Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Silver-catalysed trifluoromethylation of arenes at room temperature

Sangwon Seo a, John B. Taylor b and Michael F. Greaney *a
aSchool of Chemistry, University of Manchester, Oxford Rd, Manchester, M13 9PL, UK. E-mail: michael.greaney@manchester.ac.uk
bSyngenta, Jealott's Hill International Research Centre, Bracknell, RG42 6EY, UK

Received 11th March 2013 , Accepted 16th May 2013

First published on 17th May 2013


Abstract

A variety of heteroarenes and electron rich arenes can be trifluoromethylated at room temperature with TMSCF3, catalytic silver and PhI(OAc)2.


The trifluoromethyl group is valued for its ability to modulate the properties of diverse materials such as pharmaceuticals, agrochemicals and polymers. Aryl CF3 groups are electron-withdrawing, hydrophobic and generally very stable, all properties that can be harnessed in the design of biologically active molecules and functional materials. Synthetic methods for aryl and heteroaryl trifluoromethylation are thus critical to the discovery and production of new molecules of high value to society.1 Recent developments in metal-mediated trifluoromethylation have produced significant advances in this area,2 with common functional groups such as aryl boronic acids and halides undergoing efficient trifluoromethylation under palladium and copper catalysis.3 Metal-catalysed trifluoromethylation of unactivated C–H positions, by contrast, is significantly less developed and has great potential for accelerating medicinal and agrochemical syntheses. Despite some recent ground-breaking developments in this area,4 there is still great demand for the development of catalytic C–H trifluoromethylation methods that function under mild and simple conditions.

We were interested in developing a catalytic trifluoromethylation based on silver; in contrast to its Group 11 neighbour copper there have been few reports on silver-mediated trifluoromethylation4g,k,5 and none that we are aware of using silver catalysis. The redox catalysis of silver, comprising one electron steps between 0, +1, +2 and +3 oxidation states, has been scarcely exploited in synthesis relative to other late TMs6,7 and could offer productive catalytic pathways for trifluoromethylation. We have recently developed silver-catalysed decarboxylative C–H cross-coupling under oxidative radical conditions,8 and were keen to see if a similar approach was viable for C–H trifluoromethylation.

We started with a screen of reaction conditions based around TMSCF3 as the trifluoromethylating agent.9 The groups of Sanford, Bräse and Wang have recently demonstrated the compatibility of this reagent with stoichiometric silver salts,4g,k,5c encouraging us that it could form the basis of a catalytic system. Using 1,4-dimethoxybenzene (1a) as the substrate, we conducted an initial solvent screen using combinations of AgF, TMSCF3 and PhI(OAc)2 (Table 1). We worked at room temperature under air throughout, with the aim of developing a mild reaction with as broad a functional group tolerance as possible. The reaction proved sensitive to solvent choice with initially only MeCN from a selection of common organic solvents producing any reaction (entries 1 and 2). DMSO proved more effective still, affording the trifluoromethylated compound 2a in 51% conversion (entry 3). Fluoride was not a requirement, with Ag2CO3 being similarly effective at promoting reaction (entry 4). Alternative oxidants did not improve on PhI(OAc)2 (entries 5 and 6), and the use of a nitrogen atmosphere led to a reduction in yield (entry 7). Crucially, sub-stoichiometric amounts of silver salts proved equally effective (entries 8–10), indicating that a catalytic reaction was feasible. We settled on conditions of AgF (25 mol%) with TMSCF3 (2 equiv.) and PhI(OAc)2 (2 equiv.), at room temperature (entry 9) to take forward. The use of the more stable (and expensive) TESCF3 reagent gave only marginal improvement (entry 10), so we continued with the cheaper TMSCF3 reagent.

