Next Article in Journal
2-(4-Methylpiperazin-1-yl)-4-phenyl-6-(thiophen-2-yl)-pyridine-3-carbonitrile
Previous Article in Journal
2-[(1-Benzamido-2-methoxy-2-oxoethyl)amino]benzoic Acid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

Benzyl 3-deoxy-3-(3,4,5-trimethoxybenzylamino)-β-L-xylopyranoside

by
Fulgentius Nelson Lugemwa
Department of Chemistry, Pennsylvania State University-York, 1031 Edgecomb Avenue, York, PA 17403, USA
Molbank 2013, 2013(1), M793; https://doi.org/10.3390/M793
Submission received: 23 January 2013 / Accepted: 1 February 2013 / Published: 6 February 2013

Abstract

:
Reaction of 3,4,5-trimethoxybenzylamine and benzyl 2,3-anhydro-β-l-ribopyranoside in refluxing ethanol produced benzyl 3-deoxy-3-(3,4,5-trimethoxybenzylamino)-β-l-xylopyranoside in 72.5% yield. An attempt to synthesize the title compound by reacting neat 3,4,5-trimethoxybenzylamine with benzyl 2,3-anhydro-β-l-ribopyranoside without a solvent produced a dark brown mixture with several decomposition products. The structure of benzyl 3-deoxy-3-(3,4,5-trimethoxybenzylamino)-β-l-xylopyranoside was determined using elemental analysis, 1H-NMR and 13C-NMR and its conformation is 1C4.

Graphical Abstract

The readily available carbohydrates and their numerous, easily obtained derivatives are an unequaled source of enantiomerically pure starting materials for construction of complex asymmetric structures. But despite the greater awareness of carbohydrate synthons in recent years, the full potential of the carbohydrate chiral pool is still not fully exploited. A few simple carbohydrate analogs have been found to modulate biosynthesis of polysaccharides in live cells [1]. Specifically in those studies, benzyl 3-O-methyl-β-d-xylopyranoside and benzyl 3-deoxy-β-d-erythro-pentopyranoside were used to determine structural requirements for an enzyme involved in the biosynthesis of glycosaminoglycans. A comprehensive literature review revealed that the 3,4,5-trimethoxyphenyl moiety exists in a variety of bioactive compounds, including those with anticancer activity [2,3,4,5,6,7,8,9,10,11,12,13]. We took advantage of the facile and predictable ring opening of benzyl 2,3-anhydro-β-l-ribopyranoside with different nucleophiles to introduce 3,4,5-trimethoxyphenyl moiety at the sugar’s 3-position by reacting it with 3,4,5-trimethoxybenzylamine [14]. This led to the synthesis and structure determination of benzyl 3-deoxy-3-(3,4,5-trimethoxybenzylamino)-β-l-xylopyranoside 3.

Results and Discussion

Attempted synthesis of the title compound by heating a mixture of benzyl 2,3-anhydro-β-l-ribopyranoside and neat 3,4,5-trimethoxybenzylamine did not produce the desired product. The mixture turned brown and gave multiple spots after running thin layer chromatography. When the reaction was carried out in dimethylformamide, a polar aprotic solvent, the yield was low. Ethyl alcohol was found to be the best solvent for the reaction. The title compound was synthesized by opening the epoxide of benzyl 2,3-anhydro-β-l-ribopyranoside with 3,4,5-trimethoxybenzylamine (Scheme 1). The three broad peaks in the 1H-NMR due to one –NH at δ 2.20 ppm, and two –OH at δ 5.00 ppm and 5.26 ppm, disappeared upon D2O exchange. The chemical shifts of the sugar hydrogens, along with COSY and HMBC were used to assign C7, C1", C2", C3", C4", C5" and C7′ atoms. The coupling constant between H-1" and H-2" on the sugar ring was found to be 7.98 Hz, indicating that the protons at the 1- and 2-positions were in axial positions and that the molecule exists in solution in 1C4 conformation (Scheme 1). The coupling constant was similar to related analogs [14,15]. The coupling constant between H-2” and H-3” was found to be 9.12 Hz. The coupling constant between the pro-R and pro-S hydrogens on C7 was found to be 12.24 Hz. The 13C had five pairs of atoms with the same chemical shift. There were three pairs of carbon atoms on the 3,4,5-trimethoxybenzyl ring ( two ortho- and two meta-, and two equivalent methoxy groups) that had similar chemical shifts. On the benzyl group, chemical shifts of two pairs of carbon atoms (two ortho- and two meta-) were observed.

