), 3.73–3.83 (m, 3H, H-1a, H-1b, H-5), 3.54 (ddd, J = 10.0 Hz, J = 10.0 Hz, J = 6.0 Hz, 1H, H-2), 2.42 (br. s., 1H, OH), 1.62, 1.48, 1.39, 1.37, 1.34 (5s, 18H, C(CH3)2). 13C-NMR (101 MHz, CDCl3): δ(ppm) = 164.7 (CO), 133.8, 129.9, 129.8, 128.8 (Ph), 114.3, 109.2 (C(CH3)2), 108.5 (C-7), 99.7 (C(CH3)2), 86.4 (C-8), 85.3 (C-6), 83.2 (C-10), 80.0 (C-9), 73.2 (C-11), 73.0 (C-3), 70.7 (C-2), 69.9 (C-4), 66.4 (C-12), 61.6 (C-1), 57.4 (C-5), 29.0, 26.9, 25.6, 25.3, 24.6, 19.3 (C(CH3)2). HRESIMS m/z 587.21029 (calcd for C28H36O12Na, 587.20990).

5,7-Anhydro-6-O-acetyl-1,3:8,9:11,12-tri-O-isopropylidene-β-d-manno-β-d-manno-dodeco-6,7-diulo-2,6-pyranose-7,10-furanose (7a)

Following general procedure B 7a was obtained after quenching the reaction after 40 h as a white crystalline solid and was prepared from 5a in 70% yield. A mixture of petroleum ether/ethyl acetate 2:1 was used as the eluent for column chromatography. Crystals from 7a were grown by overlaying a saturated solution of 7a in methylene chloride with n-heptane and slowly evaporating the methylene chloride. Rf = 0.51 (petroleum ether/ethyl acetate 1:2). [ α ] D 20 = +18.5 (c = 1.0, CHCl3). M.p. 151 °C (n-hexane/ethyl acetate). 1H-NMR (400 MHz, CDCl3): δ(ppm) = 5.04 (d, J8,9 = 5.6 Hz, 1H, H-8), 4.96 (d, J4,5 = 4.9 Hz, 1H, H-5), 4.70 (dd, J9,8 = 5.6 Hz, J9,10 = 4.0 Hz, 1H, H-9), 4.37–4.46 (m, 1H, H-11), 4.00–4.13 (m, 3H, H-3, H-12a, H12b), 3.86–3.97 (m, 3H, H-1a, H-4, H-10), 3.80 (dd, J = 10.5 Hz, 1H, H-1b), 3.65–3.74 (m, 1H, H-2), 2.14 (d, JOH,4 = 9.0 Hz, 1H, OH), 2.10 (s, 3H, COCH3), 1.56, 1.55, 1.43, 1.43, 1.39, 1.36 (6s, 18H, C(CH3)2). 13C-NMR (101 MHz, CDCl3): δ(ppm) = 169.4 (CO), 114.2 (C(CH3)2), 112.9 (C-7), 109.5, 99.9 (C(CH3)2), 98.3 (C-6), 80.9 (C-5), 80.0 (C-10), 79.1 (C-8), 78.6 (C-9), 73.5 (C-11), 71.6 (C-3), 69.1 (C-2), 68.9 (C-4), 67.1 (C-12), 61.8 (C-1), 29.1, 27.1, 25.9, 25.8, 25.4 (C(CH3)2), 21.2 (COCH3), 19.2 (C(CH3)2). HRESIMS m/z 525.19318 (calcd for C23H34O12Na, 525.19425); anal. C 55.08, H 7.11, calcd for C23H34O12, C 54.97, H 6.82.

5,7-Anhydro-6-O-benzoyl-1,3:8,9:11,12-tri-O-isopropylidene-β-d-manno-β-d-manno-dodeco-6,7-diulo-2,6-pyranose-7,10-furanose (7b)

