Next Article in Journal
Application of Cyclodextrin for Cancer Immunotherapy
Next Article in Special Issue
A Sustainable Green Enzymatic Method for Amide Bond Formation
Previous Article in Journal
Pesticide Residues in Mandarins: Three-Year Monitoring Results
Previous Article in Special Issue
Research Progress on Propylene Preparation by Propane Dehydrogenation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Application of Chitosan/Poly(vinyl alcohol) Stabilized Copper Film Materials for the Borylation of α, β-Unsaturated Ketones, Morita-Baylis-Hillman Alcohols and Esters in Aqueous Phase †

1
School of Chemistry and Materials Science, Hubei Key Laboratory of Quality Control of Characteristic Fruits and Vegetables, 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.
Dedicated to the 80th anniversary celebration of Hubei Engineering University.
Molecules 2023, 28(14), 5609; https://doi.org/10.3390/molecules28145609
Submission received: 28 June 2023 / Revised: 21 July 2023 / Accepted: 21 July 2023 / Published: 24 July 2023
(This article belongs to the Special Issue Catalysis for Green Chemistry)

Abstract

:
A chitosan/poly(vinyl alcohol)-stabilized copper nanoparticle (CP@Cu NPs) was used as a heterogeneous catalyst for the borylation of α, β-unsaturated ketones, MBH alcohols, and MBH esters in mild conditions. This catalyst not only demonstrated remarkable efficiency in synthesizing organoboron compounds but also still maintained excellent reactivity and stability even after seven recycled uses of the catalyst. This methodology provides a gentle and efficient approach to synthesize the organoboron compounds by efficiently constructing carbon–boron bonds.

Graphical Abstract

1. Introduction

Organoboron compounds are a significant class of organic intermediates that are capable of reacting with various nucleophilic reagents and could be conveniently converted into various chemical bonds, such as C–C, C–O, and C–N bonds, which are very important in organic synthesis [1,2,3]. For this reason, in the past decades, various transition metals, including rhodium [4], palladium [5,6], platinum [7], cobalt [8], nickel [9], and copper [10,11,12], have been used as catalysts to synthesize the organoboron compounds. Recently, metal nanoparticles have received much attention because of their large surface area, which could increase catalytic efficiency [13,14]. However, metal nanoparticles sometimes have some drawbacks in the aqueous phase; for example, they are prone to aggregation and precipitation. To overcome these problems, an appropriate carrier was considered to load these metal nanoparticles, which could make these loaded nano-catalytic materials have several good advantages, including easy recovery and uniform dispersion [15,16].
Montmorillonite [17,18,19], activated carbon [20,21,22], and chitosan [23,24,25,26] are commonly used as carriers for heterogeneous catalytic materials in organic synthesis. Among these materials, chitosan is especially preferred due to its numerous amino and hydroxyl groups. These functional groups could coordinate well with various metals to perform a good catalytic activity. So far, several transition metals have been reported using chitosan as a support, including gold [27], silver [28], ruthenium [29], palladium [30,31,32], platinum [33], and copper [34,35,36]. Compared with these precious metals, copper has received more attention for its low price and lower toxicity.
In our previous research, we found that the borylation reactions of chalcone derivatives could be carried out smoothly, and the corresponding target products could also be obtained in good yields when CS@Cu(OH)2 [37], Cell-CuI NPs [38], or CP@Cu NPs [39] were used as catalysts (Scheme 1a). Under oxidative conditions, the resulting organoborons could give rise to the desired β-hydroxy-substituted carbonyl compounds, which are widely found in active molecules [40,41,42]. Compared with other supports, chitosan has inherent advantages due to its green property, abundance, stability, and ability of chelation [26]. With our continuous efforts in exploring applications of chitosan-supported metal catalysts, we were interested in developing a chitosan composite film of stabilized copper nanoparticles and its application for the synthesis of useful organoboron compounds. In particular, chitosan/poly(vinyl alcohol) composite films loaded with copper nanoparticles (CP@Cu NPs) were found to exhibit high reactivity, as well as excellent reusability and stability. Not only did α, β-unsaturated ketones have good reactivity in borylation reactions, but also α, β-unsaturated esters and amides could react smoothly when CP@Cu NPs was used as a catalyst in the reactions, and even after the catalyst was reused seven times, it still showed very good catalytic activity. However, in our previous work, the CP@Cu NPs were limited to the borylation reactions of 1,2-disubstituted α, β-unsaturated compounds, whereas the borylation reactions of 1,1-disubstituted unsaturated compounds were not explored, including the borylation reactions of MBH alcohols, and esters were also not involved. Morita-Baylis-Hillman alcohols or esters have aroused much attention in organic synthesis as valuable synthons and intermediates for the preparation of many important cyclic and acyclic compounds. Thus, their ready availability and condensed functional groups make them particularly attractive. In recent years, considering the importance of MBH alcohols and esters in organic synthesis, more and more research groups pay attention to their applications, especially as electrophilic reagents in borylation reactions [43,44,45,46,47,48,49,50,51,52,53]. These methods still have some shortcomings; for instance, the precious metal palladium as a catalyst is needed in reactions [43,51,52,53]. Even with copper as a catalyst, the reaction substrate range is quite limited, and only MBH alcohols or esters are compatible in these methods [44,45,46,47,48,49,50]. And more importantly, because all of the above methods are homogeneous reactions, the catalysts in reactions are difficult to be separated and reused after the reactions, which resulted in waste and heavy metal residues. Herein, in this work, we used CP@Cu NPs as a heterogeneous catalyst and 1,1-disubstituted α, β-unsaturated compounds (including α, β-unsaturated ketones, MBH alcohols, and esters) as substrates. The borylation reactions of these compounds could be achieved under mild conditions. Considering that some organoboron compounds are not very stable, the corresponding β-hydroxy-substituted carbonyl compounds were obtained via direct oxidation. Finally, the activity and stability of the catalysts were proved by the recovery experiments (Scheme 1b).

