Next Article in Journal
Isolation and Identification of an Efficient Aerobic Denitrifying Pseudomonas stutzeri Strain and Characterization of Its Nitrite Degradation
Next Article in Special Issue
Magnetically Reusable Fe3O4@NC@Pt Catalyst for Selective Reduction of Nitroarenes
Previous Article in Journal
Cu/ZSM5-Geopolymer 3D-Printed Monoliths for the NH3-SCR of NOx
Previous Article in Special Issue
Non-Solvent Synthesis of a Robust Potassium-Doped PdCu-Pd-Cu@C Nanocatalyst for High Selectively Tandem Reactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Recyclable Pd/H-MOR Catalyst for Suzuki-Miyaura Coupling and Application in the Synthesis of Crizotinib

1
College of Petroleum Chemical Industry, Changzhou University, Changzhou 213164, China
2
College of Biology and Environmental Engineering, Zhejiang Shuren University, Hangzhou 310015, China
*
Authors to whom correspondence should be addressed.
Catalysts 2021, 11(10), 1213; https://doi.org/10.3390/catal11101213
Submission received: 14 September 2021 / Revised: 30 September 2021 / Accepted: 5 October 2021 / Published: 9 October 2021
(This article belongs to the Special Issue Characterization Analysis of Heterogeneous Catalysts)

Abstract

:
In this paper, we report an effective ultrasound method for the synthesis of Pd/H-MOR, which was used as a catalyst in the Suzuki-Miyaura coupling of aryl halides with phenylboronic acid. The structure and morphology of the as-prepared catalysts were fully characterized by X-ray diffraction (XRD), N2 sorption isotherms, scanning electron microscopy (SEM), and an inductively coupled plasma-atomic emission spectrometer (ICP-AES). The advantages of Pd/H-MOR in the coupling reaction are green solvents, high yields, absence of ligands, and recyclability. The catalysts were easily reused at least ten times without significant deterioration in catalytic activity. In addition, this protocol was used in the marketed anti-tumor drug crizotinib synthesis.

Graphical Abstract

1. Introduction

Zeolites are microporous materials consisting of a crystal structure with molecular sieving characteristics [1,2,3]. Zeolites have a wide range of applications when used as carriers to support metals for cross-coupling reactions, such as C-C coupling [4,5,6,7], C-N coupling [8,9,10], C-S coupling [11], and C-H activation [12]. Mordenite zeolite (MOR), which is composed of 12-membered ring (12-MR) channels with 8-MR channels (side pockets), is one of the most vital zeolites with many industrial applications in the catalytic processes of toluene disproportionation, hydro-isomerization, and the method of producing ethanol from non-grain feedstocks [13]. However, few pieces of literature have been published on the use of metals supported on MOR zeolite as catalysts for cross-coupling reactions. Ultrasound’s unique acoustic cavitation effect can provide transient high-temperature and high-pressure extreme reaction conditions and induce weak reducing agents, such as ethylene glycol or even water, to produce highly reactive free radicals so that the reduction reaction can be completed under milder conditions [14,15,16]. At the same time, the shear oscillation generated by ultrasonic waves can effectively reduce the agglomeration of metal nanoparticles [17,18]. Suzuki coupling reactions have been reported as an efficient and less toxic method to form Csp2–Csp2 bonds [19,20]. As a large number of natural products, polymers, and pharmaceuticals have biaryl segments, much effort has been spent on the development of practical conditions for performing Suzuki couplings [21]. A typical strategy in developing heterogeneous catalysts for Suzuki coupling is to deposit Pd species onto various solid supports. Nevertheless, the problems of high reaction temperature, long reaction time [16], and an environmentally friendly solvent [4] still exist.
In this study, the preparation, characterization, and application of Pd nanoparticles/H-MOR (Pd/H-MOR) by ultrasonic treatment are reported (Figure 1). Due to the high active surface area, the contact between the reactants and the catalyst is significantly increased. Therefore, the advantages of Pd/H-MOR in the coupling reaction are green solvents, high yields, absence of ligands, and recyclability.

