Synlett 2017; 28(02): 270-274
DOI: 10.1055/s-0036-1588354
letter
© Georg Thieme Verlag Stuttgart · New York

Copper(I)-Catalyzed Enantioselective Boryl Substitution of Allyl Acylals: An Efficient Approach for Enantioenriched α-Chiral γ-Acetoxyallylboronates

Yuta Takenouchi
Division of Applied Chemistry and Frontier Chemistry Center, Faculty of Engineering, Hokkaido University, Sapporo, Hokkaido, 060-8628, Japan   Email: hajito@eng.hokudai.ac.jp
,
Ryoto Kojima
Division of Applied Chemistry and Frontier Chemistry Center, Faculty of Engineering, Hokkaido University, Sapporo, Hokkaido, 060-8628, Japan   Email: hajito@eng.hokudai.ac.jp
,
Riko Momma
Division of Applied Chemistry and Frontier Chemistry Center, Faculty of Engineering, Hokkaido University, Sapporo, Hokkaido, 060-8628, Japan   Email: hajito@eng.hokudai.ac.jp
,
Hajime Ito*
Division of Applied Chemistry and Frontier Chemistry Center, Faculty of Engineering, Hokkaido University, Sapporo, Hokkaido, 060-8628, Japan   Email: hajito@eng.hokudai.ac.jp
› Author Affiliations
Further Information

Publication History

Received: 26 September 2016

Accepted after revision: 25 October 2016

Publication Date:
21 November 2016 (online)


Abstract

A novel approach has been developed for the enantioselective synthesis of α-chiral γ-acetoxyallylboronates via the copper(I)-catalyzed γ-boryl substitution of allyl acylals. This reaction proceeded with high E/Z selectivity and enantioselectivity (E/Z = >99:1, up to 80% yield, up to 99% ee). The subsequent allylation of aldehyde with the allylboronate afforded the monoprotected anti-1,2-diol derivative with high stereoselectivity.

Supporting Information

 
  • References and Notes

    • 2a Hoffmann RW, Zeiss HJ. Angew. Chem., Int. Ed. Engl. 1979; 18: 306
    • 2b Brown HC, Bhat KS. J. Am. Chem. Soc. 1986; 108: 293
    • 2c Roush WR, Ando K, Powers DB, Palkowitz AD, Halterman RL. J. Am. Chem. Soc. 1990; 112: 6339
    • 2d Rauniyar V, Zhai H, Hall DG. J. Am. Chem. Soc. 2008; 130: 8481
    • 2e Jain P, Antilla JC. J. Am. Chem. Soc. 2010; 132: 11884
    • 3a Jadhav PK, Woerner FJ. Tetrahedron Lett. 1994; 35: 8973
    • 3b Burgess K, Chaplin DA, Henderson I, Pan YT, Elbein AD. J. Org. Chem. 2002; 57: 1103
    • 3c Ramachandran VP, Subash Chandra AJ, Reddy MV. R. J. Org. Chem. 2002; 67: 7547
    • 3d Yin N, Wang G, Qian M, Negishi E. Angew. Chem. Int. Ed. 2006; 45: 2916
    • 3e Penner M, Rauniyar V, Kaspar LT, Hall DG. J. Am. Chem. Soc. 2009; 131: 14216
    • 4a Hoffmann RW, Kemper B. Tetrahedron Lett. 1981; 22: 5263
    • 4b Wuts PG. M, Bigelow SS. J. Org. Chem. 1982; 47: 2498
    • 4c Brown HC, Jadhav PK, Bhat KS. J. Am. Chem. Soc. 1988; 110: 1535
    • 4d Ganesh P, Nicholas KM. J. Org. Chem. 1997; 62: 1737
    • 4e Hoffmann RW, Krüger J, Brückner D. New J. Chem. 2001; 102
    • 4f Muñoz-Hernández L, Soderquist JA. Org. Lett. 2009; 11: 2571
  • 5 Yamamoto E, Takenouchi Y, Ozaki T, Ito H. J. Am. Chem. Soc. 2014; 136: 16515

    • Representative examples of the routes used to synthesize the acetal and acylal substrates
    • 6a A Synthesis of Z-Allyl Dibenzyl Acetal (Scheme 6) The allyl acetal substrates were synthesized over several steps (a). The synthesis started from commercially available propargyl diethyl acetal, which was subjected to an acid-catalyzed acetal-exchange reaction with benzyl alcohol to give the corresponding dibenzyl acetal. The subsequent deprotonation of the alkyne moiety, followed by the alkylation of the alkynyl lithium and partial reduction of the carbon–carbon triple bond gave the allyl acetal substrate. Although the exchange reaction generally proceeded in high yield, the subsequent alkylation of the terminal alkyne with an alkyl halide was typically low-yielding.
    • 6b A Synthesis of Z-Allyl Acylal (Scheme 7) In contrast to the acetal substrates, the acylal substrates were much easier to prepare (b). The formylation of a terminal alkyne, followed by the gem-diacetylation of the resulting carbonyl moiety provided the corresponding propargyl acylals in moderate to high yields. The subsequent Z-selective reduction of the alkyne moiety in these propargyl acylals yielded the desired allylic substrates.
    • 7a van Heerden FR, Huyser JJ, Williams D, Holzapfel CW. Tetrahedron Lett. 1998; 39: 5281
    • 7b Trost BM, Lee CB. J. Am. Chem. Soc. 2001; 123: 3687
    • 7c Trost BM, Lee CB. J. Am. Chem. Soc. 2001; 123: 3671
    • 8a Sydnes LK, Sandberg M. Tetrahedron 1997; 53: 12679
    • 8b Sandberg M, Sydnes LK. Tetrahedron Lett. 1998; 39: 6361
    • 8c Sandberg M, Sydnes LK. Org. Lett 2000; 2: 687
  • 9 Wuts PG. M, Greene TW. Protective Groups in Organic Synthesis . Wiley Interscience; Hoboken: 2007. 4th ed.
  • 10 Kavala V, Patel BK. Eur. J. Org. Chem. 2005; 441
  • 11 Tamura K, Sugiya M, Yoshida K, Yanagisawa A, Imamoto T. Org. Lett. 2010; 12: 4400
    • 12a Ito H, Kawakami C, Sawamura M. J. Am. Chem. Soc. 2005; 127: 16034
    • 12b Ito H, Ito S, Sasaki Y, Matsuura K, Sawamura M. J. Am. Chem. Soc. 2007; 129: 14856
  • 13 Typical Procedure for the Enantioselective Boryl Substitution of Allyl Acylals CuCl (2.6 mg, 0.026 mmol), (R,R)-BenzP* (7.2 mg, 0.026 mol), B2(pin)2 (254.8 mg, 1.00 mmol), and KOt-Bu (84.3 mg, 0.75 mmol) were placed in a screw-capped test tube in a glove box under an argon atmosphere. After the vial was sealed with a screw cap containing a Teflon-coated rubber septum, the test tube was removed from the glove box and connected to a vacuum/nitrogen manifold through a needle. Then, dry DMI (1.0 mL) was added to the mixture via a syringe with stirring at r.t. After 15–30 min, acylal (Z)-1a (129.5 mg, 0.5 mmol) was added to the reaction mixture with vigorous stirring at 0 °C. After the completion of the reaction, the mixture was directly filtered through a short silica gel column with hexane–EtOAc (90:10) as the eluent. After removal of the solvents under reduced pressure, NMR yield was determined by 1H NMR analysis of the crude reaction mixture [(S,E)-2a; 79%] by using mesitylene (26.7 mg, 0.22 mmol) as the internal standard. The crude product was purified with flash chromatography (SiO2, hexane–Et2O = 100:0 to 90:10) to give the corresponding γ-acetoxyallylboronate (S,E)-2a (84.8 mg, 0.257 mmol, 52% isolated yield). 1H NMR (392 MHz, CDCl3): δ = 1.25 (s, 12 H), 1.63–1.93 (m, 3 H), 2.11 (s, 3 H), 2.52–2.71 (m, 2 H), 5.45 (dd, J = 9.4, 12.5 Hz, 1 H), 7.09 (d, J = 12.2 Hz, 1 H), 7.13–7.31 (m, 5 H). 13C NMR (99 MHz, CDCl3): δ = 20.7 (CH3), 22.9 (br, BCH), 24.6 (CH3), 24.7 (CH3), 32.8 (CH2), 35.0 (CH2), 83.4 (C), 115.4 (CH), 125.6 (CH), 128.2 (CH), 128.4 (CH), 135.2 (CH), 142.3 (C), 168.1 (C). HRMS (EI): m/z [M]+ calcd for C19H27BO4: 329.20387; found: 329.20481. [α]D 22.2 +5.4 (c 1.0, CHCl3, 95% ee). The ee value was determined by HPLC analysis [Daicel CHIRALPAK OD-3, 2-PrOH–hexane = 0.25:99.75, 0.5 mL/min, 40 °C]: t R (major) = 25.44 min; t R (minor) = 24.83 min.
  • 14 Kobayashi S, Endo T, Schneider U, Ueno M. Chem. Commun. 2010; 46: 1260
    • 15a Carosi L, Lachance H, Hall DG. Tetrahedron Lett. 2005; 46: 8981
    • 15b Rauniyar V, Hall DG. J. Am. Chem. Soc. 2004; 126: 4518
    • 15c Ishiyama T, Ahiko T.-A, Miyaura N. J. Am. Chem. Soc. 2002; 124: 12414
    • 15d Kennedy J, Hall DG. J. Am. Chem. Soc. 2002; 124: 11586