2. Results and Discussion

2.1. Catalysis of CP@Cu NPs in the Borylation Reaction of α, β-Unsaturated Ketones

The initial experiments commenced with α, β-unsaturated ketone II-1 (0.2 mmol) as a model substrate. CP@Cu NPs (10.0 mg, 9.0 mol%) was used as a catalyst by using B2(pin)2 (0.4 mmol, 2.0 equiv) as a boron source in 2.0 mL of solvents. First, various organic solvents were investigated, and no additives were added (Table 1, entries 1–7). When THF and toluene were used as solvents in this reaction, no reaction happened (Table 1, entries 1–2). Surprisingly, when ether was used as a solvent, the reaction occurred, and the target product was obtained in 8% yield (Table 1, entry 3). We continued to explore other solvents, such as MeOH, acetone, and H2O; the reaction could still happen, but the yields increased not obviously (Table 1, entries 4–6). In our previous work, we found that when we used mixed solvents in the reaction, excellent yields could be gained [37,38,39]. Considering the role of protons in this reaction, we used MeOH and H2O as mixed solvents (MeOH/H2O = 3:1), and the yield was increased to 28% yield (Table 1, entry 7). Next, we intended to examine the effects of additives in the reactions, mainly various organic bases, including 2,2-bipyridine, DMAP, 2-cyanopyridine, 2-chloropyridine, 2-bromopyridine, 2,6-dibromopyridine, and 4-picoline (Table 1, entries 8–14). We found that when we used 4-picoline as an additive in the reaction, the reaction worked very well, and the yield could obviously increase to 60% yield (Table 1, entry 14). Inspired by this result, we considered that the ratio of MeOH and H2O may have some contribution to these reactions. When we changed the ratio of the mixed solvents, different results were observed (Table 1, entries 15–18). In particular, when the ratio of MeOH to H2O was 1:1, the best result 92% yield could be obtained (Table 1, entry 16). In the organic synthesis, the reaction time is also one of the important factors that would affect the yield; therefore, we carried out the examination of the reaction time (Table 1, entries 19–21 vs. 16), and it was found that the reaction efficiency was still the highest when the reaction time was 12 h (Table 1, entry 16) and the yield was decreased whether the reaction time was shortened or prolonged. In order to study the effect of the amounts of additives on the reactions, we reduced or increased the amounts of additives and found that it actually had an effect on the reaction, and the yields were reduced to a certain extent (Table 1, entries 22–23 vs. 16). Finally, we investigated the catalyst loading in the reactions. When the catalyst loading was reduced to 4.5 mol%, the yield was still 92% (Table 1, entry 24), but when the amount of catalyst was increased to 13.5 mol%, the yield decreased to 90% (Table 1, entry 25). Therefore, we chose to use 4.5 mol% of catalyst loading to carry out the reactions from the view of economy. Thus, by a series of optimizations of the conditions, the optimal conditions of this research were 4.5 mol% CP@Cu NPs as a catalyst, 2.0 equiv of B2(pin)2 as a boron source, and 6.0 mol% of 4-picoline as an additive, and the whole reaction was conducted in 2.0 mL of mixed solvents (MeOH/H2O = 1:1) at room temperature for 12 h (Table 1, entry 24).
With the optimal experimental conditions in hand, we continued to investigate the universality of the reaction, and the results are summarized in Scheme 2. We mainly examined the effects of the substituents on the benzene ring of 1,1-disubstituted α, β-unsaturated ketones on the yields (Scheme 2). We first investigated the para-substituted functional groups on the benzene ring Ar1; when the substituents were methoxyl and methyl, the yields of the borylation were slightly decreased, and the possible reason was that both methoxyl and methyl were electron-donating groups that had an effect on the electrophilicity of the substrates (II-2a-II-3a, 81–85% yields). When the substituents were changed to fluorine and bromine, the yields were very good (II-4a, 90% yield; II-6a, 93% yield). However, when chlorine was used as a substituent, the yield was decreased to 80% yield (II-5a), mainly because it had less electron absorption than fluorine and bromine. Next, we examined the ortho-substituents on the benzene ring Ar1; the electron-donating group methoxyl had a better yield than the electron-withdrawing group bromine (II-7a, 90% yield, vs. II-8a, 57% yield). We also investigated the reactivity of meto-position of the benzene ring Ar1, and the yields of the reactions were still good (II-9a-II-10a, 80–92% yields). Last, we found that the substituents on the para-position of the benzene ring Ar2 had little effect on the yields; neither was it the electron-donating group methyl, nor the electron-withdrawing group fluorine (II-11a-II-12a, 81–88% yields, vs. II-3a, 81% yield).

