Next Article in Journal
Tris{4-[(1,3-diphenyl-1H-pyrazol-4-yl)methylene]-41-aminobiphenyl}amine
Previous Article in Journal
3-Hydroxy-4-{[(4-bromophenyl)imino]methyl}phenyl Octadecanoate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

(E)-2-((4R,5R)-5-((Benzyloxy)methyl)-2,2-dimethyl-1,3-dioxolan-4-yl)but-2-ene-1,4-diol

by
Carlos R. Carreras
1,
Celina E. García
2,
Víctor S. Martín
2,
Carlos E. Tonn
1,
David Díaz Díaz
3,4,* and
Juan Pedro Ceñal
1,*
1
INTEQUI-CONICET-Facultad de Química, Bioquímica y Farmacia, Universidad Nacional de San Luis, Chacabuco y Pedernera, 5700 San Luis, Argentina
2
Instituto Universitario de Bio-Orgánica “Antonio González”, Universidad de La Laguna, Avda. Astrofísico Francisco Sánchez, 2, 38206 La Laguna, Tenerife, Spain
3
Institut für Organische Chemie, Universität Regensburg, Universitätsstr. 31, 93040 Regensburg, Germany
4
ICMA, CSIC-Universidad de Zaragoza, Pedro Cerbuna 12, 50009 Zaragoza, Spain
*
Authors to whom correspondence should be addressed.
Molbank 2010, 2010(2), M676; https://doi.org/10.3390/M676
Submission received: 11 March 2010 / Accepted: 14 April 2010 / Published: 19 April 2010

Abstract

:
The synthesis of (E)-2-((4R,5R)-5-((benzyloxy)methyl)-2,2-dimethyl-1,3-dioxolan-4-yl)but-2-ene-1,4-diol by a one-step reduction of the appropriate 2-substituted butenolide is reported. Product characterization was carried out by IR, 1H NMR, 13C NMR, MS, elemental analysis and optical rotation.

Graphical Abstract

2-Alkylbut-2-ene-1,4-diols are present in a number of natural products and constitute highly versatile synthetic precursors [1,2,3,4,5,6,7]. For instance, biologically active 2-substituted 1,4-diacetoxy-butadienes [8,9,10], acyclic sesquiterpenoids and acyclic diterpenoids [11,12,13,14,15,16] can be easily synthesized from the above building blocks. From a synthetic point of view, 2-alkylbut-2-ene-1,4-diols are typically prepared, among different methods [17,18], by hydrostannation [9] or addition of Grignard reagents to but-2-ynediols [19].
Herein, we report the facile synthesis of (E)-2-((4R,5R)-5-((benzyloxy)methyl)-2,2-dimethyl-1,3-dioxolan-4-yl)but-2-ene-1,4-diol (3) by treatment of the corresponding 2-substituted butenolide 2 at 0 °C under reductive conditions, employing DIBAL-H as reducing agent (Scheme 1). Chiral precursor 2 was prepared by acid-mediated regioselective oxidation of the appropriate 3-substituted furan 1 as previously described in the literature [20]. All stereochemical assignments were further confirmed by differential nOe experiments (see Supplementary Files). Compound 3 represents a practical polyoxygenated chiral synthon in which each hydroxy group could be independently functionalized.

Experimental

General

1H and 13C NMR spectra were recorded at 25 °C on a Bruker Avance 500 spectrometer in CDCl3 as solvent, and chemical shifts are reported relative to Me4Si (δ = 0). Low- and high-resolution mass spectra were obtained by using a Micromass VG Autospec spectrometer. Elemental analysis was performed on a Fisons Instrument EA 1108 CHNS-O analyzer. Infrared spectra were recorded on a Bruker IFS 55 spectrophotometer for compounds dispersed on a NaCl disc. Optical rotations were determined for solutions in chloroform with a Perkin Elmer 343 polarimeter using a sodium lamp (589 nm). Thin-layer chromatography was carried out on Merck aluminium sheets coated with silica gel 60 F254. Compounds were visualized by use of 254 nm UV light and/or phosphomolybdic acid 20 wt.% solution in ethanol with heating. All solvents were purified by standard techniques [21]. Flash chromatography was performed on Merck silica gel 60 (0.040–0.063 mm, 230–400 mesh ASTM). Anhydrous magnesium sulfate was used for drying solutions.

