Metal-free directed C-H borylation of 2-(N-methylanilino)-5-fluoropyridines and 2-benzyl-5-fluoropyridines

Gaorong Wu Xiaobo Xu Shuai Wang Lu Chen Binghan Pang Tao Ma Yafei Ji

Citation:  Gaorong Wu, Xiaobo Xu, Shuai Wang, Lu Chen, Binghan Pang, Tao Ma, Yafei Ji. Metal-free directed C-H borylation of 2-(N-methylanilino)-5-fluoropyridines and 2-benzyl-5-fluoropyridines[J]. Chinese Chemical Letters, 2022, 33(4): 2005-2008. doi: 10.1016/j.cclet.2021.09.081 shu

Metal-free directed C-H borylation of 2-(N-methylanilino)-5-fluoropyridines and 2-benzyl-5-fluoropyridines

English

  • Organoborons play a major role in organic synthesis as their power in the conversion of functional groups coupled with the low toxicity and ease of handling. Among various C−H borylation strategies, one of the commonly known strategies is a metal-catalyzed directed approach [1-3]. Owing to the extensive range of applications in natural products and drugs, the development of protocols for the functionalization of N-methylanilines has attracted more attention in recent years. Traditional methods used for the borylation of N-methylanilines were focused on noble transition-metal-catalyzed, such as Ir and Pd (Scheme 1a) [4-6]. However, these works usually need ligands and harsh reaction conditions. From an economic and environmental perspective, the ability to achieve C−H borylation of N-methylanilines under mild conditions by using cheaper boron reagents and avoiding the use of transition metals would be very significant.

    Scheme 1

    Scheme 1.  Borylation of anilines and N-methylanilines.

    At present, the use of strong Lewis acids such as B(C6F5)3 or BX3 (X = Cl, Br) for the C(sp2)−H bond borylation is considered to be a very effective method [7-25]. Among these, BBr3 is particularly prominent because of its strong reactivity, low cost and easy availability. In 2010, Murakami developed a seminal method to borylate 2-aryl-pyridines using BBr3 [26]. Two years later, Fu's group had successfully developed an efficient metal-free ortho C−H borylation via sequential borylation of substituted 2-phenoxypyridines with BBr3 following esterification with pinacol [27]. After that, other N-directing groups such as pyrazoles, pyrazines, pyrimidines, pyridines, imidazolones, imidazoles and quinolines were developed [28-36]. In 2019, Shi group and Ingleson group described the ortho-C−H borylation of aniline derivatives with an acyl moiety by using BBr3 (Scheme 1b) [37, 38]. More recently, Chatani and co-workers reported the pyrimidine-directed metal-free C–H borylation of 2-pyrimidylanilines (Scheme 1c) [39]. These researches provided outstanding strategies for the borylation of anilines. Although substantial progress has been made in the development of metal-free directed C−H borylation reactions with BBr3, the borylation of N-methylanilines directed by pyridine under metal-free conditions has never been reported.

    Inspired by these achievements and given the importance of N-methylaniline derivatives, we hope to develop a new pyridine directing group to obtain N-methylanilines borates. Herein, we reported a novel route for directed C–H borylation of 2-(N-methylanilino)-5-fluoropyridines using BBr3. The reaction was also applicable to 2-benzyl-5-fluoropyridine derivatives as a substrate (Scheme 1d). Moreover, the obtained boronic esters exhibited good derivatization applications and the directing group could be removed in an acceptable yield, which provided a practical, efficient and simple pathway for synthesizing structurally diversified N-methylaniline and 2-benzyl-5-fluoropyridine derivatives.

    Our initial study focused on screening the directing groups (Scheme 2). First, a solution of 1a, DIPEA (N, N-diisopropylethylamine) (1.0 equiv.) and 2a (2.0 equiv.) in dried DCM was stirred at room temperature under a N2 atmosphere for 6 h, and then pinacol (2.0 equiv.) and Et3N (10.0 equiv.) were added and the mixture was stirred for another 2 h offering the desired compound 3a in 72% yield. Then we performed a brief survey of the electronic effects on pyridine ring. The results showed that, compared to 1a, 5-CH3 substituted substrate decreased the yield, but 5-Cl substituted substrate increased the yield, indicating that electron-withdrawing group on pyridine ring was beneficial to the reaction. Inspired by this result, stronger electron-withdrawing group 5-F and 5-CF3 were conducted. To our surprise, 3d was obtained in excellent yield (87%), but the yield of 3e was decreased, revealing the importance of proper electronic effects on pyridine ring. Finally, we investigated the influence of the position of the fluorine atom on the pyridine. 3-F and 4-F substituted substrates lowered the borylation efficiency compare to 5-F. Besides, only a trace amount of 3h was obtained when the fluorine atom at the 6-position, possibly due to the steric hindrance that prevented the formation of N−B bond [40]. Given that the fluorine atom played a crucial role in drugs and this reaction, 5-fluoropyridine was identified as the optimal directing group for further surveys.

