Next Article in Journal
5,9,11-Trihydroxy-2,2-dimethyl-3-(2-methylbut-3-en-2-yl)pyrano[2,3-a]xanthen-12(2H)-one from the Stem Bark of Calophyllum tetrapterum Miq.
Previous Article in Journal
(5-Chloroquinolin-8-yl)-2-fluorobenzoate. The Halogen Bond as a Structure Director
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

Methyl 2,3,6-tri-O-Benzoyl-4-O-(tert-butyldimethylsilyl)-β-d-galactopyranoside

School of Chemistry, National University of Ireland, Galway H91 TK33, Ireland
*
Author to whom correspondence should be addressed.
Molbank 2017, 2017(1), M935; https://doi.org/10.3390/M935
Submission received: 22 February 2017 / Revised: 15 March 2017 / Accepted: 16 March 2017 / Published: 19 March 2017
(This article belongs to the Section Organic Synthesis)

Abstract

:
Methyl 2,3,6-tri-O-benzoyl-4-O-(tert-butyldimethylsilyl)-β-d-galactopyranoside was synthesized in 47% yield by the silylation of a partially benzoylated galactose derivative, prepared from methyl β-d-galactopyranoside. The product was characterized by 1H-NMR, 13C-NMR, IR and mass spectrometry.

1. Introduction

Galactose is a monosaccharide found in many glycans that cover a wide range of applications in biomedicine and in the bioeconomy [1,2]. Galectin-3, for example, plays a crucial role in many physiological processes, including cancer development and inflammation, by its recognition of β-galactoside derivatives. Thus, the synthesis of novel β-galactoside compounds as a means of inhibiting galectin-3 has been a growing area of research in recent years [3,4]. α-Galactosides are also important, such as α-galactosylceramides, which have been shown to be ligands for natural killer T-cells and to possess potent anti-tumour activity [5].
The conversion of β-glycosides to α-glycosides by anomerisation has been an interest in our group [6,7,8,9]. In the case of bioactive compounds containing a galactose monomer, improvements in the anomerisation reactions of galactose mono- and disaccharides could potentially lead to more efficient syntheses of such compounds from a wider range of starting materials. For this reason we have been exploring factors affecting reactivity to anomerisation, including the role of protecting groups, which includes the impact of silyl derivatives. In this context, we have prepared the title compound and provide the record of its synthesis and analytical data herein.

2. Results and Discussion

Methyl β-d-galactopyranoside (1) was initially protected at the 4- and 6-position (Scheme 1). This was carried out using benzaldehyde dimethyl acetal [10], forming methyl 4,6-O-benzylidene-β-d-galactopyranoside (2) as a white solid (67%). This product was then benzoylated at the 2- and 3-positions using benzoyl chloride in the presence of pyridine [11] to give methyl 4,6-O-benzylidene-2,3-di-O-benzoyl-β-d-galactopyranoside (3) as a white solid (50%). The benzylidene group was then removed using acetic acid in water (8:2 v/v) to give methyl 2,3-di-O-benzoyl-β-d-galactopyranoside (4) as a white solid (58%) [12]. Next, regioselective benzoylation gave methyl 2,3,6-tri-O-benzoyl-β-d-galactopyranoside (5) as an off-white solid (80%). In the final step, compound 5 was silylated at the 5-position using tert-butyldimethylsilyl trifluoromethanesulfonate (TBSOTf) [13] to afford the title compound methyl 2,3,6-tri-O-benzoyl-4-O-(tert-butyldimethylsilyl)-β-d-galactopyranoside (6) as a white foam (47%). 1H-, 13C- and gCOSY NMR spectra along with IR and mass spectrometry supported the structural assignment and gave a qualitative indication of its purity. The NMR spectra obtained for (6) are provided as Supplementary Materials.
All reactions (with the exception of the deprotection step) were carried out under an inert nitrogen atmosphere using anhydrous solvents to ensure thoroughly dry reaction conditions. All reactions were monitored by TLC.

3. Materials and Methods

3.1. General Information

All analytical data for previously reported compounds (25) was found to be in accordance with data reported previously in the literature, and citations are provided. All reagents used were obtained from commercial sources and used without further purification. TLC experiments were performed using aluminium sheets pre-coated with silica gel 60 (HF254, E. Merck, Darmstadt, Germany). NMR experiments were carried out in CDCl3 using a 500 MHz spectrometer (Varian Ltd., Palo Alto, CA, USA), with the chemical shifts reported relative to internal Me4Si. NMR spectra were analysed using MestReNova software (version 11, Mestrelab Research, Santiago de Compostela, Spain). The IR spectrum was obtained using FTIR Spectrometer (Perkin Elmer Spectrum 100, Shelton, CT, USA). A mass spectrum was obtained using a Waters LCT Premier XE Spectrometer. Chromatography was performed with silica gel 60 (Sigma Aldrich, Wicklow, Ireland). Petroleum ether was the fraction with boiling point 40–60 °C.