2. Results and Discussion

These as-prepared catalysts were characterized by X-ray diffraction (XRD), N2 sorption isotherms, scanning electron microscopy (SEM), and an inductively coupled plasma-atomic emission spectrometer (ICP-AES). The XRD pattern in Figure 2A shows that Pd/H-MOR has a slightly weak diffraction peak intensity as compared to H-MOR, suggesting that the crystallinity of the zeolite is marginally reduced after the loading of metal. In addition, the diffraction peak at 2θ of 40.22° proves the palladium nanoparticles decorated on zeolite [22]. The peaks at 2θ of 90.88°, 19.78°, 25.86°, and 31.12° are attributed to the H-MOR support, confirming its structure without any changes [23]. The N2 sorption isotherms of H-MOR and Pd/H-MOR are shown in Figure 2B, which prove that Pd/H-MOR is a micro/mesoporous catalyst. A steep increase at the relative pressure P/P0 < 0.003 signifies the presence of micropores, and the hysteresis loop at the pressure P/P0 of 0.4~1 indicates slit mesopores exist in Pd/H-MOR. Pd/H-MOR has a lower low-pressure adsorption capacity due to the loaded Pd nanoparticles. The textural parameters of the supports and catalysts in Table S1 show that Pd/H-MOR has a BET surface area of 412 cm2/g and a microporous volume of 0.19 cm3/g, lower than those of support (464 cm2/g, 0.21 cm3/g). This might be caused by the blockage of the MOR zeolite micropore channel. Figure 2C,D show the SEM images of Pd/H-MOR and H-MOR, respectively. Pd/H-MOR is an irregular grain particle with the size of 4–7 μm. There is no obvious trace of Pd nanoparticle distribution on the surface of the catalyst, meaning Pd is present in the catalyst at the nanometer scale. The chemical composition of the Pd/H-MOR was determined by CP-AES, confirming that the content of Pd nanoparticles on the H-MOR is 3.17 wt%, and the Si/Al ratio in the catalyst is around 11.3.
As shown in Table 1, Pd/H-MOR was used as the catalyst for the Suzuki cross-coupling reaction of phenylboronic acid 1a with 4-iodoanisole 2a. First of all, we tested the catalytic activity of Pd/H-MOR, which was an efficient heterogeneous palladium catalyst for the Suzuki coupling reaction (Table 1, entry 1). For a comparison with the homogeneous PdCl2 catalyst, a slightly lower yield of 3a was obtained. The result means that Pd/H-MOR exhibits higher catalytic activity for the Suzuki coupling reaction (Table 1, entry 2). MOR and H-MOR were not able to catalyze the reaction (Table 1, entries 3, 4). Control experiments showed that the palladium catalyst was essential for this diphenyl formation reaction. Following from this, an effect of the base was observed and lower yields of 3a were obtained with NaHCO3. In contrast, the application of other stronger bases, such as K2CO3, Na2CO3, and Cs2CO3, were more favorable in the formation of Suzuki product 3a (Table 1, entries 1, 5–8). The effect of the ratio of solvent was explored, and the results showed that the highest yield of the desired product 3a (95%) was obtained in H2O:EtOH = 4:1 (Table 1, entries 1, 9–12). The effect of the temperature was then examined, and the result showed that 80 °C was the optimal temperature (Table 1, entries 12–15). Finally, the effect of time on the reaction was studied, and the results showed that the reaction could be completed in 0.5 h (Table 1, entries 12, 16,17). The performance of Pd/H-MOR was also compared with that of some palladium heterogeneous catalysts reported in the literature. As shown in Table 2, the Pd/H-MOR exhibited a superior catalytic activity at a relatively mild temperature and green solvent, as well as a faster conversion.
Encouraged by the above results, the heterogeneous Suzuki coupling reactions of various aryl halides with phenylboronic acid were then examined under optimized reaction conditions. As shown in Table 3, Suzuki coupling products were obtained in moderate to excellent yields, showing that this catalytic system was compatible with various substrates. Phenylboronic acid 1a was used as the substrate to test the Suzuki coupling effects of different aryl iodides. The coupling reactions of phenylboronic acid with both electron-deficient and electron-rich substituted iodobenzenes using the catalyst proceeded in high yields (Table 3, entries 1–13). For a comparison with meta-substituted or ortho-substituted substrates, when para-substituted aryl halides, including electron-withdrawing and electron-donating groups, were used as a reactant, more than 95% yield was obtained. (Table 3, entries 2, 4–7, 10–13). These results showed that m-substituted substrates had little effect on the coupling reaction under optimized reaction conditions, and o-substituted substrates were only slightly affected. The scope of the reaction was then extended to less reactive aryl bromide and chloride derivatives. The catalyst also easily catalyzes the Suzuki coupling reaction of aryl bromides with phenylboronic acid, resulting in high yields (Table 3, entries 14–16). However, Pd/H-MOR did not exhibit an excellent catalytic performance when aryl chlorides were used as the substrate, even if the reaction times were extended to 8 h (Table 3, entries 17–20). The coupling reactions using aryl boronic acids were also investigated, and the coupling products were obtained in good yields, while the electron-releasing groups in aryl boronic acid gave higher yields compared to substrates bearing electron-withdrawing groups (Table 3, entries 21–24). The Suzuki coupling reaction of heteroaryl bromide and phenylboronic acid was effective in optimized reaction conditions, producing the corresponding products (Table 3, entries 25,26).
To further showcase the abilities of the catalyst system on the synthetic utility for medicinal purposes, we turned our attention to the preparation of crizotinib, which is a tyrosine kinase inhibitor used to treat anaplastic lymphoma kinase-positive lung cancer (Scheme 1) [25]. The synthetic method was used as the catalyst for the Suzuki cross-coupling reaction of aryl bromide 4 with pinacol boronate 5. This transformation was accomplished with excellent results in the presence of the Pd/H-MOR catalyst and K2CO3 in water and ethanol at 80 °C for 6 h. Then, the intermediate was treated with 4 M HCl in 1,4-dioxane/CH2Cl2, and the crizotinib API was isolated in a 90% yield with 97% purity and <10 ppm Pd.
A significant practical application of a heterogeneous catalysis is to be able to recycle the catalyst from the reaction mixture and reuse it for subsequent reactions. Thus, the ability to recycle the Pd/H-MOR catalyst was studied for the reaction of phenylboronic acid 1 with 4-iodoanisole 2a. After the completion of the reaction, the catalyst was separated by centrifugation after each round and reused in the next catalyst reaction. As shown in Figure 3, with increasing recycle times, the catalytic activities of Pd/H-MOR catalysts were decreased slightly. The yield of the first run was 95%, and after the tenth run, its yield fell to 90%, which indicated the steadiness of the Pd/H-MOR catalyst.