    Scheme 2

    Scheme 2.  Screening of directing groups. Reaction conditions: (i) 1 (0.2 mmol), 2a (0.4 mL, 1 mol/L in DCM), DIPEA (0.2 mmol), DCM (1.5 mL), N2, 6 h; (ii) pinacol (0.4 mmol, dissolved in 1 mL dry DCM), Et3N (2.0 mmol), r.t., N2, 2 h. Isolated yields.

    Next, we evaluated the main parameters to optimize reaction conditions (Table 1). When the boron source was BCl3 and ClBcat, no target product was obtained (entries 2 and 3). To our delight, increasing the amount of 2a to 3.0 equiv., the yield of 3d was as high as 93% (entry 4). However, further increasing 2a did not give a higher yield (entry 5). Among various bases, DIPEA functioned as the best one (entries 6−8). Regrettably, we found that other boronate esters could not be accessed when different diols were used instead of pinacol (entries 9 and 10). Changing the solvent to DCE and CFC-113a (1,1,1-trichlorotrifluoroethane) provided 21% and 83% yields of 3d, respectively. But the use of Lewis basic solvent THF was inefficient (entries 11−13).

    Table 1

    Table 1.  Optimization of the reaction conditions.a
    DownLoad: CSV

    With the optimized reaction conditions in hand, the scope of 2-(N-methylanilino)-5-fluoropyridines was examined as shown in Scheme 3. To our delight, amines bearing electron-donating (Me, OTBS and Ph) or -withdrawing (F, Cl and Br) groups at the ortho-, meta- and para-positions afforded the corresponding borylation products in 57%−93% yields (3d, 3i3v). Generally, the yields of the ortho-substitution substrates were lower than the meta and para ones. The large steric hindrance group OTBS (3j) at the ortho position of the amine was also suitable to C−H borylation. In comparison with 3n3p, amines containing CH3, F and Cl groups at para-position (3q3s) delivered relatively better yields. But stronger electron-withdrawing group NO2 was not suitable for this reaction. Notably, 3,4-dichloro substituted amine 1u worked well, affording the desired products in 78% yield. Moreover, treatment of 1-naphthyl-substituted substrate 1v could afford 3v in an excellent yield under the standard reaction conditions. However, no desired products were obtained with almost all starting materials intact when amines containing N-H or N-Ph substituents instead of N-Me, indicating that the great importance of the Me group.

    Scheme 3

    Scheme 3.  Substrate scope of 2-(N-methylanilino)-5-fluoropyridines. Reaction conditions: 1d, 1i1x (0.2 mmol), 2a (0.6 mL, 1 mol/L in DCM), DIPEA (0.2 mmol), DCM (1.5 mL), N2, 6 h; then pinacol (0.6 mmol, dissolved in 1 mL dry DCM), Et3N (2.0 mmol), r.t., N2, 2 h, isolated yields.

    The strategy was not restricted to 2-(N-methylanilino)-5-fluoropyridines; 2-benzyl-5-fluoropyridines were also viable substrates for the reaction. The scope was investigated under the developed reaction conditions (Scheme 4). Most substrates exhibited high borate activity, especially 4e, 4g, 4j and 4o, and the yield of the corresponding products all exceeded 90%. In general, the compounds bearing electron-donating groups such as Me, C2H5, iPr and tBu gave the target products in higher yields than electron-withdrawing ones (such as F, Cl, Br, CN), indicating that the electron-donating groups were beneficial to the electrophilic borylation reaction by increasing the electron cloud density of the aromatic ring [41]. Similarly, among the ortho-, meta-, and para-positions substituted substrates, the borylated yields of the ortho-position substrates were relatively lower. However, the borylation of 4k became sluggish, in which the strong electronegativity affected the formation of C−B bond, thus leading to poor conversion. Subjecting compound 4n containing two methyl groups at the C3 and C4 positions could undergo C–H borylation. Besides, 1-naphthyl-substituted substrate 4o generated 5o in an excellent yield (92%).