3.2. Methyl 4,6-O-Benzylidene-β-d-galactopyranoside (2)

Methyl β-d-galactopyranoside (1) (2.01 g, 10.4 mmol) was dissolved in dry acetonitrile (35 mL) under a nitrogen atmosphere. Benzaldehyde dimethyl acetal (3.11 mL, 20.8 mmol) was added, followed by camphorsulfonic acid (480 mg, 2.08 mmol). The reaction mixture was stirred under nitrogen for 30 min at 55 °C. The reaction was quenched with triethylamine (0.3 mL). The residue was concentrated to dryness under reduced pressure and then recrystallised from ethyl acetate (EtOAc), to give 1.95 g (67%) of compound 2 as a white solid. The 1H-NMR data of 2 was in accordance with data reported previously in the literature [10].

3.3. Methyl 4,6-O-Benzylidene-2,3-di-O-benzoyl-β-d-galactopyranoside (3)

Compound 2 (1.95 g, 6.88 mmol) was dissolved in dry pyridine (10 mL) under nitrogen. 4-Dimethylaminopyridine (10 g) was added next, and the solution was then cooled to 0 °C. Benzoyl chloride (4.81 mL, 41.41 mmol) was added slowly, and the mixture was left at room temperature for 24 h. The mixture was then diluted with dichloromethane, washed with hydrochloric acid (to react with excess pyridine), dried over Na2SO4 and then concentrated to dryness under reduced pressure. Column chromatography, using a mixture of petroleum ether–EtOAc (2:1) as the mobile phase afforded 1.68 g (50%) of compound 3 as a white solid. 1H-NMR spectroscopic data for the product was in agreement with data reported previously [14].

3.4. Methyl 2,3-di-O-Benzoyl-β-d-galactopyranoside (4)

Compound 3 (1.68 g, 3.42 mmol) was dissolved in AcOH–H2O (20 mL, 8:2 v/v), heated to 60 °C and left to react for 24 h. The mixture was then diluted with EtOAc and washed through with NaHCO3 (2 × 15 mL) and brine (15 mL). Column chromatography on silica gel using petroleum ether–EtOAc (1:2) as the mobile phase afforded 805 mg (58%) of compound 4 as a white solid. 1H-NMR spectroscopic data for the product was in agreement with data reported previously [15].

3.5. Methyl 2,3,6-tri-O-Benzoyl-β-d-galactopyranoside (5)

Compound 4 (805 mg, 2.00 mmol) was selectively benzoylated at the 6-position using benzoyl chloride (0.348 mL, 3.00 mmol) and the same method outlined previously. Column chromatography on silica gel using petroleum ether–EtOAc (3:1) as the mobile phase gave 807 mg (80%) of compound 5 as an off-white solid. 1H-NMR spectroscopic data for the product was in agreement with data reported previously [16].

3.6. Methyl 2,3,6-tri-O-Benzoyl-4-O-(tert-butyldimethylsilyl)-β-d-galactopyranoside (6)

Methyl 2,3,6-tri-O-benzoyl-β-d-galactopyranoside (5) (807 mg, 1.59 mmol) was dissolved in dry pyridine (10 mL) under nitrogen. TBSOTf (0.730 mL, 3.18 mmol) and 2,6-lutidine (0.370 mL, 3.18 mmol) were added via syringe at 0 °C. The reaction mixture was then stirred under nitrogen for 24 h at 60 °C. The solution was washed with excess 1 M HCl (2 × 25 mL), dried over Na2SO4 and then concentrated to dryness under reduced pressure. Column chromatography using petroleum ether–EtOAc (4:1) as the mobile phase afforded the title compound 6 (468 mg, 47%) as a white foam; 1H-NMR (500 MHz, CDCl3) δ 8.08–8.04 (m, 2H, aromatic H), 7.98–7.92 (m, 4H, aromatic H), 7.62–7.57 (m, 1H, aromatic H), 7.52–7.44 (m, 4H, aromatic H), 7.35 (m, 4H, aromatic H), 5.82 (dd, J = 10.5, 7.9 Hz, 1H, H-2), 5.24 (dd, J = 10.5, 2.8 Hz, 1H, H-3), 4.67 (d, J = 7.9 Hz, 1H, H-1), 4.63 (dd, J = 11.1, 6.5 Hz, 1H, H-6a), 4.46 (m, 1H, H-4), 4.42 (dd, J = 11.1, 6.7 Hz, 1H, H-6b), 4.05 (m, 1H, H-5), 3.55 (s, 3H), 0.96 (s, 9H), 0.01 (s, J = 2.9 Hz, 3H), −0.10 (s, 3H); 13C-NMR (126 MHz, CDCl3) δ 166.60 (C=O), 166.35 (C=O), 165.68 (C=O), 133.55 (C-Ar), 133.42 (C-Ar), 133.20 (C-Ar), 130.03 (C-Ar), 129.85 (C-Ar), 129.83 (C-Ar), 129.80 (C-Ar), 129.36 (C-Ar), 128.67 (C-Ar), 128.57 (C-Ar), 128.45 (C-Ar), 102.29 (CH, C-1), 75.25 (CH, C-3), 72.90 (CH, C-5), 69.26 (CH, C-2), 68.55 (CH, C-4), 63.14 (CH2, C-6), 56.78 (CH3-O), 26.01 (TBS), 18.45 (TBS), −4.39 (TBS), −4.46 (TBS); FT-IR: 2931, 2857, 1720, 1602, 1451, 1264, 1069, 832, 775, 706 cm−1; HRMS (TOF MS ES+) m/z: [M + Na]+ Calc. for C34H40O9SiNa+: 643.2334. Found: 643.2397.

Supplementary Materials

1H- and 13C-NMR spectra of compound 6 are available online.
Supplementary File 1Supplementary File 2Supplementary File 3Supplementary File 4

Acknowledgments

The authors thank Science Foundation Ireland for an IvP award (grant number 12/IA/1398) co-funded by the European Regional Development Fund (grant number 14/SP/2710). We also thank the School of Chemistry at NUI Galway for financial support.