Synthesis of (E)-2-((4R,5R)-5-((benzyloxy)methyl)-2,2-dimethyl-1,3-dioxolan-4-yl)but-2-ene-1,4-diol (3)

To a stirred solution of butenolide 2 (35 mg, 0.115 mmol) in dry Et2O (5 mL) was added dropwise diisobutylaluminium hydride (DIBAL-H) (0.5 mL of a 1M solution in cyclohexane, 0.5 mmol) under a nitrogen atmosphere at 0 °C. The reaction was allowed to continue at that temperature until no more starting material was detected by TLC analysis (1 h). Then H2O (0.2 mL) was added to the mixture under vigorous stirring to destroy excess DIBAL-H. MgSO4 was directly added to the reaction mixture and filtered through a short pad of Celite. The filtrate was washed several times with Et2O until no more product could be detected by TLC analysis. The solution was concentrated and the crude material was purified by silica gel column chromatography, eluting with AcOEt:n-hexane 30:70. Product 3 (24.7 mg, 71% yield) was obtained as a colourless oil: [α]25D = + 0.99 (c 1.60, CHCl3); 1H NMR (500 MHz, CDCl3) δ/ppm = 1.45 (s, 3H), 1.48 (s, 3H), 2.91 (brs, 2H), 3.61−3.66 (m, 2H), 4.03 (ddd, J = 4.4, 4.4, 8.8 Hz, 1H), 4.14−4.27 (m, 4H), 4.34 (d, J = 8.6 Hz, 1H), 4.57 (d, J = 12.0 Hz, 1H), 4.62 (d, J = 12.0 Hz, 1 H), 5.88 (dd, J = 6.5, 6.5 Hz, 1H), 7.29−7.38 (m, 5H); 13C NMR (125 MHz, CDCl3) δ/ppm = 26.9 (q), 27.0 (q), 57.2 (t), 58.3 (t), 69.5 (t), 73.7 (t), 78.8 (d), 82.0 (d), 109.3 (s), 127.8 (d), 128.4 (d), 132.5 (d), 137.6 (s), 137.7 (d), 137.8 (s); FT-IR (thin film) νmax (cm-1) 3400, 2986, 2930, 2870, 1374, 1219, 1086, 1020, 741; MS (EI+) m/z (relative intensity %) 293 [M–CH3]+ (5.8), 232 (3.6), 214 (2.8), 190 (3.0), 172 (8.6), 149 (6.9), 117 (53.4), 111 (21.9), 107 (38), 91.0 (100); HRMS (EI+) exact mass calculated for C16H21O5 (M–CH3)+ 293.1388, found 293.1389. Elemental analysis calculated for C17H24O5: C, 66.21; H, 7.84; found: C, 65.78; H, 8.05.

Supplementary materials

Supplementary File 1Supplementary File 2Supplementary File 3Supplementary File 4

Acknowledgements

This research was supported by the Spanish MICINN co-financed by the European Regional Development Fund (CTQ2008-06806-C02-01/BQU) and the Canary Islands Government, and projects PROIPRO 2/0006 and 7301 of San Luis University (UNSL), PIP 628 CONICET, and PIT 352-ANPCyT. D.D.D. is an Experienced Research Fellow of the Alexander von Humboldt Foundation. C.G. thanks the Spanish MICINN-FSE for a Ramón y Cajal contract.