    Scheme 4

    Scheme 4.  Substrate scope of 2-benzyl-5-fluoropyridines. Reaction conditions: 4a4o (0.2 mmol), 2a (0.6 mL, 1 mol/L in DCM), DIPEA (0.2 mmol), DCM (1.5 mL), N2, 6 h; then pinacol (0.6 mmol, dissolved in 1 mL dry DCM), Et3N (2.0 mmol), r.t., N2, 2 h. Isolated yields.

    To certify the potential utility of this method, we realized the arylation of 3i via Suzuki−Miyaura coupling in 81% yield. Subsequent MeOTf-driven reductive removal of the pyridine directing group provided the N-methyl-[1,1′-biphenyl]-2-amine in an acceptable yield (Scheme 5) [42, 43]. Importantly, this convenient, easily handled, synthetically useful protocol could be executed on a gram-scale without any difficulties. The borylated compound 5a could be transformed into a number of important compounds via a Suzuki−Miyaura coupling (8) and oxidation with NaBO3·4H2O (9) in good to excellent yields [42, 44]. Moreover, the synthesis of 2-(2-chlorobenzyl)-5-fluoropyridine (10) and 2-(2-iodobenzyl)-5-fluoropyridine (11) could be achieved by deborylative chlorination and iodination (Scheme 6) [27].

    Scheme 5

    Scheme 5.  Synthetic application of product 3i and removal of the directing group. Reaction conditions: (a) 3i (0.2 mmol), PhI (0.24 mmol), Pd(PPh3)4 (0.01 mmol), Na2CO3 (0.6 mmol, 2 mol/L in H2O), PhMe (1.5 mL), reflux, N2, 12 h. (b) 6 (0.2 mmol), MeOTf (0.4 mmol), DCM (1.0 mL), r.t., N2, 24 h; then NaBH4 (1.2 mmol), MeOH (1.5 mL), r.t., air, 6 h. Isolated yields.

    Scheme 6

    Scheme 6.  Gram-scale reaction and synthetic application of product 5a. Reaction conditions: (a) 4a (6.0 mmol), 2a (18.0 mL, 1 mol/L in DCM), DIPEA (6.0 mmol), DCM (15.0 mL), N2, 6 h; then pinacol (18.0 mmol, dissolved in 10 mL dry DCM), Et3N (60.0 mmol), r.t., N2, 2 h. (b) 5a (0.2 mmol), PhI (0.24 mmol), Pd(PPh3)4 (0.01 mmol), Na2CO3 (0.6 mmol, 2 mol/L in H2O), PhMe (1.5 mL), reflux, N2, 12 h. (c) 5a (0.2 mmol), NaBO3.4H2O (0.8 mmol), THF (1.0 mL), H2O (1.0 mL), r.t., 12 h. (d) 5a (0.2 mmol), CuCl2 (0.6 mmol), MeOH (1.0 mL), H2O (1.0 mL), 90 ℃, 24 h. (e) 5a (0.2 mmol), NaI (1.0 mmol), Cu2O (0.02 mmol), NH3.H2O (0.5 mmol), EtOH, r.t., air, 20 h. Isolated yields.

    To elucidate the mechanism of this reaction, a control experiment was performed. Radical scavengers TEMPO and CHD were added into the reaction of 1i and 2a, respectively. As a result, the yield of 3i was almost unaffected, thus precluding the radical pathway (Scheme 7a) [45]. Subsequently, the kinetic isotope effect experiment between equivalent 1i and 1i' with 2a under the standard conditions for 10 min was conducted (Scheme 7b) and gave a KIE value of 2.03 by 1H NMR spectroscopic analysis (Supporting information). This result indicated that the cleavage of C(sp2)−H bond might be related to the rate-determining step [46].

    Scheme 7

    Scheme 7.  Effect of TEMPO and CHD and kinetic isotope effect experiments.

    On the basis of experimental investigations and literature precedents [9, 10, 14, 24-26, 39]. We depicted a plausible mechanism in Scheme 8. First, a Lewis acid–base adduct was initially formed between BBr3 and the N atom of 5-fluoropyridine group in 1i. Then Br transferred from A to another BBr3 and generated a borenium species B. An electrophilic aromatic substitution proceeded readily to offer a Wheland intermediate C, because of the strong electrophilic nature of the N-methylaniline ring in B. Subsequently, proton abstraction by DIPEA gave the dibromoboron complex D. Finally, under the protection of pinacol, compound 3i was obtained.