2.2. Catalysis of CP@Cu NPs in the Borylation Reaction of MBH Alcohols and Esters

MBH alcohols and esters are very important intermediates in organic synthesis, and there is not much research on the borylation reactions of these compounds at present. Therefore, in this work, we planned to use them as reaction substrates for condition optimization. The same as the above condition optimizations, we selected MBH alcohols III-1 (0.2 mmol) or MBH esters IV-1 (0.2 mmol) as a model substrate, CP@Cu NPs (5.0 mg, 4.5 mol%) as a catalyst, and B2(pin)2 (0.4 mmol, 2.0 equiv) as a boron source in 2.0 mL of solvents; the whole reaction was conducted at room temperature for 12 h, and no additives were needed (Table 2). According to the above experimental results we achieved, we believed that proton solvents were beneficial to this reaction, so we just chose methanol and water as the solvents for screening. First, when we used methanol as a solvent, both MBH alcohol III-1 and IV-1 could react smoothly, and considering that the intermediates III-1a and IV-1a were not very stable in these conditions, we directly further oxidized these intermediates to the corresponding β-hydroxy substituted products using excessive NaBO3•4H2O as an oxidant (entry1, 31% yield; entry 2, 45% yield). And when H2O was used as a solvent, both reaction yields were not improved (entry 3, 33% yield; entry 4, 20% yield). Then, based on our previous experiment results, the mixed solvents were beneficial to this reaction, and we considered using methanol and water as the mixed solvents for conditional screening. We investigated the ratio of methanol to water in a mixed solvent and found that the highest yield could be obtained while the ratio of methanol to water was 2:1, and the target products could be obtained in 93% (entry 7 for III-1b) and 92% (entry 8 for IV-1b) isolated yields. Finally, we investigated the catalyst loading and found that the yields of the reactions did not change much. From the economic point of view, 4.5 mol% of catalyst loading was still the best choice in the reactions. Therefore, the optimal conditions for this reaction were MBH alcohols III-1 (0.2 mmol) or MBH esters IV-1 (0.2 mmol) as a model substrate, CP@Cu NPs (5.0 mg, 4.5 mol%) as a catalyst, and B2(pin)2 (0.4 mmol, 2.0 equiv) as a boron source in 2.0 mL of mixed solvent (MeOH/H2O = 2:1), and the whole reaction was conducted at room temperature for 12 h (entry 7 for III-1b, entry 8 for IV-1b).
With the optimized conditions in hand, we investigated the universality of the borylation reactions of MBH alcohols and esters, and the results are summarized in Scheme 3. We first examined the reactions as group R1 was different; when R1 was 4-methylphenyl, 4-ethylphenyl, 4-isopropylphenyl, 4-tert-butylphenyl, and 4-methoxyphenyl, the effect of the substituent on the reaction was not significant, whether they were MBH alcohols (III-2b-III-6b, 85–98% yields) or esters (IV-2b-IV-6b, 80–91% yields). However, when group R1 was 4-fluorophenyl, 4-chlorophenyl, 4-bromophenyl, and 4-trifluoromethylphenyl, the reaction results were not very good compared with the model reaction. Especially for MBH alcohols, when R1 was 4-bromophenyl, the target product was not detected (III-9b), and when R1 was 4-trifluoromethylphenyl, the yield was not good because of its strong electron absorption (III-10b, 53% yield). When R1 consisted of 3-substituted benzene rings, the electronic effect of the benzene ring had a great influence on the reactions. When the benzene ring was connected with electron-donating groups, such as 3-methyl and 3-methoxy, the yields were better, but for the electron-deficient group, for example, 3-bromophenyl, the yield had a great influence (III-13b, 50% yield; IV-13b, 39% yield). However, when R1 was 2-substituted phenyl, the electronic effect of the aromatic rings had no effect on the reactions, no matter whether they were electron-absorbing substituents or electron-giving substituents (III-14b-III-16b, 93–98% yields; IV-14b-IV-16b, 87–95% yields). For the disubstituted benzenes of R1, no matter whether they were electron-absorbing substituents or electron-giving substituents, the reactions could still have good yields (III-17b-III-20b, 71–97% yields; IV-17b-IV-20b, 67–80% yields). To our delight, when R1 was the 2-thiophene substituent, the reactions could still take place, and the target products could be obtained in medium yields (III-21b, 43% yield; IV-21b, 62% yield). Next, we continued to investigate the reactions when R1 consisted of alkyl groups. From the reaction results, we found that the alkyl substituents could occur smoothly, and the target products could be synthesized in medium to excellent yields (III-22b-III-27b, 48–99% yields; IV-22b-IV-27b, 40–66% yields). Finally, we investigated the reaction of R2 and found that when R2 was ethyl, the reaction activity was still very good, and the corresponding target product could be obtained with good yields (III-28b, 86% yield; IV-28b, 99% yield).

2.3. Recycling Experiments of CP@Cu NPs in Borylation Reactions

The main advantage of heterogeneous catalysis in organic synthesis was that the catalyst in the system could be easily recovered and reused. Such a type of operation could not only increase the catalytic efficiency of the catalyst and reduce the cost of the reactions but also avoid the heavy metal residue to the environment. In this work, to assess the reusability and stability of the CP@Cu NPs in borylation reactions, we used MBH alcohols III-1 as a substrate and CP@Cu NPs as a catalyst. After the completion of the reaction, the catalyst CP@Cu NPs could be recycled with a simple operation. The results showed that the activity of the catalyst stayed very well, and the yield of the product could also still be up to 84% even in the seventh experiment, which confirmed that the catalyst could be recyclable (Figure 1). Notably, the yields of the eighth and ninth cycles were still 83% and 82%, respectively. The slight decrease in the yields that was observed in the recycling experiments was probably due to the formation of a byproduct, which may be absorbed onto the surface of CP@Cu NPs. It must also be mentioned that the catalyst could be reactivated by washing with 10% aq. NaOH solution and dried again after the reaction. By using this process, an average of ~90% yield could be obtained after each cycle. Furthermore, ICP tests of recycled catalyst were carried out, and almost no detectable copper leaching was observed. These results strongly indicated that the CP@Cu NPs was a highly active heterogeneous catalyst for this borylation process.

3. Materials and Methods

3.1. Materials

Chitosan/poly(vinyl alcohol) composite film-supported copper nanoparticles (CP@Cu NPs) were prepared, according to the procedures reported [39]. The characterization of CP@Cu NPs was described in the supplementary materials. Bis(pinacolato)diboron (B2(pin)2, AR), methanol (MeOH, AR), ethanol (EtOH, AR), acetone (AR), tetrahydrofuran (THF, AR), ether (Et2O, AR), 2,2-bipyridine (AR), 4-dimethylaminopyridine (DMAP, AR), 2-cyanopyridine (AR), 2-chloropyridine (AR), 2-bromopyridine (AR), and 4-picoline (AR) were obtained commercially from Energy Chemical (Shanghai, China).