Author Contributions

J.B. and A.R. conceived and designed the experiments; J.B. performed the experiments and drafted the manuscript; J.B., A.R. and P.M. analysed the data; P.M. is principal investigator and project director and contributed to the preparation of the manuscript; J.B. prepared the draft of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, R.; Jin, C.; Tong, Z.; Lu, J.; Tan, L.; Tian, L.; Chang, Q. Optimization Extraction, Characterization and Antioxidant Activities of Pectic Polysaccharide from Tangerine Peels. Carbohydr. Polym. 2016, 136, 187–197. [Google Scholar] [CrossRef] [PubMed]
  2. Choi, E.J.; Kim, J.W.; Kim, S.J.; Seo, S.O.; Lane, S.; Park, Y.C.; Jin, Y.S.; Seo, J.H. Enhanced Production of 2,3-Butanediol in Pyruvate Decarboxylase-Deficient Saccharomyces Cerevisiae through Optimizing Ratio of Glucose/galactose. Biotechnol. J. 2016, 11, 1424–1432. [Google Scholar] [CrossRef] [PubMed]
  3. Glinsky, V.V.; Kiriakova, G.; Glinskii, O.V.; Mossine, V.V.; Mawhinney, T.P.; Turk, J.R.; Glinskii, A.B.; Huxley, V.H.; Price, J.E.; Glinsky, G.V. Synthetic Galectin-3 Inhibitor Increases Metastatic Cancer Cell Sensitivity to Taxol-Induced Apoptosis In Vitro and In Vivo. Neoplasia 2009, 11, 901–909. [Google Scholar] [CrossRef] [PubMed]
  4. Öberg, C.T.; Leffler, H.; Nilsson, U.J. Arginine Binding Motifs: Design and Synthesis of Galactose-Derived Arginine Tweezers as Galectin-3 Inhibitors. J. Med. Chem. 2008, 51, 2297–2301. [Google Scholar] [CrossRef] [PubMed]
  5. Motoki, K.; Kobayashi, E.; Uchida, T.; Fukushima, H.; Koezuka, Y. Antitumour Activities of α-, β-Monoglycosylceramides and Four Diastereomers of an α-Galactosylceramide. Bioorg. Med. Chem. Lett. 1995, 5, 705–710. [Google Scholar] [CrossRef]
  6. Pilgrim, W.; O’Reilly, C.; Murphy, P.V. Synthesis of α-O- and α-S-Glycosphingolipids Related to Sphingomonous Cell Wall Antigens Using Anomerisation. Molecules 2013, 18, 11198–11218. [Google Scholar] [CrossRef] [PubMed]
  7. Pilgrim, W.; Murphy, P.V. SnCl4- and TiCl4-Catalyzed Anomerization of Acylated O- and S-Glycosides: Analysis of Factors That Lead to Higher α:β Anomer Ratios and Reaction Rates. J. Org. Chem. 2010, 75, 6747–6755. [Google Scholar] [CrossRef] [PubMed]
  8. O’Brien, C.; Polakova, M.; Pitt, N.; Tosin, M.; Murphy, P.V. Glycosidation-Anomerisation Reactions of 6,1-Anhydroglucopyranuronic Acid and Anomerisation of β-d-Glucopyranosiduronic Acids Promoted by SnCl4. Chem. Eur. J. 2007, 13, 902–909. [Google Scholar] [CrossRef] [PubMed]
  9. Farrell, M.; Zhou, J.; Murphy, P.V. Regiospecific Anomerisation of Acylated Glycosyl Azides and Benzoylated Disaccharides by Using TiCl4. Chem. Eur. J. 2013, 19, 14836–14851. [Google Scholar] [CrossRef] [PubMed]
  10. Angles d’Ortoli, T.; Widmalm, G. Synthesis of the Tetrasaccharide Glycoside Moiety of Solaradixine and Rapid NMR-Based Structure Verification Using the Program CASPER. Tetrahedron 2016, 72, 912–927. [Google Scholar] [CrossRef]
  11. Esmurziev, A.; Simic, N.; Sundby, E.; Hoff, B.H. 1H- and 13C-NMR data of methyl tetra-O-benzoyl-d-pyranosides in acetone-d6. Magn. Reson. Chem. 2009, 47, 449–452. [Google Scholar] [CrossRef] [PubMed]
  12. Abdu-Allah, H.H.M.; Tamanaka, T.; Yu, J.; Zhuoyuan, L.; Sadagopan, M.; Adachi, T.; Tsubata, T.; Kelm, S.; Ishida, H.; Kiso, M. Design, Synthesis, and Structure—Affinity Relationships of Novel Series of Sialosides as CD22-Specific Inhibitors. J. Med. Chem. 2008, 51, 6665–6681. [Google Scholar] [CrossRef] [PubMed]
  13. Durrat, F.; Texier-Nogues, I.; Robert, V. Saccharidic Fluorescent Substrates, Their Process of Preparation and Their Use. Patent No. 8,076,466, 13 December 2011. [Google Scholar]
  14. Pádár, P.; Bokros, A.; Paragi, G.; Forgó, P.; Kele, Z.; Howarth, N.M.; Kovács, L. Single Diastereomers of Polyhydroxylated 9-Oxa-1-azabicyclo[4.2.1]nonanes from Intramolecular 1,3-Dipolar Cycloaddition of ω-Unsaturated Nitrones. J. Org. Chem. 2006, 71, 8669–8672. [Google Scholar] [CrossRef] [PubMed]
  15. Kihlberg, J.; Frejd, T.; Jansson, K.; Sundin, A.; Magnusson, G. Synthetic Receptor Analogues: Preparation and Calculated Conformations of the 2-Deoxy, 6-O-Methyl, 6-Deoxy, and 6-Deoxy-6-Fluoro Derivatives of Methyl 4-O-α-d-Galactopyranosyl-β-d-Galactopyranoside (Methyl β-d-Galabioside). Carbohydr. Res. 1988, 176, 271–286. [Google Scholar] [CrossRef]
  16. Jiang, J.; Biggins, J.B.; Thorson, J.S. A General Enzymatic Method for the Synthesis of Natural and “Unnatural” UDP- and TDP-Nucleotide Sugars. J. Am. Chem. Soc. 2000, 122, 6803–6804. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of 6 from methyl β-d-galactopyranoside.
Scheme 1. Synthesis of 6 from methyl β-d-galactopyranoside.
Molbank 2017 m935 sch001

Share and Cite

MDPI and ACS Style

Bennett, J.; Roux, A.; Murphy, P.V. Methyl 2,3,6-tri-O-Benzoyl-4-O-(tert-butyldimethylsilyl)-β-d-galactopyranoside. Molbank 2017, 2017, M935. https://doi.org/10.3390/M935

AMA Style

Bennett J, Roux A, Murphy PV. Methyl 2,3,6-tri-O-Benzoyl-4-O-(tert-butyldimethylsilyl)-β-d-galactopyranoside. Molbank. 2017; 2017(1):M935. https://doi.org/10.3390/M935

Chicago/Turabian Style

Bennett, Jack, Amélie Roux, and Paul V. Murphy. 2017. "Methyl 2,3,6-tri-O-Benzoyl-4-O-(tert-butyldimethylsilyl)-β-d-galactopyranoside" Molbank 2017, no. 1: M935. https://doi.org/10.3390/M935

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