Next Article in Journal
The Phase Inversion Mechanism of the pH-Sensitive Reversible Invert Emulsion
Next Article in Special Issue
Nickel-Catalyzed Three-Component Unsymmetrical Bis-Allylation of Alkynes with Alkenes: A Density Functional Theory Study
Previous Article in Journal
Development and Validation of LC-MS/MS Method for the Determination of 1-Methyl-4-Nitrosopiperazine (MNP) in Multicomponent Products with Rifampicin—Analytical Challenges and Degradation Studies
Previous Article in Special Issue
Copper-Promoted Intramolecular Oxidative Dehydrogenation for Synthesizing Dihydroisocoumarins and Isocoumarins
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Copper Complexes with N,N,N-Tridentate Quinolinyl Anilido-Imine Ligands: Synthesis and Their Catalytic Application in Chan−Lam Reactions

1
National Experimental Chemistry Teaching Center, Hebei Key Laboratory of Organic Functional Molecules, College of Chemistry and Materials Science, Hebei Normal University, Shijiazhuang 050024, China
2
Auckland Cancer Society Research Centre, Faculty of Medical and Health Sciences, Maurice Wilkins Centre, The University of Auckland, Auckland 1142, New Zealand
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(21), 7406; https://doi.org/10.3390/molecules28217406
Submission received: 28 September 2023 / Revised: 21 October 2023 / Accepted: 30 October 2023 / Published: 3 November 2023
(This article belongs to the Special Issue Advances in Transition-Metal-Catalyzed Synthesis)

Abstract

:
The treatment of 2-(ArNC(H))C6H4-HNC9H6N with n-BuLi and the subsequent addition of CuCl2 afforded the anilido-aldimine Cu(II) complexes 1-5 Cu[{2-[ArN=C(H)]C6H4}N(8-C9H6N)]Cl (Ar = 2,6-iPr2C6H3 (1), 2,4,6-(CH3)3C6H2 (2), 4-OCH3C6H4 (3), 4-BrC6H4 (4), 4-ClC6H4 (5)), respectively. All the copper complexes were fully characterized by IR, EPR and HR-MS spectra. The X-ray diffraction analysis reveals that 2 and 4 are mononuclear complexes, and the Cu atom is sitting in a slightly square-planar geometry. These Cu(II) complexes have exhibited excellent catalytic activity in the Chan–Lam coupling reactions of benzimidazole derivatives with arylboronic acids, achieving the highest yields of up to 96%.

Graphical Abstract

1. Introduction

The construction of C-N bonds is one of the most widely practiced reactions in synthetic chemistry. The name reactions, such as the Buchwald–Hartwig [1,2], Chan–Lam [3,4] and Ullmann [5] reactions, have been extensively explored for C-N bond formation. These reactions have a wide range of applications, such as in pesticide synthesis and pharmaceutical and material chemistry [6]. Among these reactions, the Cu-promoted Chan–Lam coupling reaction features the advantages of low cost and operational simplicity and is an efficient synthetic method to construct various carbon-heteroatom bonds. In 1998, this name reaction was first reported to construct C-X bonds through the coupling between arylboronic acids and different nucleophiles in the presence of Cu(OAc)2 as a catalyst (Scheme 1a) [7,8,9]. Since then, Cu and other transition metal-catalyzed Chan–Lam reactions have contributed greatly to C-N bond formation [10,11,12,13].
In addition to metal salts, well-defined copper complexes were also rapidly developed as catalysts for these reactions. Representative examples can be found in Scheme 1b [14,15,16,17,18,19,20,21]. In 2000, Collman and co-workers used 10 mol% of simple Cu complex [Cu(OH)·TMEDA]2Cl2 (TMEDA = N,N,N′,N′-Tetramethylethylenediamine) to produce a variety of N-arylimidazoles with good-to-excellent yields by cross-coupling arylboronic acids with imidazole compounds [14]. In 2018, Schaper et al. synthesized a sulfonato-imino copper(II) complex that can efficiently catalyze the N-arylation or N-alkylation of amines and anilines with arylboronic acids without the need for any bases [18]. Recently, Emerson’s group reported a new tetradentate NHC-copper(II) complex, which showed good catalytic activity in the Chan–Lam coupling of aniline and phenylboronic acid [20]. More recently, Jia et al. developed an efficient method for the copper-promoted Chan–Lam coupling of 1H-imidazole derivatives with arylboronic acids using an N, O-bidentate Cu complex (8.0 mol%) as a catalyst [21]. Although considerable progress has been made in this field, there are still limitations, such as long reaction times, high catalyst loading, or the use of a base. Thus, developing novel and efficient catalysts for Chan–Lam coupling under mild conditions is in high demand.
Scheme 1. Classical and selected Cu complexes for Chan–Lam reactions [15,16,17,19,20,21].
Scheme 1. Classical and selected Cu complexes for Chan–Lam reactions [15,16,17,19,20,21].
Molecules 28 07406 sch001
Moreover, it is well-accepted that ligands play an important role in catalysis. They can improve the solubility of transition-metal complexes in organic media and modify catalytic activity via the precise control of their electronic and geometric properties [22]. Recently, our group has been focusing on synthesizing anilido-imine ligand-supported transition-metal complexes. The corresponding Ni [23], Fe [24], Cr [25] and Cu [26] complexes exhibit moderate to high activities in olefin polymerization and the atom transfer radical polymerization (ATRP) reaction, but studies on organic transformation are less reported. Herein, we report the synthesis of several copper complexes bearing N,N,N-tridentate anilido-imine ligands, which displayed excellent catalytic performance in the Chan–Lam coupling of arylboronic acids with benzimidazole derivatives under base-free conditions (Scheme 1c).