3.2. Synthesis of α, β-Unsaturated Ketones II

In step 1, a mixture of substituted phenylacetonitrile (10 mmol), substituted phenylboronic acid (20 mmol), Pd (OAc)2 (112.3 mg, 0.5 mmol), 2,2′-dipyridine (156.2 mg, 1.0 mmol), TFA (11.4 g, 100 mmol), and H2O (4 mL) were added into THF (20 mL). Then the mixture was refluxed under nitrogen atmosphere for 2–3 days. The residue was extracted with EtOAc (20 mL) three times. After evaporation of solvent, the crude mixture was purified by flash column chromatograph to afford the intermediate compounds.
In step 2, to the compounds (5 mmol) obtained from step 1, formaldehyde (0.60 g, 20 mmol), piperidine (42.1 mg, 0.5 mmol), AcOH (60.1 mg, 1.0 mmol), and MeOH (5 mL) were added. The mixture was then refluxed for 6 h. After the completion of this reaction, evaporation was carried out to remove MeOH. The residue was washed with CH2Cl2 to collect the organic layer, which was washed with brine, dried over Na2SO4, and concentrated in vacuo. The desired ketones II [54,55] were obtained by further purification by silica gel chromatography.

3.3. Synthesis of MBH Alcohols III

Different substituted benzaldehyde (10 mmol), methyl acrylate (1.72 g, 20 mmol) or ethyl acrylate (2.00 g, 20 mmol), and DABCO (1.12 g,10 mmol) were successively added into a 50 mL flask under air. After stirring for 3–7 days at room temperature, the reaction mixture was filtered, and the filtrate was extracted with EtOAc (20 mL) three times. The crude mixture was purified by silica gel chromatography to afford the desired MBH alcohols III [56].

3.4. Synthesis of MBH Esters IV

Different substituted MBH alcohols III (10 mmol), acetic anhydride (1.23 g, 12 mmol), 4-DMAP (122.2 mg, 1 mmol), and DCM (10 mL) were successively added to a 50 mL flask under air. The reaction was monitored by TLC. After completion of reaction, the mixture was filtered, and the filtrate was extracted with EtOAc (20 mL) three times. Then the crude mixture was purified by silica gel chromatography to afford the corresponding MBH esters IV [56].

3.5. Analytical Methods

The purification of products was accomplished by using flash column chromatography on silica gel (200–300 mesh) or preparative TLC. Nuclear magnetic resonance (NMR) spectra were recorded on a Bruker Avance III 400 MHz spectrometer (Karlsruhe, Germany) operating at 400 MHz 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.

3.6. Copper-Catalyzed Borylation Reactions

The reaction procedure is depicted in Scheme 1b. Nano-sized copper loaded onto the membrane material could enable the borylation reaction of α, β-unsaturated ketones, MBH alcohols, and esters under very mild conditions. Because of the difference in reactivity, the borylation of α, β-unsaturated ketones with B2(pin)2 required the additional bases, whereas the borylation of MBH alcohols and esters could be conducted smoothly without the bases.

3.6.1. Borylation Reactions of α, β-Unsaturated Ketones

At room temperature, 0.2 mmol of α, β-unsaturated ketones II, 0.4 mmol of B2(pin)2, 6 mol% of base, 10.0 mg of CP@Cu NPs as a catalyst, and 2.0 mL of solvent were added in the reaction system. The whole reactions were stirred at room temperature for 12 h, and after completion of the reaction, the mixture was filtered, and the desired products II-a were obtained by being purified with column chromatography and characterized by 1H NMR and 13C NMR (see supplementary materials).

3.6.2. Borylation Reactions of MBH Alcohols and Esters

At room temperature, a reaction mixture containing 0.2 mmol of MBH alcohols or esters (III or IV), 0.4 mmol of B2(pin)2, 10.0 mg of CP@Cu NPs, and 2.0 mL of solvent were prepared, and the whole reactions were stirred at room temperature. After the completion of the reactions, the organic phase was separated, and the crude intermediates III-a or IV-a were added to a mixture of THF-H2O containing an excess of sodium perborate, and the mixture continued to be stirred for 4 h. When the reaction finished, the desired products III-b or IV-b were obtained, purified by column chromatography, and characterized by 1H NMR and 13C NMR (see supplementary materials).

3.7. General Procedure for ICP Test of CP@Cu NPs

Chitosan/poly(vinyl alcohol) composite film supported copper nanoparticles (CP@Cu NPs) (~20 mg) were placed in a clean test tube and heated with H2SO4 (1 mL) at 200 °C. After 30 min, several drops of concentrated HNO3 were added carefully, and the tube was shaken occasionally. HNO3 was continuously added until a clear solution was obtained, and the excess amount of HNO3 was allowed to evaporate under heating. After the solution was cooled to room temperature, 1 mL of aqua regia was added carefully. The effervescence of gas was observed, and the solution became clearer. The solution was then transferred to a volumetric flask and increased up to 50 mL with water, which was submitted for ICP analysis.

3.8. General Procedure for the Sample Preparation for ICP Analysis to Determine Metal Leaching

After the reaction was finished, the reaction mixture was filtered. The filtrate obtained was concentrated and diluted with 10 mL of THF. Then, 50% v/v of the crude THF solution (5 mL) was then passed through a membrane filter (0.25 or 0.45 µm) into a clean test tube. After the evaporation of the solvent, the solid obtained in the test tube was heated to 200 °C, and 1.0 mL of concentrated H2SO4 was added. Following a procedure similar to that described above, concentrated HNO3 was added at regular intervals until the resulting solution was clear. After the solution was cooled to room temperature, 1 mL of aqua regia was added carefully. The effervescence of the gas was observed, and the solution became clearer. The solution was then transferred to a volumetric flask and increased up to 50 mL with water, which was submitted for ICP analysis.

4. Conclusions

In summary, we reported the preparation of a chitosan-loaded copper catalyst (CP@Cu NPs) and its application in the borylation of α, β-unsaturated ketones, MBH alcohols, and esters with B2(pin)2 as a boron source. The whole reaction conditions were very mild, and no additives were even needed in the borylation of the MBH alcohols and esters. It demonstrated that the substrate scope of this newly developed method was very broad (more than 40 examples) with very high activity of the catalyst (up to 99% yield). Remarkably, a single heterogeneous catalyst could efficiently catalyze three types of substrates including the borylation of α, β-unsaturated ketones, MBH alcohols, and esters. Moreover, this newly developed strategy could largely solve the recovery of copper catalysts, providing a green and economic way for the efficient synthesis of organoboron compounds in the aqueous phase.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28145609/s1.