Following general procedure B 7b was obtained after quenching the reaction after 40 h as a white amorphous solid and was prepared from 5b in 63% yield. A mixture of petroleum ether/ethyl acetate 2:1 was used as the eluent for column chromatography. Rf = 0.60 (petroleum ether/ethyl acetate 1:2). [ α ] D 20 = +15.7 (c = 1.0, CHCl3). 1H-NMR (400 MHz, acetone-d6): δ(ppm) = 8.04–8.10 (m, 2H, Ph), 7.68–7.74 (m, 1H, Ph), 7.53–7.59 (m, 2H, Ph), 5.16 (d, J8,9 = 5.6 Hz, 1H, H-8), 5.10 (d, J5,4 = 4.5 Hz, 1H, H-5), 4.82 (dd, J9,8 = 5.6 Hz, J9,10 = 3.7 Hz, 1H, H-9), 4.36 (ddd, J = 6.5 Hz, J = 6.5 Hz, J = 5.1 Hz, 1H, H-11), 4.15–4.26 (m, 2H, H-3, OH), 3.99–4.06 (m, 3H, H-4, H-10, H-12a), 3.80–3.97 (m, 4H, H-1a, H-1b, H-2, H-12b), 1.52, 1.51, 1.37, 1.35, 1.34, 1.28 (6s, 18H, C(CH3)2). 13C-NMR (101 MHz, acetone-d6): δ(ppm) = 165.7 (CO), 135.0, 130.8, 130.2, 129.7 (Ph), 114.2 (C(CH3)2), 113.5 (C-7), 109.5, 100.2 (C(CH3)2), 99.5 (C-6), 83.3 (C-5), 80.1 (C-10), 79.6 (C-9), 79.6 (C-8), 74.5 (C-11), 72.1 (C-3), 70.7 (C-2), 69.4 (C-4), 67.1 (C-12), 62.4 (C-1), 29.5, 27.2, 26.3, 26.2, 25.7, 19.3 (C(CH3)2). HRESIMS m/z 587.20977 (calcd for C28H36O12Na, 587.20990); anal. C 59.64, H 6.86, calcd for C28H36O12, C 59.57, H 6.43.

5,7-Anhydro-1,3:8,9:11,12-tri-O-isopropylidene-β-d-manno-β-d-manno-dodeco-6,7-diulo-2,6-pyranose-7,10-furanose (8)

From 7a: 7a (50 mg, 0.10 mmol) was dissolved in methanol (2 mL) and ammonia (7 N in methanol, 280 μL, 1.99 mmol) was added at ambient temperature. TLC indicated complete transformation of the starting material after 5 h and the volatile components were removed. Neat 8 (46 mg, 0.099 mmol, 99%) remained as a white crystalline solid.
From 7b: 7b (20 mg, 0.0354 mmol) was dissolved in methanol (1 mL) and ammonia (7 N in methanol, 101 μL, 0.708 mmol) was added at ambient temperature. TLC indicated complete transformation of the starting material after 5 h and the volatile components were removed. Subsequently, the crude product was purified by column chromatography (petroleum ether/ethyl acetate 1:2) and 8 (13 mg, 0.282 mmol, 80%) was isolated as a white crystalline solid.
Rf = 0.32 (petroleum ether/ethyl acetate 1:3). [ α ] D 20 = +17.5 (c = 1.0, CHCl3). M.p. 122 °C (n-hexane). 1H-NMR (600 MHz, DMSO-d6): δ(ppm) = 7.09 (s, 1H, OH-6), 5.16 (d, JOH,4 = 5.7 Hz, 1H, OH-4), 4.75 (d, J8,9 = 5.9 Hz, 1H, H-8), 4.57–4.63 (m, 2H, H-5, H-9), 4.22 (ddd, J11,10 = 7.3 Hz, J11,12a = 6.4 Hz, J11,12b = 5.1, 1H, H-11), 4.01 (dd, J12a,12b = 8.4 Hz, J12a,11 = 6.4 Hz, 1H, H-12a), 3.92 (dd, J12b,12a = 8.4 Hz, J12b,11 = 5.1 Hz, 1H, H-12b), 3.78–3.85 (m, 2H, H-1a, H-3), 3.76 (dd, J10,11 = 7.3 Hz, J10,9 = 3.7 Hz, 1H, H-10), 3.68 (dd, J1b,1a = 10.5 Hz, J1b,2 = 10.5 Hz, 1H, H-1b), 3.52 (ddd, J = 9.7 Hz, J4,OH = 5.1 Hz, J = 4.4 Hz, 1H, H-4), 3.40 (ddd, J2,1b = 9.9 Hz, J = 9.9 Hz, J = 5.8 Hz, 1H, H-2), 1.45, 1.39, 1.32, 1.31, 1.28, 1.26 (6s, 18H, C(CH3)2).13C-NMR (101 MHz, DMSO-d6): δ(ppm) = 112.5 (C-7), 112.3, 108.2, 98.8 (C(CH3)2), 97.4 (C-6), 83.9 (C-5), 78.2 (C-9), 77.0 (C-8), 76.9 (C-10), 72.7 (C-11), 71.1 (C-3), 67.8 (C-4), 66.8 (C-2), 66.1 (C-12), 61.1 (C-1), 28.9,26.6, 25.8, 25.5, 25.2, 19.0 (C(CH3)2). HRESIMS m/z 483.18377 (calcd for C21H32O11Na, 483.18368); anal. C 54.58, H 7.41, calcd for C21H32O11, C 54.78, H 7.00.

