Next Article in Journal
Efficacy of the Immobilized Kocuria flava Lipase on Fe3O4/Cellulose Nanocomposite for Biodiesel Production from Cooking Oil Wastes
Previous Article in Journal
Auramine O UV Photocatalytic Degradation on TiO2 Nanoparticles in a Heterogeneous Aqueous Solution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Power and High-Performance Catalyst for Suzuki Coupling Reaction

by
Mansour Binandeh
*,
Mohammad Ali Nasseri
and
Ali Allahresani
Basic of Science, Chemistry Department, University of Birjand, Birjand 9717434765, Iran
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(9), 976; https://doi.org/10.3390/catal12090976
Submission received: 19 May 2022 / Revised: 20 June 2022 / Accepted: 28 June 2022 / Published: 31 August 2022
(This article belongs to the Section Catalysis in Organic and Polymer Chemistry)

Abstract

:
In this project, the aim was to carry out the Suzuki reaction using a new and unique catalyst with a base of Fe3O4 magnetic nanoparticles coated using a new ligand. In fact, this magnetic catalyst has an inhomogeneous surface to create a connection between the organic and aqueous phases so that the carbon–carbon coupling reaction is completely performed on its large surface. The structure of this catalyst uses two metals, nickel and cobalt, which are coated on a bed of amino linkers and propel the Suzuki coupling reaction at high speed on the catalyst surface. The products obtained are from ideal and optimal conditions with an efficiency of over 98%. The catalyst has a recovery power of over 96% and has enough power to perform several coupling reactions several times. Lastly, the magnetic nanocatalyst is easily separated from the reaction medium by an external field and has 100% power when performing other reactions.

1. Introduction

For many years, in the field of nanotechnology, special attention has been paid to nanoparticles and its applications, among which magnetic nanoparticles have many practical advantages. Their several advantages include a large catalytic surface to create a heterogeneous surface to accelerate a variety of chemical reactions, easy synthesis of inexpensive materials such as iron salts, and easy separation from the reaction solution by an external magnetic field with high loading of various ligands (organic and metal) [1].
These nanoparticles have many applications in the chemical and medical industries as nanocatalysts in the reactions of atoms pairing with each other, as well as for separating and purifying medical enzymes and biomolecules in vitro. One of the reactions that has been considered by researchers for many years is the carbon–carbon coupling reaction, which is obtained from the reaction of two aryl halide reactants with phenyl boronic acid, following Akira Suzuki’s efforts to expand this coupling for which he won the Nobel Prize in 2010 [2,3,4,5,6].
The importance of this type of coupling both in the chemical industry and in the medical industry is due to the synthesis of important and vital structures that allow the production of useful and biologically active compounds for the production of drug biomolecules, attracting the attention of many scientists [7,8,9,10,11,12].
To evaluate the performance of magnetic nanocatalysts, special attention should be paid to the surface of nanoparticles because their surface may be slightly oxidized in the presence of air, and their efficiency may be reduced. For this purpose, the surface can be covered with insulating and very useful materials such as silica (SiO2) to prevent excess oxidation of their surface and to prevent clumping or loss of their reactivity surfaces, which is also a biodegradation advantage. Coating also confers plasticity to nanoparticles in targeted drug transfers so that they can be easily separated from the reaction medium by an external field without causing toxicity at the transport site. Applications of targeted drug delivery, adsorption, and release, as well as the antibacterial properties of magnetic nanocatalysts, have been presented in previous reports [13,14,15,16,17].
In the Suzuki reaction, carbon–carbon coupling has been used to create di-aryl products that contain very important ligands with many applications in the medical industry, such as creating pharmaceutical products with anti-inflammatory, antiplatelet, and antiviral properties [18,19,20,21].
Today, all efforts to create structures with amino groups and oxygen positioned in catalytic structures are made using ligands to create a suitable space for loading a variety of intermediate metals (which have the catalytic power to accelerate the Suzuki coupling reaction on the magnetic nanoparticle substrate). The existence of such functional groups has greatly helped to accelerate the Suzuki coupling reaction with the loaded metal to synthesize the resulting compounds with a high percentage of efficiency in the development of drugs or biological compounds [22,23].
In this project, the main goal was to introduce a potential catalyst of the Suzuki reaction with much better reaction conditions than other catalysts and a much higher percentage of products obtained by the material. This nanocatalyst is made of magnetic nanoparticles that have magnetic properties (controlled by an external field), very good particle size, and a very large catalyst surface to advance the carbon–carbon coupling reaction. Nanoparticles ideally have a size of about 100 nm, and the efficiency of the products obtained at a temperature of 70 °C is more than 98%. The bulk ligand tested on a magnetic nanoparticle substrate by IR, NMR, and EDX analysis is a new ligand used in the construction of catalysts. Due to the optimal conditions and higher efficiency, this nanocatalyst has a better performance than other catalysts, with an estimated 25% higher power. An overview of the Suzuki reaction is shown in this project, which, in addition to showing the conversion of reagents into a product (coupling), provides 1H-NMR analysis to identify and confirm the resulting product along with the reaction conditions (Scheme 1).