Author Contributions

Conceptualization, L.Z.; methodology, B.L., W.W. (Wu Wen) and W.W. (Wei Wen); formal analysis, H.G., C.F. and Y.Z.; resources, B.L., Y.Z. and L.Z.; writing—original draft preparation, B.L.; writing—review and editing, Y.Z. and L.Z.; supervision, L.Z.; funding acquisition, Y.Z. and L.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Key Project of Science and Technology Research of Hubei Provincial Department of Education (D20222701), the National Natural Science Foundation of China (Nos. 22108065, 22278118, 21774029), Key Research and Development Program of Hubei Province (2022BAD083), Central Leading Local Science and Technology Development Special Project of Hubei Province (2022BGE258), Excellent Young and Middle-aged Science and Technology Innovation Team Project of Hubei University (No. T201816).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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]
  2. 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]
  3. 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]
  4. Kabalka, G.W.; Das, B.C.; Das, S. Rhodium-catalyzed 1,4-addition reactions of diboron reagents to electron deficient olefins. Tetrahedron Lett. 2002, 43, 2323. [Google Scholar] [CrossRef]
  5. Abu, A.H.; Goldberg, I.; Srebnik, M. Addition reactions of bis(pinacolato)diborane(4) to carbonyl enones and synthesis of (pinacolato)2BCH2B and (pinacolato)2BCH2CH2B by insertion and coupling. Organometallics 2001, 20, 3962. [Google Scholar] [CrossRef]
  6. Bonet, A.; Gulyas, H.; Koshevoy, I.O.; Estevan, F.; Sanau, M.; Úbeda, M.A.; Fernandez, E. Tandem β-boration/arylation of α, β-unsaturated carbonyl compounds by using a single palladium complex to catalyse both steps. Chem.–A Eur. J. 2010, 16, 6382. [Google Scholar] [CrossRef]
  7. Lawson, Y.; Norman, N.; Rice, C.; Marder, T. Platinum catalysed 1,4-diboration of α, β-unsaturated ketones. Chem. Commun. 1997, 2051. [Google Scholar] [CrossRef]
  8. Kajiwara, T.; Terabayashi, T.; Yamashita, M.; Nozaki, K. Syntheses, structures, and reactivities of borylcopper and-zinc compounds: 1, 4-silaboration of an α, β-unsaturated ketone to form a γ-siloxyallylborane. Angew. Chem. 2008, 120, 6708. [Google Scholar] [CrossRef]
  9. Hirano, K.; Yorimitsu, H.; Oshima, K. Nickel-catalyzed β-boration of α, β-unsaturated esters and amides with bis(pinacolato)diboron. Org. Lett. 2007, 9, 5031. [Google Scholar] [CrossRef] [PubMed]
  10. Ito, H.; Yamanaka, H.; Tateiwa, J.; Hosomi, A. Boration of an α, β-enone using a diboron promoted by a copper(I)–phosphine mixture catalyst. Tetrahedron Lett. 2000, 41, 6821. [Google Scholar] [CrossRef]
  11. Gawande, M.B.; Goswami, A.; Felpin, F.; Asefa, T.; Huang, X.; Silva, R.; Zou, X.; Zboril, R.; Varma, R.S. Cu and Cu-based nanoparticles: Synthesis and applications in catalysis. Chem. Rev. 2016, 116, 3722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Kitanosono, T.; Xu, P.; Kobayashi, S. Heterogeneous versus homogeneous copper(II) catalysis in enantioselective conjugate-addition reactions of boron in water. Chem.–Asian J. 2014, 9, 179. [Google Scholar] [CrossRef] [PubMed]
  13. Yan, N.; Xiao, C.; Kou, Y. Transition metal nanoparticle catalysis in green solvents. Coord. Chem. Rev. 2010, 254, 1179. [Google Scholar] [CrossRef]
  14. Nørskov, J.K.; Bligaard, T.; Hvolbæk, B.; Abild-Pedersen, F.; Chorkendorff, I.; Christensen, C.H. The nature of the active site in heterogeneous metal catalysis. Chem. Soc. Rev. 2008, 37, 2163. [Google Scholar] [CrossRef]
  15. Klemm, D.; Heublein, B.; Fink, H.P.; Bohn, A. Cellulose: Fascinating biopolymer and sustainable raw material. Angew. Chem. Int. Ed. 2005, 44, 3358. [Google Scholar] [CrossRef]
  16. Rezayat, M.; Blundell, R.K.; Camp, J.E.; Walsh, D.A.; Thielemans, W. Green one-step synthesis of catalytically active palladium nanoparticles supported on cellulose nanocrystals. ACS Sustain. Chem. Eng. 2014, 2, 1241. [Google Scholar] [CrossRef]
  17. Pinnavaia, T.J. Intercalated clay catalysts. Science 1983, 220, 365. [Google Scholar] [CrossRef]
  18. Kumar, D.D.; Jyoti, B.B.; Pollov, S.P. Recent advances in metal nanoparticles stabilization into nanopores of montmorillonite and their catalytic applications for fine chemicals synthesis. Catal. Rev. 2015, 57, 257. [Google Scholar] [CrossRef]
  19. Bhat, A.H.; Rangreez, T.A.; Chisti, H. Wastewater treatment and biomedical applications of montmorillonite based nanocomposites: A review. Curr. Anal. Chem. 2022, 18, 269. [Google Scholar] [CrossRef]
  20. Duan, D.; Chen, D.; Huang, L.; Zhang, Y.; Zhang, Y.; Wang, Q.; Xiao, G.; Zhang, W.; Lei, H.; Ruan, R. Activated carbon from lignocellulosic biomass as catalyst: A review of the applications in fast pyrolysis process. J. Anal. Appl. Pyrolysis 2021, 158, 105246. [Google Scholar] [CrossRef]
  21. Domínguez, C.M.; Ocón, P.; Quintanilla, A.; Casas, J.A.; Rodriguez, J.J. Highly efficient application of activated carbon as catalyst for wet peroxide oxidation. Appl. Catal. B Environ. 2013, 140, 663. [Google Scholar] [CrossRef]
  22. Matatov-Meytal, U.; Sheintuch, M. Activated carbon cloth-supported Pd–Cu catalyst: Application for continuous water denitrification. Catal. Today 2005, 102, 121. [Google Scholar] [CrossRef]
  23. Lee, M.; Chen, B.; Den, W. Chitosan as a natural polymer for heterogeneous catalysts support: A short review on its applications. Appl. Sci. 2015, 5, 1272. [Google Scholar] [CrossRef] [Green Version]