Experimental

General

1H and 13C-NMR spectra were obtained using a Varian Gemini 400 NMR and were recorded at 400 MHz and 100 MHz respectively. The composition of the reaction mixtures was monitored by TLC using glass plates coated with silica gel 60F254 -Analtech, Inc. 75 Blue Hen Drive, Newark, DE, USA; detection was effected by observation under UV light (254 nm), then spraying with 5% sulfuric acid in methanol and charring with heat. Elemental analyses were performed by the Robertson Microlit Laboratories Inc., Legdewood, NJ, USA. Optical rotations were measured using an Autopol IV polarimeter -Rudolf Research Analytical, 55 Newburgh Road, Hackettstown, NJ, USA. Melting points were determined using Mel-Temp Laboratory Devices, MA, USA, and are not corrected. All reagents and chemicals were obtained from Aldrich Chemical Company (Milwaukee, WI, USA) and were used without further purification.
Benzyl 2,3-anhydro-β-l-ribopyranoside (1) was obtained from l-arabinose in five steps using a previously reported synthetic route [14]. To a mixture of benzyl 2,3-anhydro-β-l-ribopyranoside 1 (0.15 g, 0.68 mmol) and 3,4,5-trimethoxybenzylamine 2 (180 mL, 0.91 mmol) was added ethyl alcohol (3 mL). After refluxing the mixture for 16 h and cooling at room temperature for 12 h, white crystals (needles) formed. Recrystallization from hexane/ethyl acetate mixture (3:2, v/v) produced a pure compound (0.206 g, 72%, m.p. 158–160 °C); [α]D26 +50° (c 1, CHCl3).
C22H29NO7
Calculated:C 62.99; H, 6.97; N, 3.34; O, 26.70
Found:C 62.89; H, 7.01; N, 3.29; O, 26.65
1H-NMR (400 MHz, Me2SO-d6) δ 2.20 (bs, 1H, –NH), 2.41 (t, J = 9.12, 7.98 Hz, 1H, H-3), 3.21 (m, 2H), 3.45 (bs, 1H), 3.65 (s, 3H, –OCH3), 3.75 (b, 1H), 3.80 (s, 6H, 2-OCH3), 3.97 (m, 2H), 4.31 (d, J = 7.98 Hz, 1H, H-1), 4.61 (d, J = 12.24 Hz, 1H, –OCH2Ar), 4.80 (d, J = 12.24 Hz, 1H, –OCH2Ar), 5.00 (bs, 1H, –OH), 5.26 (bs, 1H, –OH).
13C-NMR (100 MHz, Me2SO-d6), δ 53.22 (C-7′), 56.61 (–OCH3), 60.81 (–OCH3), 65.11 (C-3″), 67.37 (C-5″), 70.14 (C-4″), 70.41 (C-7), 73.00 (C-2″), 103.90 (C-1″), 105.80, 128.23, 128.40, 129.00, 136.80, 138.23, 138.93, 153.52.

Supplementary materials

Supplementary File 1Supplementary File 2Supplementary File 3

Acknowledgements

Funds from Pennsylvania State University-York Advisory Board Grant were partly used for this work.

