Next Article in Journal
Valorization of Pimenta racemosa Essential Oils and Extracts: GC-MS and LC-MS Phytochemical Profiling and Evaluation of Helicobacter pylori Inhibitory Activity
Previous Article in Journal
1-Tetradecanol, Diethyl Phthalate and Tween 80 Assist in the Formation of Thermo-Responsive Azoxystrobin Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of Chitosan-Composite-Film-Supported Copper Nanoparticles and Their Application in 1,6-Hydroboration Reactions of p-Quinone Methides

1
School of Chemistry and Materials Science, Hubei Engineering University, Xiaogan 432000, China
2
School of Materials Science and Engineering, Hubei University, Wuhan 430062, China
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(22), 7962; https://doi.org/10.3390/molecules27227962
Submission received: 7 October 2022 / Revised: 10 November 2022 / Accepted: 15 November 2022 / Published: 17 November 2022
(This article belongs to the Section Applied Chemistry)

Abstract

:
Here, we describe the preparation of copper nanoparticles that are stabilized on a chitosan composite film (CP@Cu). This material could catalyze the 1,6-hydroboration reactions of p-quinone methides with B2pin2 as a boron source under mild conditions. This reaction exhibited very good functional group compatibility, and the organoboron compounds that were formed could easily be converted into corresponding hydroxyl products with good to excellent yields. This newly developed methodology provides an efficient and sequential pathway for the synthesis of gem-disubstituted methanols.

Graphical Abstract

1. Introduction

Not only do organoboron compounds exist in a wide range of active molecules, natural products, and materials [1,2,3], but they are also the key intermediates in the synthesis of many functional chemicals [4,5]. For these reasons, in recent years, a series of methodologies—especially transition metal catalysis—have been developed for the synthesis of organoboron compounds [6,7,8], which have become more and more important [9,10,11,12]. Copper catalysts are increasingly favored by organic chemists due to their low cost, lower toxicity, and solid performance [13,14]. In previous work, copper-catalyzed hydroboration reactions of unsaturated compounds were widely studied, and this is also a common method for constructing C-B bonds [15,16,17,18]. However, examples reported here involved the use of strong bases and specifically designed ligands, which reduced the reaction economy. Thus, it is necessary to explore alternative highly active and sustainable copper-catalyzed hydroboration reactions of unsaturated compounds.
Metallic nanoparticles have been widely used in various reactions with the continuous development of organic synthetic chemistry over recent decades [19,20,21,22]. However, to the best of our knowledge, though they are some of the most widely used metallic nanoparticles, copper nanoparticles are rarely used to catalyze the hydroboration of unsaturated compounds [23,24]. In particular, with p-quinone methides, as a class of intermediates with a wide range of applications in organic synthesis [25,26,27,28], the 1,6-hydroboration products can be well transformed into gem-disubstituted methanols under certain conditions; they are widely spread throughout nature and are the core skeletons of many biologically active molecules and natural products [29,30,31,32,33]. So far, there have only been a few reports in the literature on the copper-catalyzed 1,6-boron addition reaction of p-quinone methides, and these reports were mainly focused on Cu(I)-catalyzed reactions [34,35]. In our previous work, we used Cu(OH)2 as a catalyst to investigative the 1,6-hydroboration reaction of p-quinone methides, and good functional group compatibilities and reaction yields were obtained (Scheme 1a) [36]. Based on the results of our previous research, we found that when chitosan-supported copper nanoparticles were used as a heterogeneous catalyst, construct C-B bonds [37] and C-Si bonds could be constructed with a high efficiency [38]. Therefore, in this work, we hope to use chitosan-supported copper nanoparticles as a heterogeneous catalyst and p-quinone methides as substrates to study the 1,6-hydroboration reactions. In comparison with previous work, the biggest advantage of this work is the avoidance of the participation of bases in the reactions and the recycling of the catalyst to increase its utilization rate (Scheme 1b).