  24. Sahu, P.K.; Sahu, P.K.; Gupta, S.K.; Agarwal, D.D. Chitosan: An efficient, reusable, and biodegradable catalyst for green synthesis of heterocycles. Ind. Eng. Chem. Res. 2014, 53, 2085. [Google Scholar] [CrossRef]
  25. Zheng, K.; Xiao, S.; Li, W.; Wang, W.; Chen, H.; Yang, F.; Qin, C. Chitosan-acorn starch-eugenol edible film: Physico-chemical, barrier, antimicrobial, antioxidant and structural properties. Int. J. Biol. Macromol. 2019, 135, 344. [Google Scholar] [CrossRef]
  26. El, K.A. Chitosan as a sustainable organocatalyst: A concise overview. ChemSusChem 2015, 8, 217. [Google Scholar]
  27. Mironenko, A.Y.; Sergeev, A.; Nazirov, A.; Modin, E.; Voznesenskiy, S.; Bratskaya, S.Y. H2S optical waveguide gas sensors based on chitosan/Au and chitosan/Ag nanocomposites. Sens. Actuators B Chem. 2016, 225, 348. [Google Scholar] [CrossRef]
  28. Murugadoss, A.; Chattopadhyay, A. A ‘green’ chitosan–silver nanoparticle composite as a heterogeneous as well as micro-heterogeneous catalyst. Nanotechnology 2007, 19, 015603. [Google Scholar] [CrossRef]
  29. Baig, R.N.; Nadagouda, M.N.; Varma, R.S. Ruthenium on chitosan: A recyclable heterogeneous catalyst for aqueous hydration of nitriles to amides. Green Chem. 2014, 16, 2122. [Google Scholar] [CrossRef]
  30. Rafiee, F. Recent advances in the application of chitosan and chitosan derivatives as bio supported catalyst in the cross coupling reactions. Curr. Org. Chem. 2019, 23, 390. [Google Scholar] [CrossRef]
  31. Pei, X.; Zheng, X.; Liu, X.; Lei, A.; Zhang, L.; Yin, X. Facile fabrication of highly dispersed Pd catalyst on nanoporous chitosan and its application in environmental catalysis. Carbohydr. Polym. 2022, 286, 119313. [Google Scholar] [CrossRef]
  32. Zeng, M.; Zhang, X.; Shao, L.; Qi, C.; Zhang, X. Highly porous chitosan microspheres supported palladium catalyst for coupling reactions in organic and aqueous solutions. J. Organomet. Chem. 2012, 704, 29. [Google Scholar] [CrossRef]
  33. Xie, H.; Yue, H.; Zhang, W.; Hu, W.; Zhou, X.; Prinsen, P.; Luque, R. A chitosan modified Pt/SiO2 catalyst for the synthesis of 3-poly (ethylene glycol) propyl ether-heptamethyltrisiloxane applied as agricultural synergistic agent. Catal. Commun. 2018, 104, 118. [Google Scholar] [CrossRef]
  34. Shen, C.; Xu, J.; Yu, W.; Zhang, P. A highly active and easily recoverable chitosan@copper catalyst for the C–S coupling and its application in the synthesis of zolimidine. Green Chem. 2014, 16, 3007. [Google Scholar] [CrossRef]
  35. Mounir, C.; Ahlafi, H.; Aazza, M.; Moussout, H.; Mounir, S. Kinetics and langmuir–hinshelwood mechanism for the catalytic reduction of para-nitrophenol over Cu catalysts supported on chitin and chitosan biopolymers. React. Kinet. Mech. Catal. 2021, 134, 285. [Google Scholar] [CrossRef]
  36. Guo, Y.; Dai, M.; Zhu, Z.; Chen, Y.; He, H.; Qin, T. Chitosan modified Cu2O nanoparticles with high catalytic activity for p-nitrophenol reduction. Appl. Surf. Sci. 2019, 480, 601. [Google Scholar] [CrossRef]
  37. 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]
  38. Zhou, L.; Han, B.; Zhang, Y.; Li, B.; Wang, L.; Wang, J.; Wang, X.; Zhu, L. Cellulosic CuI nanoparticles as a heterogeneous, recyclable catalyst for the borylation of α, β-unsaturated acceptors in aqueous media. Catal. Lett. 2021, 151, 3220. [Google Scholar] [CrossRef]
  39. 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]
  40. Chi, W.W.; John, T.G.H.; David, O.H.; Richard, J.R. Tropic acid biosynthesis: The incorporation of (RS)-phenyl[2-18O,2-2H]lactate into littorine and hyoscyamine in datura stramonium. Chem. Commun. 1998, 1045. [Google Scholar]
  41. Ushimaru, R.; Ruszczycky, M.W.; Liu, H.-W. Changes in regioselectivity of H atom abstraction during the hydroxylation and cyclization reactions catalyzed by hyoscyamine 6β-Hydroxylase. J. Am. Chem. Soc. 2019, 141, 1062. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, K.; Ren, B.-H.; Liu, X.-F.; Wang, L.-L.; Zhang, M.; Ren, W.-M.; Lu, X.-B.; Zhang, W.-Z. Direct and selective electrocarboxylation of styrene oxides with CO2 for accessing β-hydroxy acids. Angew. Chem. Int. Ed. 2022, 61, e202207660. [Google Scholar]
  43. Guillaume, D.; Nicklas, S.; Kálmán, J.S.; Varinder, K.A. Direct synthesis of functionalized allylic boronic esters from allylic alcohols and inexpensive reagents and catalysts. Synthesis 2008, 14, 2293. [Google Scholar]
  44. Veeraraghavan, R.; Debarshi, P.; Debanjan, B.; Amit, S.; Venkat, R.R. Novel functionalized trisubstituted allylboronates via hosomi−miyaura borylation of functionalized allyl acetates. Org. Lett. 2004, 6, 481. [Google Scholar]
  45. Soumya, M.; Srinivas, R.G.; Robert, S.C. Total synthesis of the eupomatilones. J. Org. Chem. 2007, 72, 8724. [Google Scholar]
  46. Veeraraghavan, R.; Debarshi, P.; Hari, N.G.N.; Matthew, W.; Sadie, S.; Michele, T.Y.; Huangbing, W.; Max, S. Tailored a-methylene-c-butyrolactones and their effects on growth suppression in pancreatic carcinoma cells. Bioorg. Med. Chem. Lett. 2010, 20, 6620. [Google Scholar]
  47. Veeraraghavan, R.; Hari, N.G.N.; Pravin, G. Syntheses of β- and γ-fluorophenyl cis- and trans-α-methylene-γ-butyrolactones. Tetrahedron Lett. 2014, 55, 5722. [Google Scholar]
  48. Henry, J.H.; Nader, S.A.; Rusha, P.; Mohamed, N.S.; Veeraraghavan, R. β,γ-Diaryl α-methylene-γ-butyrolactones as potent antibacterials against methicillin-resistant Staphylococcus aureus. Bioorg. Chem. 2020, 104, 104183. [Google Scholar]
  49. Daniel, P.D.; Chong-Lei, J.; Peng, L.; Kay, M.B. Thiol reactivity of n-aryl α-methylene-γ-lactams: Influence of the guaianolide structure. J. Org. Chem. 2022, 87, 11204. [Google Scholar]