2. Results and Discussion

2.1. Synthesis and Characterization of Ligands and Cu(II) Complexes

The tridentate anilido-imine ligands L1H, L2H, L4H, L5H and Cu(II) complex 1 were prepared according to the literature method [27] and our recent reports [26,28]. The ligand L3H and other Cu complexes 2-5 were synthesized following the analogous methods. Cu(II) complex 1 was prepared using the literature method [26]. The other Cu complexes (2-5) were synthesized following the analogous methods. The reactions of ligands L2H-L5H with n-BuLi at −78 °C followed by adding CuCl2 afforded complexes 2-5 as brown powders in 55–68% yields, respectively (Scheme 2). These complexes are air- and moisture-stable and soluble in common organic solvents (e.g., THF, CH2Cl2 and toluene). Complexes 2-5 were all characterized by HR-MS, EPR (electron paramagnetic resonance) and IR spectroscopy. The HR-MS peaks at 427.1104 (2), 415.0740 (3), 462.9740 (4) and 419.0245 (5) match well with the calculated masses of the cationic species [M-Cl]+ (Figures S1–S4). The EPR spectra of complexes 2-5 exhibit the values of g around 2.04–2.09 and g around 2.11–2.16 (Figure S5), which are comparable to the EPR data of some reported mononuclear Cu complexes [29] and our previous work [30]. In their IR spectra, a strong vibration band at 1602–1613 cm−1 can be assigned to the imine group. The signals between 3100 and 3500 cm−1 of the N-H group in the ligands disappeared in the IR spectra of complexes 2-5, indicating the Cu-N bond formation through the deprotonation of the –NH group.

2.2. Description of the Crystal Structures of Complexes 2 and 4

Crystals 2 and 4 suitable for single-crystal X-ray diffraction were grown from their solutions in CH2Cl2/hexane. X-ray single-crystal diffraction analyses show that crystals 2 and 4 are monoclinic with space groups of Cc and P21/c, respectively. The crystal data are summarized in Table 1. Their molecular structures are depicted in Figure 1 and Figure 2, with selected bond lengths and angles. The unit cell of 4 contains two crystallographically-independent molecules with similar connectivity and only one molecule is depicted in Figure 2. The central metal Cu is coordinated by three nitrogen atoms and one chlorine atom, and the environment around the metal can be described as a distorted square-planar. The Cu-Cl bond lengths (2.2248(12) Å for 2 and 2.2529(13) Å for 4) are relatively longer than those of 2.1997(13)–2.2001(13) Å in complex [CuCl2(Mes-BIAN)] OEt2 [31]. The Cu-Namino bond lengths of 2 and 4 are 1.927(3) Å and 1.956(4) Å, respectively, which are similar to the Cu-Namino bond length (1.939(3) Å) in copper complex 1 [26]. The Cu-Nimino bond lengths (1.973(3) Å for 2 and 1.980(4) Å for 4) are slightly longer than that of 1 (1.967(3) Å) but much shorter than the bond length of 2.026(1) Å in the sulfonato-imine copper complex [18].

2.3. Catalytic Activity

In order to investigate the catalytic properties of copper complexes 1-5 in the Chan–Lam coupling reaction, benzimidazole 6a and phenylboronic acid 7a were set as the model substrates. The reaction parameters, such as solvents, reaction time, temperature and catalyst loading, were optimized using complex 1 as a catalyst, and the catalytic results are shown in Table 2. At first, the solvent effect was investigated using 5 mol% of complex 1 under an air atmosphere and base-free conditions. The results showed that the best catalytic activity was achieved when methanol was used as the solvent to give the target product 8a with 90% yield at 50 °C for 20 h (Table 2, entry 1). Using other solvents, such as iPrOH, 1,4-Dioxane, DMF, CH3CN, THF, and toluene, under the same conditions gave only trace or low yields (Table 2, entries 2–7). Next, the reaction time and temperature were screened. When the reaction time was shortened from 20 h to 16 h and 12 h, the yields of 8a were 92% and 90%, respectively (Table 2, entries 8 and 9). Further reducing the reaction time to 8 h slightly decreased the yield to 83% (Table 2, entry 10). Since the yields were almost the same at 20 h and 12 h, 12 h was considered the optimized reaction time. Upon reducing the reaction temperature from 50 °C to 40 °C, the yield of 8a increased to 92% (Table 2, entry 11). When the reaction was performed either at 60 °C or room temperature, the yields of desired products dropped slightly (Table 2, entries 12 and 13). Thus, the best reaction temperature was selected to be 40 °C. Afterward, the effect of catalyst loading on the reaction was explored. Changing the catalyst loading of 1 from 5 mol% to 3 mol%, 2 mol%, and 1 mol% resulted in yields of 93%, 84%, and 71%, respectively (Table 2, entries 14–16). Therefore, the most suitable catalyst loading was decided to be 3 mol%. Subsequently, the catalytic activities of catalysts 2-5 were tested under the optimized conditions, and complex 3 showed the highest activity, furnishing product 8a with 96% yield (Table 2, entries 17–20). The CuCl2 was also tested in the Chan–Lam coupling of 6a and 7a to afford the corresponding product with only 48% yield, showing the important role of the ligand in the catalytic reaction (Table 2, entry 21). The control experiment indicated that the coupling reaction did not occur in the absence of any catalyst (Table 2, entry 22). Finally, when the coupling reaction was performed under an inert atmosphere, no product 8a was observed, indicating the key role of air in the catalytic reaction (Table 2, entry 23).
We next examined the substrate generality of this coupling reaction (Table 3). The phenylboronic acids with electron-deficient and electron-rich groups efficiently proceeded to afford the desired products (8a8p) in 75–96% yields. Various substituents, including methyl, methoxyl, fluoro, chloro, bromo, phenyl, aceto, ester and cyano, were tolerated, indicating good functional group compatibility. The steric effect has relatively little influence on this coupling reaction, which can be seen from 8b8d and 8g8h. However, heteroaryl boronic acids such as furan-2-ylboronic acid and thiophen-2-ylboronic acid were not suitable for the present catalytic system, and only trace amounts of the desired products—8q and 8r—were obtained. Gratifyingly, disubstituted 1,4-phenylenediboronic acid was also smoothly coupled with benzimidazole 6a to produce the corresponding 8s in 76% yield. Further, the reaction of 6 with different monosubstituted and disubstituted groups with phenylboronic acid gave the corresponding products in good-to-excellent yields (8t–8y). The reaction between dimethylbenzimidazole and F- or OMe-substituted phenylboronic acid also afforded the corresponding products, 8x and 8y, in 88% and 90% yields, respectively.
Taking previous reports [19,32,33,34] into account, a plausible mechanism for this Cu-catalyzed Chan–Lam reaction is depicted in Scheme 3. Initially, phenylboronic acid 7a undergoes a transmetallation reaction with CuII complex A to produce CuII species B and ClB(OH)2. Species B reacts with benzimidazole 6a to form a CuII intermediate C, which lowers the CuIII/CuII reduction potential [35]. Then, intermediate C undergoes a disproportionation process [18,36,37,38] in the presence of A to afford the high-valence CuIII intermediate D and CuI species E, and a molecule of HCl is released. Subsequently, intermediate D yields the product 1-phenylbenzimidazole 8a and generates another molecule of intermediate E through a reductive elimination step. Finally, intermediate E is converted to CuII complex A in the presence of air, HCl and ClB(OH)2 to finish the catalytic cycle.

