Synlett 2017; 28(20): 2871-2875
DOI: 10.1055/s-0036-1589069
letter
© Georg Thieme Verlag Stuttgart · New York

Hypervalent Iodine-Based Activation of Triphenylphosphine for the Functionalization of Alcohols

Jasmin Eljo
Department of Chemistry, University of Waterloo, 200 University Ave W., Waterloo, ON, N2L3G1, Canada   graham.murphy@uwaterloo.ca
,
Myriam S. Carle
Department of Chemistry, University of Waterloo, 200 University Ave W., Waterloo, ON, N2L3G1, Canada   graham.murphy@uwaterloo.ca
,
Department of Chemistry, University of Waterloo, 200 University Ave W., Waterloo, ON, N2L3G1, Canada   graham.murphy@uwaterloo.ca
› Author Affiliations
This work was supported by the University of Waterloo and the Natural Sciences and Engineering Research Council (NSERC) of Canada (Discovery Grant 418602-2013).
Further Information

Publication History

Received: 18 April 2017

Accepted after revision: 06 June 2017

Publication Date:
12 July 2017 (online)


Dedicated to Professor Victor Snieckus on the occasion of his 80th birthday

Abstract

The use of hypervalent iodine reagents as a general tool for the activation of PPh3 and its application to the functionalization of alcohols is reported. Combination of PPh3 with PhICl2 or TolIF2 gives dihalophosphoranes that are characterized by 31P NMR, however, with PhIOAc2, PhI(OTFA)2, or the cyclic chloro(benzoyloxy)iodane, no phosphoranes were observed. Reaction of these iodanes with PPh3 in the presence of primary, secondary, or tertiary alcohols results in either halogenation or acyl-transfer products in moderate to high yield.

Supporting Information

 
  • References and Notes

    • 1a Tian J. Gao WC. Zhou DM. Zhang C. Org. Lett. 2012; 14: 3020
    • 1b Zhang C. Liu SS. Sun B. Tian J. Org. Lett. 2015; 17: 4106
  • 2 Appel R. Angew. Chem., Int. Ed. Engl. 1975; 14: 801
    • 3a Zhdankin VV. Stang PJ. Chem. Rev. 2002; 102: 2523
    • 3b Wirth T. Hypervalent Iodine Chemistry: Modern Developments in Organic Synthesis. Springer; Berlin: 2003
    • 3c Zhdankin VV. Stang PJ. Chem. Rev. 2008; 108: 5299
    • 3d Yoshimura A. Zhdankin VV. Chem. Rev. 2016; 116: 3328
    • 3e Zhdankin VV. Hypervalent Iodine Chemistry: Preparation Structure, and Synthetic Applications of Polyvalent Iodine Compounds. Wiley; Chichester: 2014
  • 5 Wright DW. Russell GA. Phenyliodine(III) Dichloride In e-EROS . Wiley; Hoboken, NJ: 2001
    • 6a For syntheses of Ph3PCl2 with various reagents such as Cl2, CCl4, (COCl)2, COCl2, or C2Cl6, see: Michaelis A. Michaelis A. v. Soden H. Justus Liebigs Ann. Chem. 1885; 229: 295
    • 6b Appel R. Angew. Chem., Int. Ed. Engl. 1975; 14: 801
    • 6c Appel R. Geisler K. Scholer H. Chem. Ber. 1977; 110: 376

    • For NMR studies on the synthesis of 2 with C2Cl6, see:
    • 6d Yin Q. Ye Y. Tang G. Zhao YF. Spectrochim. Acta, Part A 2006; 63: 192
    • 6e Wiley GA. Stine WR. Tetrahedron Lett. 1967; 2321
    • 6f Dillon KB. Lynch RJ. Reeve RN. Waddington TC. J. Chem. Soc. Dalton 1976; 1243
    • 6g Appel R. Scholer H. Chem. Ber. 1977; 110: 2382
    • 7a Kirk KL. Org. Process. Res. Dev. 2008; 12: 305
    • 7b Bégué J.-P. Bonnet-Delpon D. Bioorganic and Medicinal Chemistry of Fluorine . John Wiley and Sons; Hoboken, NJ: 2008: 365
    • 7c Ojima I. Fluorine in Medicinal Chemistry and Chemical Biology . Wiley-Blackwell; Chichester: 2009: 624