Table 1 Ag-catalysed trifluoromethylation: reaction optimisation

Entrya Catalyst (equiv.) Oxidant Solvent Yieldb (%)
a 1,4-Dimethoxybenzene 1a (0.3 mmol), TMSCF3 (1.2 mmol), oxidant (0.6 mmol), F source or base, solvent (1.0 mL), room temperature, 20 h. b Yields determined by 19F NMR using 4-fluoroanisole as the internal standard. c THF, 1,4-dioxane, MeOH, (CF3)2CHOH, DCE, DCM. d Under N2. e Slow addition of AgF to the stirring mixture of 1a, TMSCF3 and PhI(OAc)2 in DMSO. f 2 equiv. of TMSCF3. g Isolated yield. h TESCF3 (2 equiv.) instead of TMSCF3.
1 AgF (1) PhI(OAc)2 Solventc 0
2 AgF (1) PhI(OAc)2 MeCN 26
3 AgF (1) PhI(OAc)2 DMSO 51
4 Ag2CO3 (0.5) PhI(OAc)2 DMSO 48
5 AgF (1) PhI(TFA)2 DMSO 5
6 AgF (1) K2S2O8 DMSO 20
7d AgF (1) PhI(OAc)2 DMSO 35
8e AgF (0.25) PhI(OAc)2 DMSO 55
9e,f AgF (0.25) PhI(OAc)2 DMSO 55 (58g)
10e,h AgF (0.25) PhI(OAc)2 DMSO 60


Substrate scope investigations established that the procedure was effective for a variety of electron rich arenes with broad substrate scope tolerance (Table 2). For unsymmetrical substrates isomeric mixtures were generally observed, with regioselectivities consistent with radical SArH addition (vide infra). Importantly, the reaction was compatible with halogen groups, illustrating an orthogonal reactivity to conventional C–X trifluoromethylations whereby neighbouring C–H bonds undergo preferential reaction. The useful building blocks 2f, 2g, 2h and 2i were prepared in this fashion. Electron-withdrawing groups such as aldehyde (2j), ketone (2k) and ester (2l) were likewise tolerated without problem. Importantly, dialkylanilines could be trifluoromethylated, a key class of building block that has rarely featured in C–H trifluoromethylation reports.4i,10 A slight preference for ortho over para selectivity was observed for simple dimethylamine (2m), with bromo substitution also tolerated (2n) along with N-acylation (2o). We were pleased to observe that the reaction was also effective for un-activated arenes (2p, 2q and 2r), although these substrates did require an excess of the arene and the reaction temperature raised to 70 °C.

Table 2 Ag-catalysed trifluoromethylation: substrate scopea,b
a 1 (0.3 mmol), TMSCF3 (0.6 mmol), PhI(OAc)2 (0.6 mmol), AgF (0.075 mmol), DMSO (1.0 mL), room temperature, 20 h. b Isolated yields. For isomer mixtures, the minor regioisomeric position is labeled with *. c Yields determined by 19F NMR using 4-fluoroanisole as the internal standard. d Reaction conducted at 70 °C, 5–10 equiv. of arene.


The reaction could be extended to heteroarenes with N–Me pyrroles in particular being excellent substrates (2s, 2t). Electron-withdrawing groups on the heteroarene nucleus were well-tolerated (2t), but on nitrogen less so (N–Boc, 2u). Furans (2v), thiophenes (2w, 2x) and indoles (2y) were all productive, indicating that the method is viable for the major classes of π-excessive heterocycle. π-Deficient heteroarenes, by contrast, were not generally effective in the reaction but could be efficiently captured by masking the azine nucleus with electron-donating groups (2aa).

We next turned to the trifluoromethylation of more complex, biologically active molecules – a major driver for the development of new methods in this area. Introduction of the CF3 group at unactivated C–H positions represents a very versatile approach to fluorine incorporation for modulation of biological activity,4c,d,j demanding mild reaction conditions that are tolerant of functional groups and reasonable stoichiometries with respect to the (often valuable) C–H substrate. Accordingly, we extended the reaction to trifluoromethylate some more complex molecules in the agrochemistry field, an area where the CF3 group is particularly prevalent. We could successfully incorporate the CF3 group into the commercial herbicides pyriftalid11 and napropamide12 (Scheme 1). The functional group tolerance of the reaction was illustrated by sulfide, lactone and α-hydroxyamide functionality all being stable to the reaction conditions (Scheme 1, 3 and 4).