3. Materials and Methods

3.1. General Considerations

All manipulations (except the catalytic reactions) under a nitrogen atmosphere were carried out using a Schlenk line. THF was distilled from Na and benzophenone under N2 before use. The other solvents for the catalytic reactions, CuCl2 and other reagents were obtained from commercial suppliers and used without further purification. IR spectra were recorded as KBr disks on a Thermo Fisher iS50 spectrometer (Thermo Fisher, Waltham, MA, USA). Mass spectroscopy was performed with an AB SCIEX 3200 Q-TRAP mass spectrometer (AB SCIEX, Framingham, MA, USA). The melting points were determined on an X-5 melting point apparatus (Beijing Tech Instrument Co., Ltd., Beijing, China). High-resolution mass spectra (HR-MS) were acquired using an Agilent 6210 ESI-TOF mass spectrometer (Agilent technology Co., Ltd, Santa Clara, CA, USA). 1H NMR and 13C{1H} NMR spectra were recorded on Zhongke-Niujin Quantum-I 400 MHz spectrometer (Zhongke-Niujin Co., Ltd., Wuhan, China). The EPR spectra were obtained at 77K in CH2Cl2 (0.01 M) solution using a Bruker-A200 Electron Spin Paramagnetic Resonance instrument (Bruker Corporation, Karlsruhe, Germany).

3.2. X-ray Crystallographic Studies

Diffraction data of 2 and 4 were collected on an Oxford Diffraction SuperNova dual source diffractometer with graphite-monochromated Cu-Kα radiation (λ = 1.54178 Å). The structures were solved by direct methods and refined by full-matrix least-squares on F2. All non-hydrogen atoms were refined anisotropically. The hydrogen atoms were introduced in calculated positions with the displacement factors of the host carbon atoms. Structure solution and refinement were performed using the SHELXL package [39].

3.3. Synthesis of Ligand L3H

A solution of n-BuLi in hexane (1.6 M, 8.3 mL, 13.2 mmol) was added to a THF (20 mL) solution of 8-aminoquinoline (1.73 g, 12.0 mmol) at −78 °C under nitrogen atmosphere. The reaction was allowed to warm up to room temperature and stirred overnight. The solution was transferred into a THF (20 mL) solution of ortho-C6H4F(C(H)=N-4-OCH3C6H4) (2.29 g, 10.0 mmol) at room temperature and then heated at 50 °C for 8 h. After cooling to room temperature, the reaction mixture was distributed in CH2Cl2 and H2O. The aqueous layer was extracted with CH2Cl2 (15 mL × 2). The combined organic phase was dried over anhydrous MgSO4 and filtered. The solvent was removed in vacuo to give the crude product, which was recrystallized from methanol-hexane to afford L3H as a yellow solid. Yield: 2.12 g (60%). 1H NMR (400 MHz, CDCl3) δ 12.78 (s, 1H), 8.96 (d, J = 2.8 Hz, 1H), 8.74 (s, 1H), 8.13 (d, J = 8.2 Hz, 1H), 7.89 (d, J = 8.0 Hz, 2H), 7.52 (t, J = 6.7 Hz, 1H), 7.51–7.42 (m, 4H), 7.37 (dd, J = 17.3, 8.2 Hz, 2H), 6.96 (dd, J = 14.8, 8.1 Hz, 3H), 3.84 (s, 3H) ppm. 13C{1H} NMR (101 MHz, CDCl3) δ 159.1, 158.3, 148.2, 143.7, 143.6, 140.6, 139.2, 135.9, 134.5, 131.1, 129.2, 126.8, 122.4, 121.8, 121.5, 118.7, 118.6, 115.1, 114.3, 111.8, 55.4 ppm. MS (ESI, m/z): 354.2 [M + H]+.