    Scheme 8

    Scheme 8.  Proposed mechanism.

    In summary, we firstly reported the method for C−H borylation of 2-(N-methylanilino)-5-fluoropyridines and 2-benzyl-5-fluoropyridines under metal-free conditions. The obtained borylated products could be converted to various useful intermediates through common synthetic methods. Importantly, the directing group could be removed to enrich synthetic applications. Considering its simplicity and versatility, we expect that this novel borylation approach could show great promise in the screening of potential pharmaceuticals of 2-(N-methylanilino)-5-fluoropyridines and 2-benzyl-5-fluoropyridines.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    We gratefully thank the National Natural Science Foundation of China (No. 21676088) for financial support.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2021.09.081.


    1. [1]

      M.Y. Liu, S.B. Hong, W. Zhang, W. Deng, Chin. Chem. Lett. 26(2015) 373-376. doi: 10.1016/j.cclet.2014.12.004

    2. [2]

      A.K. Cook, S.D. Schimler, A.J. Matzger, M.S. Sanford, Science 351(2016) 1421-1424. doi: 10.1126/science.aad9289

    3. [3]

      T.J. Gong, Y.Y. Jiang, Y. Fu, Chin. Chem. Lett. 25(2014) 397-400. doi: 10.1016/j.cclet.2014.01.006

    4. [4]

      M.E. Hoque, M.M.M. Hassan, B. Chattopadhyay, J. Am. Chem. Soc. 143(2021) 5022-5037. doi: 10.1021/jacs.0c13415

    5. [5]

      M.R. Smith, R. Bisht, C. Haldar, et al., ACS Catal. 8(2018) 6216-6223. doi: 10.1021/acscatal.8b00641

    6. [6]

      T.A. Boebel, J.F. Hartwig, J. Am. Chem. Soc. 130(2008) 7534-7535. doi: 10.1021/ja8015878

    7. [7]

      Y.L. Liu, G. Kehr, C.G. Daniliuc, G. Erker, Chem. Eur. J. 23(2017) 12141-12144. doi: 10.1002/chem.201701771

    8. [8]

      V. Bagutski, A. Del Grosso, J.A. Carrillo, et al., J. Am. Chem. Soc. 135(2013) 474-487. doi: 10.1021/ja3100963

    9. [9]

      J. Lv, B. Zhao, Y. Yuan, Y. Han, Z. Shi, Nat. Commun. 11(2020) 1316. doi: 10.1038/s41467-020-15207-x

    10. [10]

      G. Wu, X. Fu, Y. Wang, et al., Org. Lett. 22(2020) 7003-7007. doi: 10.1021/acs.orglett.0c02552

    11. [11]

      F. Focante, I. Camurati, D. Nanni, R. Leardini, L. Resconi, Organometallics 23(2004) 5135-5141. doi: 10.1021/om0499308

    12. [12]

      Z.J. Wang, X. Chen, L. Wu, et al., Angew. Chem. Int. Ed. 60(2021) 8500-8504. doi: 10.1002/anie.202016573

    13. [13]

      V. Iashin, D. Berta, K. Chernichenko, et al., Chem. Eur. J. 26(2020) 13873-13879. doi: 10.1002/chem.202001436

    14. [14]

      S. Rej, N. Chatani, J. Am. Chem. Soc. 143(2021) 2920-2929. doi: 10.1021/jacs.0c13013

    15. [15]

      F.A. Davis, M.J.S. Dewar, J. Am. Chem. Soc. 90(1968) 3511-3515. doi: 10.1021/ja01015a039

    16. [16]

      C. Cazorla, T.S. De Vries, E. Vedejs, Org. Lett. 15(2013) 984-987. doi: 10.1021/ol303203m

    17. [17]

      Z. Zhao, Z. Chang, B. He, et al., Chem. Eur. J. 19(2013) 11512-11517. doi: 10.1002/chem.201301815

    18. [18]

      J. Chen, R.A. Lalancette, F. Jäkle, Organometallics 32(2013) 5843-5851. doi: 10.1021/om400426g

    19. [19]

      D.L. Crossley, I.A. Cade, E.R. Clark, et al., Chem. Sci. 6(2015) 5144-5151. doi: 10.1039/C5SC01800E

    20. [20]

      A.C. Shaikh, D.S. Ranade, S. Thorat, et al., Chem. Commun. 51(2015) 16115-16118. doi: 10.1039/C5CC06351E

    21. [21]

      M. Yusuf, K. Liu, F. Guo, R.A. Lalancette, F. Jäkle, Dalton Trans 45(2016) 4580-4587. doi: 10.1039/C5DT05077D

    22. [22]

      D.L. Crossley, I. Vitorica-Yrezabal, M.J. Humphries, M.L. Turner, M.J. Ingleson, Chem. Eur. J. 22(2016) 12439-12448. doi: 10.1002/chem.201602010

    23. [23]

      K. Liu, R.A. Lalancette, F. Jäkle, J. Am. Chem. Soc. 139(2017) 18170-18173. doi: 10.1021/jacs.7b11062

    24. [24]

      S.A. Iqbal, K. Yuan, J. Cid, J. Pahl, M.J. Ingleson, Org. Biomol. Chem. 19(2021) 2949-2958. doi: 10.1039/D1OB00018G

    25. [25]

      J. Lv, B. Zhao, Y. Han, Y. Yuan, Z. Shi, Chin. Chem. Lett. 32(2021) 691-694. doi: 10.1016/j.cclet.2020.06.028

    26. [26]

      N. Ishida, T. Moriya, T. Goya, M. Murakami, J. Org. Chem. 75(2010) 8709-8712. doi: 10.1021/jo101920p

    27. [27]

      L. Niu, H. Yang, R. Wang, H. Fu, Org. Lett. 14(2012) 2618-2621. doi: 10.1021/ol300950r

    28. [28]

      V. Mukundam, S. Sa, A. Kumari, R. Das, K. Venkatasubbaiah, J. Mater. Chem. C 7(2019) 12725-12737. doi: 10.1039/C9TC04309H

    29. [29]

      C. Zhu, Z.H. Guo, A.U. Mu, et al., J. Org. Chem. 81(2016) 4347-4352. doi: 10.1021/acs.joc.6b00238

    30. [30]

      S. Tanaka, Y. Saito, T. Yamamoto, T. Hattori, Org. Lett. 20(2018) 1828-1831. doi: 10.1021/acs.orglett.8b00335

    31. [31]

      Y. Li, B. Pang, H. Meng, et al., Tetrahedron Lett. 60(2019) 151286. doi: 10.1016/j.tetlet.2019.151286

    32. [32]

      Y. Li, H. Meng, T. Liu, et al., Adv. Mater. 31(2019) 1904585. doi: 10.1002/adma.201904585

    33. [33]

      C.R. Opie, H. Noda, M. Shibasaki, N. Kumagai, Chem. Eur. J. 25(2019) 4648-4653. doi: 10.1002/chem.201900715

    34. [34]

      S. Olsen, M.S. Baranov, N.S. Baleeva, et al., PCCP 18(2016) 26703-26711. doi: 10.1039/C6CP02423H

    35. [35]

      M.S. Baranov, K.M. Solntsev, N.S. Baleeva, et al., Chem. Eur. J. 20(2014) 13234-13241. doi: 10.1002/chem.201403678

    36. [36]

      K. Dhanunjayarao, S. Sa, B.P.R. Aradhyula, K. Venkatasubbaiah, Tetrahedron 74(2018) 5819-5825. doi: 10.1016/j.tet.2018.08.019

    37. [37]

      J. Lv, X. Chen, X.S. Xue, et al., Nature 575(2019) 336-340. doi: 10.1038/s41586-019-1640-2

    38. [38]

      S.A. Iqbal, J. Cid, R.J. Procter, et al., Angew. Chem. Int. Ed. 58(2019) 15381-15385. doi: 10.1002/anie.201909786

    39. [39]

      S. Rej, A. Das, N. Chatani, Chem. Sci. 12(2021) 11447-11454. doi: 10.1039/D1SC02937A

    40. [40]

      S.A. Iqbal, J. Pahl, K. Yuan, M.J. Ingleson, Chem. Soc. Rev. 49(2020) 4564-4591. doi: 10.1039/C9CS00763F

    41. [41]

      F.G. Oliveira, F.L. Rodrigues, A.V.B. de Oliveira, D.V.L.M. Marçal, P.M. Esteves, Struct. Chem. 28(2017) 545-553. doi: 10.1007/s11224-017-0915-1

    42. [42]

