Isolation and characterization of copper(III) trifluoromethyl complexes and reactivity studies of aerobic trifluoromethylation of arylboronic acids

Song-Lin Zhang* and Wen-Feng Bie
The Key Laboratory of Food Colloids and Biotechnology, Ministry of Education, School of Chemical and Material Engineering, Jiangnan University, Wuxi 214122, Jiangsu Province, China. E-mail: slzhang@jiangnan.edu.cn; Fax: +86-510-85917763; Tel: +86-510-85917763

Received 21st April 2016 , Accepted 20th July 2016

First published on 22nd July 2016


Abstract

The isolation, characterization and reactivity of transition metal trifluoromethyl complexes are fundamental and challenging topics in trifluoromethylation chemistry. We report herein the synthesis and isolation of two new complexes [(phen)CuI(PPh3)2]+[CuIII(CF3)4] (2) and (phen)CuIII(CF3)3 (3) as well as a known complex (bpy)CuIII(CF3)3 (4) at room temperature. 2 and 3 have been fully characterized using 1H, 19F, 31P NMR, elemental analyses and X-ray crystallography. Reactivity studies indicate that 2 is unreactive toward arylboronic acids. In contrast, 3 and 4 can react with various aryl and heteroaryl boronic acids to deliver trifluoromethylated arenes in good to quantitative yields under mild conditions. The presence of a fluoride additive in DMF under aerobic conditions is crucial to these reactions. This study provides fundamental information about the structure and reactivity of elusive Cu(III) trifluoromethyl complexes that have been proposed as relevant reactive intermediates in many trifluoromethylation reactions.


Copper-mediated trifluoromethylation reactions have been greatly developed and constitute one important branch of organofluorine chemistry.1–3 Trifluoromethylation of aryl halides,4,5 boronic acids,6 arenes7 and others8,9 with various CF3 sources promoted by copper salts have been intensively studied during the past few decades. Despite the significant progress achieved in developing efficient reaction methodologies, the isolation and reactivity properties of possible copper trifluoromethyl intermediates are surprisingly scarcely described but are however essential to the mechanistic understanding of these reactions.10 Very recently, the characterization and reactivity properties of several CuI–CF3 complexes have been reported that typically contains N-heterocyclic carbene (NHC) or phenanthroline (phen) ligand.11

In contrast, extremely rare studies have been reported on the isolation, characterization and reactivity properties of high-valent CuIII CF3 complexes which have been frequently proposed as reactive intermediates for oxidative or electrophilic trifluoromethylation reactions.12,13 In 1989, Burton et al. reported a pioneering study on the synthesis and X-ray crystal structure of a square planar (CF3)2CuIII(dithiocarbamate) complex by oxidation of [CuI(CF3)2] using thiuramdisulphide.13a This Cu(III) trifluoromethyl complex could react with aryl iodides to give trifluoromethylated arenes at 90–100 °C. As far as we know, this is the first example of the isolation and reactivity of well-characterized Cu(III) trifluoromethyl complex. Later, Naumann et al. reported anionic square planar [CuIII(CF3)4] complexes with Ph4P+ or Bu4N+ counterions via oxidation of in situ Cu(I) trifluoromethyl by halogen.13b Similar ion–pair complexes were obtained by Grushin by reaction of CuCl or in situ formed “CuI–CF3” with CF3SiMe3 using air as the oxidant.13c Heating the ion–pair complexes with bipyridine (bpy) in acetic acid gave neutral (bpy)CuIII(CF3)3.

Herein, we report the preparation and isolation of CuIII trifluoromethyl complexes 2–4 by reaction of CuI precursors with CF3SiMe3 in the presence of AgF (Scheme 1). Reactivity studies show that these Cu(III) trifluoromethyl complexes are able to efficiently trifluoromethylate arylboronic acids in up to quantitative yields even at room temperature. Considering the predominant role of phen ligand in copper trifluoromethylation chemistry,3 the isolation and reactivity properties of the phen-containing CuIII trifluoromethyl complexes should be fundamental to this important area.


image file: c6ra10302b-s1.tif
Scheme 1 Syntheses and reactivity of CuIII–CF3 complexes 2–4.

Reaction of (phen)CuI(PPh3) (Br) (1)14 with 3 equivalents of CF3SiMe3 in the presence of AgF in CH2Cl2 at room temperature allows the concurrent access to 2 and 3 as pure yellow solids in 30% and 31% yields respectively (Scheme 1a, see ESI for details). Alternatively, complex 3 can be synthesized and isolated by a different method of reaction of CuI/phen with CF3SiMe3 using AgF as oxidant in a higher yield of 63% (Scheme 1b, see ESI for details). In both reactions, AgF is crucial, which should probably accelerate the transmetalation of CF3SiMe3 and act as a one-electron oxidant to convert the Cu(I) to higher oxidation states. When bpy was used instead of phen in Scheme 1b, a neutral (bpy)CuIII(CF3)3 (4) was obtained in an isolated high yield of 71%.

19F NMR spectrum of complex 2 shows a resonance at −34.6 ppm with singlet splitting, consistent with previous reported [CuIII(CF3)4] species.13b,c 19F NMR spectrum of complex 3 shows two signals at −24.4 (septet) and −37.4 (quartet) ppm, with a ratio of ca. 1[thin space (1/6-em)]:[thin space (1/6-em)]2 with the signal at −37.4 ppm being the major one. 1H NMR of complex 2 clearly shows a ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]2 of phen/PPh3 and the typical splitting of the phen ligand in Cu(I) complexes (see ESI for more details). 1H NMR spectrum of complex 3 shows only resonances of phen ligand which are greatly low-field shifted compared to those in complex 2. 31P NMR spectrum of 2 shows a weak signal at 3.0 ppm, confirming the existence of coordinated PPh3 ligand. Complex 4 shows a similar structure to 3, with two fluorine signals at −24.0 (septet) and −36.1 (quartet) ppm in 19F NMR and protons of bpy in 1H NMR, which is consistent with previously reported data.13c

The structures of 2 and 3 were further confirmed by X-ray crystallographic diffraction analyses (Fig. 1 and 2).15 Complex 2 is composed of a square planar [CuIII(CF3)4] anion and a tetrahedral [(phen)CuI(PPh3)2]+ countercation (Fig. 1). The Cu–CF3 bond lengths in [CuIII(CF3)4] are averaged to about 1.936(7) Å. Complex 3 adopts an approximately trigonal bipyramidal geometry with phen, Cu and one CF3 ligand lying in the equatorial plane and the other two CF3s occupying the apical positions (Fig. 2). The equatorial Cu–CF3 bond is shorter than apical Cu–CF3 bonds (1.938(5) Å for equatorial vs. 1.953(5), 1.961(5) Å for apical Cu–CF3 bonds). These bond lengths are shorter than Cu–CF3 bonds in related Cu(I) trifluoromethyl complexes reported. For example, Cu–CF3 bonds in (PPh3)3CuI–CF3, phenCuI(PPh3)CF3, CuI(CF3)2, and (NHC)CuI–CF3 are reported to be 2.025(7), 1.985(1), 1.970(6) and 2.022(4) Å.11a,b,d This should partly be attributed to the different coordination geometries and the more electrophilic nature of the Cu(III) center in 3. The trigonal bipyramidal geometry of 3 is unusual considering that transition metal complexes with d8 electronic configuration, such as Pd(II), Ni(II) and Cu(III), often adopt square planar geometry.


image file: c6ra10302b-f1.tif
Fig. 1 ORTEP drawing of complex 2 with thermal ellipsoids at 50% probability. All the hydrogen atoms and one solvent molecule are omitted for clarity. Disorders of fluorine atoms are observed.

image file: c6ra10302b-f2.tif
Fig. 2 ORTEP drawing of complex 3 with thermal ellipsoids at 30% probability. All the hydrogen atoms are omitted for clarity.

As to the mechanism for the concurrent formation of 2 and 3 in Scheme 1a, it is proposed that a Cu(II) precursor A is initially formed.16 A disproportionates into 3 + B + PPh3 which is in equilibrium with 2 + phen by ligand exchange (Scheme 2). Evidence supporting this mechanistic scheme include the observation of Cu(II) intermediates and Ag mirror during workup, 19F NMR detection of FSiMe3 at −157.9 ppm and the access to X-ray crystal structure of PPh3-ligated dimeric AgBr (5) (for more details, refer to ESI).17 Additional evidence supporting the equilibrium between 2 and 3 comes from the observation that addition of phen to acetic acid solution of 2 led to the efficient formation of 3 in 45% isolated yield (see ESI for details). Possibly, the addition of excess phen to the solution of 2 drives the equilibrium to the side of 3 (Scheme 2).


image file: c6ra10302b-s2.tif
Scheme 2 Plausible mechanism for the concurrent formation of 2 and 3.

Reactivity properties of complexes 2–4 were further studied to evaluate their reaction with arylboronic acids.18 After initial screening of the reaction conditions (Table 1, entries 1–4), it was found that reaction of complex 3 with 2 equivalents of para-methoxyphenyl boronic acid (6a) gave the trifluoromethylated product 7a in 65% yield in the presence of 2 equivalents of AgF in DMF at 80 °C for 18 hours (entry 4).19 A breakthrough was achieved with the use of KF additive, which led to the quantitative trifluoromethylation of 5a even at room temperature in 3 hours (entries 5–10). These results should be the first examples of room-temperature quantitative trifluoromethylation of arylboronic acids by isolated CuIII trifluoromethyl complexes. Reducing the reaction time to 1 hour still led to a 90% yield while the reaction can reach a 85% yield in only 0.5 hour (entries 11 & 12).

Table 1 Reactivity studies of complex 3 with arylboronic acida

image file: c6ra10302b-u1.tif

Entry Additive T (°C) Time (h) Yieldb (%)
a Reaction conditions: 3 (0.1 mmol), 6a (0.2 mmol), additive (0.2 mmol), 4,4′-difluorobiphenyl (0.2 mmol, internal standard), DMF (1 mL), under dry O2 atmosphere.b Yields determined by 19F NMR spectroscopy relative to 6a.
1 80 18 Trace
2 KI 80 18 23
3 Cs2CO3 80 18 12
4 AgF 80 18 65
5 KF 80 18 97
6 KF 80 10 99
7 KF 50 18 99
8 KF RT 18 99
9 KF RT 10 99
10 KF RT 3 99
11 KF RT 1 90
12 KF RT 0.5 85


With the optimized reaction conditions identified, we next examined reaction of 3 with a variety of substituted aryl boronic acids with diverse electronic properties (Table 2). The reactions show a complicated effect of electronic properties and substitution position of the substituents. Generally, electron-rich substituents are beneficial to the reactions. For substrates with strongly electron-withdrawing substituents, using AgF at a little higher 50 °C (in replace of KF at RT) is needed to achieve better yields. Additionally, para-substituted boronic acids are generally more reactive than meta- and ortho-substituted analogs (please see 7a vs. 7l, 7m; 7b vs. 7k). Finally, heteroaryl boronic acids such as 2-benzothiophenyl, 2-benzofuryl, and 3-pyridyl boronic acids can also be trifluoromethylated to produce the desired trifluoromethylated heteroarenes 7o–q in moderate to good yields. These results imply that 3 may be potential reactive intermediate in some copper/phen-mediated oxidative and electrophilic trifluoromethylation reactions.

Table 2 Substrate scope for reaction of 3 with various arylboronic acidsa

image file: c6ra10302b-u2.tif

a Reaction yields were determined by 19F NMR spectroscopy using 4,4′-difluorobiphenyl (−117.0 ppm) as internal standard.b Values in parentheses are isolated yields using column chromatography.c AgF at 50 °C in the presence of 4 Å molecular sieves.
image file: c6ra10302b-u3.tif


Under similar conditions, complex 4 also shows good reactivity with various aryl boronic acids, as summarized in Table 3. In sharp contrast to 3 and 4, the ion–pair complex 2 shows little reactivity (Table S1 in ESI). The reasons for this reactivity difference are currently not well understood.

Table 3 Substrate scope for reaction of 4 with various arylboronic acidsa

image file: c6ra10302b-u4.tif

a Reaction yields were determined by 19F NMR spectroscopy using 4,4′-difluorobiphenyl (−117.0 ppm) as internal standard.b Values in parentheses are isolated yields using column chromatography.
image file: c6ra10302b-u5.tif


Although at the current stage, the detailed reaction pathway remains still elusive for the aerobic trifluoromethylation using these Cu(III) CF3 complexes, a plausible mechanism is suggested as in Scheme 3 based on the experimental observations. A key Cu(III) intermediate i is proposed to account for the formation of Ar–CF3 via reductive elimination, which should result from ligand exchange of initial complex 3 with Ar ligand from ArB(OH)2 in the presence of F salt and water impurity. Accordingly, intermediate ii (or iii by comproportionation of ii with 3) is formed which is then oxidized by oxygen in the presence of water to generate a new Cu(III) intermediate iv. Ligand exchange of Ar with iv should produce intermediate i′ that can give a second Ar–CF3 after reductive elimination. This mechanism proposal can rationalize the indispensable role of F salt and oxygen for this reaction, and the optimal reaction stoichiometry of 1[thin space (1/6-em)]:[thin space (1/6-em)]2 of 3 with ArB(OH)2. In addition, the observation of better reactivity of electron-rich arylboronic acids and para-substituted arylboronic acids (please refer to Table 2) can also be explained by the more facile transmetalation of Ar of ArB(OH)2 to copper center in critical intermediate i and i′ for electronically more rich and sterically less demanding Ar group. Further studies are underway on the elucidation of the detailed mechanism and the preparation and reactivity of CuIII CF3 complexes with other ancillary ligands.


image file: c6ra10302b-s3.tif
Scheme 3 Proposed mechanism for the aerobic trifluoromethylation of arylboronic acids by CuIII CF3 complex 3.

In summary, CuIII trifluoromethyl complexes 2–4 with representative phen or bpy ligand were prepared and isolated at room temperature via oxidation of CuI precursors by AgF in the presence of CF3SiMe3 and have been fully characterized in both solid state and solution phase. Reactivity studies show that 3 and 4 are highly reactive with aryl and heteroaryl boronic acids under mild conditions. Up to quantitative yields can be achieved at room temperature for trifluoromethylation of boronic acids using 3. In contrast, 2 shows little reactivity. The isolation and reactivity of these CuIII CF3 complexes provides valuable information for copper trifluoromethylation chemistry.

Acknowledgements

This study was supported by the National Natural Science Foundation of China (No. 21472068, 21202062) and the Natural Science Foundation of Jiangsu (No. BK2012108).

Notes and references

  1. For organofluorine chemistry and applications, see: (a) P. Kirsch, Modern Fluoroorganic Chemistry, Wiley-VCH, Weinheim, Germany, 2004 Search PubMed; (b) S. Purser, P. R. Moore, S. Swallow and V. V. Gouverneur, Chem. Soc. Rev., 2008, 37, 320 RSC; (c) Y. Jiang, H. Yu, Y. Fu and L. Liu, Sci. China: Chem., 2015, 58, 673 CrossRef CAS.
  2. For general reviews, see: (a) M. Schlosser, Angew. Chem., Int. Ed., 2006, 45, 5432 CrossRef CAS PubMed; (b) J.-A. Ma and D. Cahard, Chem. Rev., 2008, 108, PR1 CrossRef CAS PubMed; (c) V. V. Grushin, Acc. Chem. Res., 2010, 43, 160 CrossRef CAS PubMed; (d) T. Furuya, A. S. Kamlet and T. Ritter, Nature, 2011, 473, 470 CrossRef CAS PubMed; (e) T. Liang, C. N. Neumann and T. Ritter, Angew. Chem., Int. Ed., 2013, 52, 8214 CrossRef CAS PubMed; (f) C. Zhang, Org. Biomol. Chem., 2014, 12, 6580 RSC; (g) S. Barata-Vallejo, S. M. Bonesi and A. Postigo, Org. Biomol. Chem., 2015, 13, 11153 RSC.
  3. For reviews on copper-mediated trifluoromethylation reactions, see: (a) O. A. Tomashenko and V. V. Grushin, Chem. Rev., 2011, 111, 4475 CrossRef CAS PubMed; (b) S. Roy, B. T. Gregg, G. W. Gribble, V.-D. Le and S. Roy, Tetrahedron, 2011, 67, 2161 CrossRef CAS; (c) T. Liu and Q. Shen, Eur. J. Org. Chem., 2012, 6679 CrossRef CAS; (d) X. Liu and X. Wu, Synlett, 2013, 1882 CAS; (e) P. Chen and G. Liu, Synthesis, 2013, 2919 CAS.
  4. For some examples of copper-mediated trifluoromethylation of aryl and heteroaryl halides, see: (a) Z.-Y. Yang and D. J. Burton, J. Fluorine Chem., 2000, 102, 89 CrossRef CAS; (b) F. Cottet and M. Schlosser, Eur. J. Org. Chem., 2004, 3793 CrossRef CAS; (c) 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; (d) 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 PubMed; (e) A. Zanardi, M. A. Novikov, E. Martin, J. Benet-Buchholz and V. V. Grushin, J. Am. Chem. Soc., 2011, 133, 20901 CrossRef CAS PubMed; (f) M. M. Kremlev, A. I. Mushta, W. Tyrra, Y. L. Yagupolskii, D. Naumann and A. Möller, J. Fluorine Chem., 2012, 133, 67 CrossRef CAS; (g) M. Huiban, M. Tredwell, S. Mizuta, Z. Wan, X. Zhang, T. L. Collier, V. Gouverneur and J. Passchier, Nat. Chem., 2013, 5, 941 CrossRef CAS PubMed; (h) A. Lishchynskyi, M. A. Novikov, E. Martin, E. C. Escudero-Adan, P. Novak and V. V. Grushin, J. Org. Chem., 2013, 78, 11126 CrossRef CAS PubMed.
  5. For catalytic copper trifluoromethylation of aryl halides, see: (a) Q.-Y. Chen and S.-W. Wu, J. Chem. Soc., Chem. Commun., 1989, 705 RSC; (b) M. Oishi, H. Kondo and H. Amii, Chem. Commun., 2009, 1909 RSC; (c) H. Kondo, M. Oishi, K. Fujikawa and H. Amii, Adv. Synth. Catal., 2011, 383, 1247 CrossRef; (d) T. Knauber, F. Arikan, G.-V. Roschenthaler and L. J. Goossen, Chem.–Eur. J., 2011, 17, 2689 CrossRef CAS PubMed; (e) Y. Li, T. Chen, H. Wang, R. Zhang, K. Jin, X. Wang and C. Duan, Synlett, 2011, 1713 CrossRef CAS; (f) I. Popov, S. Lindeman and O. Daugulis, J. Am. Chem. Soc., 2011, 133, 9286 CrossRef CAS PubMed; (g) Z. Gonda, S. Kovacs, C. Weber, T. Gati, A. Meszaros, A. Kotschy and Z. Novak, Org. Lett., 2014, 16, 4268 CrossRef CAS PubMed; (h) Y. Miyake, S. Ota, M. Shibata, K. Nakajima and Y. Nishibayashi, Org. Biomol. Chem., 2014, 12, 5594 RSC.
  6. For selected copper-mediated trifluoromethylation of arylboronic acids, see: (a) L. Chu and F.-L. Qing, Org. Lett., 2010, 12, 5060 CrossRef CAS PubMed; (b) T. D. Senecal, A. T. Parsons and S. L. Buchwald, J. Org. Chem., 2011, 76, 1174 CrossRef CAS PubMed; (c) T. Liu and Q. Shen, Org. Lett., 2011, 13, 2342 CrossRef CAS PubMed; (d) J. Xu, D.-F. Luo, B. Xiao, Z.-J. Liu, T.-J. Gong, Y. Fu and L. Liu, Chem. Commun., 2011, 47, 4300 RSC; (e) C.-P. Zhang, J. Cai, C.-B. Zhou, X.-P. Wang, X. Zheng, Y.-C. Gu and J.-C. Xiao, Chem. Commun., 2011, 47, 9516 RSC; (f) X. Jiang, L. Chu and F.-L. Qing, J. Org. Chem., 2012, 77, 1251 CrossRef CAS PubMed; (g) B. A. Khan, A. E. Buba and L. J. Goossen, Chem.–Eur. J., 2012, 18, 1577 CrossRef CAS PubMed; (h) N. D. Litvinas, P. S. Fier and J. F. Hartwig, Angew. Chem., Int. Ed., 2012, 51, 536 CrossRef CAS PubMed; (i) T. Liu, X. Shao, Y. Wu and Q. Shen, Angew. Chem., Int. Ed., 2012, 51, 540 CrossRef CAS PubMed; (j) P. Novák, A. Lishchynskyi and V. V. Grushin, Angew. Chem., Int. Ed., 2012, 51, 7767 CrossRef PubMed; (k) Y. Ye and M. S. Sanford, J. Am. Chem. Soc., 2012, 134, 9034 CrossRef CAS PubMed; (l) Y. Ye, S. A. Kunzi and M. S. Sanford, Org. Lett., 2012, 14, 4979 CrossRef CAS PubMed; (m) J. Xu, B. Xiao, C.-Q. Xie, D.-F. Luo, L. Liu and Y. Fu, Angew. Chem., Int. Ed., 2012, 51, 12551 CrossRef CAS PubMed; (n) N. Nebra and V. V. Grushin, J. Am. Chem. Soc., 2014, 136, 16998 CrossRef CAS PubMed.
  7. Copper-mediated trifluoromethylation of arenes, see: (a) R. Koller, K. Stanek, D. Stolz, R. Aardoom, K. Niedermann and A. Togni, Angew. Chem., Int. Ed., 2009, 48, 4332 CrossRef CAS PubMed; (b) L. Chu and F.-L. Qing, J. Am. Chem. Soc., 2012, 134, 1298 CrossRef CAS PubMed; (c) R. Shimizu, H. Egami, T. Nagi, J. Chae, Y. Hamashima and M. Sodeoka, Tetrahedron Lett., 2010, 51, 5947 CrossRef CAS; (d) M. S. Wiehn, E. V. Vinogradova and A. Togni, J. Fluorine Chem., 2010, 131, 951 CrossRef CAS; (e) S. Cai, C. Chen, Z. Sun and C. Xi, Chem. Commun., 2013, 49, 4552 RSC; (f) S. Jana, A. Ashokan, S. Kumar, A. Verma and S. Kumar, Org. Biomol. Chem., 2015, 13, 8411 RSC.
  8. (a) H. Egami and M. Sodeoka, Angew. Chem., Int. Ed., 2014, 53, 8293 CrossRef PubMed; (b) E. Merino and C. Nevado, Chem. Soc. Rev., 2014, 43, 6598 RSC.
  9. Copper-mediated trifluoromethylation of alkynes, see: (a) L. Chu and F.-L. Qing, J. Am. Chem. Soc., 2010, 132, 7262 CrossRef CAS PubMed; (b) K. Zhang, X.-L. Qiu, Y. Huang and F.-L. Qing, Eur. J. Org. Chem., 2012, 58 CrossRef; (c) D.-F. Luo, J. Xu, Y. Fu and Q.-X. Guo, Tetrahedron Lett., 2012, 53, 2769 CrossRef CAS; (d) Y.-L. Ji, J.-J. Kong, J.-H. Lin, J.-C. Xiao and Y.-C. Gu, Org. Biomol. Chem., 2014, 12, 2903 RSC.
  10. D. M. Wiemers and D. J. Burton, J. Am. Chem. Soc., 1986, 108, 832 CrossRef CAS.
  11. For well-characterized CuI–CF3 complexes, see: (a) G. G. Dubinina, H. Furutachi and D. A. Vicic, J. Am. Chem. Soc., 2008, 130, 8600 CrossRef CAS PubMed; (b) G. G. Dubinina, J. Ogikubo and D. A. Vicic, Organometallics, 2008, 27, 6233 CrossRef CAS; (c) H. Morimoto, T. Tsubogo, N. D. Litvinas and J. F. Hartwig, Angew. Chem., Int. Ed., 2011, 50, 3793 CrossRef CAS PubMed; (d) O. A. Tomashenko, E. C. Escudero-Adan, M. M. Belmonte and V. V. Grushin, Angew. Chem., Int. Ed., 2011, 50, 7655 CrossRef CAS PubMed; (e) Z. Weng, R. Lee, W. Jia, Y. Yuan, W. Wang, X. Feng and K.-W. Huang, Organometallics, 2011, 30, 3229 CrossRef CAS; (f) A. I. Konovalov, J. Benet-Buchholz, E. Martin and V. V. Grushin, Angew. Chem., Int. Ed., 2013, 52, 11637 CrossRef CAS PubMed; (g) L. I. Panferova, F. M. Miloserdov, A. Lishchynskyi, M. M. Belmonte, J. Benet-Buchholz and V. V. Grushin, Angew. Chem., Int. Ed., 2015, 54, 5218 CrossRef CAS PubMed.
  12. (a) A. J. Hickman and M. S. Sanford, Nature, 2012, 484, 177 CrossRef CAS PubMed; (b) A. Casitas and X. Ribas, Chem. Sci., 2013, 4, 2301 RSC.
  13. For CuIII–CF3 complexes, see: (a) M. A. Willert-Porada, D. J. Burton and N. C. Baenziger, J. Chem. Soc., Chem. Commun., 1989, 1633 RSC. The crystal structure was later redetermined by Marsh with corrected space group, see: R. E. Marsh, Acta Crystallogr., 1997, B53, 317 CAS; (b) D. Naumann, T. Roy, K.-F. Tebbe and W. Crump, Angew. Chem., Int. Ed. Engl., 1993, 32, 1482 CrossRef; (c) A. M. Romine, N. Nebra, A. I. Konovalov, E. Martin, J. Benet-Buchholz and V. V. Grushin, Angew. Chem., Int. Ed., 2015, 54, 2745 CrossRef CAS PubMed; (d) J. A. Schlueter, U. Geiser, J. M. Williams, H. H. Wang, W.-K. Kwok, J. A. Fendrich, K. D. Carlson, C. A. Achenbach, J. D. Dudek, D. Naumann, T. Roy, J. E. Schirber and W. R. Bayless, J. Chem. Soc., Chem. Commun., 1994, 1599 RSC The crystal structure was later redetermined, see: U. Geiser, J. A. Schlueter, J. M. Williams, D. Naumann and T. Roy, Acta Crystallogr., 1995, B51, 789 Search PubMed.
  14. For the preparation of phenCu(PPh3)Br (1), see: R. K. Gujadhur, C. G. Bates and D. Venkataraman, Org. Lett., 2001, 3, 4315 CrossRef CAS PubMed.
  15. CCDC 1402421 (3), 1402422 (2) and 1402423 (5) contain the supplementary crystallographic data for this manuscript. Please see also ESI for some details on crystallographic studies..
  16. During the work-up process, a small amount of green impurity was observed which might imply the involvement of Cu(II) intermediates in the reaction process.
  17. For the formation of intermediate A and crystal structure of PPh3-ligated AgBr (5), please refer to ESI..
  18. For detailed reaction procedure and 19F NMR determination of reaction yields, please refer to ESI..
  19. Reaction of 1 eq. or 3 eq. of 6a with complex 3 led to lower yields.

Footnote

Electronic supplementary information (ESI) available: Experimental details, spectroscopic characterization data, X-ray crystallographic study, 19F NMR monitoring of reaction course and general procedures of reaction of 2–4 with arylboronic acids. CCDC 1402421–1402423. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6ra10302b

This journal is © The Royal Society of Chemistry 2016