3.4. Synthesis of Complex 2

A solution of n-BuLi (1.6 mol/L) in hexane (0.7 mL, 1.1 mmol) was added dropwise to a THF (20 mL) solution of ligand L2H (0.37 g, 1.0 mmol) at −78 °C under nitrogen atmosphere. After stirring for 1 h, CuCl2 (0.13 g, 1.0 mmol) was added in one portion. The mixture was gradually warmed to room temperature and stirred overnight. Evaporation of the solvent in vacuo gave the crude product, which was extracted with 15 mL of CH2Cl2 and then filtered through Celite. The solution was evaporated under reduced pressure to a volume of ~3 mL. The resulting residue was recrystallized from CH2Cl2/hexane to give complex 2 as a brown powder. Yield: 0.30 g (65%). Single crystals for X-ray analysis were grown from CH2Cl2/hexane at room temperature. M.p. 229–230 °C. IR (KBr disk, cm−1): 2899 (w), 1613 (s), 1598 (s), 1568 (m), 1502 (m), 1456 (m), 1437 (m), 1377 (m), 1326 (m), 1188 (s), 1160 (s), 849 (m), 824 (m), 761 (m), 744 (m). HR-MS (ESI-TOF): calcd for C25H22ClCuN3, [M-Cl]+ 427.1104, found 427.1101.

3.5. Synthesis of Complex 3

Following the procedure as described for complex 1 using L3H as the ligand, complex 3 was obtained as a brown solid. Yield: 0.31 g (68%). M.p. 211–212 °C. IR (KBr disk, cm−1): 2911 (w), 1612 (s), 1597 (s), 1570 (m), 1507 (s), 1500 (s), 1470 (s), 1440 (s), 1379 (s), 1333 (m), 1245 (m), 1180 (m), 1164 (s), 828 (m), 803 (w), 763 (m), 741 (m). HR-MS (ESI-TOF): calcd for C23H18ClCuN3O, [M-Cl]+ 415.0740, found 415.0748.

3.6. Synthesis of Complex 4

Following the procedure as described for complex 1 using L4H as the ligand, complex 4 was obtained as a brown solid. Yield: 0.31 g (61%). M.p. 252–253 °C. IR (KBr disk, cm−1): 2924 (w), 1603 (s), 1571 (m), 1535 (m), 1501 (m), 1458 (s), 1436 (s), 1383 (m), 1224 (w), 1183 (m), 1157 (s), 831 (m), 818 (m), 760 (m), 746 (m). HR-MS (ESI-TOF): calcd for C22H15BrClCuN3, [M-Cl]+ 462.9740, found 462.9738.

3.7. Synthesis of Complex 5

Following the procedure as described for complex 1 using L5H as the ligand, complex 5 was obtained as a brown solid. Yield: 0.25 g (55%). M.p. 266–267 °C. IR (KBr disk, cm−1): 2923 (w), 1602 (m), 1570 (m), 1536 (s), 1500 (m), 1457 (m), 1436 (m), 1386 (m), 1222 (w), 1180 (m), 1159 (m), 830 (w), 822 (w), 750 (w), 739 (w). HR-MS (ESI-TOF): calcd for C22H15Cl2CuN3, [M-Cl]+ 419.0245, found 419.0242.

3.8. General Procedure for the Cu-Catalyzed Chan-Lam Coupling Reactions

A mixture of benzimidazoles 6 (0.20 mmol), arylboronic acid 7 (0.60 mmol), Cu catalyst 3 (3 mol%) and CH3OH (1.0 mL) were added in a 10 mL reaction tube. The reaction mixture was then heated at 40 °C in an oil bath for 12 h under the air atmosphere. After the reaction was completed, the reaction was cooled to room temperature and the volatiles were removed in vacuo. The resulting residue was purified by column chromatography on Al2O3 using CH2Cl2 and petroleum ether as eluent to give the product 8, which was further identified and confirmed by NMR (Figures S6–S54).

4. Conclusions

In summary, we have successfully synthesized and characterized a range of copper complexes containing amine-imine ligands. These Cu(II) complexes have proven to be effective catalysts in the Chan–Lam coupling reaction, facilitating the formation of C-N bonds without the need for a base. This catalytic system offers mild reaction conditions, ease of operation, and excellent compatibility with various functional groups. Copper amine-imine complexes of this nature hold the promise of broadening the scope of efficient C-N bond formation. Ongoing research in our laboratory is exploring further applications for these catalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28217406/s1, Figures S1–S4: HR-MS spectra of copper complexes 2-5; Table S1 and Figure S5: the EPR spectra and g values of copper complexes 2-5; Figures S6–S54: NMR spectra of catalytic products.