2. Results and Discussion

2.1. General Overview of Nanocatalysts

The aim of this project was to design a unique catalyst with very successful performance in most possible reactions that is able to successfully undergo IR analyses after synthesis. The main structure of the catalyst is based on a shell/core structure with a core of magnetite nanoparticles and a shell of a bulky ligand that is eventually attached to the silicate structure. The important thing about these magnetic nanoparticles is the coordination of the particles that make up their structure, in which are in the range of 0 to 100 nm, typically less than 20 nm in medical applications (as widely used in previous projects). Another important point to consider is the very large catalytic surface suitable for a variety of chemical and even biological reactions, as well as their unique property of being easily separated from the reaction medium by an external magnetic field.

2.2. Analysis Tools of Detection

2.2.1. SEM and TEM Analyses

Synthetic nanocatalysts should be identified using a series of devices including SEM and TEM analyses to determine nanoparticle morphology and nanoparticle diameter, as well as produce images of their inner and outer surface for evidence of the surface crystalline order. These analyses are performed within the dimensions of 50–500 nm, and the shape information of the nanoparticles is shown schematically, allowing the presence of the coated metal and ligands on the surface of the nanoparticles to be identified (Figure 1).

2.2.2. FTIR Analysis

FTIR analysis is important for identifying the functional groups in a compound, revealing the presence of functional groups and their links. This analysis is typically reported between 400 and 4000 cm−1. In the nanocatalyst synthesized in this project, the iron–oxygen bond in the range of 500 cm−1 revealed the core structure, i.e., magnetic nanoparticle, presenting a very large peak that gradually became more regular and sharper with the coating of the nanocatalyst and ligand. The loose structure or shell can subsequently be identified, consisting of a silicate in the form of a silyl–oxygen–silyl bond with a strong peak around 1000 cm−1. The presence of the silica coating prevents excessive oxidation of the nanocatalyst surface and grants it good biocompatibility and biodegradability. The C20H19N3O ligand to be placed on the silica coating, resulted in a bulge next to the silicate peak, thus generating a bifurcation. Peaks related to nitrogen were located in the area of oxygen peaks, but could also be independently identified around 3300 cm−1. Hydrogen–carbon bonds of ring structures generated a peak around 3100 cm−1. If we compare the spectra, we can come to the conclusion that the peaks became more regular and the structure of the nanocatalyst became more complete, when considering the presence of carbon–hydrogen, carbon–carbon, and free amine bonds (Figure 2).

2.2.3. EDX Analysis

EDX analysis allows characterizing the nanoparticles as a function of the percentage of elements contained. This is also a valuable tool for identifying links between elements, whereby overlapping peaks indicate that two elements are related, regardless of their size. Indistinguishable peaks indicate the presence of a covalent or van der Waals bond (such as a hydrogen or electrostatic bond). Results are expressed as the percentage of an element in units of keV (Figure 3).

2.2.4. The XRD spectrum of Fe3O4@L/Co/Ni

X-ray diffraction allows analyzing the number of electron transmissions using X-rays in terms of the energy levels in orbital layers and their electron capacities. Fe3O4 presented peaks at 2θ = 30.1°, 35.4°, 43.2°, 53.7°, 56.9°, and 62.9° with bands visible at 220, 311, 422, 511, and 440 cm−1 (Figure 3), indicating that Fe3O4 was successfully synthesized without an effect on its crystalline structure. However, two new peaks were observed at 2θ = 44° and 56°, corresponding to bands at 111 and 200 cm−1, which confirmed the complexation of Co to the surface of the Fe@L/Co/Ni magnetic nanoparticles with a silica substrate. Furthermore, peaks at 2θ = 44° and 56° for Co and 34° and 57° for Ni, according to the Scherrer equation, suggested that the crystalline nanoparticles of Co/Ni (0) were about 9 nm in size (Figure 3).

2.2.5. VSM Analysis

VSM is used to measure the amount of magnetometer saturation according to the type of magnetic nanoparticle (para-, dia-, or ferromagnetic), as well as the amount of magnetism in the magnetic field. Therefore, for this purpose, analysis was performed in terms of emu/g in the units of the magnetic field (kOe). The magnetic nanoparticles (Fe3O4) presented results of 57 emu/g at 25–100 °C and pH = 8–12, in contrast to 21 emu/g for silica/(C20H19N3O) groups and 13 emu/g for Co/Ni fixed to magnetic nanoparticles. Lastly, after synthesis of the Fe3O4/L/Co/Ni nanocomposite, the degree of magnetization was 31 emu/g. This decrease was due to the magnetic nanoparticles becoming superparamagnetic, indicating their excellent quality for the Suzuki coupling reaction, with a 27% improvement compared to previous studies (Figure 3).