  50. Xuan, Q.; Wei, Y.; Chen, J.; Song, Q. An expedient E-stereoselective synthesis of multi-substituted functionalized allylic boronates from Morita-Baylis-Hillman alcohols. Org. Chem. Front. 2017, 4, 1220. [Google Scholar] [CrossRef]
  51. George, W.K.; Venkataiah, B.; Dong, G. Pd-catalyzed cross-coupling of baylis-hillman acetate adducts with bis(pinacolato)diboron: An efficient route to functionalized allyl borates. J. Org. Chem. 2004, 69, 5807. [Google Scholar]
  52. George, W.K.; Bollu, V. The total synthesis of eupomatilones 2 and 5. Tetrahedron Lett. 2005, 46, 7325. [Google Scholar]
  53. Feng, J.; Lei, X.; Bao, R.; Li, Y.; Xiao, C.; Hu, L.; Tang, Y. Enantioselective and collective total syntheses of xanthanolides. Angew. Chem. Int. Ed. 2017, 56, 16323. [Google Scholar] [CrossRef]
  54. Zhang, G.; Fan, Q.; Zhao, Y.; Ding, C. Copper-promoted oxidative intramolecular C–H amination of hydrazones to synthesize 1H-indazoles and 1H-pyrazoles using a cleavable directing group. Eur. J. Org. Chem. 2019, 33, 5801. [Google Scholar] [CrossRef]
  55. Zhao, P.; Wu, S.; Ke, C.; Liu, X.; Feng, X. Chiral lewis acid-catalyzed enantioselective cyclopropanation and C–H insertion reactions of vinyl ketones with α-diazoesters. Chem. Commun. 2018, 54, 9837. [Google Scholar] [CrossRef]
  56. Xia, B.; Xu, J.; Xiang, Z.; Cen, Y.; Hu, Y.; Lin, X.; Wu, Q. Stereoselectivity-tailored, metal-free hydrolytic dynamic kinetic resolution of Morita-Baylis-Hillman acetates using an engineered lipase–organic base cocatalyst. ACS Catal. 2017, 7, 4542. [Google Scholar] [CrossRef]
Scheme 1. Copper-catalyzed borylation reactions of α, β-unsaturated compounds.
Scheme 1. Copper-catalyzed borylation reactions of α, β-unsaturated compounds.
Molecules 28 05609 sch001
Scheme 2. Screening substrate scope of CP@Cu NPs in the borylation reaction of α, β-unsaturated ketones. Reaction conditions: II (0.2 mmol), B2(pin)2 (2.0 equiv), CP@Cu NPs (5.0 mg, 4.5 mol%), 4-picoline (6.0 mol%) in 2.0 mL of mixed solvents (MeOH/H2O =1:1) at room temperature for 12 h.
Scheme 2. Screening substrate scope of CP@Cu NPs in the borylation reaction of α, β-unsaturated ketones. Reaction conditions: II (0.2 mmol), B2(pin)2 (2.0 equiv), CP@Cu NPs (5.0 mg, 4.5 mol%), 4-picoline (6.0 mol%) in 2.0 mL of mixed solvents (MeOH/H2O =1:1) at room temperature for 12 h.
Molecules 28 05609 sch002
Scheme 3. Screening substrate expansion scope of CP@Cu NPs in the borylation reaction of MBH alcohols and esters. Reaction conditions: III-1 and IV-1 (0.2 mmol), B2Pin2 (0.4 mmol), CP@Cu (5.0 mg), 2.0 mL of mixed solvents with MeOH/H2O =3:1 for III-1 reaction and MeOH/H2O =2:1 for IV-1 reaction, at room temperature for 12 h. N.D. = no detection.
Scheme 3. Screening substrate expansion scope of CP@Cu NPs in the borylation reaction of MBH alcohols and esters. Reaction conditions: III-1 and IV-1 (0.2 mmol), B2Pin2 (0.4 mmol), CP@Cu (5.0 mg), 2.0 mL of mixed solvents with MeOH/H2O =3:1 for III-1 reaction and MeOH/H2O =2:1 for IV-1 reaction, at room temperature for 12 h. N.D. = no detection.
Molecules 28 05609 sch003
Figure 1. The recycle experiments.
Figure 1. The recycle experiments.
Molecules 28 05609 g001
Table 1. Optimization of CP@Cu NPs in the borylation reaction of α, β-unsaturated ketones a.
Table 1. Optimization of CP@Cu NPs in the borylation reaction of α, β-unsaturated ketones a.
Molecules 28 05609 i001
EntriesCP@Cu NPsSolvents (2.0 mL)AdditivesTime (h)NMR Yields (%)
1(9.0 mol%)THF-12N.R.
2(9.0 mol%)Toluene-12N.R.
3(9.0 mol%)Et2O-128
4(9.0 mol%)MeOH-1216
5(9.0 mol%)Acetone-128
6(9.0 mol%)H2O-1219
7(9.0 mol%)MeOH/H2O = 3:1-1228
8(9.0 mol%)MeOH/H2O = 3:12,2-bipyridine1235
9(9.0 mol%)MeOH/H2O = 3:1DMAP1233
10(9.0 mol%)MeOH/H2O = 3:12-cyanopyridine1254
11(9.0 mol%)MeOH/H2O = 3:12-chloropyridine1257
12(9.0 mol%)MeOH/H2O = 3:12-bromopyridine1250
13(9.0 mol%)MeOH/H2O = 3:12,6-dibromopyridine1248
14(9.0 mol%)MeOH/H2O = 3:14-picoline1260
15(9.0 mol%)MeOH/H2O = 2:14-picoline1272
16 b(9.0 mol%)MeOH/H2O = 1:14-picoline1292
17(9.0 mol%)MeOH/H2O = 1:24-picoline1289
18(9.0 mol%)MeOH/H2O = 1:44-picoline1280
19(9.0 mol%)MeOH/H2O = 1:14-picoline479
20(9.0 mol%)MeOH/H2O = 1:14-picoline887
21 b(9.0 mol%)MeOH/H2O = 1:14-picoline1690
22 b,c(9.0 mol%)MeOH/H2O = 1:14-picoline1290
23 b,d(9.0 mol%)MeOH/H2O = 1:14-picoline 1290
24 b(4.5 mol%)MeOH/H2O = 1:14-picoline1292
25 b(13.5 mol%)MeOH/H2O = 1:14-picoline1290
Reaction conditions: a II-1 (0.2 mmol), B2Pin2 (0.4 mmol), CP@Cu NPS (10.0 mg, 9.0 mol%), additives (6.0 mol%, 0.012 mmol), solvents (2.0 mL) at room temperature. b Isolated yield. c 4-picoline (5.0 mol%). d 4-picoline (7.0 mol%). N.R. = no reaction.
Table 2. Optimization of CP@Cu NPs in the borylation reaction of MBH alcohols and esters a.
Table 2. Optimization of CP@Cu NPs in the borylation reaction of MBH alcohols and esters a.
Molecules 28 05609 i002
EntriesSubstratesCP@Cu NPsSolvents (2.0 mL)NMR Yields (%)
1III-15.0 mgMeOH31
2IV-15.0 mgMeOH45
3III-15.0 mgH2O33
4IV-15.0 mgH2O20
5III-15.0 mgMeOH/H2O =3:194
6IV-15.0 mgMeOH/H2O =3:190
7 bIII-15.0 mgMeOH/H2O =2:193
8 bIV-15.0 mgMeOH/H2O =2:192
9III-15.0 mgMeOH/H2O =1:190
10IV-15.0 mgMeOH/H2O =1:186
11III-15.0 mgMeOH/H2O =1:289
12IV-15.0 mgMeOH/H2O =1:283
13III-15.0 mgMeOH/H2O =1:388
14IV-15.0 mgMeOH/H2O =1:385
15III-110.0 mgMeOH/H2O =2:193
16IV-110.0 mgMeOH/H2O =2:191
Reaction conditions: a III-1 and IV-1 (0.2 mmol), B2(pin)2 (0.4 mmol), CP@Cu (5.0 mg, 4.5 mol%), solvents (2.0 mL) at room temperature for 12 h. b Isolated yields.
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