A radical mechanism is implicated for the trifluoromethylation reaction,13 as radical quenching reactions using TEMPO and galvinoxyl radical both shut down the reaction, with the TEMPO–CF3 adduct being clearly observed in the crude 19F NMR. The electrophilic CF3 radical usually (but with some exceptions)4c,j displays a marked preference for electron rich substrates, as seen here, underlining the likelihood of a radical pathway. A possible mechanism is shown in Scheme 2 whereby TMSCF3 is oxidised to the CF3 radical, followed by SArH addition, then a second one electron oxidation and proton loss to give the product 2. Control experiments to investigate the role of silver in the first step of the proposed mechanism indicated that AgF alone was insufficiently oxidizing to generate CF3˙ (mixing AgF with TMSCF3 in the presence of TEMPO in DMSO at room temperature gave only trace quantities of TEMPO–CF3). PhI(OAc)2 alone was moderately effective (44% NMR yield of TEMPO–CF3) and the combination of PhI(OAc)2 and AgF highly effective (91% NMR yield).14 The background oxidizing activity of the hypervalent iodine reagent could be quantified in the trifluoromethylation of 1,4-dimethoxybenzene 1a in the absence of any silver salt, producing a low conversion to the trifluoromethylated product 2a (26% NMR yield).


Silver-catalysed radical trifluoromethylation.
Scheme 2 Silver-catalysed radical trifluoromethylation.

Alternative mechanisms were investigated by treating dimethoxyanisole 1a with in situ prepared AgCF35b in both MeCN and DMSO as solvents. No reaction could be observed in each case, suggesting organometallic AgCF3 intermediates are not participating under our reaction conditions. A further control experiment with Togni's reagent4a in DMSO at room temperature gave no reaction, ruling out simple SEAr attack on an electrophilic CF3 source. Finally, we considered the possibility of initial arene oxidation by PhI(OAc)2, followed by CF3 anion addition to a cationic arene intermediate. Extensive work by Kita has demonstrated the C–H functionalization of electron rich arenes using PhI(TFA)2 in the presence of stoichiometric BF3·OEt2 and nucleophiles.15 It seems the present conditions are not sufficiently oxidizing to enable an analogous pathway, as a control reaction in the absence of TMSCF3 gave no reaction, where some degree of homocoupling would be expected if this mechanism was in operation.

In conclusion, we have developed a silver-catalysed trifluoromethylation system for electron rich aromatic and heteroaromatic substrates. The reaction works at room temperature under air, does not require excessive stoichiometries of substrate or reagent, and is operationally simple to carry out. The application of this chemistry to new trifluoromethylation substrates will be the subject of future work in our laboratory.

We thank Syngenta, the University of Manchester and the EPSRC for funding (Leadership Fellowship to M.F.G.), and the EPSRC mass spectrometry service at the University of Swansea.