3. Experimental Materials

The starting materials were commercially available and, except for the solvents, were used without further purification. H-MOR zeolite was provided by Nankai University Catalyst Co., Ltd. (Tianjin, China). Palladium (II) chloride (PdCl2, 59.5%) was provided by J&K Scientific Ltd. (Shanghai, China). Ethanol was provided by Anhui Ante Food Co., Ltd. (Suzhou, Anhui, China). Potassium hydroxide was provided by Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Other materials were of an analytical grade and used as received.

3.1. Characterization

X-ray diffraction (XRD) measurements were performed at room temperature by using a Rigaku Ultimate VI X-ray diffractometer (40 kV, 40 mA) (Tokyo, Japan) using CuKα (λ = 1.5406 Å). The N2 sorption isotherms at the temperature of liquid nitrogen were measured using Micromeritics ASAP 2020 (Norcross, GA, USA). The Pd loading of the sample was determined by an inductively coupled plasma-atomic emission spectrometer (ICP-AES) with a Perkin–Elmer 3300DV emission spectrometer (Norwalk, CA, USA). Scanning electron microscopy (SEM) experiments were performed on a Hitachi SU-1510 electron microscope (Tokyo, Japan). 1H and 13C NMR spectra were measured with a Bruker Advance 400 spectrometer (New York, NY, USA) by using CDCl3 or DMSO-d6 as solvents and TMS as the internal standard.

3.2. Preparation of Pd/H-MOR Catalyst

We added 0.5 g of H-MOR zeolite to 50 mL of ethanol in a 100 mL round-bottom flask. This mixture was infiltrated and dispersed by an ultrasonic bath for 30 min. After this period of ultrasonic treatment, PdCl2 (0.075 g) was added to the solution and dispersed by ultrasonic for 20 min. Then, KOH (0.048 g with 10 mL H2O) was added to this solution by a constant pressure dropping funnel for 10 min. This mixture was dispersed by an ultrasonic bath for 1 h. After ultrasonic treatment, the product was centrifuged, washed, and kept constant at 100 °C overnight.