Li, B.; Wen, W.; Wen, W.; Guo, H.; Fu, C.; Zhang, Y.; Zhu, L. Application of Chitosan/Poly(vinyl alcohol) Stabilized Copper Film Materials for the Borylation of α, β-Unsaturated Ketones, Morita-Baylis-Hillman Alcohols and Esters in Aqueous Phase. Molecules 2023, 28, 5609. https://doi.org/10.3390/molecules28145609

AMA Style

Li B, Wen W, Wen W, Guo H, Fu C, Zhang Y, Zhu L. Application of Chitosan/Poly(vinyl alcohol) Stabilized Copper Film Materials for the Borylation of α, β-Unsaturated Ketones, Morita-Baylis-Hillman Alcohols and Esters in Aqueous Phase. Molecules. 2023; 28(14):5609. https://doi.org/10.3390/molecules28145609

Chicago/Turabian Style

Li, Bojie, Wu Wen, Wei Wen, Haifeng Guo, Chengpeng Fu, Yaoyao Zhang, and Lei Zhu. 2023. "Application of Chitosan/Poly(vinyl alcohol) Stabilized Copper Film Materials for the Borylation of α, β-Unsaturated Ketones, Morita-Baylis-Hillman Alcohols and Esters in Aqueous Phase" Molecules 28, no. 14: 5609. https://doi.org/10.3390/molecules28145609

Article Metrics

Back to TopTop