      M. Alessi, A.L. Larkin, K.A. Ogilvie, et al., J. Org. Chem. 72(2007) 1588-1594. doi: 10.1021/jo0620359

    43. [43]

      W. Miura, K. Hirano, M. Miura, Org. Lett. 18(2016) 3742-3745. doi: 10.1021/acs.orglett.6b01762

    44. [44]

      X. Zou, H. Zhao, Y. Li, et al., J. Am. Chem. Soc. 141(2019) 5334-5342. doi: 10.1021/jacs.8b13756

    45. [45]

      F. Michailidou, N. Klöcker, N.V. Cornelissen, et al., Angew. Chem. Int. Ed. 60(2021) 480-485. doi: 10.1002/anie.202012623

    46. [46]

      J.Y. Zhou, J.B. Wang, Sci. China Chem. 46(2016) 573-578. doi: 10.1360/N032015-00254

  • Scheme 1  Borylation of anilines and N-methylanilines.

    Scheme 2  Screening of directing groups. Reaction conditions: (i) 1 (0.2 mmol), 2a (0.4 mL, 1 mol/L in DCM), DIPEA (0.2 mmol), DCM (1.5 mL), N2, 6 h; (ii) pinacol (0.4 mmol, dissolved in 1 mL dry DCM), Et3N (2.0 mmol), r.t., N2, 2 h. Isolated yields.

    Scheme 3  Substrate scope of 2-(N-methylanilino)-5-fluoropyridines. Reaction conditions: 1d, 1i1x (0.2 mmol), 2a (0.6 mL, 1 mol/L in DCM), DIPEA (0.2 mmol), DCM (1.5 mL), N2, 6 h; then pinacol (0.6 mmol, dissolved in 1 mL dry DCM), Et3N (2.0 mmol), r.t., N2, 2 h, isolated yields.

    Scheme 4  Substrate scope of 2-benzyl-5-fluoropyridines. Reaction conditions: 4a4o (0.2 mmol), 2a (0.6 mL, 1 mol/L in DCM), DIPEA (0.2 mmol), DCM (1.5 mL), N2, 6 h; then pinacol (0.6 mmol, dissolved in 1 mL dry DCM), Et3N (2.0 mmol), r.t., N2, 2 h. Isolated yields.

    Scheme 5  Synthetic application of product 3i and removal of the directing group. Reaction conditions: (a) 3i (0.2 mmol), PhI (0.24 mmol), Pd(PPh3)4 (0.01 mmol), Na2CO3 (0.6 mmol, 2 mol/L in H2O), PhMe (1.5 mL), reflux, N2, 12 h. (b) 6 (0.2 mmol), MeOTf (0.4 mmol), DCM (1.0 mL), r.t., N2, 24 h; then NaBH4 (1.2 mmol), MeOH (1.5 mL), r.t., air, 6 h. Isolated yields.

    Scheme 6  Gram-scale reaction and synthetic application of product 5a. Reaction conditions: (a) 4a (6.0 mmol), 2a (18.0 mL, 1 mol/L in DCM), DIPEA (6.0 mmol), DCM (15.0 mL), N2, 6 h; then pinacol (18.0 mmol, dissolved in 10 mL dry DCM), Et3N (60.0 mmol), r.t., N2, 2 h. (b) 5a (0.2 mmol), PhI (0.24 mmol), Pd(PPh3)4 (0.01 mmol), Na2CO3 (0.6 mmol, 2 mol/L in H2O), PhMe (1.5 mL), reflux, N2, 12 h. (c) 5a (0.2 mmol), NaBO3.4H2O (0.8 mmol), THF (1.0 mL), H2O (1.0 mL), r.t., 12 h. (d) 5a (0.2 mmol), CuCl2 (0.6 mmol), MeOH (1.0 mL), H2O (1.0 mL), 90 ℃, 24 h. (e) 5a (0.2 mmol), NaI (1.0 mmol), Cu2O (0.02 mmol), NH3.H2O (0.5 mmol), EtOH, r.t., air, 20 h. Isolated yields.

    Scheme 7  Effect of TEMPO and CHD and kinetic isotope effect experiments.

    Scheme 8  Proposed mechanism.

    Table 1.  Optimization of the reaction conditions.a

    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  511
  • HTML全文浏览量:  69
文章相关
  • 发布日期:  2022-04-15
  • 收稿日期:  2021-09-01
  • 接受日期:  2021-09-24
  • 修回日期:  2021-09-22
  • 网络出版日期:  2021-09-30
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章