Author Contributions

Conceptualization, Z.H. (Zhiqiang Hao), J.L. and G.-L.L.; methodology, X.Z. and J.Y.; investigation, X.Z., J.Y. and Z.H. (Zhangang Han); resources, Z.H. (Zhangang Han); writing—original draft preparation, X.Z. and Z.H. (Zhiqiang Hao); writing—review and editing, G.-L.L.; supervision, Z.H. (Zhiqiang Hao) and J.L. funding acquisition, Z.H. (Zhiqiang Hao), J.L. and G.-L.L.; project administration, J.L. and G.-L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Hebei Natural Science Foundation of China (No. B2021205007), the Science Research Project of Hebei Education Department (No. ZD2022110), the Graduate Student Innovation Funding Project of Hebei Normal University (CXZZSS2022057) and Health Research Council of New Zealand (HRC-22-821).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to express their gratitude to Guang Zeng of the State Key Laboratory of Catalysis at the Dalian Institute of Chemical Physics, Chinese Academy of Sciences, for providing support in EPR, and to the reviewers for their valuable suggestions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hu, H.; Gilliam, A.M.; Qu, F.; Shaughnessy, K.H. Enolizable ketones as activators of palladium(II) precatalysts in amine arylation reactions. ACS Catal. 2020, 10, 4127–4135. [Google Scholar] [CrossRef]
  2. Lu, C.-J.; Xu, Q.; Feng, J.; Liu, R.-R. The asymmetric Buchwald-Hartwig amination reaction. Angew. Chem. Int. Ed. 2023, 62, e202216863. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, J.-Q.; Li, J.-H.; Dong, Z.-B. A review on the latest progress of Chan-Lam coupling reaction. Adv. Synth. Catal. 2020, 362, 3311–3331. [Google Scholar] [CrossRef]
  4. Jia, X.; Tong, X. Recent progress on Chan-Lam coupling reactions catalyzed by copper(II) complexes. Chin. J. Org. Chem. 2022, 42, 2640–2658. [Google Scholar] [CrossRef]
  5. Chen, Z.; Jiang, Y.; Zhang, L.; Guo, Y.; Ma, D. Oxalic diamides and tert-butoxide: Two types of ligands enabling practical access to alkyl aryl ethers via Cu-catalyzed coupling reaction. J. Am. Chem. Soc. 2019, 141, 3541–3549. [Google Scholar] [CrossRef]
  6. Roh, J.; Vavrová, K.; Hrábálek, A. Synthesis and functionalization of 5-substituted tetrazoles. Eur. J. Org. Chem. 2012, 2012, 6101–6118. [Google Scholar] [CrossRef]
  7. Chan, D.M.T.; Monaco, K.L.; Wang, R.-P.; Winters, M.P. New N- and O-arylations with phenylboronic acids and cupric acetate. Tetrahedron Lett. 1998, 39, 2933–2936. [Google Scholar] [CrossRef]
  8. Evans, D.A.; Katz, J.L.; West, T.R. Synthesis of diaryl ethers through the copper-promoted arylation of phenols with arylboronic acids. An expedient synthesis of thyroxine. Tetrahedron Lett. 1998, 39, 2937–2940. [Google Scholar] [CrossRef]
  9. Lam, P.Y.S.; Clarkt, C.G.; Saubernt, S.; Adamst, J.; Winters, M.P.; Chan, D.M.T.; Combs, A. New aryl/heteroaryl C-N bond cross-coupling reactions via arylboronic acid/cupric acetate arylation. Tetrahedron Lett. 1998, 39, 2941–2944. [Google Scholar] [CrossRef]
  10. Fu, T.; Qiao, H.; Peng, Z.; Hu, G.; Wu, X.; Gao, Y.; Zhao, Y. Palladium-catalyzed air-based oxidative coupling of arylboronic acids with H-phosphine oxides leading to aryl phosphine oxides. Org. Biomol. Chem. 2014, 12, 2895–2902. [Google Scholar] [CrossRef]
  11. Kumar, K.A.; Kannaboina, P.; Rao, D.N.; Das, P. Nickel-catalyzed Chan-Lam cross-coupling: Chemoselective N-arylation of 2-aminobenzimidazoles. Org. Biomol. Chem. 2016, 14, 8989–8997. [Google Scholar] [CrossRef] [PubMed]
  12. Kumari, K.; Kumar, S.; Singh, K.N.; Drew, M.G.B.; Singh, N. Synthesis and characterization of new square planar heteroleptic cationic complexes [Ni(II) β-oxodithioester-dppe]+; their use as a catalyst for Chan-Lam coupling. New J. Chem. 2020, 44, 12143–12153. [Google Scholar] [CrossRef]
  13. Shi, W.; Shi, Y.; Xü, M.; Zou, G.; Wu, X.-Y. Chemoselective Chan-Lam coupling by directly using copper powders via mechanochemical metal activation for catalysis. Mol. Catal. 2022, 528, 112472. [Google Scholar] [CrossRef]
  14. Collman, J.P.; Zhong, M. An efficient diamine copper complex-catalyzed coupling of arylboronic acids with imidazoles. Org. Lett. 2000, 2, 1233–1236. [Google Scholar] [CrossRef]