Notes and references

  1. (a) A. Studer, Angew. Chem., Int. Ed., 2012, 51, 8950 CrossRef CAS; (b) O. A. Tomashenko and V. V. Grushin, Chem. Rev., 2011, 111, 4475 CrossRef CAS; (c) T. Furuya, A. S. Kamlet and T. Ritter, Nature, 2011, 473, 470 CrossRef CAS.
  2. Recent examples of stoichiometric metal-mediated trifluoromethylation: (a) K. A. McReynolds, R. S. Lewis, L. K. G. Ackerman, G. G. Dubinina, W. W. Brennessel and D. A. Vicic, J. Fluorine Chem., 2010, 131, 1108 CrossRef CAS; (b) Y. Ye, N. D. Ball, J. W. Kampf and M. S. Sanford, J. Am. Chem. Soc., 2010, 132, 14682 CrossRef CAS; (c) T. D. Senecal, A. T. Parsons and S. L. Buchwald, J. Org. Chem., 2011, 76, 1174 CrossRef CAS; (d) N. D. Ball, J. B. Gary, Y. Ye and M. S. Sanford, J. Am. Chem. Soc., 2011, 133, 7577 CrossRef CAS; (e) C.-P. Zhang, Z.-L. Wang, Q.-Y. Chen, C.-T. Zhang, Y.-C. Gu and J.-C. Xiao, Angew. Chem., Int. Ed., 2011, 50, 1896 CrossRef CAS; (f) H. Morimoto, T. Tsubogo, N. D. Litvinas and J. F. Hartwig, Angew. Chem., Int. Ed., 2011, 50, 3793 CrossRef CAS; (g) O. A. Tomashenko, E. C. Escudero, M. M. Belmonte and V. V. Grushin, Angew. Chem., Int. Ed., 2011, 50, 7655 CrossRef CAS; (h) Y. Ye, S. A. Künzi and M. S. Sanford, Org. Lett., 2012, 14, 4979 CrossRef CAS.
  3. Cu and Pd-catalysed trifluoromethylation of functionalised arenes: (a) M. Oishi, H. Kondo and H. Amii, Chem. Commun., 2009, 1909 RSC; (b) E. J. Cho, T. D. Senecal, T. Kinzel, Y. Zhang, D. A. Watson and S. L. Buchwald, Science, 2010, 328, 1679 CrossRef CAS; (c) R. Shimizu, H. Egami, T. Nagi, J. Chae, Y. Hamashima and M. Sodeoka, Tetrahedron Lett., 2010, 51, 5947 CrossRef CAS; (d) T. Knauber, F. Arikan, G.-V. Roeschenthaler and L. J. Goossen, Chem.–Eur. J., 2011, 17, 2689 CrossRef CAS; (e) J. Xu, D.-F. Luo, B. Xiao, Z.-J. Liu, T.-J. Gong, Y. Fu and L. Liu, Chem. Commun., 2011, 47, 4300 RSC; (f) T. Liu and Q. Shen, Org. Lett., 2011, 13, 2342 CrossRef CAS; (g) Y. Ye and M. S. Sanford, J. Am. Chem. Soc., 2012, 134, 9034 CrossRef CAS; (h) T. Schareina, X.-F. Wu, A. Zapf, A. Cotte, M. Gotta and M. Beller, Top. Catal., 2012, 55, 426 CrossRef CAS.
  4. Recent C–H Trifluoromethylation of arenes: (a) M. S. Wiehn, E. V. Vinogradova and A. Togni, J. Fluorine Chem., 2009, 131, 951 CrossRef; (b) X. Wang, L. Truesdale and J.-Q. Yu, J. Am. Chem. Soc., 2010, 132, 3648 CrossRef CAS; (c) Y. Ji, T. Brueckl, R. D. Baxter, Y. Fujiwara, I. B. Seiple, S. Su, D. G. Blackmond and P. S. Baran, Proc. Natl. Acad. Sci. U. S. A., 2011, 108, 14411 CrossRef CAS; (d) D. A. Nagib and D. W. C. MacMillan, Nature, 2011, 480, 224 CrossRef CAS; (e) R. N. Loy and M. S. Sanford, Org. Lett., 2011, 13, 2548 CrossRef CAS; (f) X. Mu, S. Chen, X. Zhen and G. Liu, Chem.–Eur. J., 2011, 17, 6039 CrossRef CAS; (g) Y. Ye, S. H. Lee and M. S. Sanford, Org. Lett., 2011, 13, 5464 CrossRef CAS; (h) L. Chu and F.-L. Qing, J. Am. Chem. Soc., 2012, 134, 1298 CrossRef CAS; (i) E. Mejia and A. Togni, ACS Catal., 2012, 2, 521 CrossRef CAS; (j) Y. Fujiwara, J. A. Dixon, F. O'Hara, E. D. Funder, D. D. Dixon, R. A. Rodriguez, R. D. Baxter, B. Herlé, N. Sach, M. R. Collins, Y. Ishihara and P. S. Baran, Nature, 2012, 492, 95 CrossRef CAS; (k) A. Hafner and S. Bräse, Angew. Chem., Int. Ed., 2012, 51, 3713 CrossRef CAS; (l) X.-G. Zhang, H.-X. Dai, M. Wasa and J.-Q. Yu, J. Am. Chem. Soc., 2012, 134, 11948 CrossRef CAS; (m) X. Wu, L. Chu and F.-L. Qing, Tetrahedron Lett., 2013, 54, 249 CrossRef CAS.