3.3. General Procedure for the Suzuki Coupling Reactions

Aryl halides (1 mmol), aryl boronic acid (1.3 mmol), K2CO3 (2 mmol), Pd/H-MOR (10 mg), 8 mL H2O, and 2 mL ethanol were placed in a reaction flask and stirred at 80 °C under air. After the reaction was complete, it was cooled to room temperature. The mixture was then extracted with ethyl acetate, dried, filtered, and concentrated in vacuo. The crude product was purified by column chromatography on silica gel, using pure petroleum to afford the product.

3.4. General Procedure for Catalyst Recovery

The 4-iodoanisole (2.5 mmol), phenylboronic acid (3.25 mmol), K2CO3 (5 mmol), and Pd/H-MOR (50 mg) were mixed in H2O and ethanol (25 mL, 4:1). The mixture was stirred at 80 °C under air. After the reaction was complete, the catalyst was centrifugated, washed, vacuum dried, and used in the next run. The ingredients involved in the cyclic reaction are based on the Pd/H-MOR obtained by separation and are fed in proportion.

4. Conclusions

In summary, a simple and green sonochemical method was applied to palladium nanoparticles deposited on MOR zeolite with ultrasonic treatment. Pd/H-MOR (3.17 wt% of Pd) was an efficient heterogeneous catalyst for Suzuki coupling of aryl halides and phenylboronic acid. Reaction conditions, including the base, temperature, time, catalyst type, solvent ratio, and substrate scope, were optimized. Furthermore, the catalyst was recycled and reused up to 10 times. This work introduces the advantages of green solvents, short reaction times, high yields, absence of ligands, and reusability of the catalyst. Moreover, this catalyst easily synthesized the marketed drug crizotinib (anti-tumor). This work expands the application’s scope of inorganic material MOR as the catalyst carrier as well as promoting the development of green chemistry.

Supplementary Materials

The following are available online at www.mdpi.com/article/10.3390/catal11101213/s1, Table S1: Textural parameters of the synthesized sample. 1H and 13C NMR patterns of products.

Author Contributions

Conceptualization, C.S. and J.J.; investigation, E.Z. and L.Z.; supervision, K.Z. and H.X.; writing—original draft, E.Z. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support from Zhejiang Provincial Natural Science Foundation of China (No. LGF19B060002, LY17B020005) and the National Natural Science Foundation of China (No.21302171). This work is also supported by the Zhejiang Shuren University Basic Scientific Research Special Funds (2020XZ011).