  15. Liu, B.; Liu, B.; Zhou, Y.; Chen, W. Copper(II) hydroxide complexes of N-heterocyclic carbenes and catalytic oxidative amination of arylboronic acids. Organometallics 2010, 29, 1457–1464. [Google Scholar] [CrossRef]
  16. Gogoi, A.; Sarmah, G.; Dewan, A.; Bora, U. Unique copper-salen complex: An efficient catalyst for N-arylations of anilines and imidazoles at room temperature. Tetrahedron Lett. 2014, 55, 31–35. [Google Scholar] [CrossRef]
  17. Duparc, V.H.; Schaper, F. Sulfonato-diketimine copper(II) complexes: Synthesis and application as catalysts in Chan−Evans−Lam couplings. Organometallics 2017, 36, 3053–3060. [Google Scholar] [CrossRef]
  18. Duparc, V.H.; Bano, G.L.; Schaper, F. Chan−Evans−Lam couplings with copper iminoarylsulfonate complexes: Scope and mechanism. ACS Catal. 2018, 8, 7308–7325. [Google Scholar] [CrossRef]
  19. Akatyev, N.; Il’in, M., Jr.; Peregudova, S.; Peregudov, A.; Buyanovskaya, A.; Kudryavtsev, K.; Dubovik, A.; Grinberg, V.; Orlov, V.; Pavlov, A.; et al. Chan-Evans-Lam C-N coupling promoted by a dinuclear positively charged Cu(II) complex. catalytic performance and some evidence for the mechanism of CEL reaction obviating Cu(III)/Cu(I) catalytic cycle. ChemCatChem 2020, 12, 3010–3021. [Google Scholar] [CrossRef]
  20. Cope, J.D.; Sheridan, P.E.; Galloway, C.J.; Awoyemi, R.F.; Stokes, S.L.; Emerson, J.P. Synthesis and characterization of a tetradentate, N-heterocyclic carbene copper(II) complex and its use as a Chan-Evans-Lam coupling catalyst. Organometallics 2020, 39, 4457–4464. [Google Scholar] [CrossRef]
  21. Jia, X.; He, J. Three copper(II) complexes derived from 2-methylquinoline and cyclic secondary amines: Synthesis and catalytic application in C-N bond forming reactions. Appl. Organomet. Chem. 2022, 36, e6743. [Google Scholar] [CrossRef]
  22. Zhang, L.; Cheng, Z.; Shi, S.; Li, Q.; Zhu, X. AGET ATRP of methyl methacrylate catalyzed by FeCl3/iminodiacetic acid in the presence of air. Polymer 2008, 49, 3054–3059. [Google Scholar] [CrossRef]
  23. Hao, Z.; Yang, N.; Gao, W.; Xin, L.; Luo, X.; Mu, Y. Nickel complexes bearing N,N,N-tridentate quinolinyl anilido-imine ligands: Synthesis, characterization and catalysis on norbornene addition polymerization. J. Organomet. Chem. 2014, 749, 350–355. [Google Scholar] [CrossRef]
  24. Hao, Z.; Han, Y.; Gao, W.; Xin, L.; Mu, Y. Iron(II) complexes bearing anilido-imine ligands: Synthesis and catalysis on ATRP of methyl methacrylate. Polyhedron 2014, 83, 236–241. [Google Scholar] [CrossRef]
  25. Hao, Z.; Xu, B.; Gao, W.; Han, Y.; Zeng, G.; Zhang, J.; Li, G.; Mu, Y. Chromium complexes with N,N,N-tridentate quinolinyl anilidoimine ligand: Synthesis, characterization, and catalysis in ethylene polymerization. Organometallics 2015, 34, 2783–2790. [Google Scholar] [CrossRef]
  26. Hao, Z.; Ma, A.; Xu, B.; Gao, W.; Mu, Y. Cu(II) complexes with anilido-imine ligands: Synthesis, characterization and catalysis on reverse atom transfer radical polymerization of styrene. Polyhedron 2017, 126, 276–281. [Google Scholar] [CrossRef]
  27. Gao, W.; Cui, D.; Liu, X.; Zhang, Y.; Mu, Y. Rare-earth metal bis(alkyl)s supported by a quinolinyl anilido-imine ligand: Synthesis and catalysis on living polymerization of ε-caprolactone. Organometallics 2008, 27, 5889–5893. [Google Scholar] [CrossRef]
  28. Zhou, X.; Xu, H.; Hao, Z.; Han, Z.; Lu, G.-L.; Lin, J. Nickel complexes bearing N,N,N-tridentate quinolinyl anilido-imine ligands: Synthesis, characterization and catalysis in Suzuki-Miyaura cross-coupling reaction. Polyhedron 2023, 246, 116676. [Google Scholar] [CrossRef]
  29. Kongkaew, M.; Sitthisuwannakul, K.; Nakarajouyphon, V.; Pornsuwan, S.; Kongsaeree, P.; Sangtrirutnugul, P. Benzimidazole–triazole ligands with pendent triazole functionality: Unexpected formation and effects on copper-catalyzed aerobic alcohol oxidation. Dalton Trans. 2016, 45, 16810–16819. [Google Scholar] [CrossRef]
  30. Gao, T.; Meng, L.; Zeng, G.; Hao, Z.; Han, Z.; Feng, Q.; Lin, J. Copper(II) complexes supported by 8-hydroxyquinoline-imine ligands: Synthesis, characterization and catalysis in aerobic alcohols oxidation. Polyhedron 2022, 224, 115984. [Google Scholar] [CrossRef]