2. Results and Discussion

The initial experiments commenced with p-quinone methide 1a as a model substrate. CP@Cu (6 mol%) was used as a catalyst by using B2Pin2 (1.2 equiv.) as the boron source in the reactions. Firstly, the various organic solvents were investigated, and considering the role of protons, the whole reaction was started with ethanol (2.5 mL) as the solvent. However, no reaction happened (Table 1, Entry 1). When DCM and THF were used as solvents and MeOH (2 equiv.) was used as an additive, reactions were still not observed (Table 1, Entries 2–3). To our delight, when MeCN was used as the solvent and MeOH (2 equiv.) was used as an additive, the reaction was able to take place, and the desired product 2a was smoothly obtained. It was confirmed that further oxidation of 2a gave the corresponding gem-disubstituted methanol product 3a by using NaBO3·4H2O as an oxidant. (Table 1, Entry 4). Continuing to use acetone as the solvent and MeOH (2 equiv.) as an additive further promoted the occurrence of the reaction, and the target product 3a could be obtained with 70% yield (Table 1, Entry 5). Since water was a green solvent in this reaction, we added water (2 equiv.) to the reaction as an additive; unfortunately, the reaction did not happen (Table 1, Entry 6). As far as we know, in organic synthesis reactions, the use of mixed solvents could sometimes greatly improve the efficiency of the whole reaction. Therefore, in order to further improve the yield, we considered using acetone and H2O as mixed solvents to carry out the reaction, and the ratios of the solvents were screened (Table 1, Entries 7–10). When the ratio of acetone to H2O was 4:1, the reaction had the highest rate of conversion, and it occurred almost completely; the final target product could be obtained with 98% yield (Table 1, Entry 7). When the weight of H2O in the mixed solvents was continually increased, it was found that the conversion rate of the reaction decreased as the proportion of water increased; even when the ratio of acetone to H2O was reversed to 1:4, the reaction hardly occurred, and only trace amounts of product could be detected (Table 1, Entry 10). In order to verify the importance of the CP@Cu in the reactions, we performed a control experiment, and no reaction occurred without any CP@Cu, which proved that the catalyst was indispensable in these reactions (Table 1, Entry 11). Finally, the reaction time was also investigated. Even if the reaction time was shortened to 1 h, the reaction still occurred efficiently and produced the target product with 93% yield (Table 1, Entry 12). Thus, through a series of optimizations of the conditions, the optimal conditions in this research were found to be 6 mol% of CP@Cu as a catalyst and 1.2 equiv. of B2Pin2 as a boron source, and the whole reaction was conducted in 2.5 mL of mixed solvents (acetone:H2O = 4:1) at room temperature for 2 h (Table 1, Entry 7).
With the optimal conditions in hand, we continued to examine the universality of the reaction, and the results are summarized in Figure 1. Firstly, the effects of substituents at the ortho-position of the benzene ring on the reaction were investigated. For electron-donating substituents, such as methyl and methoxy, the desired target product could be obtained with excellent reaction yields (3b3c, 96–98% yields). The whole reaction could still proceed smoothly and achieved the corresponding products with satisfactory yields when the more conjugated 2-substituted naphthyl was selected as the substituent instead of phenyl (3d, 92% yield). Although the electron-withdrawing substituents at the ortho-position had a certain effect on the reaction, the desired product was still obtained with a good yield (3e, 74% yield).
Next, we investigated the reactivity of the substituents at the meta-position of the benzene ring. From the reaction results summarized in Figure 1, the electron-donating substituents had a good effect on the reaction, and the desired products could be obtained with an almost equivalent yield (3f3g, 96–97% yields). However, when an electron-withdrawing substituent was used, such as fluorine, the reaction yield was reduced to some extent (3h, 77% yield). To our delight, when the benzene ring had multiple substituents, such as naphthyl, dimethoxy, or even trimethoxy, the reaction could still occur well, and good to excellent reaction yields could be obtained (3i3k, 80–96% yields). We also investigated the reactivity of para-substituents on the benzene ring; both electron-donating substituents (methyl, isopropyl, tert-butyl, methoxy, and benzyloxy) and electron-withdrawing substituents (fluorine, chlorine, bromine) had little effect on the reaction (3l3s, 91–96% yields). Finally, we investigated thiophene, and although the target product could be obtained only with a moderate yield, the reaction still proved that the catalyst had good functional group compatibility (3t, 58% yield).
Considering that this could be a heterogeneous catalyst in this reaction, it is necessary to identify the reusability and stability of catalyst. It was demonstrated that when the reaction was completed, the CP@Cu catalyst could be easily recycled with a simple operation. The catalytic activity stayed almost the same after experimenting with recycling the catalyst six times, and the yield was still up to 96% even in the sixth experiment, so the catalyst has the advantage of being recyclable (Figure 2).
The Cu nanoparticles supported on a chitosan–PVA composite are shown in Figure 3. As observed, the dark spherical particles in the red circles are the CuNPs, which are uniformly dispersed in the CP matrix. The particle sizes ranged from 2 to 4 nm, showing that the Cu nanoparticles were uniformly distributed on the chitosan–PVA composite. In addition, no aggregation of CuNPs was noticed, which confirmed that the CP matrix is a good stabilizing agent for the synthesis of CuNPs. The good dispersion of CuNPs into the CP matrix enhanced their performance during the catalytic process [39].
The full-scan XPS spectrum showed that the major elements of the Cu nanoparticles supported on the chitosan–PVA composite were O, C, and N (Figure 4a). This is consistent with the chemical structures of chitosan and PVA, which are rich in the functional groups of –NH–C=O, –NH2, and –OH. The presence of a Cu 2p peak in the Cu-loaded PVA–CS nanofiber membrane proved the adsorption of Cu(II) onto the adsorbent. The spectrum of the adsorbent after copper adsorption showed peaks at 932.67, 933.85, and 934.74 eV, which corresponded to Cu 2p (Figure 4b). The C 1s spectra of the Cu nanoparticles supported on the chitosan–PVA composite showed peaks at 284.48, 285.88, and 287.81 eV after the adsorption of Cu(II), indicating the involvement of the functional groups in the adsorption of Cu(II) onto the adsorbent (Figure 4c) [40,41].