References

  1. Lugemwa, F.N.; Sarkar, A.; Esko, J.D. Unusual β-d-xylosides that prime glycosaminoglycan in animal cells. J. Biol. Chem. 1996, 271, 19159–19165. [Google Scholar] [CrossRef] [PubMed]
  2. Lee, H.-Y.; Chang, J.-Y.; Nien, C.-Y.; Kuo, C.-C.; Shih, K.-H.; Wu, C.-H.; Chang, C.-Y.; Lai, W.-Y.; Liou, J.-P. 5-Amino-2 aroylquinolines as highly potent tubulin polymerization inhibitor. Part 2. The impact of bridging groups at position C-2. J. Med. Chem. 2011, 54, 8517–8525. [Google Scholar] [CrossRef] [PubMed]
  3. Stevenson, R.J.; Denny, W.A.; Tercel, M.; Pruijn, F.B.; Ashoorzadeh, A. Nitro seco analogues of the duocarmycins containing sulfonate leaving groups as hypoxia-activated prodrugs for cancer therapy. J. Med. Chem. 2012, 55, 2780–2802. [Google Scholar] [CrossRef] [PubMed]
  4. Xu, J.; Li, Z.; Luo, J.; Yang, F.; Liu, T.; Liu, M.; Qiu, W.-W.; Tang, J. Synthesis and biological evaluation of heterocyclic ring-fused betulinic acid derivatives as novel inhibitors of osteoclast differentiation and bone resorption. J. Med. Chem. 2012, 55, 3122–3134. [Google Scholar] [CrossRef] [PubMed]
  5. Bryan, M.C.; Whittington, D.A.; Doherty, E.M.; Falsey, J.R.; Chend, A.C.; Emkey, R.; Brake, R.L.; Lewis, R.T. Rapid development of piperidone carboxamides as potent and selective anaplastic lymphoma kinase inhibitors. J. Med. Chem. 2012, 55, 1698–1705. [Google Scholar] [CrossRef] [PubMed]
  6. Solinas, A.; Faure, H.; Roudaut, H.; Traiffort, E.; Schoenfelder, A.; Mann, A.; Manetti, F.; Taddei, M.; Ruat, M. Acylthiourea, acylurea, and acylguanidine derivatives with potent hedgehog inhibiting activity. J. Med. Chem. 2012, 55, 1559–1571. [Google Scholar] [CrossRef] [PubMed]
  7. Tripodi, F.; Pagliarin, R.; Fumagalli, G.; Bigi, A.; Fusi, P.; Orsini, F.; Frattini, M.; Coccetti, P. Synthesis and biological evaluation of 1,4-diaryl-2-azetidinones as specific anticancer Agents: Activation of adenosine monophosphate activated protein kinase and induction of apoptosis. J. Med. Chem. 2012, 55, 2112–2124. [Google Scholar] [CrossRef] [PubMed]
  8. Pellicani, R.Z.; Stefanachi, A.; Niso, M.; Carotti, A.; Leonetti, F.; Nicolotti, O.; Perrone, R.; Berardi, F.; Cellamare, S.; Colabufo, N.A. Potent galloyl-based selective modulators targeting multidrug resistance associated protein 1 and P-glycoprotein. J. Med. Chem. 2012, 55, 424–436. [Google Scholar] [CrossRef] [PubMed]
  9. Flynn, B.L.; Gill, G.S.; Grobelny, D.W.; Chaplin, J.H.; Paul, D.; Leske, A.F.; Lavranos, T.C.; Chalmers, D.K.; Charman, S.A.; Kostewicz, E.; et al. Discovery of 7-hydroxyl-6-methoxy-2-methyl-3-(3,4,5-trimethoxybenzoyl)benzo[b]furan (BNC105), a tubulin polymerization inhibitor with potent antiproliferative and tumor vascular disrupting properties. J. Med. Chem. 2011, 54, 6014–6027. [Google Scholar] [CrossRef] [PubMed]
  10. Theeramunkong, S.; Caldarelli, A.; Massarotti, A.; Aprile, S.; Caprioglio, D.; Zaninetti, R.; Teruggi, A.; Pirali, T.; Grosa, G.; Tron, G.C.; et al. Regioselective Suzuki coupling of dihaloheteroaromatic compounds as a rapid strategy to synthesize potent rigid combretastation analogues. J. Med. Chem. 2011, 54, 4977–4986. [Google Scholar] [CrossRef] [PubMed]
  11. Romagnoli, R.; Baraldi, P.G.; Brancale, A.; Ricci, A.; Hamel, E.; Bortolozzi, R.; Basso, G.; Viola, G. Convergent synthesis and biological evaluation of 2-amino-4(3′,4′,5′-trimethoxyphenyl)-5-aryl thiazoles as microtubule targeting agents. J. Med. Chem. 2011, 54, 5144–5153. [Google Scholar] [CrossRef] [PubMed]
  12. Lu, Y.; Li, C.-M.; Wang, Z.; Chen, J.; Mohler, M.L.; Li, W.; Dalton, J.T.; Miller, D.D. Design, synthesis, and SAR studies of 4-substituted methoxylbenzoyl-aryl-thiazoles analogues as potent and orally bioavailable anticancer agents. J. Med. Chem. 2011, 54, 4678–4693. [Google Scholar] [CrossRef] [PubMed]
  13. Combes, S.; Barbier, P.; Douillard, S.; McLeer-Florin, A.; Bourgarel-Rey, V.; Pierson, J.-T.; Fedorov, A.Y.; Finet, J.-P.; Boutonnat, J.; Peyrot, V. Synthesis and biological evaluation of 4-arylcoumarin analogues of combretastatins. Part 2. J. Med. Chem. 2011, 54, 3153–3162. [Google Scholar] [CrossRef] [PubMed]
  14. Lugemwa, F.N.; Denison, L. Synthesis of 3-substituted xylopyranosides from 2,3-anhydropentosides. J. Carbohydr. Chem. 1997, 16, 1433–1441. [Google Scholar] [CrossRef]
  15. Lugemwa, F.N.; Denison, L. Benzyl 3-deoxy-3-(phenethylamino) β-l-xylopyranoside. Molecules 1998, 3, M76. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of benzyl 3-deoxy-3-(3,4,5-trimethoxybenzylamino)-β-l-xylopyranoside.
Scheme 1. Synthesis of benzyl 3-deoxy-3-(3,4,5-trimethoxybenzylamino)-β-l-xylopyranoside.
Molbank 2013 m793 sch001

Share and Cite

MDPI and ACS Style

Lugemwa, F.N. Benzyl 3-deoxy-3-(3,4,5-trimethoxybenzylamino)-β-L-xylopyranoside. Molbank 2013, 2013, M793. https://doi.org/10.3390/M793

AMA Style

Lugemwa FN. Benzyl 3-deoxy-3-(3,4,5-trimethoxybenzylamino)-β-L-xylopyranoside. Molbank. 2013; 2013(1):M793. https://doi.org/10.3390/M793

Chicago/Turabian Style

Lugemwa, Fulgentius Nelson. 2013. "Benzyl 3-deoxy-3-(3,4,5-trimethoxybenzylamino)-β-L-xylopyranoside" Molbank 2013, no. 1: M793. https://doi.org/10.3390/M793

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