  5. (a) W. E. Tyrra, J. Fluorine Chem., 2001, 112, 149 CrossRef CAS; (b) M. M. Kremlev, A. I. Mushta, W. E. Tyrra, D. Naumann, H. T. M. Fischer and Y. L. Yagupolskii, J. Fluorine Chem., 2007, 128, 1385 CrossRef CAS; (c) Z. Weng, R. Lee, W. Jia, Y. Yuan, W. Wang, X. Feng and K.-W. Huang, Organometallics, 2011, 30, 3229 CrossRef CAS.
  6. Silver in organic chemistry, ed. M. Harmata, Wiley, NY, 2010 Search PubMed.
  7. For silver-catalysed fluorination, see: P. Tang, T. Furuya and T. Ritter, J. Am. Chem. Soc., 2010, 132, 12150 CrossRef CAS.
  8. (a) S. Seo, M. Slater and M. F. Greaney, Org. Lett., 2012, 14, 2650 CrossRef CAS; (b) S. Seo, J. B. Taylor and M. F. Greaney, Chem. Commun., 2012, 48, 8270 RSC.
  9. G. K. S. Prakash and A. K. Yudin, Chem. Rev., 1997, 97, 757 CrossRef CAS.
  10. T. Akiyama, K. Kato, M. Kajitani, Y. Sakaguchi, J. Nakamura, H. Hayashi and A. Sugimori, Bull. Chem. Soc. Jpn., 1988, 61, 3531 CrossRef CAS.
  11. C. Luthy, H. Zondler, T. Rapold, G. Seifert, B. Urwyler, T. Heinis, H. C. Steinrucken and J. Allen, Pest Manage. Sci., 2001, 57, 205 CrossRef CAS.
  12. Stauffer Chemical Company, Br. Pat., GB1066606, 1967 Search PubMed.
  13. Examples of radical trifluoromethylation: (a) A. Gregorcic and M. Zupan, J. Org. Chem., 1979, 44, 4120 CrossRef CAS; (b) C. Wakselman and M. Tordeux, Chem. Commun., 1987, 1701 RSC; (c) H. Sawada, M. Nakayama, M. Yoshida, T. Yoshida and N. Kamigata, J. Fluorine Chem., 1990, 46, 423 CrossRef CAS; (d) B. R. Langlois, E. Laurent and N. Roidot, Tetrahedron Lett., 1991, 32, 7525 CrossRef CAS; (e) K. L. Kirk, M. Nishida, S. Fujii and H. Kimoto, J. Fluorine Chem., 1992, 59, 197 CrossRef CAS; (f) C. Lai and T. E. Mallouk, Chem. Commun., 1993, 1359 RSC; (g) N. Kamigata, T. Ohtsuka, T. Fukushima, M. Yoshida and T. Shimizu, J. Chem. Soc., Perkin Trans. 1, 1994, 1339 RSC; (h) J.-B. Tommasino, A. Brondex, M. Médebielle, M. Thomalla, B. R. Langlois and T. Billard, Synlett, 2002, 1697 CrossRef CAS; (i) T. Kino, Y. Nagase, Y. Ohtsuka, K. Yamamoto, D. Uraguchi, K. Tokuhisa and T. Yamakawa, J. Fluorine Chem., 2010, 131, 98 CrossRef CAS.
  14. ArI(OAc)2 could potentially generate Ag(II) as the oxidant for TMSCF3in situ under the reaction conditions. For in situ silver(II) generation from catalytic Ag(I) and a stoichiometric oxidant in decarboxylation chemistry, see: J. M. Anderson and J. K. Kochi, J. Am. Chem. Soc., 1970, 92, 1651 CrossRef CAS.
  15. T. Dohi, M. Ito, N. Yamaoka, K. Morimoto, H. Fujioka and Y. Kita, Tetrahedron, 2009, 65, 10797 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: Synthesis and characterisation data for all new compounds. See DOI: 10.1039/c3cc41829d

This journal is © The Royal Society of Chemistry 2013