  31. Fliedel, C.; Rosa, V.; Santos, C.I.M.; Gonzalez, P.J.; Almeida, R.M.; Gomes, C.S.B.; Gomes, P.T.; Lemos, M.A.N.D.A.; Aullón, G.; Weltere, R.; et al. Copper(II) complexes of bis(aryl-imino)-acenaphthene ligands: Synthesis, structure, DFT studies and evaluation in reverse ATRP of styrene. Dalton Trans. 2014, 43, 13041–13054. [Google Scholar] [CrossRef] [PubMed]
  32. Li, C.; Zhang, K.; Ma, H.; Wu, S.; Huang, Y.; Duan, Y.; Luo, Y.; Yan, J.; Yang, G. An optimized Ni-catalyzed Chan-Lam type coupling: Enantioretentive access to chiral N-aryl sulfinamides. Chem. Eur. J. 2022, 28, e202202190. [Google Scholar] [CrossRef] [PubMed]
  33. King, A.E.; Ryland, B.L.; Brunold, T.C.; Stahl, S.S. Kinetic and spectroscopic studies of aerobic copper(II)-catalyzed methoxylation of arylboronic esters and insights into aryl transmetalation to copper(II). Organometallics 2012, 31, 7948–7957. [Google Scholar] [CrossRef] [PubMed]
  34. Vantourout, J.C.; Li, L.; Bendito-Moll, E.; Chabbra, S.; Arrington, K.; Bode, B.E.; Isidro-Llobet, A.; Kowalski, J.A.; Nilson, M.G.; Wheelhouse, K.M.P.; et al. Mechanistic insight enables practical, scalable, room temperature Chan-Lam N-arylation of N-aryl sulfonamides. ACS Catal. 2018, 8, 9560–9566. [Google Scholar] [CrossRef]
  35. Phillips, A.J.; Uto, Y.; Wipf, P.; Reno, M.J.; Williams, D.R. Synthesis of functionalized oxazolines and oxazoles with DAST and Deoxo-Fluor. Org. Lett. 2000, 2, 1165–1168. [Google Scholar] [CrossRef]
  36. King, A.E.; Brunold, T.C.; Stahl, S.S. Mechanistic study of copper-catalyzed aerobic oxidative coupling of arylboronic esters and methanol: Insights into an organometallic oxidase reaction. J. Am. Chem. Soc. 2009, 131, 5044–5045. [Google Scholar] [CrossRef]
  37. Vantourout, J.C.; Miras, H.N.; Isidro-Llobet, A.; Sproules, S.; Watson, A.J.B. Spectroscopic studies of the Chan-Lam amination: A mechanism-inspired solution to boronic ester reactivity. J. Am. Chem. Soc. 2017, 139, 4769–4779. [Google Scholar] [CrossRef]
  38. West, M.J.; Fyfe, J.W.B.; Vantourout, J.C.; Watson, A.J.B. Mechanistic development and recent applications of the Chan-Lam amination. Chem. Rev. 2019, 119, 12491–12523. [Google Scholar] [CrossRef]
  39. Sheldrick, G.M. A short history of SHELX. Acta Cryst. 2008, A64, 112–122. [Google Scholar] [CrossRef]
Scheme 2. Synthesis of complexes 1-5.
Scheme 2. Synthesis of complexes 1-5.
Molecules 28 07406 sch002
Figure 1. Molecular structure of Cu complex 2 (CCDC: 2248465) with thermal ellipsoids drawn at a 50% probability level. Hydrogen atoms are omitted for clarity. Selected bond lengths (Å): Cu(1)-N(1) 1.927(3), Cu(1)-N(2) 1.990(3), Cu(1)-N(3) 1.973(3), Cu(1)-Cl(1) 2.2248(12), N(1)-C(8) 1.389(5), N(2)-C(13) 1.367(5), N(3)-C(7) 1.288(5), N(3)-C(17) 1.442(5); Selected bond angles (◦):C(1)-N(1)-C(8) 122.1(3), N(1)-Cu(1)-Cl(1) 159.97(11), N(1)-C(1)-C(6) 119.2(4), N(3)-Cu(1)-N(2) 158.12(11), C(8)-N(1)-Cu(1) 114.2(3).
Figure 1. Molecular structure of Cu complex 2 (CCDC: 2248465) with thermal ellipsoids drawn at a 50% probability level. Hydrogen atoms are omitted for clarity. Selected bond lengths (Å): Cu(1)-N(1) 1.927(3), Cu(1)-N(2) 1.990(3), Cu(1)-N(3) 1.973(3), Cu(1)-Cl(1) 2.2248(12), N(1)-C(8) 1.389(5), N(2)-C(13) 1.367(5), N(3)-C(7) 1.288(5), N(3)-C(17) 1.442(5); Selected bond angles (◦):C(1)-N(1)-C(8) 122.1(3), N(1)-Cu(1)-Cl(1) 159.97(11), N(1)-C(1)-C(6) 119.2(4), N(3)-Cu(1)-N(2) 158.12(11), C(8)-N(1)-Cu(1) 114.2(3).
Molecules 28 07406 g001
Figure 2. Molecular structure of Cu complex 4 (CCDC: 2248467) with thermal ellipsoids drawn at a 50% probability level. Hydrogen atoms are omitted for clarity. Selected bond lengths (Å) Cu(1)-N(1) 1.956(4), Cu(1)-N(2) 1.985(4), Cu(1)-N(3) 1.980(4), Cu(1)-Cl(1) 2.2529(13), N(1)-C(8) 1.393(5); N(2)-C(13) 1.355(6); N(3)-C(7) 1.295(6), N(3)-C(17) 1.427(6); Selected bond angles (◦):C(1)-N(1)-C(8) 122.6(4), N(1)-Cu(1)-Cl(1) 161.71(12), N(1)-Cu(1)-N(3) 90.90(15), N(1)-C(1)-C(6) 120.0(4), N(3)-Cu(1)-Cl(1) 94.30(11).