References and Notes

  1. Gasparski, C.M.; Herrinton, P.M.; Overman, L.E.; Wolfe, J.P. Synthesis of 3-acyltetrahydrofurans from formaldehyde acetals of allylic diols. Tetrahedron Lett. 2000, 41, 9431–9435. [Google Scholar] [CrossRef]
  2. Commeiras, L.; Santelli, M.; Parrain, J.-L. First total synthesis of (±)-taxifolial a and (±)-iso-caulerpenyne. Org. Lett. 2001, 3, 1713–1715. [Google Scholar] [CrossRef] [PubMed]
  3. Murakami, H.; Matsui, Y.; Ozawa, F.; Yoshifuji, M. Cyclodehydration of cis-2-butene-1,4-diol with active methylene compounds catalyzed by a diphosphinidenecyclobutene-coordinated palladium complex. J. Organomet. Chem. 2006, 691, 3151–3156. [Google Scholar] [CrossRef]
  4. Miura, T.; Takahashi, Y.; Murakami, M. Rhodium-catalysed substitutive arylation of cis-allylic diols with arylboroxines. Chem. Commun. 2007, 595–597. [Google Scholar] [CrossRef] [PubMed]
  5. Clarke, P.A.; Rolla, G.A.; Cridland, A.P.; Gill, A.A. An improved synthesis of (2E,4Z)-6-(benzyloxy)-4-bromohexa-2,4-dien-1-ol. Tetrahedron 2007, 63, 9124–9128. [Google Scholar] [CrossRef] [Green Version]
  6. Musolino, M.G.; Apa, G.; Donato, A.; Pietropaolo, R.; Frusteri, F. Supported palladium catalysts for the selective conversion of cis-2-butene-1,4-diol to 2-hydroxytetrahydrofuran: Effect of metal particle size and support. Appl. Catal. A 2007, 325, 112–120. [Google Scholar] [CrossRef]
  7. Aponick, A.; Biannic, B. Gold-catalyzed dehydrative cyclization of allylic diols. Synthesis 2008, 3356–3359. [Google Scholar] [CrossRef]
  8. Commeiras, L.; Santelli, M.; Parrain, J.-L. total synthesis of (+) and (-)-furocaulerpin. Synlett 2002, 743–746. [Google Scholar] [CrossRef]
  9. Commeiras, L.; Parrain, J.-L. Concise enantioselective synthesis of furocaulerpin. Tetrahedron Asymmetry 2004, 15, 509–517. [Google Scholar] [CrossRef]
  10. Commeiras, L.; Bourdron, J.; Douillard, S.; Barbier, P.; Vanthuyne, N.; Peyrot, V. total synthesis of terpenoids isolated from caulerpale algae and their inhibition- of tubulin assembly. Synthesis 2006, 166–181. [Google Scholar]
  11. Commeiras, L.; Santelli, M.; Parrain, J.-L. On the construction of 2-substituted 1,4-diacetoxybutadiene moiety: Application to the synthesis of caulerpenyne. Tetrahedron Lett. 2003, 44, 2311–2314. [Google Scholar] [CrossRef]
  12. Commeiras, L.; Valls, R.; Santelli, M.; Parrain, J.-L. Efficient synthesis of (+/-)-dihydrorhipocephalin, a bioactive terpenoid from Caribbean marine algae of the Genera Penicillus and Udotea. Synlett 2003, 1719–1721. [Google Scholar]
  13. Gross, H.; Konig, G.M. Terpenoids from marine organisms: unique structures and their pharmacological potential. Phytochem. Rev. 2006, 5, 115–141. [Google Scholar] [CrossRef]
  14. Maimone, T.J.; Baran, P.S. Modern synthetic approaches to terpenes. Nat. Chem. Biol. 2007, 3, 396–407. [Google Scholar] [CrossRef] [PubMed]
  15. Gademann, K.; Portmann, C. Secondary metabolites from cyanobacteria: Complex structures and powerful bioactivities. Curr. Org. Chem. 2008, 12, 326–341. [Google Scholar] [CrossRef]
  16. Rezanka, T.; Siristova, L.; Sigler, K. Antiviral Sesqui-, Di- and Sesterterpenes. Anti-Infect. Agents Med. Chem. 2009, 8, 169–192. [Google Scholar] [CrossRef]
  17. Paul, M.; Domingo, R. Synthesis of 1H-Cyclopropa[b]naphthalenes via Trapping of o-benzoquinodimethanes. Helv. Chim. Acta 1985, 68, 975–980. [Google Scholar]
  18. Myriam, S.; Narciso, C.; Manuel, R.-C.; Ester, I.; Tore, D.; Denis, T.; Albert, B.; Michel, R. Isoprenoid biosynthesis in Escherichia coli via the methylerythritol phosphate pathway: Enzymatic conversion of methylerythritol cyclodiphosphate into a phosphorylated derivative of (E)-2-methylbut-2-ene-1,4-diol. Tetrahedron Lett. 2002, 43, 1413–1415. [Google Scholar]
  19. Yoshio, I.; Kohji, W.; Tsuneaki, H. A new synthesis of trans-2-Substituted-2-butene-1,4-diols from 2-Butyne-1,4-diol via nucleophilic addition of gragnard reagents. Chem. Lett. 1984, 5, 765–768. [Google Scholar]
  20. Ceñal, J.P.; Carreras, C.R.; Tonn, C.E.; Padrón, J.I.; Ramírez, M.A.; Díaz, D.D.; García-Tellado, F.; Martín, V.S. Acid-mediated highly regioselective oxidation of substituted furans: A simple and direct entry to substituted butenolides. Synlett 2005, 1575–1578. [Google Scholar]
  21. Armarego, W.L.F.; Perrin, D.D. Purification of Laboratory Chemicals, 4th ed.; Butterworth-Heinemann: Oxford, UK, 1996. [Google Scholar]
Scheme 1. Synthesis of 2-alkylbut-2-ene-1,4-diol 3.
Scheme 1. Synthesis of 2-alkylbut-2-ene-1,4-diol 3.
Molbank 2010 m676 sch001

Share and Cite

MDPI and ACS Style

Carreras, C.R.; García, C.E.; Martín, V.S.; Tonn, C.E.; Díaz, D.D.; Ceñal, J.P. (E)-2-((4R,5R)-5-((Benzyloxy)methyl)-2,2-dimethyl-1,3-dioxolan-4-yl)but-2-ene-1,4-diol. Molbank 2010, 2010, M676. https://doi.org/10.3390/M676

AMA Style

Carreras CR, García CE, Martín VS, Tonn CE, Díaz DD, Ceñal JP. (E)-2-((4R,5R)-5-((Benzyloxy)methyl)-2,2-dimethyl-1,3-dioxolan-4-yl)but-2-ene-1,4-diol. Molbank. 2010; 2010(2):M676. https://doi.org/10.3390/M676

Chicago/Turabian Style

Carreras, Carlos R., Celina E. García, Víctor S. Martín, Carlos E. Tonn, David Díaz Díaz, and Juan Pedro Ceñal. 2010. "(E)-2-((4R,5R)-5-((Benzyloxy)methyl)-2,2-dimethyl-1,3-dioxolan-4-yl)but-2-ene-1,4-diol" Molbank 2010, no. 2: M676. https://doi.org/10.3390/M676

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