3. Materials and Methods

3.1. Materials

Chitosan (degree of deacetylation ≥95%, viscosity 100–200 MPa·s) was purchased directly from Aladdin (Shanghai, China), poly (vinyl alcohol) (Mw 120 kDa) was purchased from Sinopharm Chemical Reagent Co., Ltd. (Beijing, China), B2pin2 was purchased from Energy Chemical (Shanghai, China), and CuCl2·2H2O was purchased from Sinopharm Chemical Reagent Co., Ltd. (Beijing, China). All p-quinone methides were obtained smoothly through reactions between aldehydes and variously substituted 2,6-di-tert-butylphenol. Chitosan/poly (vinyl alcohol)-composite-film-supported copper nanoparticles (CP@Cu NPs) were prepared according to the method reported in the literature [39].

3.2. Analytical Methods

Nuclear magnetic resonance (NMR) spectra were recorded on a Bruker Avance III 400 MHz spectrometer (Karlsruhe, Germany) operating at 400 for 1H and 100 MHz for 13C NMR in CDCl3 unless otherwise noted. CDCl3 served as the internal standard (δ = 7.26 ppm) for 1H NMR and (δ = 77.0 ppm) for 13C NMR. The data for 1H NMR are reported as follows (See the Supplementary Materials): chemical shift (ppm, scale), multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet and/or multiplet resonances, br = broad), coupling constant (Hz), and integration. The data for 13C NMR are reported in terms of the chemical shift (ppm, scale), multiplicity, and coupling constant (Hz). The purification of products was accomplished by using flash column chromatography on silica gel (200–300 mesh) or preparative TLC. The weight percentage and metal leaching of copper were determined by inductively coupled plasma–optical emission spectroscopy (ICP-OES) (PerkinElmer, Waltham, MA, USA) analysis.
An X-ray photoelectron spectroscopy (XPS) analysis was performed with a Thermo Fisher ESCALAB250Xi spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) by using monochromatized Al-Ka radiation at a detection angle of 30°. The photon energy was 1486.6 eV. A pass energy of 30 eV was used for high-resolution scans in a valence band analysis. The test area size was 500 um. The binding energy of all spectra was determined by using binding energy correction with respect to polluting carbon (C 1s, 284.6 or 284.8eV). The spectra were collected over a range of 0–1486.6 eV, and the high-resolution spectra of C 1s and Cu 2p regions were provided. The Shirley background and Gaussian/Lorentzian functions were used to fit the peaks.
Transmission electron microscopy (TEM) was used to observe the morphology on a Jeol 2100f instrument (JEOL, Tokyo, Japan). Samples were prepared for TEM analysis by placing a drop of a sample of a particle suspension on a copper grid and quickly wicking away the solution with filter paper.