4. Conclusions

In summary, we have demonstrated that the stability of 2,5-anhydro sugars depends on their chemical environment. In the case of epoxide 6a and 6b, where an ester group is adjacent to the epoxide group, reversible rearrangements to highly annulated oxetane-bridged disaccharides 7a and 7b occur. In further studies, these compounds serve as precursors for the synthesis of symmetrical dodeco-6,7-diuloses.

Supplementary Materials

The following are available online, 1H-NMR and 13C-NMR spectra of all compounds, H,H-NOESY-NMR spectra of compounds 4a and 4b and X-ray data of compounds 5a and 7a.

Author Contributions

Synthesis, conceptualization, investigation, writing—original draft preparation, formal analysis, M.B.; X-ray crystallography, C.M.-M.; conceptualization, investigation, writing—review and editing, validation, supervision, T.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

We thank Petra Schülzle (elemental analysis), Dorothee Wistuba, Claudia Kruse and Peter Haiss (mass spectra) for their contributions to this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hadzic, P.; Popsavin, M.; Borozan, S. Alkylating ability of carbohydrate oxetanes: Practical synthesis of bolaform skeleton derivative. J. Serb. Chem. Soc. 2015, 80, 1273–1278. [Google Scholar] [CrossRef]
  2. Kawana, M.; Kuzuhara, H.; Emoto, S. The Use of Grignard Reagents in the Synthesis of Carbohydrates. III. The One-way Anomerization of Methyl Glycofuranosides and the Opening of Their Furanose Rings. Bull. Chem. Soc. Jpn. 1981, 54, 1492–1504. [Google Scholar] [CrossRef]
  3. Hadzic, P.; Vukojevic, N.; Popsavin, M.; Canadi, J. Nucleophilic opening of the 3,5-anhydro ring in 1,2-O-cyclohexylidene-d-xylofuranose. J. Serb. Chem. Soc. 2001, 66, 1–8. [Google Scholar] [CrossRef]
  4. Telli, F.Ç.; Ay, K.; Murat, G.; Kök, G.; Salman, Y. Acid promoted intramolecular formation of 3,5-anhydro-1,4-furano-7-ulose derivatives via the Wittig-cyclization procedure and their antimicrobial properties. Med. Chem. Res. 2013, 22, 2253–2259. [Google Scholar] [CrossRef]
  5. Ong, K.-S.; Whistler, R.L. Photochemical addition of acetone to d-glucal triacetate and subsequent oxetane ring cleavage. J. Org. Chem. 1972, 37, 572–574. [Google Scholar] [CrossRef]
  6. Brunckova, J.; Crich, D. Intramolecular hydrogen atom abstraction: The β-oxygen effect in the Norrish type II photoreaction. Tetrahedron 1995, 51, 11945–11952. [Google Scholar] [CrossRef]
  7. Lichtenthaler, F.W.; Weimer, T.; Immel, S. [4+2] and [2+2] Photocycloadditions of 1,2-diketones to glycal and hydroxyglycal esters. Tetrahedron Asymmetry 2004, 15, 2703–2709. [Google Scholar] [CrossRef]
  8. Craig, D.; Tierney, J.P.; Williamson, C. Template-directed intramolecular C-glycosidation. Stereoselective synthesis of bicyclic ketooxetanes from anomeric sulfones. Tetrahedron Lett. 1997, 38, 4153–4156. [Google Scholar] [CrossRef]
  9. Craig, D.; Munasinghe, V.R.N.; Tierney, J.P.; White, A.J.P.; Williams, D.J.; Williamson, C. Template-directed intramolecular C-glycosidation. cation-mediated synthesis of ketooxetanes from thioglycosides. Tetrahedron 1999, 55, 15025–15044. [Google Scholar] [CrossRef]
  10. Schulze, O.; Voss, J.; Adiwidjaja, G. Convenient preparation of 3,5-anhydro- and 2,5-anhydropentofuranosides, and 5,6-anhydro-d-glucofuranose by use of the Mitsunobu reaction. Carbohydr. Res. 2005, 340, 587–595. [Google Scholar] [CrossRef] [PubMed]