Acknowledgments

Many thanks to all co-authors, and Huimin Luan from Zhejiang University for catalyst characterization. And love from my family.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sen, T.; Jana, S.; Koner, S.; Patra, A. Efficient energy transfer between confined dye and Y-zeolite functionalized Au nanoparticles. J. Phys. Chem. C 2010, 114, 19667–19672. [Google Scholar] [CrossRef]
  2. Kumari, S.; Das, B.; Ray, S. An insight into the catalytic activity of palladium schiff-base complexes towards the Heck coupling reaction: Routes via encapsulation in zeolite-Y. Dalton Trans. 2019, 48, 15942–15954. [Google Scholar] [CrossRef] [PubMed]
  3. Xu, H.; Zhu, J.; Qiao, J.; Yu, X.; Sun, N.; Bian, C.; Li, J.; Zhu, L. Solvent-free synthesis of Aluminosilicate SSZ-39 zeolite. Micropor. Mesopor. Mat. 2021, 312, 110736. [Google Scholar] [CrossRef]
  4. Cui, T.; Ke, W.; Zhang, W.; Wang, H.; Li, X.; Chen, J. Encapsulating palladium nanoparticles inside mesoporous MFI zeolite nanocrystals for shape-selective catalysis. Angew. Chem. Int. Ed. 2016, 55, 9178–9182. [Google Scholar] [CrossRef] [PubMed]
  5. Miao, J.; Huang, B.; Liu, H.; Cai, M. A phosphine-free, atom-efficient cross-coupling reaction of triorganoindiums with acyl chlorides catalyzed by immobilization of palladium (0) in MCM-41. RSC Adv. 2017, 7, 42570–42578. [Google Scholar] [CrossRef] [Green Version]
  6. Wang, Y.; Liao, J.; Xie, Z.; Zhang, K.; Wu, Y.; Zuo, P.; Zhang, W.; Li, J.; Gao, Z. Zeolite-enhanced sustainable Pd-catalyzed C-C cross-coupling reaction: Controlled release and capture of palladium. ACS Appl. Mater. Interfaces 2020, 12, 11419–11427. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, H.; Fu, W.; Zhang, L.; Shen, R.; Cai, G.; Chen, Q.; Tang, T. Basic Cu/ETS-10-N catalyst induced the chemisorption and alkyl radical formation of the substrates enhanced the cinnamic acid decarboxylative cross-coupling reaction activity. Ind. Eng. Chem. Res. 2019, 58, 7014–7024. [Google Scholar] [CrossRef]
  8. Garniera, T.; Saklya, R.; Danela, M.; Chassaing, S.; Pale, P. Chan-Lam-Type C-N cross-coupling reactions under base- and ligand-free CuI-zeolite catalysis. Synthesis 2017, 49, 1223–1230. [Google Scholar]
  9. Moglie, Y.; Buxaderas, E.; Mancini, A.; Alonso, F.; Radivoy, G. Amide bond formation catalyzed by recyclable copper nanoparticles supported on zeolite Y under mild conditions. ChemCatChem 2019, 11, 1487–1494. [Google Scholar] [CrossRef] [Green Version]
  10. Jin, Y.; Cao, Y.; Fang, G.; Ruan, F.; Ke, Q. Synergistic effect between cerium species and micropores in CeAlPO-5 molecular sieves. ChemCatChem 2019, 11, 3178–3181. [Google Scholar] [CrossRef]
  11. Mitrofanov, A.Y.; Murashkina, A.V.; Martín-García, I.; Alonso, F.; Beletskaya, I.P. Formation of C-C, C-S and C-N bonds catalysed by supported copper nanoparticles. Catal. Sci. Technol. 2017, 7, 4401–4412. [Google Scholar] [CrossRef] [Green Version]
  12. Vercammen, J.; Bocus, M.; Neale, S.; Bugaev, A.; Tomkins, P.; Hajek, J.; Van Minnebruggen, S.; Soldatov, A.; Krajnc, A.; Mali, G.; et al. Shape-selective C-H activation of aromatics to biarylic compounds using molecular palladium in zeolites. Nat. Catal. 2020, 3, 1002–1009. [Google Scholar] [CrossRef]
  13. Li, L.; Wang, Q.; Liu, H.; Sun, T.; Fan, D.; Yang, M.; Tian, P.; Liu, Z. Preparation of spherical mordenite zeolite assemblies with excellent catalytic performance for dimethyl ether carbonylation. ACS Appl. Mater. Interfaces 2018, 10, 32239–32246. [Google Scholar] [CrossRef] [PubMed]