Figure 2. Molecular structure of Cu complex 4 (CCDC: 2248467) with thermal ellipsoids drawn at a 50% probability level. Hydrogen atoms are omitted for clarity. Selected bond lengths (Å) Cu(1)-N(1) 1.956(4), Cu(1)-N(2) 1.985(4), Cu(1)-N(3) 1.980(4), Cu(1)-Cl(1) 2.2529(13), N(1)-C(8) 1.393(5); N(2)-C(13) 1.355(6); N(3)-C(7) 1.295(6), N(3)-C(17) 1.427(6); Selected bond angles (◦):C(1)-N(1)-C(8) 122.6(4), N(1)-Cu(1)-Cl(1) 161.71(12), N(1)-Cu(1)-N(3) 90.90(15), N(1)-C(1)-C(6) 120.0(4), N(3)-Cu(1)-Cl(1) 94.30(11).
Molecules 28 07406 g002
Scheme 3. Proposed catalytic mechanism.
Scheme 3. Proposed catalytic mechanism.
Molecules 28 07406 sch003
Table 1. Summary of the crystal data for complexes 2 and 4.
Table 1. Summary of the crystal data for complexes 2 and 4.
Complex24
FormulaC25H22ClCuN3C22H15BrClCuN3
Fw463.44500.27
T (K)199.99 (10)199.99 (10)
Crystal systemmonoclinicmonoclinic
Space groupCcP21/c
a (Å)17.2718 (8)15.5617 (4)
b (Å)15.2065 (6)9.4245 (2)
c (Å)8.1558 (3)27.3576 (7)
α (°)9090
β (°)98.532 (4)96.439 (2)
γ (°)9090
Volume (Å3)2118.36 (15)3986.99 (17)
Z48
Dcalc (g/cm3)1.4531.667
μ (mm−1)2.7365.220
F (000)956.01992.0
Crystal size (mm)0.14 × 0.12 × 0.10.14 × 0.12 × 0.1
2Θ range for data collection (°)7.784 to 146.8965.716 to 147.15
Reflections collected424214,755
Data/restraints/parameters2723/2/2747793/0/505
Goodness-of-fit on F21.0741.008
R1, wR2 [I > 2σ(I)]0.0327, 0.08880.0533, 0.1325
R1, wR2 (all data)0.0339, 0.08990.0758, 0.1496
Max. peak (e·Å−3)0.490.58
Mini. peak (e·Å−3)−0.71−0.71
CCDC22484652248467
Table 2. Optimization of Chan–Lam coupling reaction a.
Table 2. Optimization of Chan–Lam coupling reaction a.
Molecules 28 07406 i001
EntryCat. (mol%)SolventT (°C)Time (h)Yield (%) b
11 (5)MeOH502090
21 (5)iPrOH50206
31 (5)1,4-Dioxane502040
41 (5)DMF5020trace
51 (5)CH3CN5020trace
61 (5)THF502046
71 (5)Toluene502025
81 (5)MeOH501692
91 (5)MeOH501290
101 (5)MeOH50883
111 (5)MeOH401292
121 (5)MeOHrt1281
131 (5)MeOH601288
141 (3)MeOH401293
151 (2)MeOH401284
161 (1)MeOH401271
172 (3)MeOH401291
183 (3)MeOH401296
194 (3)MeOH401294
205 (3)MeOH401290
21CuCl2 (3)MeOH401248
22-MeOH4012-
23 c3 (3)MeOH4012-
a Reaction conditions: 6a (0.20 mmol), 7a (0.60 mmol), Cat. (3 mol%), MeOH (1.0 mL), 40 °C, 12 h, under air. b Isolated yield. c Under an N2 atmosphere.
Table 3. Substrate scope of Chan–Lam coupling reaction catalyzed by 3 a.
Table 3. Substrate scope of Chan–Lam coupling reaction catalyzed by 3 a.
Molecules 28 07406 i002
Molecules 28 07406 i003Molecules 28 07406 i004Molecules 28 07406 i005Molecules 28 07406 i006
8a, 96%8b, 96%8c, 95%8d, 91%
Molecules 28 07406 i007Molecules 28 07406 i008Molecules 28 07406 i009Molecules 28 07406 i010
8e, 89%8f, 94%8g, 93%8h, 85%
Molecules 28 07406 i011Molecules 28 07406 i012Molecules 28 07406 i013Molecules 28 07406 i014
8i, 96%8j, 88%8k, 91%8l, 83%
Molecules 28 07406 i015Molecules 28 07406 i016Molecules 28 07406 i017Molecules 28 07406 i018
8m, 75%8n, 77%8o, 84%8p, 87%
Molecules 28 07406 i019Molecules 28 07406 i020Molecules 28 07406 i021
8q, trace8r, trace8s b, 76%
Molecules 28 07406 i022Molecules 28 07406 i023Molecules 28 07406 i024Molecules 28 07406 i025
8t/8t’ (5/6-Me), 95%8u/8u’ (5/6-F), 79%8v, 93%8w, 58%
Molecules 28 07406 i026Molecules 28 07406 i027
8x, 88%8y, 90%
a Reaction conditions: 6 (0.20 mmol), 7 (0.60 mmol), 3 (3 mol%), MeOH (1.0 mL), 40 °C, 12 h, under air, isolated yield. b benzimidazole 6a (0.20 mmol), 1,4-phenylenediboronic acid 7s (0.20 mmol).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhou, X.; Yang, J.; Hao, Z.; Han, Z.; Lin, J.; Lu, G.-L. Copper Complexes with N,N,N-Tridentate Quinolinyl Anilido-Imine Ligands: Synthesis and Their Catalytic Application in Chan−Lam Reactions. Molecules 2023, 28, 7406. https://doi.org/10.3390/molecules28217406

AMA Style

Zhou X, Yang J, Hao Z, Han Z, Lin J, Lu G-L. Copper Complexes with N,N,N-Tridentate Quinolinyl Anilido-Imine Ligands: Synthesis and Their Catalytic Application in Chan−Lam Reactions. Molecules. 2023; 28(21):7406. https://doi.org/10.3390/molecules28217406

Chicago/Turabian Style

Zhou, Xiaoyu, Jiaxin Yang, Zhiqiang Hao, Zhangang Han, Jin Lin, and Guo-Liang Lu. 2023. "Copper Complexes with N,N,N-Tridentate Quinolinyl Anilido-Imine Ligands: Synthesis and Their Catalytic Application in Chan−Lam Reactions" Molecules 28, no. 21: 7406. https://doi.org/10.3390/molecules28217406

Article Metrics

Back to TopTop