3.3. General Procedure for the Preparation of CP@Cu NPs

According to a report in the literature [39], 200 mg of chitosan powder was dissolved in 10 mL of acetic acid solution (2%, v/v) and stirred at room temperature for 5 h. At the same time, 400 mg of poly (vinyl alcohol) was dissolved in 10 mL of water and stirred at 80 °C for 12 h. The two solutions obtained were mixed and stirred at room temperature for another 0.5 h; then, 32 μL of glutaraldehyde solution (25%, w/w) was added, and stirring was continued for 5 min. In order to form the chitosan/poly (vinyl alcohol) composite film, the mixed solution described above was transferred to a Petri dish and dried at 40 °C for 12 h. After completion of this procedure, 0.1 mol/L of NaOH solution was added to the above composite film and allowed to soak for 5 min; then, this was washed until it was neutral by using water and dried over 12 h at 40 °C. After immersing the composite film in 0.2 mol/L CuCl2 solution for 2.5 h, the excess Cu2+ and Cl were removed by washing with water, and then drying took place at 40 °C for 12 h. Finally, the chitosan/poly (vinyl alcohol)-composite-film-supported copper nanoparticles (CP@Cu NPs) were obtained by reducing with 0.05 mol/L of NaBH4 solution, and they were then submitted for ICP analysis. The copper loading of the CP@Cu NPs was found to be 1.78 mmol/g.

3.4. Recycling and Reuse of CP@Cu NPs

To demonstrate the recyclability of the CP@Cu NPs, the addition of a boron conjugate was repeated six times with the same composite film. The initial amount of catalyst was 5 mg (6 mol % Cu loading). Reactions were carried out under standard conditions. After the completion of the reaction, the catalyst was filtered off, washed with acetone, and then dried at 50 °C before the next run.

4. Conclusions

In conclusion, we have reported the preparation of a copper nanoparticle stabilized on chitosan composite film (CP@Cu) and its application for catalyzing the 1,6-hydroboration reaction of p-quinone methides with B2pin2 as a boron source. The conditions of the whole reaction were very mild, and no additional bases were needed. This newly developed methodology showed very good functional group compatibility and reactivity (20 examples, up to 98% yield). The organoboron products that were formed could be easily and directly oxidized to the corresponding hydroxyl products with good to excellent yields. In addition, the recycling experiments evidenced that this catalyst still showed good reactivity after being recycled six times (>96% yield), which proved that the catalyst had good reusability and stability.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27227962/s1, characterization data, and spectra for the 1H and 13C NMR of products 3a3t. Citation: [34,36].