  14. Thokchom, B.; Qiu, P.; Cui, M.; Park, B.; Pandit, A.B.; Khim, J. Magnetic Pd@Fe3O4 composite nanostructure as recoverable catalyst for sonoelectrohybrid degradation of ibuprofen. Ultrason. Sonochem. 2017, 34, 262–272. [Google Scholar] [CrossRef] [PubMed]
  15. Padilla, R.H.; Priecel, P.; Lin, M.; Lopez-Sanchez, J.A.; Zhong, Z. A versatile sonication-assisted deposition-reduction method for preparing supported metal catalysts for catalytic applications. Ultrason. Sonochem. 2017, 35, 631–639. [Google Scholar] [CrossRef] [PubMed]
  16. Tadjarodi, A.; Dehghani, M.; Imani, M. Green Synthesis and characterization of palladium nanoparticles supported on zeolite Y by sonochemical method, powerful and efficient catalyst for Suzuki-Miyaura coupling of aryl halides with phenylboronic acid. Appl. Organomet. Chem. 2018, 32, e4594. [Google Scholar] [CrossRef]
  17. Ezoddin, M.; Majidi, B.; Abdi, K. Ultrasound-assisted supramolecular dispersive liquid-liquid microextraction based on solidification of floating organic drops for preconcentration of palladium in water and road dust samples. J. Mol. Liq. 2015, 209, 515–519. [Google Scholar] [CrossRef]
  18. Jin, Q.; Lin, C.; Chang, Y.; Yang, C.; Yeh, C. Roles of textural and surface properties of nanoparticles in ultrasound-responsive systems. Langmuir 2018, 34, 1256–1265. [Google Scholar] [CrossRef] [PubMed]
  19. Miyaura, N.; Yamada, K.; Suzuki, A. A new stereospecific cross-coupling by the palladium-catalyzed reaction of 1-alkenylboranes with 1-alkenyl or 1-alkynyl halides. Tetrahedron Lett. 1979, 20, 3437–3440. [Google Scholar] [CrossRef] [Green Version]
  20. Miyaura, N.; Yamada, K.; Suginome, H.; Suzuki, A. Novel and convenient method for the stereo- and regiospecific synthesis of conjugated alkadienes and alkenynes via the palladium-catalyzed cross-coupling reaction of 1-alkenylboranes with bromoalkenes and bromoalkynes. J. Am. Chem. Soc. 1985, 107, 972–980. [Google Scholar] [CrossRef]
  21. Ke, W.; Cui, T.; Yu, Q.; Wang, M.; Lv, L.; Wang, H.; Jiang, Z.; Li, X.; Chen, J. Mesoporous H-ZSM-5 nanocrystals with programmable number of acid sites as “Solid Ligands” to activate Pd nanoparticles for C-C coupling reactions. Nano Res. 2017, 11, 874–881. [Google Scholar] [CrossRef]
  22. Dehghani, M.; Tadjarodi, A.; Chamani, S. Synthesis and characterization of magnetic zeolite Y-palladium-nickel ferrite by ultrasonic irradiation and investigating its catalytic activity in Suzuki-Miyaura cross-coupling reactions. ACS Omega. 2019, 4, 10640–10648. [Google Scholar] [CrossRef] [PubMed]
  23. Xu, R.; Pang, W. Chemistry-Zeolites and Porous Materials; Science Press: Beijing, China, 2004; p. 44. [Google Scholar]
  24. Shao, Y.; Zeng, H. Nanowire networks of metal-organosilicates as reversible Pd(II) reservoirs for Suzuki coupling reactions. ACS Appl. Nano Mater. 2021. [Google Scholar] [CrossRef]
  25. Magano, J.; Dunetz, J.R. Large-Scale applications of transition metal-catalyzed couplings for the synthesis of pharmaceuticals. Chem. Rev. 2011, 111, 2177–2250. [Google Scholar] [CrossRef] [PubMed]
Figure 1. General procedure for the synthesis of Pd/H-MOR catalyst.
Figure 1. General procedure for the synthesis of Pd/H-MOR catalyst.
Catalysts 11 01213 g001
Figure 2. (A) XRD patterns of Pd/H-MOR (a) and H-MOR (b); (B) N2 sorption isotherm of Pd/H-MOR (a) and H-MOR (b); SEM images (scale bar 10 μm) of Pd/H-MOR (C) and H-MOR (D).
Figure 2. (A) XRD patterns of Pd/H-MOR (a) and H-MOR (b); (B) N2 sorption isotherm of Pd/H-MOR (a) and H-MOR (b); SEM images (scale bar 10 μm) of Pd/H-MOR (C) and H-MOR (D).