  11. Bayer, M.; Bächle, F.; Ziegler, T. Synthesis and Pd-catalyzed coupling of 1-C-stannylated glycals. J. Carbohydr. Chem. 2018, 37, 347–369. [Google Scholar] [CrossRef]
  12. Oddo, A.; Holl, R. Design and stereoselective synthesis of a C-aryl furanoside as a conformationally constrained CHIR-090 analogue. Carbohydr. Res. 2012, 359, 59–64. [Google Scholar] [CrossRef] [PubMed]
  13. Mochizuki, T.; Shiozaki, M. Synthesis of (6R,7R)-d-gluco-l-gulo-6,7-dodecodiulose-(6,2),(7,11). Chem. Lett. 1997, 26, 801–802. [Google Scholar] [CrossRef]
  14. Jarowicki, K.; Kilner, C.; Kocienski, P.J.; Komsta, Z.; Milne, J.E.; Wojtasiewicz, A.; Coombs, V. A synthesis of 1-lithiated glycals and 1-tributylstannyl glycals from 1-phenylsulfinyl glycals via sulfoxide-lithium ligand exchange. Synthesis 2008, 2008, 2747–2763. [Google Scholar]
  15. Marín, I.; Matheu, M.I.; Díaz, Y.; Castillón, S. Stereoselective Tandem Epoxidation-Alcoholysis/Hydrolysis of Glycals with Molybdenum Catalysts. Adv. Synth. Catal. 2010, 352, 3407–3418. [Google Scholar] [CrossRef]
  16. Borowski, D.; Zweiböhmer, T.; Ziegler, T. 1,2-Annulated Sugars: Synthesis of Polyhydroxylated 2,10-Dioxadecalins with β-manno Configuration. Eur. J. Org. Chem. 2016, 2016, 5248–5256. [Google Scholar] [CrossRef]
  17. Borowski, D.; Maichle-Mössmer, C.; Ziegler, T. Synthetic Adventures with 2-C-Branched Carbohydrates: 4-C-Formyl Branched Octoses with Structural Analogy to Bradyrhizose. Eur. J. Org. Chem. 2019, 2019, 2653–2670. [Google Scholar] [CrossRef]
  18. Bellucci, G.; Catelani, G.; Chiappe, C.; D’Andrea, F. A simple and highly diastereoselective preparation of glycal epoxides using the MCPBA-KF complex. Tetrahedron Lett. 1994, 35, 8433–8436. [Google Scholar] [CrossRef]
  19. Camps, F.; Coll, J.; Messeguer, A.; Pujol, F. m-Chloroperoxybenzoic acid-potassium fluoride system: Study of its stability and reaction with. alpha.-methylstyrene. J. Org. Chem. 1982, 47, 5402–5404. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of the isopropylidene-protected oxetane-bridged disaccharide 8 and X-ray structures of 5a and 7a.
Scheme 1. Synthesis of the isopropylidene-protected oxetane-bridged disaccharide 8 and X-ray structures of 5a and 7a.
Molbank 2020 m1108 sch001
Figure 1. Proposed mechanisms for the rearrangement from epoxide 6a to oxetane 7a and their 1H-NMR spectra in CDCl3 to illustrating the reaction progress.
Figure 1. Proposed mechanisms for the rearrangement from epoxide 6a to oxetane 7a and their 1H-NMR spectra in CDCl3 to illustrating the reaction progress.
Molbank 2020 m1108 g001

Share and Cite

MDPI and ACS Style

Bayer, M.; Maichle-Mössmer, C.; Ziegler, T. Unexpected Formation of Oxetanes during the Synthesis of Dodeco-6,7-diuloses. Molbank 2020, 2020, M1108. https://doi.org/10.3390/M1108

AMA Style

Bayer M, Maichle-Mössmer C, Ziegler T. Unexpected Formation of Oxetanes during the Synthesis of Dodeco-6,7-diuloses. Molbank. 2020; 2020(1):M1108. https://doi.org/10.3390/M1108

Chicago/Turabian Style

Bayer, Marius, Cäcilia Maichle-Mössmer, and Thomas Ziegler. 2020. "Unexpected Formation of Oxetanes during the Synthesis of Dodeco-6,7-diuloses" Molbank 2020, no. 1: M1108. https://doi.org/10.3390/M1108

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