Author Contributions

L.Z. conceived and designed the experiments; S.C., W.W. and B.L. performed the experiments. W.L. helped with the characterization of some new compounds; X.Z., Z.Z. and Y.Z. contributed reagents/materials/analysis tools; L.Z. reviewed and modified the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was founded by the National Natural Science Foundation of China (Nos. 21774029, 22108065), the Hubei University Excellent Young and Middle-aged Science and Technology Innovation Team Project (No. T201816), the Opening Fund of Hubei Key Laboratory of Processing and Application of Catalytic materials, and Huanggang Normal University (No. 202023404).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors acknowledge the financial support from the National Natural Science Foundation of China (Nos. 21774029, 22108065), the Hubei University Excellent Young and Middle-aged Science and Technology Innovation Team Project (No. T201816), the Opening Fund of Hubei Key Laboratory of Processing and Application of Catalytic materials, and Huanggang Normal University (No. 202023404).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Coghi, P.S.; Zhu, Y.; Xie, H.; Hosmane, N.S.; Zhang, Y. Organoboron compounds: Effective antibacterial and antiparasitic agents. Molecules 2021, 26, 3309. [Google Scholar] [CrossRef] [PubMed]
  2. Roccaro, A.M.; Vacca, A.; Ribatti, D. Bortezomib in the treatment of cancer. Recent Pat. Anti-cancer Drug Discov. 2006, 1, 397. [Google Scholar] [CrossRef] [PubMed]
  3. Miao, J.; Wang, Y.; Liu, J.; Wang, L. Organoboron molecules and polymers for organic solar cell applications. Chem. Soc. Rev. 2022, 51, 153. [Google Scholar] [CrossRef]
  4. Sacramento, M.; Costa, G.P.; Barcellos, A.M.; Perin, G.; Lenardão, E.J.; Alves, D. Transition-metal-free C-S, C-Se, and C-Te bond formation from organoboron compounds. Chem. Rec. 2021, 21, 2855. [Google Scholar] [CrossRef]
  5. McClary, C.A.; Taylor, M.S. Applications of organoboron compounds in carbohydrate chemistry and glycobiology: Analysis, separation, protection, and activation. Carbohydr. Res. 2013, 381, 112. [Google Scholar] [CrossRef]
  6. Yang, T.; Tang, N.; Wan, Q.; Yin, S.-F.; Qiu, R. Recent progress on synthesis of N,N′-chelate organoboron derivatives. Molecules 2021, 26, 1401. [Google Scholar] [CrossRef]
  7. Tian, Y.-M.; Guo, X.-N.; Braunschweig, H.; Radius, U.; Marder, T.B. Photoinduced borylation for the synthesis of organoboron compounds. Chem. Rev. 2021, 121, 3561. [Google Scholar] [CrossRef] [PubMed]
  8. Wen, Y.; Deng, C.; Xie, J.; Kang, X. Recent synthesis developments of organoboron compounds via metal-free catalytic borylation of alkynes and alkenes. Molecules 2018, 24, 101. [Google Scholar] [CrossRef] [Green Version]
  9. Whyte, A.; Torelli, A.; Mirabi, B.; Zhang, A.; Lautens, M. Copper-catalyzed borylative difunctionalization of π-systems. ACS Catal. 2020, 10, 11578. [Google Scholar] [CrossRef]
  10. Hu, J.; Ferger, M.; Shi, Z.; Marder, T.B. Recent advances in asymmetric borylation by transition metal catalysis. Chem. Soc. Rev. 2021, 50, 13129. [Google Scholar] [CrossRef]
  11. Ishiyama, T.; Miyaura, N. Metal-catalyzed reactions of diborons for synthesis of organoboron compounds. Chem. Rec. 2004, 3, 271. [Google Scholar] [CrossRef]