      For syntheses of Ph3PF2 with various reagents such as XeF2, SF4, IF5, DAST, COF2, or HgF2, see:
    • 8a Kobayash Y. Akashi C. Chem. Pharm. Bull. 1968; 16: 1009
    • 8b Frohn HJ. Maurer H. J. Fluorine Chem. 1986; 34: 73
    • 8c Holmes RR. Chandrasekhar V. Day RO. Harland JJ. Payne JS. Holmes JM. Phosphorus Sulfur Relat. Elem. 1987; 30: 409
    • 8d Gupta OD. Shreeve JM. J. Chem. Soc., Chem. Commun. 1984; 416
    • 8e Doxsee KM. Hanawalt EM. Weakley TJ. R. Inorg. Chem. 1992; 31: 4420
  • 9 Caputo CB. Winkelhaus D. Dobrovetsky R. Hounjet LJ. Stephan DW. Dalton Trans. 2015; 44: 12256
  • 10 We believe its increased hydrolytic stability (relative Ph3PCl2) is due to the increased strength of P–F (490 kJ/mol) over P-Cl (326 kJ/mol) bonds.
  • 11 Iranpoor N. Firouzabadi H. Davan EE. Tetrahedron Lett. 2013; 54: 1813
  • 12 Gallos J. Varvoglis A. Chim. Chron., New Ser. 1987; 16: 87
  • 13 Makowiec S. Rachon J. Heteroat. Chem. 2003; 14: 352
  • 14 McNulty J. Capretta A. Laritchev V. Dyck J. Robertson AJ. Angew. Chem. Int. Ed. 2003; 42: 4051
  • 15 Dormoy J.-R. Castro B. Triphenylphosphine Dichloride . In e-EROS . Wiley; Hoboken, NJ: 2001
  • 16 Kobayashi Y. Akashi C. Chem. Pharm. Bull. 1968; 16: 1009
    • 17a Mohandas TP. Mamman AS. Nair PM. Tetrahedron 1983; 39: 1187
    • 17b Wicha J. Zarecki A. Kocór M. Tetrahedron Lett. 1973; 14: 3635
  • 18 Taniguchi T. Hirose D. Ishibashi H. ACS Catal. 2011; 1: 1469
  • 19 Carle MS. Shimokura GK. Murphy GK. Eur. J. Org. Chem. 2016; 3930
  • 20 Reacting Ph3PF2 (4, 1.6 equiv) with 4-nitrophenethyl alcohol for 45 min at 160 °C using 10 mol% TiF3 as an activating agent produced alkyl fluoride 5a in 66% yield.
  • 21 The major competing reaction pathway in these experiments was alcohol dimerization, giving the ether. It is proposed that this arises from the alcohol outcompeting fluoride as the nucleophile. See ref. 16.
  • 22 Zhang and co-workers reported phenethyl acetate to be the major byproduct of a reaction between 5, phenethyl alcohol, hexanoic acid, DMAP, and PPh3. No indication of phenethyl trifluoroacetate was reported for the analogous reaction using iodane 7. See Table S1 in the supporting information of ref. 1a.
  • 23 Only trace amounts of 15a or 15b were visible by 1H NMR analysis of the crude reaction mixtures after 10 min. Prolonging the reaction times to 2–3 h did not significantly change the product ratios, suggesting that the reaction pathways are competing, as opposed to 14 reacting with the benzoate byproduct of the reaction to give 15. Concentrating the reaction mixtures before column chromatography appears to have increased the ratio of the benzoate products.
  • 24 Zhdankin VV. Koposov AY. Litvinov DN. Ferguson MJ. McDonald R. Luu T. Tykwinski RR. J. Org. Chem. 2005; 70: 6484
  • 25 Sample Experimental Procedure Into a conical flask was added the 4-phenyl-2-butanol (0.063 g, 0.42 mmol, 1 equiv) and CH2Cl2 (0.5 mL), and to this was added PPh3 (1.1 equiv) and [bis(trifluoro-acetoxy)iodo]benzene (1.1 equiv). The reaction mixture was stirred at r.t. for 30 min, concentrated by rotary evaporation and subjected to column chromatography, where the stationary phase as acidified with 0.2% TFA in hexane. The crude reaction mixture was purified via column chromatography (10% EtOAc/hexane) to give 13c (0.066 g, 64% yield) as a colorless oil.19F NMR (282 MHz, CDCl3): δ = –75.58. 1H NMR (300 MHz, CDCl3): δ = 7.34–7.16 (m, 5 H), 5.11 (m, 1 H), 2.74–2.60 (m, 2 H), 2.15–1.87 (m, 2 H), 1.39 (d, J = 6.3 Hz, 3 H);13C NMR (75 MHz, CDCl3): δ = 157.1 (q, J = 41.5 Hz), 140.5, 128.6, 128.3, 126.3, 114.6 (q, J = 286.1 Hz), 75.8, 37.1, 31.4, 19.5.