2.3. General Overview of Suzuki Reaction

In this section, an attempt was made to apply the synthesized nanocatalysts (analyzed by SEM, TEM, FTIR, etc.) in the Suzuki coupling reaction (Scheme 2), i.e., carbon–carbon bonding. This reaction was performed under optimal conditions, i.e., using the minimum amount of catalyst sample, at the right temperature, with adequate levels of salt and cheap base, resulting in products with an efficiency of over 98%. The purpose of this study was to optimize the reaction conditions and synthesize products with suitable efficiency compared to the literature.

2.4. The Amount of Nanocatalyst

According to previous studies, most nanocatalysts used in this coupling reaction are typically above 0.02 g, which is not ideal for the standardization of a reaction. Thus, in order to increase the efficiency of the catalyst and prevent wasting of the sample, the least amount of catalyst should be used. Therefore, the amount of catalyst used in this reaction was 0.2.% mol, i.e., an extraordinarily low amount of ~0.002 g. As shown in the table, this amount of catalyst was enough to complete the reaction (Table 1). By examining the Suzuki reaction using magnetite nanoparticles (without ligand), as well as with ligand and cobalt/nickel nanoparticles (alone), the efficiency of the obtained products was much lower over time (even with a sample size of 20 mg) compared to when the nanocatalyst was used.

2.5. Solvent Type

Although solvents are not always needed for reactions, some organic reactants only dissolve in organic solvents. However, since the tendency of two reactants to react with each other in different solvents (whether organic or inorganic) should be tested, the best conditions involve the use of a harmless green solvent. Therefore, the role of the solvent is very effective in bringing the two phases closer, providing the exchange of functional groups and the creation of a covalent bond between the two organic reactants. On the other hand, because a strong base is needed to help remove the boronic group from both sides of the reactants in this reaction, a compound with a new covalent bond is formed as a coupling. As shown in Table 1, the solvent used was 1 cc of DMF/H2O for C–C (single bond) with 0.2 mL of KI. Water was selected as a green solvent with optimal results at 70 °C, with a high ability to combine the reactants.

2.6. Salt

Salts are used as a substrate to adjust the pH of the reaction medium and mediate the ion exchange between two reactants. Potassium carbonate salt (0.02 g) was found to be ideal as it did not ionize itself or cause a series of unwanted reactions (Table 1).

2.7. Time

The reaction involving the magnetic nanoparticles, salt, and solvent creates a level of inhomogeneity between the catalyst and the reactants, which decreases with increasing temperature. Although some reactions could proceed at room temperature, a substantial amount of catalyst was required; thus, the ideal temperature for this reaction was 70 °C (Table 1). According Table 1, the nanocatalyst presented a high efficiency of 98% in a reaction time of (0.5–1) h.

2.8. Correct Identification of the Obtained Products

In order to identify the products, a TLC solution was taken, after 5 min, and hexane, ethanol, and ethyl acetate were used as solvents to elute nonpolar and polar products. As the product composition percentage was 100%, the organic product could be easily separated from the aqueous phase by an organic solvent using a decanter funnel (i.e., ethyl acetate). Then, for further studies, a tablet was generated using potassium bromide and analyzed in an IR device. It can be concluded that the proposed nanocatalyst (last row) was ideal compared to other catalysts in the literature (Table 2 and Table 3) [24,25,26,27,28,29,30,31,32,33,34,35,36,37,38].

2.9. Coupling Reaction Process

This coupling reaction process involved a one-step reaction with two metals: one coupled to the boronic group of the phenylboronic compound (nucleophile) and the other coupled to the iodide of the arylphenyl compound (electrophile), thereby yielding a Suzuki carbon–carbon bonding product. This reaction was performed under 1 mmol of potassium iodide, using water as the solvent, at a temperature 70 °C; the final product of this reaction can be used in the medical industry as an organic bio-compound. Scheme 3 provides an overview of the process.

2.10. Comparison of Mechanism, Recovery, and Reuse of Nanocatalysts

A mechanism showing the general process of the reaction is proposed in Scheme 4. In this design, the reaction mechanism is divided into two parts, as explained previously. The magnetic nanocatalyst features two metals, cobalt and nickel, on its catalytic substrate. Both cobalt and nickel acetate have two electron capacities (through oxidation, the two empty orbitals to bond to the two reactants, i.e., a nucleophile or an electrophile). Cobalt/nickel(II) attack the reactants separately before being coupled in the final transformation of two aryl carbon sp2 via a sigma bond (sp3). The metals can be separated from the reaction by reducing their acetate group, and the Suzuki coupling product along with its various derivatives can be separated and purified. This reaction is continuously repeated to synthesize a product with high efficiency and optimal quality. In the previous section, this nanocatalyst was shown to be superior to others in the literature. Furthermore, the current nanocatalyst had almost maximum power after being reused 10 times, showing only a 5% reduction, as revealed by SEM, TEM, and IR analyses. Figure 4 also shows that the structural and functional quality of the catalyst was maintained.

3. Materials and Methods

All reagents used met the necessary standards, and the solvent used was deionized water at 18 MΩ/cm. C20H19N3O2 (molecular weight 333.89 g/mol, 99.999% purity), C4H6NiO4 (molecular weight 176.8 g/mol, 99% purity), C4H6CoO4, (cobalt(II) acetate, molecular weight 177.02124 g/mol, 99% purity), and CPTES (molecular weight 198.72, 97% purity) were purchased from Sigma Aldrich (St. Louis, MO, USA). FeCl2·4H2O, Fe(Cl)3·6H2O, deionized water, argon gas, NaOH (34% aqueous solution), TEOS, HCl, and methanol were purchased from Sinopharm Chemical Reagent Co. (Shanghai, China).
Powder XRD of the prepared catalyst was performed using a Philips PW 1830 X-ray diffractometer with a Cu Kα source (λ = 1.5418 Å) in the Bragg angle range 10–80° at 25 °C. FTIR spectra were obtained using an FTIR spectrometer (Vector 22, Bruker) in the range 400–4000 cm−1 at room temperature. SEM analysis was conducted using a VEGA//TESCAN KYKY-EM 3200 microscope (acceleration voltage of 26 kV). TEM experiments were conducted using a Philips EM 208 electron microscope. EDX analysis of the catalyst was conducted using a VEGA3 XUM/TESCAN. TGA was performed using a Stanton Red Craft STA-780 (London, UK). NMR spectra were obtained using a Bruker DRX-400 instrument (300.1 MHz for 1H-NMR, 75.4 MHz for 13C-NMR). The spectra were obtained using CHCL3-d1 as a solvent. Magnetic measurements were carried out using a VSM instrument (MDK, model 7400). Melting points were evaluated using an Electrothermal 9100 apparatus.

3.1. Nanoparticle Synthesis Method

The chemical coprecipitation method [13,14,15,16,17,18] was used due to its simplicity and the high yield of synthetic products. This study was based on the creation of a heterogeneous catalyst in the aqueous phase, initiating phase transfer of the organic phase (reactors). Magnetic nanoparticles were obtained in the form of iron oxide, which was achieved by combining two iron salts (II) and (III) in a ratio of 1:2.
M 2 + + 8 OH + Fe 3 + MFe 2 O 4 + 4 H 2 O
MFe 2 O 4 M = Fe , L Fe 3 O 4 @ L + Co / Ni Fe 3 O 4 @ L Co / Ni

3.2. Synthesis of Fe3O4 Magnetic Nanoparticles

The synthesis of magnetite magnetic nanoparticles was based on a core/shell nuclear structure consisting of iron oxide composed of iron(II) (0.9 g, 0.2 mol.%) and iron(III) (1.7 g, 0.5 mol.%) salts in 300 cc of water, which was stirred for 2 h. This reaction was scaled from room temperature to 65 °C than 2 h, at which point 20 mol.% of sodium hydroxide was added, and the reaction was continued for another 2 h. continued. The reaction solution was separated from the nanoparticles using an external magnetic field and repeatedly rinsed with ethanol and distilled water several times, before being placed in an oven at 70 °C. After drying completely, the resulting powder (burnt brown color) was collected.

3.3. Nanoparticle Core/Shell Structure with Metal Coating

After the synthesis of the magnetic nanoparticles, the surface was coated with silicate nanoparticles. To do this, 0.25 g of the synthesized catalyst was first weighed in 50 mL of distilled water on a sonication device for 1.5 h. Then, in a separate reaction vessel, 4 cc of tetraorthosilicate was combined with 5 cc of ethanol and added to the previous solution. During this reaction, 5 cc of 10% sodium hydroxide solution was added to the reaction solution, and the sonication step was continued for 2 h. The reaction was further continued for another 2.5 h at room temperature, i.e., 25 °C, thus yielding the core/shell product containing silica nanoparticles. However, in this reaction, magnetite nanoparticles were synthesized first before adding the desired ligand (via electrostatic bonding with the silicate). The reaction is performed sequentially, whereby the first 0.25 g of the core/shell nanocatalyst was dissolved in 50 cc of distilled water, before dissolving another 1 mmol. Then, 25 cc of ethanol was added to the previous solution, and the whole solution is refluxed at 75 °C for 24 h. After 1 day, the product was washed and dried several times with ethanol and distilled water and collected after oven drying at 65 °C. In another step, 0.25 g of the core/shell catalyst was added to 25 cc of distilled water before being combined with 0.002 g of cobalt(II) acetate and nickel(II) acetate nanoparticles dissolved in 25 cc of ethanol. Finally, the whole solution was refluxed together at 75 °C for 1.5 days. The synthesized nanocatalysts were repeatedly washed with ethanol and double-distilled water and dried in an oven at 55 °C. Scheme 5 shows complete reaction process.

4. Conclusions

The main purpose of this project was to design a new nanocatalyst to optimize the carbon–carbon coupling reaction. For this purpose, a magnetic nanocatalyst was used, characterized by external field control properties, a wide catalyst surface for the coupling reaction, easy separation from the reaction medium, a low fabrication cost, and the inclusion of two metals, which accelerated the reaction process. Synthetic catalysts show great promise for application in the chemical and medical industries.

Author Contributions

Conceptualization, methodology, investigation, writing—original draft preparation, writing—review and editing, M.B.; supervision, M.A.N. and A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

This project was carried out in the Birjand University Chemistry Lab, Iran.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Khanmoradi, M.; Nikoorazm, M.; Ghorbani-Choghamarani, A. Synthesis and Characterization of Pd Schiff Base Complex Immobilized onto Functionalized Nanoporous MCM-41 and its Catalytic Efficacy in the Suzuki, Heck and Stille Coupling Reactions. Catal. Lett. 2017, 147, 1114. [Google Scholar] [CrossRef]
  2. Yang, J.; Liu, S.; Zheng, J.F.; Zhou, J. Room-Temperature Suzuki–Miyaura Coupling of Heteroaryl Chlorides and Tosylates. Eur. J. Org. Chem. 2012, 31, 6248. [Google Scholar] [CrossRef]
  3. Tamoradi, T.; Ghorbani-Choghamarani, A.; Ghadermazi, M. Synthesis of a new Pd(0)-complex supported on magnetic nanoparticles and study of its catalytic activity for Suzuki and Stille reactions and synthesis of 2,3-dihydroquinazolin-4(1H)-one derivatives. Polyhedron 2018, 145, 120. [Google Scholar] [CrossRef]
  4. Li, X.H.; Baar, M.; Blechert, S.; Antonietti, M. Facilitating room-temperature Suzuki coupling reaction with light: Mott-Schottky photocatalyst for C-C-coupling. Sci. Rep. 2013, 3, 1743. [Google Scholar] [CrossRef]
  5. Ghorbani-Choghamarani, A.; Naghipour, A.; Heidarizadi, F.; Shirkhani, R. Efficient synthesis of aryl amines from aryl halides in the presence of melamine-supported CuO nanocatalyst in wáter. Inorg. Chim. Acta 2016, 446, 97. [Google Scholar] [CrossRef]
  6. Samarasimharseddy, M.; Prabhu, G.; Vishwanatha, T.M.; Sureshbabu, V.V. Synthesis and characterization of a Pd (0) Schiff base complex anchored on magnetic nanoporous MCM-41 as a novel and recyclable catalyst for the Suzuki and Heck reactions under green conditions. Synthesis 2013, 45, 1201. [Google Scholar]
  7. Hajjami, M.; Tahmasbi, B. Four-component synthesis of polyhydroquinolines under catalyst- and solvent-free conventional heating conditions: Mechanistic studies. RSC Adv. 2015, 5, 59194. [Google Scholar] [CrossRef]
  8. Veisi, H.; Ghadermazi, M.; Naderi, A. Ligand-free Mizoroki–heck reaction using reusable modified graphene oxide-supported Pd(0) nanoparticles. Appl. Organomet. Chem. 2016, 30, 341. [Google Scholar] [CrossRef]
  9. Tobisu, M.; Xu, T.; Shimasaki, T.; Chatani, N. Nickel-catalyzed Suzuki–Miyaura reaction of aryl fluorides. J. Am. Chem. Soc. 2011, 133, 19505. [Google Scholar] [CrossRef]
  10. Li, J.H.; Liang, Y.; Xie, Y.Y. An efficient Stille cross-coupling reaction catalyzed by Pd (OAc)2/DAB-Cy catalytic system. Tetrahedron 2005, 61, 7289. [Google Scholar] [CrossRef]
  11. Molnar, A. Efficient, selective, and recyclable palladium catalysts in carbon-carbon coupling reactions. Chem. Rev. 2011, 111, 2251. [Google Scholar] [CrossRef]
  12. Miyaura, N.; Suzuki, A. Palladium-catalyzed cross-coupling reactions of organoboron compounds. Chem. Rev. 1995, 95, 2457. [Google Scholar] [CrossRef]
  13. Binandeh, M. Frequency of high-performance magnetic nanoparticles of mid-ampicillin, as antibacterial agents. J. Antimicrob. Agents 2018, 4, 162–170. [Google Scholar]
  14. Binandeh, M.; Karimi, F.; Rostamnia, S. Superior performance of magnetic nanoparticles for entrapment and fixation of bovine serum albumin in-vitro. J. Health Sci. 2020, 30, 599–606. [Google Scholar]
  15. Binandeh, M.; Karimi, F.; Rostamnia, S. Application of magnetic nanoparticles by comparing the absorbance and stabilization of biomolecules DNA-C, L by the electrophoretic detection. Int. J. Health Sci. 2021, 15, 3–8. [Google Scholar]
  16. Binandeh, M.; Karimi, F.; Rostamnia, S. Use the best of MNPS-IHSP nanoparticles with coating of ampicillin antibiotic, as bactericidal properties. J. Microb. Biochem. Technol. 2019, 11, 416–423. [Google Scholar]
  17. Binandeh, M.; Karimi, F.; Rostamnia, S. MNPs-IHSPN nanoparticles in multi-application with absorption of bio drugs in vitro. Biochem. Biophys. Rep. 2021, 28, 101159–101167. [Google Scholar] [CrossRef]
  18. Kim, J.; Swager, T.M. Control of conformational and interpolymer effects in conjugated polymers. Nature 2001, 411, 1030. [Google Scholar] [CrossRef]
  19. Fang, Y.-Q.; Karisch, R.; Lautens, M. Efficient syntheses of KDR kinase inhibitors using a Pd-catalyzed tandem C−N/suzuki coupling as the key step. J. Org. Chem. 2007, 72, 1341. [Google Scholar] [CrossRef]
  20. Bringmann, G.; Rüdenauer, S.; Bruhn, T.; Benson, L.; Brun, R. Total synthesis of the antimalarial naphthylisoquinoline alkaloid 5-epi-4′-O-demethylancistrobertsonine C by asymmetric Suzuki cross-coupling. Tetrahedron 2008, 64, 5563. [Google Scholar] [CrossRef]
  21. Hassan, J.; Sevignon, M.; Gozzi, C.; Schulz, E.; Lemaire, M. Aryl−Aryl Bond Formation One Century after the Discovery of the Ullmann Reaction. Chem. Rev. 2002, 102, 1359. [Google Scholar] [CrossRef]
  22. Al Zoubi, W.; Kandil, F.; Chebani, K.; Zoubi, W. Active transport of metal ions by using Schif bases. Phys. Sci. Res. Int. 2014, 2, 12–23. [Google Scholar]
  23. Hassan, A.F.; Elhadidy, H. Effect of Zr4+ doping on characteristics and sonocatalytic activity of TiO2/carbon nanotubes composite catalyst for degradation of chlorpyrifos. J. Phys. Chem. Solids 2019, 129, 180–187. [Google Scholar] [CrossRef]
  24. Rajabi, F.; Thiel, W.R. An Efficient Palladium N-Heterocyclic Carbene Catalyst Allowing the Suzuki–Miyaura Cross-Coupling of Aryl Chlorides and Arylboronic Acids at Room Temperature in Aqueous Solution. Adv. Synth. Catal. 2014, 356, 1873. [Google Scholar] [CrossRef]
  25. Yaşar, S.; Çekirdek, S.; Özdemir, I. New Bisbenzimidazolin-2-Ylidene Salts as N -Heterocyclic Dicarbene Precursors: Synthesis, Characterization, and Involvement in Palladium-Catalyzed Suzuki Reactions. Heteroatom Chem. 2014, 25, 157. [Google Scholar] [CrossRef]
  26. Keleş, M.; Yılmaz, M.K. Synthesis, characterization and catalytic activity of new aminomethyldiphosphine–Pd (II) complexes for Suzuki cross-coupling reaction. Appl. Organometal. Chem. 2014, 28, 91. [Google Scholar] [CrossRef]
  27. Li, B.; Guan, Z.; Wang, W.; Yang, X.; Hu, J.; Tan, B.; Li, T. Magnetic Microporous Organic Polymer of 1,1′-Bi-2-naphthol Supported Palladium for Suzuki Coupling Reactions. Adv. Mater. 2012, 24, 3390. [Google Scholar] [CrossRef]
  28. Targhan, H.; Hassanpour, A.; Sohrabnezhad, S.; Bahrami, K. Palladium Nanoparticles Immobilized with Polymer Containing Nitrogen-Based Ligand: A Highly Efficient Catalyst for Suzuki–Miyaura and Mizoroki–Heck Coupling Reactions. Catal. Lett. 2020, 150, 660. [Google Scholar] [CrossRef]
  29. Zhu, Z.; Liu, J.; Dong, S.; Chen, B.; Wang, Z.; Tang, R.Y.; Li, Z. Copper-Catalyzed Cross-Coupling of Benzylic Bromides with Arylboronic Acids: Synthesis of Diarylalkanes and Preliminary Antifungal Evaluation Against Magnaporthe Grisea. Asian J. Org. Chem. 2020, 9, 631. [Google Scholar] [CrossRef]
  30. Mohammadi, M.; Ghorbani-Choghamarani, A. l-Methionine–Pd complex supported on hercynite as a highly efficient and reusable nanocatalyst for C–C cross-coupling reactions. New J. Chem. 2020, 44, 2919. [Google Scholar] [CrossRef]
  31. Seyedi, N.; Saidi, K.; Sheibani, H. Recyclable Pd/CuFe2O4 nanowires: A highly active catalyst for C–C couplings and synthesis of benzofuran derivatives. Catal. Lett. 2018, 148, 277. [Google Scholar] [CrossRef]
  32. Kaboudin, B.; Abedi, Y.; Yokomatsu, T. CuII–β-Cyclodextrin Complex as a Nanocatalyst for the Homo- and Cross-Coupling of Arylboronic Acids under Ligand- and Base-Free Conditions in Air: Chemoselective Cross-Coupling of Arylboronic Acids in Water. Eur. J. Org. Chem. 2011, 2011, 6656. [Google Scholar] [CrossRef]
  33. Islam, M.; Mondal, S.; Mondal, P.; Roy, A.S.; Tuhina, K.; Mobarok, M.; Paul, S.; Salam, N.; Hossain, D. An Efficient Recyclable Polymer Supported Copper(II) Catalyst for C–N Bond Formation by N-Arylation. Catal. Lett. 2011, 141, 1171. [Google Scholar] [CrossRef]
  34. Singh, R.; Allam, B.K.; Raghuvanshi, D.S.; Singh, K.N. A Facile and Efficient Synthesis of Diaryl Amines or Ethers under Microwave Irradiation at Presence of KF/Al2O3 without Solvent and Their Anti-Fungal Biological Activities against Six Phytopathogens. Tetrahedron 2013, 69, 1038. [Google Scholar] [CrossRef]
  35. Cai, L.; Qian, X.; Song, W.; Liu, T.; Tao, X.; Li, W.; Xie, X. Effects of solvent and base on the palladium-catalyzed amination: PdCl2 (Ph3P) 2/Ph3P-catalyzed selective arylation of primary anilines with aryl bromides. Tetrahedron 2014, 70, 4754. [Google Scholar] [CrossRef]
  36. Liu, J.; Yang, D.; Yang, X.; Nie, M.; Wu, G.; Wang, Z.; Gong, P. Design, synthesis and biological evaluation of novel 4-phenoxyquinoline derivatives containing 3-oxo-3, 4-dihydroquinoxaline moiety as c-Met kinase inhibitors. Bioorg. Med. Chem. 2017, 25, 4475. [Google Scholar] [CrossRef]
  37. Antilla, J.C.; Buchwald, S.L. Copper-catalyzed coupling of arylboronic acids and amines. Org. Lett. 2001, 3, 2077. [Google Scholar] [CrossRef]
  38. Kantam, M.L.; Venkanna, G.T.; Sridhar, C.; Sreedhar, B.; Choudary, B.M. An Efficient Base-Free N-Arylation of Imidazoles and Amines with Arylboronic Acids Using Copper-Exchanged Fluorapatit. J. Org. Chem. 2006, 71, 9522. [Google Scholar] [CrossRef]
Scheme 1. Suzuki coupling reaction using magnetic nanoparticle catalyst.
Scheme 1. Suzuki coupling reaction using magnetic nanoparticle catalyst.
Catalysts 12 00976 sch001
Figure 1. TEM and SEM analyses with a graph representing the order of nanoparticles.
Figure 1. TEM and SEM analyses with a graph representing the order of nanoparticles.
Catalysts 12 00976 g001
Figure 2. FTIR analysis: (a) Fe3O4; (b) Fe3O4@L.
Figure 2. FTIR analysis: (a) Fe3O4; (b) Fe3O4@L.
Catalysts 12 00976 g002
Figure 3. EDX, VSM, and XRD spectra.
Figure 3. EDX, VSM, and XRD spectra.
Catalysts 12 00976 g003
Scheme 2. Suzuki coupling reaction.
Scheme 2. Suzuki coupling reaction.
Catalysts 12 00976 sch002
Scheme 3. Full reaction scheme of Suzuki coupling.
Scheme 3. Full reaction scheme of Suzuki coupling.
Catalysts 12 00976 sch003
Scheme 4. Mechanism of Suzuki coupling reaction.
Scheme 4. Mechanism of Suzuki coupling reaction.
Catalysts 12 00976 sch004
Figure 4. IR, SEM, and TEM analyses of Fe@L/Co/Ni nanoparticles after being reused 10 times.
Figure 4. IR, SEM, and TEM analyses of Fe@L/Co/Ni nanoparticles after being reused 10 times.
Catalysts 12 00976 g004
Scheme 5. Synthesis of Fe@L/Co/Ni magnetic nanoparticles.
Scheme 5. Synthesis of Fe@L/Co/Ni magnetic nanoparticles.
Catalysts 12 00976 sch005
Table 1. Suzuki coupling reaction results.
Table 1. Suzuki coupling reaction results.
EntrySaltSolventBaseCat. (mg)Temperature (°C)Time (min)Yield (%) aTONTOF (h−1)
1NaSCNCH3CNDABCO10709080686.8
2Cs2CO3THFKF10110100857413.5
3K2CO3DMSOKF510080925427
4NaHCO3DMF/H2ONaOH5707090598.9
5K2CO3DMF/H2OKI0.755060968241
6K2CO3DMF/H2OKI0.57055968643
7K2CO3DMF/H2OKI18045989247
8K2CO3DMF/H2OKI1.510035999449.5
9K2CO3DMF/H2OKI27060999748.5
10-Not cat.--100120-8944.5
11K2CO3Fe3O4KI2010022 htrace9045
12K2CO3Fe/LKI2010020208743.5
13-Co/Ni (OAc)2·H2O-207014658241
14K2CO3Fe/L/Co/NiKI27060994824
a Yield refer isolated products.
Table 2. Comparison of Fe@L/Co/Ni nanocatalyst with other catalysts.
Table 2. Comparison of Fe@L/Co/Ni nanocatalyst with other catalysts.
EntryCatalystConditionsTime (h)Yield (%) aReferences
1Cu2–β-CD (0.01 mmol), K2CO3 (3 equiv.)DMF (1 mL), 90 °C4890[24]
2Cu(OAc)2 (5–20 mol %), myristic acid (10–40 mol%)Toluene (2 mL), r.t.2492[25]
3CuFAP (100 mg)Methanol (4 mL), r.t.390[26]
4Cu-BIA-Si-Fe3O4 (1.5 mol%), KF (0.2 mmol)DMSO (4 mL), 100 °C2.596[27]
5Cu-BIA-Si- Fe3O4 (1.5 mol%) K2CO3 (1 mmol)DMSO, 80 °C295[28]
6Palladium N-heterocyclic carbene (0.02 mmol), K2CO3 (1.2 eq)i-PrOH (4 mL), r.t.698[29]
7Pd(OAc)2/LHX (15 mol%)DMF/H2O (2 mmol), 50 °C393[30]
8Aminomethyldiphosphine–Pd(II) (1.2 mmol)DMF/H2O (3/3 mL), 80 °C795[31]
9Pd (1.0 mol) K3PO4·3H2O (1.5 mmol)H2O/ethanol
(2 mL), 80 °C
398[32]
10Cross-linked poly(ITC-HPTPy)-Pd (0.23 mol%) K2CO3 (1.5 mmol)H2O/ethanol (3 mL),
80 °C
298[33]
11CuCl2 (10 mol%), 4,40-di-tert-butyl-2,20-bipyridine (10 mol%), DCM (1.0 mL)Cs2CO3 (2.0 equiv), r.t.2478[34]
12Hercynite@L-methionine-Pd (0.1 mol%) K2CO3 (3 mmol)ethanol (3 mL), 80 °C1.580[35]
13GO/Fe3O4/Pd nanocomposite (0.36 mol) K2CO3 (3 mmol)H2O/ethanol (3 mL), 80 °C1080[36]
14Cu(II)–β-cyclodextrin (0.1 eq)DMF, 90 °C1080[37]
15Cu-PAR (0.05 g, 0.0129 mmol) KOH (1 mmol)DMSO, 120 °C1494[38]
16Fe@L/Co/Ni (0.2 mol%)DMF/KI, 70 °C0.5–199present
a Yield refer isolated products, ref. [24,25,26,27,28,29,30,31,32,33,34,35,36,37,38].
Table 3. Derivatives provided of Suzuki coupling reaction.
Table 3. Derivatives provided of Suzuki coupling reaction.
EntryArylhalide + Phenyl Boronic AcidProductTime (min)Yield (%) aTONTOF (h−1)FoundReported
Melting Point °C
1Ortho-NO2PhI Catalysts 12 00976 i001451008743.5102–104102.8
2Para-NO2PhI Catalysts 12 00976 i00250998643102–104103.2
3NO2PhI Catalysts 12 00976 i00360989547.5OilOil
42,4,6-tri-NO2PhI Catalysts 12 00976 i00460997135.5345.68345.7
52,4-di-NO2PhF Catalysts 12 00976 i00555986834201.9202
6Para-NH2PhCl Catalysts 12 00976 i0061.5 h975829100.12100.2
72,6-di-NO2PhF Catalysts 12 00976 i00760986733.5201.8201.9
8Para-CO2HPhCl Catalysts 12 00976 i00860986130.5134.77134.8
92-CH2OHPhI Catalysts 12 00976 i009559710150.588.9589
103-CH2OHPhBr Catalysts 12 00976 i01060971145788.9889.2
114-CH2OHPhI Catalysts 12 00976 i011709612160.589.189.5
a Yield refer isolated products.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Binandeh, M.; Nasseri, M.A.; Allahresani, A. High-Power and High-Performance Catalyst for Suzuki Coupling Reaction. Catalysts 2022, 12, 976. https://doi.org/10.3390/catal12090976

AMA Style

Binandeh M, Nasseri MA, Allahresani A. High-Power and High-Performance Catalyst for Suzuki Coupling Reaction. Catalysts. 2022; 12(9):976. https://doi.org/10.3390/catal12090976

Chicago/Turabian Style

Binandeh, Mansour, Mohammad Ali Nasseri, and Ali Allahresani. 2022. "High-Power and High-Performance Catalyst for Suzuki Coupling Reaction" Catalysts 12, no. 9: 976. https://doi.org/10.3390/catal12090976

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