  12. Bisht, R.; Haldar, C.; Hassan, M.M.M.; Hoque, M.E.; Chaturvedi, J.; Chattopadhyay, B. Metal-catalysed C-H bond activation and borylation. Chem. Soc. Rev. 2022, 51, 5042. [Google Scholar] [CrossRef] [PubMed]
  13. Jin, S.; Li, J.; Liu, K.; Ding, W.Y.; Wang, S.; Huang, X.; Li, X.; Yu, P.; Song, Q. Enantioselective Cu-catalyzed double hydroboration of alkynes to access chiral gem-diborylalkanes. Nat. Commun. 2022, 13, 3524. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, I.-H.; Kanai, M.; Shibasaki, M. Copper(I)-secondary diamine complex-catalyzed enantioselective conjugate boration of linear β,β-disubstituted enones. Org. Lett. 2010, 12, 4098. [Google Scholar] [CrossRef] [PubMed]
  15. Xi, Y.; Hartwig, J.F. Mechanistic studies of copper catalyzed asymmetric hydroboration of alkenes. J. Am. Chem. Soc. 2017, 139, 12758. [Google Scholar] [CrossRef]
  16. Hoang, G.L.; Takacs, J.M. Enantioselective γ-borylation of unsaturated amides and stereoretentive Suzuki–Miyaura cross-coupling. Chem. Sci. 2017, 8, 4511. [Google Scholar] [CrossRef] [Green Version]
  17. Jang, W.J.; Song, S.M.; Moon, J.H.; Lee, J.Y.; Yun, J. Copper-catalyzed enantioselective hydroboration of unactivated 1,1-disubstituted alkenes. J. Am. Chem. Soc. 2017, 139, 13660. [Google Scholar] [CrossRef]
  18. Lee, H.; Lee, B.Y.; Yun, J. Copper(I)-taniaphos catalyzed enantiodivergent hydroboration of bicyclic alkenes. Org. Lett. 2015, 17, 764. [Google Scholar] [CrossRef]
  19. Narayanan, R. Recent advances in noble metal nanocatalysts for Suzuki and Heck cross-coupling reactions. Molecules 2010, 15, 2124. [Google Scholar] [CrossRef]
  20. Hosseini-Sarvari, M.; Razmi, Z. Palladium supported on zinc oxide nanoparticles as efficient heterogeneous catalyst for Suzuki-Miyaura and Hiyama reactions under normal laboratory conditions. Helv. Chim. Acta. 2015, 98, 805. [Google Scholar] [CrossRef]
  21. Li, W.; Yan, F.; Cai, S.; Ding, L.; Li, B.; Zhang, B.; Zhang, Y.; Zhu, L. Platinum nanoparticles as recyclable heterogeneous catalyst for selective methylation of amines and imines with formic acid: Indirect utilization of CO2. J. Saudi. Chem. Soc. 2022, 26, 101421. [Google Scholar] [CrossRef]
  22. Rai, R.K.; Gupta, K.; Tyagi, D.; Mahata, A.; Behrens, S.; Yang, X.; Xu, Q.; Pathak, B.; Singh, S.K. Access to highly active Ni-Pd bimetallic nanoparticle catalysts for C-C coupling reactions. Catal. Sci. Technol. 2016, 6, 5567. [Google Scholar] [CrossRef]
  23. Wen, W.; Han, B.; Yan, F.; Ding, L.; Li, B.; Wang, L.; Zhu, L. Borylation of α,β-unsaturated acceptors by chitosan composite film supported copper nanoparticles. Nanomaterials 2018, 8, 326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Zhou, X.-F.; Sun, Y.-Y.; Wu, Y.-D.; Dai, J.-J.; Xu, J.; Huang, Y.; Xu, H.-J. Borylation and selective reduction of α,β-unsaturated ketones under mild conditions catalyzed by Cu nanoparticles. Tetrahedron 2016, 72, 5691. [Google Scholar] [CrossRef]
  25. Singh, G.; Pandey, R.; Pankhade, Y.A.; Fatma, S.; Anand, R.V. Construction of oxygen- and nitrogen-based heterocycles from p-quinone methides. Chem. Rec. 2021, 21, 4150. [Google Scholar] [CrossRef]
  26. Mane, B.B.; Waghmode, S.B. Iron-catalyzed ring opening of cyclopropanols and their 1,6-conjugate addition to p-quinone methides. J. Org. Chem. 2021, 86, 17774. [Google Scholar] [CrossRef]
  27. Shirsath, S.R.; Chandgude, S.M.; Muthukrishnan, M. Iron catalyzed tandem ring opening/1,6-conjugate addition of cyclopropanols with p-quinone methides: New access to γ,γ-diaryl ketones. Chem. Commun. 2021, 57, 13582. [Google Scholar] [CrossRef]
  28. Sun, Z.; Sun, B.; Kumagai, N.; Shibasaki, M. Direct catalytic asymmetric 1,6-conjugate addition of amides to p-quinone methides. Org. Lett. 2018, 20, 3070. [Google Scholar] [CrossRef]
  29. Francesco, B.; Gabriella, R.; Fabio, C.; Marialorena, C.; Roberta, G.; Serafinella, P.C. Topical tacalcitol as neoadjuvant for photodynamic therapy of acral actinic keratoses: An intra-patient randomized study. Photodiagnosis Photodyn. Ther. 2020, 31, 101803. [Google Scholar]
  30. Nanna, K.; Vibeke, B.; Sebastian, R.; Peter, H.; Morten, H. Pharmacokinetics of nebulized and oral procaterol in asthmatic and non-asthmatic subjects in relation to doping analysis. Drug Test. Anal. 2016, 8, 1056. [Google Scholar]
  31. Francesco, L.G.; Colin, G.E. Use of isoxsuprine hydrochloride as a tocolytic agent in the treatment of preterm labour: A systematic review of previous literature. Arzneimittelforschung 2010, 60, 415. [Google Scholar]
  32. Chiyohiko, S.; Yoshie, M.; Rie, S.; Katsuyuki, S.; Taichi, N.; Masahito, M. Tulobuterol patch maintains diaphragm muscle contractility for over twenty-four hours in a mouse model of sepsis. Tohoku J. Exp. Med. 2009, 218, 271. [Google Scholar]
  33. Fabiola, O.; Jesus, G.-R.; Julia, C.; Alvaro, C.; Aranzazu, H. Spectro- electrochemical determination of isoprenaline in a pharmaceutical sample. Sensors 2020, 20, 5179. [Google Scholar]
  34. Lou, Y.; Cao, P.; Jia, T.; Zhang, Y.; Wang, M.; Liao, J. Copper-catalyzed enantioselective 1,6-boration of para-quinone methides and efficient transformation of gem-diarylmethine boronates to triarylmethanes. Angew. Chem. Int. Ed. 2015, 54, 12134. [Google Scholar] [CrossRef] [PubMed]
  35. JaravaBarrera, C.; Parra, A.; López, A.; Cruz-Acosta, F.; Collado-Sanz, D.; Cárdenas, D.J.; Tortosa, M. Copper-catalyzed borylative aromatization of p-quinone methides: Enantioselective synthesis of dibenzylic boronates. ACS Catal. 2016, 6, 442. [Google Scholar] [CrossRef]
  36. Li, W.; Wen, W.; Chen, S.; Ding, L.; He, B.; Zhang, Y.; Li, B.; Zhu, L. Copper(II)-catalyzed 1,6-hydroboration reactions of p-quinone methides under ligand-Free conditions: A sequential methodology to gem-disubstituted methanols. Cat. Lett. 2022. [Google Scholar] [CrossRef]
  37. Xu, P.; Li, B.; Wang, L.; Qin, C.; Zhu, L. A green and recyclable chitosan supported catalyst for the borylation of α,β-unsaturated acceptors in water. Catal. Commun. 2016, 86, 23. [Google Scholar] [CrossRef]
  38. Zhu, L.; Li, B.; Wang, S.; Wang, W.; Wang, L.; Ding, L.; Qin, C. Recyclable heterogeneous chitosan supported copper catalyst for silyl conjugate addition to α,β-unsaturated acceptors in water. Polymers 2018, 10, 385. [Google Scholar] [CrossRef] [Green Version]
  39. De Souza, J.F.; da Silva, G.T.; Fajardo, A.R. Chitosan-based film supported copper nanoparticles: A potential and reusable catalyst for the reduction of aromatic nitro compounds. Carbohydr. Polym. 2017, 161, 187. [Google Scholar] [CrossRef]
  40. Wu, R.X.; Zheng, F.G.; Li, W.W.; Zhong, L.B.; Zheng, Y.M. Electrospun chitosan nanofiber membrane for adsorption of Cu(II) from aqueous solution: Fabrication, characterization and performance. J. Nanosci. Nanotechnol. 2018, 18, 5624. [Google Scholar] [CrossRef]
  41. Kaiser, M.; Mudasir, A.; Suhail, A.; Saiqa, I. Synthesis, characterization, kinetics, and thermodynamics of EDTA-modified chitosan-carboxymethyl cellulose as Cu(II) ion adsorbent. ACS Omega 2019, 17, 17425. [Google Scholar]
Scheme 1. Cu(II)-catalyzed 1,6-hydroboration of p-quinone methides.
Scheme 1. Cu(II)-catalyzed 1,6-hydroboration of p-quinone methides.
Molecules 27 07962 sch001
Figure 1. Screening of the scope of the substrates. Reaction conditions: a 1a (0.15 mmol), B2Pin2 (0.18 mmol), CP@Cu (5.0 mg, 0.009 mmol), acetone:H2O = 4:1 (2.5 mL) in room temperature for 2 h. b 12 h.
Figure 1. Screening of the scope of the substrates. Reaction conditions: a 1a (0.15 mmol), B2Pin2 (0.18 mmol), CP@Cu (5.0 mg, 0.009 mmol), acetone:H2O = 4:1 (2.5 mL) in room temperature for 2 h. b 12 h.
Molecules 27 07962 g001
Figure 2. The recycling experiments.
Figure 2. The recycling experiments.
Molecules 27 07962 g002
Figure 3. TEM images of (a) CP@Cu at 60 nm and (b) CP@Cu at 20 nm.
Figure 3. TEM images of (a) CP@Cu at 60 nm and (b) CP@Cu at 20 nm.
Molecules 27 07962 g003
Figure 4. High-resolution XPS spectra of (a) CP@Cu after metal adsorption; (b) high-resolution Cu 2p spectra; (c) high-resolution C 1s spectra after adsorption.
Figure 4. High-resolution XPS spectra of (a) CP@Cu after metal adsorption; (b) high-resolution Cu 2p spectra; (c) high-resolution C 1s spectra after adsorption.
Molecules 27 07962 g004
Table 1. Reaction condition optimizations a.
Table 1. Reaction condition optimizations a.
Molecules 27 07962 i001
EntriesCP@CuSolventAdditiveT (h)Yields (%) b
16 (mol%)EtOH-2N.R.
26 (mol%)DCMMeOH (2 equiv.)2N.R.
36 (mol%)THFMeOH (2 equiv.)2N.R.
46 (mol%)MeCNMeOH (2 equiv.)234
56 (mol%)AcetoneMeOH (2 equiv.)270
66 (mol%)AcetoneH2O (2 equiv.)2N.R.
76 (mol%)Acetone:H2O (4:1)-298
86 (mol%)Acetone:H2O (2:1)-284
96 (mol%)Acetone:H2O (1:2)-259
106 (mol%)Acetone:H2O (1:4)-2<5
11-Acetone:H2O (4:1)-2N.R.
126 (mol%)Acetone:H2O (4:1)-193
Reaction conditions: a 1a (0.15 mmol), B2Pin2 (0.18 mmol), CP@Cu (5.0 mg, 0.009 mmol), solvents (2.5 mL) at room temperature. b Isolated yield. N.R. = No reaction.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, S.; Wen, W.; Zhao, X.; Zhang, Z.; Li, W.; Zhang, Y.; Li, B.; Zhu, L. Preparation of Chitosan-Composite-Film-Supported Copper Nanoparticles and Their Application in 1,6-Hydroboration Reactions of p-Quinone Methides. Molecules 2022, 27, 7962. https://doi.org/10.3390/molecules27227962

AMA Style

Chen S, Wen W, Zhao X, Zhang Z, Li W, Zhang Y, Li B, Zhu L. Preparation of Chitosan-Composite-Film-Supported Copper Nanoparticles and Their Application in 1,6-Hydroboration Reactions of p-Quinone Methides. Molecules. 2022; 27(22):7962. https://doi.org/10.3390/molecules27227962

Chicago/Turabian Style

Chen, Shuhan, Wei Wen, Xue Zhao, Zelang Zhang, Weishuang Li, Yaoyao Zhang, Bojie Li, and Lei Zhu. 2022. "Preparation of Chitosan-Composite-Film-Supported Copper Nanoparticles and Their Application in 1,6-Hydroboration Reactions of p-Quinone Methides" Molecules 27, no. 22: 7962. https://doi.org/10.3390/molecules27227962

Article Metrics

Back to TopTop