Synlett 2008(1): 25-28  
DOI: 10.1055/s-2007-990921
LETTER
© Georg Thieme Verlag Stuttgart · New York

Intramolecular Ring-Opening Reactions of 1-(2-Methoxyphenyl)-6-oxabicyclo[3.2.0]heptanes: Spirocyclic Dihydrobenzofurans from Fused Bicyclic Oxetanes

Richard J. Boxalla, Richard S. Grainger*b, Caterina S. Aricòb, Leigh Ferrisc
a Department of Chemistry, King’s College London, Strand, London WC2R 2LS, UK
b School of Chemistry, University of Birmingham, Edgbaston, Birmingham B15 2TT, UK
Fax: +44(121)4144403; e-Mail: r.s.grainger@bham.ac.uk;
c AstraZeneca, Silk Road Business Park, Macclesfield, Cheshire, SK10 2NA, UK
Further Information

Publication History

Received 5 July 2007
Publication Date:
03 December 2007 (online)

Abstract

Treatment of fused oxetanes with Et2AlCl, TMSCl, acetyl chloride, or ethereal hydrochloric acid leads to the formation of spirocyclic dihydrobenzofurans through intramolecular attack of an oxygen atom of a proximal phenolic methyl ether.

    References and Notes

  • 1 Linderman RJ. In Comprehensive Heterocyclic Chemistry II   Vol. 1B:  Katritzky AR. Rees CW. Scriven EFV. Padwa A. Elsevier; Oxford: 1996. 
  • 2 Yamaguchi M. Hirao I. Tetrahedron Lett.  1984,  25:  4549 
  • 3a Bach T. Kather K. J. Org. Chem.  1996,  61:  7642 
  • 3b Bach T. Kather K. Krämer O. J. Org. Chem.  1998,  63:  1910 
  • 4a Ng FW. Lin H. Tan Q. Danishefsky SJ. Tetrahedron Lett.  2002,  43:  545 
  • 4b Ng FW. Lin H. Danishefsky SJ. J. Am. Chem. Soc.  2002,  124:  9812 
  • 5 Corey EJ. Raju N. Tetrahedron Lett.  1983,  24:  5571 
  • 6 Capek K. Vydra T. Sedmera P. Carbohydr. Res.  1987,  168:  c1 
  • 7 Khan N. Morris TH. Smith EH. Walsh R. J. Chem. Soc., Perkin Trans. 1  1991,  865 
  • 8 Itoh A. Hirose Y. Kashiwagi H. Masaki Y. Heterocycles  1994,  38:  2165 
  • In taxoids, the close proximity of the C-2 hydroxy group with the oxetane ring allows for a particularly facile intramolecular SN2 displacement at C-20 leading to the formation of tetrahydrofurans:
  • 9a Farina V. Huang S. Tetrahedron Lett.  1992,  33:  3979 
  • 9b Wahl A. Guéritte-Voegelein F. Guénard D. Le Goff M.-T. Potier P. Tetrahedron  1992,  48:  6965 
  • 9c Klein LL. Tetrahedron Lett.  1993,  34:  2047 
  • 9d Nicolaou KC. Renaud J. Nantermet PG. Couladouros EA. Guy RK. Wrasidlo W. J. Am. Chem. Soc.  1995,  117:  2409 
  • 9e Danishefsky SJ. Masters JJ. Young WB. Link JT. Snyder LB. Magee TV. Jung DK. Isaacs RCA. Bornmann WG. Alaimo CA. Coburn CA. Di Grandi MJ. J. Am. Chem. Soc.  1996,  118:  2843 
  • 9f Ojima I. Lin S. Inoue T. Miller ML. Borella CP. Geng X. Walsh JT. J. Am. Chem. Soc.  2000,  122:  5343 
  • 10 Bach T. Schröder J. J. Org. Chem.  1999,  64:  1265 
  • 11 Boxall RJ. Ferris L. Grainger RS. Synlett  2004,  2379 
  • 12 Grainger RS. Patel A. Chem. Commun.  2003,  1072 
  • 13 Kitagawa Y. Itoh A. Hashimoto S. Yamamoto H. Nozaki H. J. Am. Chem. Soc.  1977,  99:  3864 
  • 16 Joshi BV. Rao TS. Reese CB. J. Chem. Soc., Perkin Trans. 1  1992,  2537 
  • 17 Thurner A. Faigl F. Tõke L. Mordini A. Valacchi M. Reginato G. Czira T. Tetrahedron  2001,  57:  8173 
  • 18a Kočovský P. Pour M. J. Org. Chem.  1990,  55:  5580 
  • 18b Polt R. Wijayaratne T. Tetrahedron Lett.  1991,  32:  4831 
  • 19 Ring opening by chloride on activated 15 can lead to two possible regioisomeric chloroacetates. The stucture of 16 was confirmed by a long-range COSY experiment, and configuration at the chloro-substituted stereocenter was possible through NOESY correlation. The stereochemical outcome of this reaction suggests it also likely proceeds via an SN2-type ring opening of the activated ether with chloride nucleophile, where the regiochemical outcome is presumably a result of the preferred attack by chloride at a secondary carbon over a primary neopentyl-like carbon. For an analogous example in oxetane ring opening using HCl, see: Ceccherelli P. Curini M. Marcotullio MC. J. Chem. Soc., Perkin Trans. 1  1985,  2173 
  • 20a Benedetti MOV. Monteagudo ES. Burton G. J. Chem. Res., Synop.  1990,  248 
  • 20b Hernández R. Velázquez SM. Suárez E. J. Org. Chem.  1994,  59:  6395 
  • 20c Abad A. Agulló C. Cuñat AC. García AB. Giménez-Saiz C. Tetrahedron  2003,  59:  9523 
  • 21 Kleinwächter P. Schlegel B. Dörfelt H. Gräfe U. J. Antibiot.  2001,  54:  526 
  • 22 Srikrishna A. Lakshmi BV. Tetrahedron Lett.  2005,  46:  7029 
14

Typical Experimental for Oxetane Ring-Opening Reaction with HCl To a solution of oxetane 1a (50 mg, 0.18 mmol) in Et2O (5 mL) at 0 °C HCl (1 M in Et2O, 1.80 mL, 1.80 mmol) was added over a period of 10 min. The resulting solution was stirred for 18 h at r.t. The resulting mixture was poured into H2O (10 mL) and the aqueous layer was extracted with Et2O (2 × 10 mL). The combined organic phases were washed with brine (10 mL), dried over MgSO4, filtered and concentrated in vacuo. The crude product was purified via column chromatography (hexane-EtOAc, 8:2) to furnish spirocycle 8a as a colorless oil (37 mg, 78%). R f = 0.2 (hexane-EtOAc, 8:2); mp 88-90 °C. IR (neat): νmax = 3445, 2945, 2865, 1651, 1497, 1466, 1199 cm-1. 1H NMR (360 MHz, CDCl3): δ = 6.65 (1 H, s, CCHCOCH3), 6.58 (1 H, s, CH2OCCHCCH3), 5.07 (1 H, dd, J = 6.8, 1.8 Hz, CHOH), 3.93 (1 H, d, J = 11.0 Hz, COCH2C), 3.79 (3 H, s, ArOCH3), 3.65 (1 H, d, J = 11.0 Hz, COCH2C), 2.19 (3 H, s, ArCH3), 2.16-2.11 (1 H, m, HOCHCH 2CH2), 1.97-1.94 (1 H, m, HOCHCH 2CH2), 1.61-1.50 (1 H, m, HOCHCH2CH 2), 1.49-1.44 (1 H, m, HOCHCH2CH 2), 1.26 (1 H, br, OH), 1.07 [3 H, s, C(CH3)2], 1.03 [3 H, s, C(CH3)2]. 13C NMR (100 MHz, CDCl3): δ = 17.0 (q), 24.6 (q), 26.7 (q), 32.2 (t), 40.9 (t), 44.9 (s), 56.9 (q), 64.9 (s), 65.2 (t), 92.8 (d), 108.6 (d), 111.7 (d), 125.3 (s), 128.2 (s), 152.1 (s), 156.1 (s). ESI-HRMS: m/z calcd for C16H22O3Na: 285.1461; found: 285.1457; LRMS (EI): m/z (%) = 262 (100) [M+], 231 (5), 205 (33), 192 (6), 175 (33), 163 (9).

15

Typical Experimental Procedure for Oxetane Ring-Opening Reaction with AcCl To a solution of oxetane 1a (50 mg, 0.18 mmol) in DCE (10 mL) at r.t., AcCl (0.13 mL, 1.80 mmol) was added. The resulting solution was stirred overnight at r.t. poured into H2O (10 mL) and the aqueous layer was extracted with CH2Cl2 (2 × 10 mL). The combined organic phases were dried over MgSO4, filtered, and reduced in vacuo. The crude product was purified via column chromatography (hexane-EtOAc, 9:1) to furnish spirocycle 9a as a colorless oil (45 mg, 81%). R f = 0.2 (hexane-EtOAc, 9:1). IR (neat): νmax = 2950, 2869, 2252, 2105, 1739, 1651, 1490, 1464, 1223, 1047 cm-1. 1H NMR (360 MHz, CDCl3): δ = 6.66 (1 H, s, CCHCOCH3), 6.54 (1 H, s, CH2OCCHCCH3), 4.96 (1 H, dd, J = 6.8, 1.6 Hz, CHOCOCH3), 4.45 (1 H, d, J = 11.1 Hz, COCH2C), 4.16 (1 H, d, J = 11.2 Hz, COCH2C), 3.77 (3 H, s, ArOCH3), 2.18 (3 H, s, ArCH3), 2.15-2.12 (1 H, m, COCHCH 2CH2), 1.97 (3 H, s, OCOCH3), 1.99-1.91 (1 H, m, COCHCH 2CH2), 1.63-1.55 (1 H, m, COCHCH2CH 2), 1.52-1.48 (1 H, m, COCHCH2CH 2), 1.10 [3 H, s, C(CH3)2], 1.06 [3 H, s, C(CH3)2]. 13C NMR (125 MHz, CDCl3): δ = 16.5 (q), 20.9 (q), 24.6 (q), 26.4 (q), 31.9 (t), 39.6 (t), 44.9 (s), 56.4 (q), 61.7 (s), 67.1 (t), 92.3 (d), 109.0 (d), 111.2 (d), 125.9 (s), 127.3 (s), 151.4 (s), 154.8 (s), 171.0 (s). ESI-HRMS: m/z calcd for C18H24O4Na: 327.1567; found: 327.1562. LRMS (EI): m/z (%) = 304 (100) [M+], 262 (21), 175 (66), 160 (8), 115 (8), 91 (7), 69 (5), 43 (16).