Catalysts 11 01213 g002
Scheme 1. Application of the catalyst in the synthesis of crizotinib.
Scheme 1. Application of the catalyst in the synthesis of crizotinib.
Catalysts 11 01213 sch001
Figure 3. Recycling of Pd/H-MOR in the Suzuki coupling.
Figure 3. Recycling of Pd/H-MOR in the Suzuki coupling.
Catalysts 11 01213 g003
Table 1. Optimization of reaction conditions a.
Table 1. Optimization of reaction conditions a.
Catalysts 11 01213 i001
EntryCatalystBaseSolventTemp. (°C)Yield b (%)
1Pd/H-MORK2CO3EtOH8092
2PdCl2K2CO3EtOH8094 c
3MORK2CO3EtOH80NR
4H-MORK2CO3EtOH80NR
5Pd/H-MORNa2CO3EtOH8092
6Pd/H-MORNaHCO3EtOH8027
7Pd/H-MORKOHEtOH8090
8Pd/H-MORCsCO3EtOH8092
9Pd/H-MORK2CO3H2O8078
10Pd/H-MORK2CO3H2O:EtOH (1:1)8093
11Pd/H-MORK2CO3H2O:EtOH (3:2)8094
12Pd/H-MORK2CO3H2O:EtOH (4:1)8095
13Pd/H-MORK2CO3H2O:EtOH (4:1)60NR
14Pd/H-MORK2CO3H2O:EtOH (4:1)70NR
15Pd/H-MORK2CO3H2O:EtOH (4:1)9095
16Pd/H-MORK2CO3H2O:EtOH (4:1)8095 d
17Pd/H-MORK2CO3H2O:EtOH (4:1)8095 e
a The reaction conditions: phenylboronic acid 1 (1.3 mmol), 4-iodoanisole 2a (1 mmol), 10 mg catalysts, and base (2 mmol) in 10 mL solvent were reacted for 2 h under air. b Isolated yield. c 0.5 mg PdCl2. b Isolated yield. d 0.5 h. e 1 h.
Table 2. Comparison of Pd/H-MOR with some recently reported Pd catalysts for the reaction of Suzuki coupling.
Table 2. Comparison of Pd/H-MOR with some recently reported Pd catalysts for the reaction of Suzuki coupling.
Catalysts 11 01213 i002
NumberRXArCatalystReaction ConditionsYield (%)Ref.
1Hp-IC6H4-CH3Pd/H-MORH2O:EtOH(4:1)/K2CO3/80 °C/0.5 h98This work
2Hp-IC6H4-OCH3Pd@mnc-S1MeOH/K2CO3/80 °C/0.5 h93[4]
3Hp-IC6H4-CH3Pd-OSEtOH/K2CO3/80 °C/0.5 h93[24]
4Hp-IPhZ-Y-Pd NPsH2O:EtOH(1:1)/K2CO3/Reflux/1.5 h88[16]
Table 3. Suzuki coupling between phenylboronic acid and aryl halides a.
Table 3. Suzuki coupling between phenylboronic acid and aryl halides a.
Catalysts 11 01213 i003
EntryRXArYield b (%)
1Hp-IC6H4-OCH395 (3a)
2Hp-IC6H4-NH295 (3b)
3Hp-IC6H4-OH96 (3c)
4Hp-IC6H4-CH398 (3d)
5Hp-IC6H4-NO298 (3e)
6Hp-IC6H4-CHO97 (3f)
7Hp-IC6H4-COCH395 (3g)
8Hp-IC6H4-Cl95 (3h)
9HIPh95 (3i)
10Hm-IC6H4-NO294 (3j)
11Hm-IC6H4-COCH393 (3k)
12Ho-IC6H4-NH286 (3l)
13Ho-IC6H4-CH384 (3m)
14Hp-BrC6H4-CH379 (3d)
15Hp-BrC6H4-CHO92 (3f)
16HBrPh92 (3i)
17Hp-ClC6H4-NH248 c (3b)
18Hp-ClC6H4-CHO50 c (3f)
19Hp-ClC6H4-COCH352 c (3g)
20Hp-ClPh51 c (3i)
214-CH3Op-IC6H4-OCH396 (3n)
224-CH3p-IC6H4-OCH395 (3o)
234-Fp-IC6H4-OCH392 (3p)
243-NO2p-IC6H4-OCH386 (3q)
25Ho-Brpyridine83 (3r)
26Ho-Brquinoline81 (3s)
a The reaction conditions: boronic acids 1 (1.3 mmol), aryl halides 2 (1 mmol), Pd/H-MOR (10 mg), and K2CO3 (2 mmol) in 10 mL solvent (H2O:EtOH = 4:1) were reacted for 0.5 h under air. b Isolated yield. c 8 h.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhou, E.; Jin, J.; Zheng, K.; Zhang, L.; Xu, H.; Shen, C. Novel Recyclable Pd/H-MOR Catalyst for Suzuki-Miyaura Coupling and Application in the Synthesis of Crizotinib. Catalysts 2021, 11, 1213. https://doi.org/10.3390/catal11101213

AMA Style

Zhou E, Jin J, Zheng K, Zhang L, Xu H, Shen C. Novel Recyclable Pd/H-MOR Catalyst for Suzuki-Miyaura Coupling and Application in the Synthesis of Crizotinib. Catalysts. 2021; 11(10):1213. https://doi.org/10.3390/catal11101213

Chicago/Turabian Style

Zhou, Enmu, Jianzhong Jin, Kai Zheng, Letian Zhang, Hao Xu, and Chao Shen. 2021. "Novel Recyclable Pd/H-MOR Catalyst for Suzuki-Miyaura Coupling and Application in the Synthesis of Crizotinib" Catalysts 11, no. 10: 1213. https://doi.org/10.3390/catal11101213

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop