Home Journals Progress in Chemistry
Progress in Chemistry

Abbreviation (ISO4): Prog Chem      Editor in chief: Jincai ZHAO

About  /  Aim & scope  /  Editorial board  /  Indexed  /  Contact  / 
Review

Design, Synthesis and Application of Magnetic Nanoparticle Catalytic Materials Based on Multientate Palladium Compounds

  • Yunhua Ma ,
  • Han Shao ,
  • Tenglong Lin ,
  • Qinyue Deng , *
Expand
  • College of Materials and Chemistry, University of Shanghai for Science and Technology, Shanghai, 200093, China

Received date: 2023-02-01

  Revised date: 2023-03-24

  Online published: 2023-05-30

Supported by

The Shanghai Young Teachers Training and Support Program(slg20035)

Abstract

Catalyst loading is one of the effective strategies for green catalysis. Palladium (Pd) catalysts supported by magnetic nanoparticles (MNPs) have been widely studied and used in organic synthesis due to their good dispersibility, high catalytic activity, rapid separation under the action of an external magnetic field, and efficient recovery. The MNPs-supported polydentate Pd compound catalyst (MNPs@L-Pd) shows better catalytic activity and stability than the MNPs-supported Pd nanoparticle catalyst (MNPs@PdNP). This is mainly because the introduction of the modified ligand in MNPs@L-Pd can regulate the electronic effect and steric hindrance of the catalyst metal center to achieve the regulation of its activity, on the other hand, it makes the stable chemical bond between the catalyst metal center and the magnetic material to achieve the regulation of stability. This paper mainly focuses on MNPs@L-Pd, the preparation of MNPs@L-Pd based on different ligands and coordination methods and its application in C-X(Cl, Br, I) activation reaction in the past 10 years are reviewed from the aspects of catalyst stability and activity, and the prospect of these reactions are also presented.

Contents

1 Introduction

2 Palladium-catalyzed system based on bidentate coordination mode

2.1 N-Pd-N coordination bond catalytic system

2.2 O-Pd-N coordination bond catalytic system

2.3 P-Pd-P coordination bond catalytic system

2.4 S-Pd-N coordination bond catalytic system

2.5 Se-Pd-N coordination bond catalytic system

3 Palladium-catalyzed system based on tridentate coordination mode

4 Palladium-catalyzed system based on tetradecentate coordination mode

5 Palladium-catalyzed system based on multidentate coordination mode

6 Palladium-catalyzed system based on Pd-C covalent bonds

7 Conclusion and outlook

Cite this article

Yunhua Ma , Han Shao , Tenglong Lin , Qinyue Deng . Design, Synthesis and Application of Magnetic Nanoparticle Catalytic Materials Based on Multientate Palladium Compounds[J]. Progress in Chemistry, 2023 , 35(9) : 1369 -1388 . DOI: 10.7536/PC230115

1 Introduction

Human production and life are inseparable from chemical reactions, which are related to health, environment and energy[1]. The development of catalytic technology makes it possible to achieve more accurate, efficient and controllable chemical reactions[2]. Among them, transition metal catalysis has been widely used in the activation of various C-C/H bonds and C-heteroatom bonds[3~6]. Palladium compounds are one of the most important transition metal catalysts in organic synthesis[7~10]. In the homogeneous palladium catalyst system, the catalyst can be effectively dispersed in the reactant system and generally has high catalytic activity, but there is also a problem that the catalyst molecules are easy to polymerize, resulting in the reduction of the activity of the catalyst[11]. The palladium catalyst used in the reaction is relatively expensive, and most of the efficient catalytic systems require carefully designed and synthesized ligand modification, which makes the cost of the catalyst higher. The catalyst is not easy to recycle, resulting in a waste of resources and metal residues in the system and products, which limits its wide application in the chemical industry, especially in the pharmaceutical field[12].
In the past decade, the heterogeneous strategies of different homogeneous catalysts, such as loading, organometallic polymer, hypercrosslinking, self-loading, CTF and COF, have been widely and systematically studied[13,14]. It is proved that the efficient and stable immobilization of homogeneous metal catalyst can realize the independent dispersion of catalyst molecules, effectively solve the problem of catalyst molecule polymerization, and also realize the recovery of catalyst, effectively solve the problems of cost and environmental hazards. And the dual optimization of the stability and the activity of the catalyst can be realized through the design of the catalyst structure. Immobilization is a strategy to achieve catalyst heterogeneity by immobilizing the dominant catalytic fragments on polymers, graphene, silica, zeolites, magnetic nanomaterials and other carriers through different linking methods (mainly chemical methods)[15][16][17,18].
Supported catalysts are also favored because of their simple synthesis and easy structural characterization. Among them, MNPs are widely used as carriers for the loading of palladium compound catalysts because of their nano-size, good dispersion of homogeneous catalytic system, high specific surface area for catalyst loading, and magnetic property for simple separation and recovery by external magnetic field. The reactivity and stability of MNPs @ L-Pd are not only related to the inherent characteristics of the support and the link between the active species and the support, but also closely related to the activity and stability of the dominant fragment of the catalytic center.
Therefore, in addition to selecting a suitable support, it is particularly important to design a dominant catalyst fragment with a suitable ligand that modulates the steric hindrance or electronic effect of the catalytic center.
In order to explore the construction strategy of a more efficient and stable catalytic system, based on the coordination characteristics of Pd atoms, researchers designed ligands to coordinate C, N, O, P and other elements as the main coordination atoms with palladium (Fig.Develop a series of MNPs @ L-Pd with different ligands (such as phosphine ligands, nitrogen-rich ligands, Schiff bases, palladacycles, nitrogen heterocyclic carbenes (NHCs), etc.) And bonding modes, and achieve efficient and recoverable catalysis for a variety of organic synthesis reactions. In this paper, the representative research progress in the past decade is reviewed, focusing on the activity and stability of catalysts with different ligand structures in organic reactions, and the prospects are put forward.
图1 催化剂优势片段配位方式示意图

Fig.1 Schematic diagram of the coordination mode of the dominant fragment of the catalyst

2 Palladium catalytic system based on bidentate coordination mode

2.1 N-Pd-N coordination bond catalytic system

In 2011, Phan et al. prepared the phosphine-free catalyst MNP-1 (A \ B) with an average particle size of 30 ~ 40 nm by functionalizing the amine group of superparamagnetic nanoparticles synthesized by microemulsion method and complexing the immobilized Schiff base formed by condensation of 2-acetylpyridine with Pd (II) (Figure 1)[19]. Among them, MNP-1-B achieved its efficient Suzuki coupling reaction with aryl halides and arylboronic acids at 0.1 mol% catalytic amount. The catalyst can be stably reused for 11 times, and the leaching rate of the reaction system is detected by ICP-MS, which shows that the catalyst has good stability in the Suzuki reaction. In addition, through the comparative experiments of catalyst A and B, it was found that the catalytic activity in the Suzuki reaction was significantly improved when the tether in the catalyst structure was extended. This may be because the elongation of the tether facilitates the access of the reactants to the catalytic active center.
图式1 MNP-1的合成及其在Suzuki偶联反应中的应用[19]

Scheme 1 Synthesis of MNP-1 and its application in Suzuki coupling reaction[19]

In 2013, Alizadeh et al. Prepared a bidentate magnetic nanocatalyst MNP-2 with an average particle size of 40 ~ 45 nm by surface modification of MNP with a biguanide group for Pd (II) loading, and achieved an efficient aqueous Suzuki coupling reaction of aryl halides with different aryl boronic acids at a catalytic amount of 0.14 mol% (Scheme 2)[20]. However, when the catalytic amount was increased to 1 mol%, only 45% yield could be achieved for the coupling reaction involving aryl chloride. In the template reaction, the catalyst can be stably reused for at least 8 times.
图式2 MNP-2的合成及其在Suzuki偶联反应中的应用[20]

Scheme 2 Synthesis of MNP-2 and its application in Suzuki coupling reaction[20]

Similarly, in 2016, Bian et al. Prepared a magnetic nanocatalyst MNP-3 (Scheme 3) with an average particle size of 14.3 nm by immobilizing palladium on the surface of a metformin-functionalized magnetic polymer nanocomposite, and applied it to the Suzuki coupling reaction of aryl halides and boronic acid derivatives[21]. Various aryl iodides and aryl bromides with phenylboronic acid can be almost quantitatively converted into the corresponding biphenyls under the optimal conditions with a catalytic amount of 0.01 mol%. Even the reaction of aryl chloride with arylboronic acid can still give 74% ~ 91% yield. The catalyst has very good catalytic activity and has the best TOF value among the MNP supported catalysts at that time. In addition, the catalyst can be stably recycled for at least 12 times, and has good recyclability.
图式3 MNP-3的合成及其在Suzuki偶联反应中的应用[21]

Scheme 3 Synthesis of MNP-3 and its application in Suzuki coupling reaction[21]

In 2015, Gu GuÉnin et al. Prepared a novel nanocatalyst MNP-4 (Fig. 4) by chelating the CAT-Pro ligand obtained by coupling proline and dopamine on the surface of bare magnet γ-Fe2O3 NPs, and applied it to the Suzuki coupling reaction of 4-iodonitrobenzene with 4-methylphenylboronic acid[22]. When the catalytic amount was 0. 1 mol%, biphenyl was obtained in 97% yield after 30 min. It is proved that the catalytic system has good tolerance to inert arylboronic acid substrates with electron-withdrawing groups and even substrates with active functional groups by expanding the substrates. In addition, the catalyst can be stably reused for 8 times, and the nanoparticles after the 8th cycle were analyzed by TEM and EDX, which showed that the morphology and structure of MNP-4 did not change significantly before and after the reaction, and the leaching rate was small.
图式4 MNP-4的合成及其在Suzuki偶联反应中的应用[22]

Scheme 4 Synthesis of MNP-4 and its application in Suzuki coupling reaction[22]

In 2019, Bodaghifard attempted to synthesize a novel magnetic nanocatalyst, MNP-5, using melamine as a ligand[23]. Aryanasab et al. Then prepared the magnetic nanocatalyst MNP-6 by linking melamine to the surface of the prepared epoxy-functionalized MNPs through the covalent bond of epoxy-amine coupling chemistry, and applied it to the Heck coupling reaction of various aryl halides with olefins[24]. The results show that both catalysts have good catalytic activity and functional group tolerance. The catalytic efficiency of MNP-5 was not significantly reduced after 8 times of recycling. However, MNP-6 was cycled up to 6 times. It can be seen that the formation of multidentate coordinated palladium compounds is beneficial to the stability of the catalyst, and the appropriate extension of the length of the tethering group is beneficial to the activity of the catalyst (Figure 5).
图式5 MNP-5和MNP-6的合成及其在Heck偶联反应中的应用[23,24]

Scheme 5 Synthesis of MNP-5 and MNP-6 and their application in Heck coupling reaction[23,24]

In 2022, Dehbanipour et al. Developed a magnetic nanocatalyst MNP-7 (Fig. 6) with 3,5-bis (2-benzothiazolyl) pyridine (BTP) as a ligand and used it for the reaction of benzyl alcohol derivatives with 1,2-phenylenediamine or 2-aminothiophene to synthesize benzothiazole and benzimidazole[25]. The reaction of various aromatic benzyl alcohols with 1,2-phenylenediamine gave benzimidazoles and benzothiazoles in high yields of more than 90% at 80 ℃ with a Pd loading of 0. 09 mol%, showing good catalytic activity. This newly synthesized catalyst system has good stability and can be easily separated and recovered and reused at least five times in both reactions without loss of its activity.
图式6 MNP-7的合成及其在交叉偶联反应中的应用[25]

Scheme 6 Synthesis of MNP-7 and its application in cross-coupling reaction[25]

In the same year, Xu et al. Successfully prepared palladium nanocatalyst MNP-8 (Fig. 7) by immobilizing palladium (II) complex on the surface of 2-aminopyridine modified magnetic nanoparticles, and used it in the C-S and C-Se cross-coupling reactions of aryl halides[26]. The catalytic system is applicable to various substituted aryl halides and generates the corresponding diaryl sulfides and diaryl selenides in moderate to excellent yields in the appropriate time in green solvents at 100 ° C. The catalyst can be recovered by simple magnetic separation and reused for 7 times without significant decrease in catalytic activity, and the morphology and structure of the catalyst are not significantly damaged.
图式7 MNP-8的合成及其在C-S和C-Se交叉偶联反应中的应用[26]

Scheme 7 Synthesis of MNP-8 and its application in cross-coupling of C-S and C-Se[26]

In 2023, Cai et al. Successfully prepared magnetic nanocatalyst MNP-9 (Fig. 8) by immobilizing palladium (II) complex on the surface of silica-coated MNP with phenanthroline as ligand, and used it to develop a mild and efficient synthetic method for cyanation of benzoxazole, benzothiazole and aryl halide[27]. Under mild conditions, the catalytic system synthesized benzoxazoles, benzothiazoles, and cyanides in good to excellent yields, and the substrate functionality was well tolerated. In addition, the nanocatalyst can be easily recovered by an applied magnetic field, and the catalytic activity is maintained without a significant drop even after seven cycles.
图式8 MNP-9的合成及其在交叉偶联反应和氰基化反应中的应用[27]

Scheme 8 Synthesis of MNP-9 and its application in cross-coupling and cyanidation reactions[27]

Dendrimers can form stable and high-density coordination with transition metals through internal and external functional groups due to their unique three-dimensional structure and highly branched structure, and have good dispersion. In addition, dendrimers have a large number of cavities due to their three-dimensional structure, which can absorb and concentrate the reactants and make the reaction more efficient, and they have the recyclability of heterogeneous catalysts by being immobilized on solid supports. Therefore, dendrimers are considered to be good candidate supports for heterogeneous catalysis. In addition, rigid aromatic dendrites and dendrimers can maintain structural rigidity and shape persistence under high pressure and temperature, thereby maintaining the location of functional groups and their interaction with the carrier molecules or NPs[28].
In 2018, Bian et al. Prepared a magnetic dendrimer nanocomposite through the stepwise reaction of methacrylate and 3,3-diaminodipropylamine on the surface of amino-functionalized Fe3O4@SiO2, and immobilized PdCl2 on the composite to obtain catalyst MNP-10 with an average particle size of about 16.1 nm[29]. The nanocatalyst could effectively catalyze solvent-free Heck reaction, Suzuki reaction in EtOH/H2O solvent with 98% yield under the optimized conditions with only 0.009 mol% catalytic amount. After five consecutive runs of MNP-10, there was still an appreciable yield of 88%, and the amount of palladium leached from the fifth run solution based on AAS determination was only 0.06 ppm (Figure 9).
图式9 MNP-10的合成及其在Heck和Suzuki交叉偶联反应中的应用[29]

Scheme 9 Synthesis of MNP-10 and its application in Heck and Suzuki cross-coupling reactions[29]

In the following year, Bronstein et al. Immobilized pyridine-phenylene dendrimers with hydroxyl or carboxyl groups on the surface of MNPs through ether bonds (or amide bonds), and further interacted with Pd(OAc)2, so that Pd in the system existed in the form of Pd2+ and Pd0 species and palladium nanoparticles (MNP-11-A/B)[30]. Moreover, the combination of hydrophobic dendrimer and hydrophilic support leads to the amphiphilic nature of the nanocomposite catalyst, which effectively catalyzes the Suzuki-Miyaura cross-coupling reaction of 4-Br-anisole and phenylboronic acid in ethanol/water mixture. The target cross-coupling products were obtained with 92.0% and 96.4% conversion and 97.1% and 97.9% yield within 5 min for MNP-11-A and B, respectively. The high catalytic activity was observed even with a very small amount of catalyst, which is most likely due to the effective stabilization of the Pd2+ and Pd0 species by the dendrimer and the rigidity of the dendrimer facilitating substrate access to the active site, but this catalyst could be stably reused only 3 times (Scheme 10).
图式10 MNP-11的合成及其在Suzuki交叉偶联反应中的应用[30]

Scheme 10 Synthesis of MNP-11 and its application in Suzuki cross-coupling reactions[30]

Compared with pyridine-phenylene dendrimers, polyamide (PAMAM) dendrimers have a particularly well-defined structure, which is beneficial to the immobilization of metal nanoparticles. In addition, the generation of dendrimers can control the size and metal loading of nanoparticles, while the solubility in common solvents can be adjusted by the nature of the dendritic end groups. In 2022, Nasseri et al. Reported a recoverable magnetic nanoscale Pd complex (MNP-12) with zero-generation PAMAM dendrimer G0 as ligand and applied to Pd-catalyzed C-C coupling[31]. The catalytic system showed unusually high activity for Suzuki-Miyaura and Mizoroki-Heck coupling reactions in aqueous environment, and even showed good catalytic activity for less active aryl chlorides. Compared with other effective hydrophilic/amphiphilic catalysts in C-C cross-coupling reactions, the catalytic system shows high activity in water, and can effectively catalyze aryl chloride coupling even under relatively mild conditions (90 deg C),Moreover, it is easy to be recovered and reused for 6 times, the catalytic activity is not significantly reduced, and the morphology and structure of the catalyst are not significantly damaged (Figure 11).
图式11 MNP-12的合成及其在C-C交叉偶联反应中的应用[31]

Scheme 11 Synthesis of MNP-12 and its application in C-C cross-coupling reactions[31]

2.2 O-Pd-N coordination bond catalytic system

In 2013, Bian et al. Prepared magnetic nanoparticles coated with olefinic acid by chemical coprecipitation method and anchored palladium (II) on the surface to synthesize catalyst MNP-13, which could effectively catalyze Suzuki and Heck reactions of various aryl iodides and aryl bromides with arylboronic acid and acrylic acid in H2O[32]. In addition, MNP-13 can be separated and recycled by a magnet, and can be continuously used for six times without significant decrease in catalytic activity (Fig. 12). In 2015, they synthesized a magnetic polymer nanocomposite by radical polymerization, and immobilized Pd (II) on the material to synthesize a nitrogen-oxygen chelated palladium coordination compound catalyst MNP-14 with an average particle size of 13.3 nm, which can effectively catalyze the Heck coupling of acrylic acid and aryl halides at a low catalytic amount[33]. At the same time, the catalyst also has good catalytic activity for the Suzuki coupling reaction of aryl bromides and phenylboronic acid, and even shows certain reactivity for aryl chlorides. The catalyst can be stably recycled for 6 times in the Heck coupling template reaction (Figure 13).
图式12 MNP-13的合成及其在C-C偶联反应中的应用[32]

Scheme 12 Synthesis of MNP-13 and its application in C-C cross-coupling reactions[32]

图式13 MNP-14的合成及其在C-C偶联反应中的应用[33]

Scheme 13 Synthesis of MNP-14 and its application in C-C cross-coupling reactions[33]

In 2020, Mamaghani et al. Designed and synthesized MNP-15, which can efficiently catalyze Heck coupling and selectively obtain trans-isomer products. After 10 consecutive cycles of catalysis, it still maintains a stable structure and catalytic activity, indicating that the bonding between MNPs and palladium is strong (Figure 14)[34][35,36].
图式14 MNP-15的合成及其在Heck偶联反应中的应用[34]

Scheme 14 Synthesis of MNP-15 and its application in Heck coupling reactions[34]

Nanocellulose is the most abundant biopolymer in the earth, which is extracted from plants, bacteria and algae. It has good hydrophobicity, biodegradability, economy, biocompatibility and extensive chemical functionalization ability.These properties of cellulose make it one of the most perfect coated supports for MNPs, as it not only has the ability to stabilize nanoparticles in solution, but also facilitates functionalization to produce biopolymer-based catalysts. In 2021, Mohammadnia et al. Reported a 5-carboxyoxindole-functionalized cell@Fe3O4 nanoparticle-supported Pd catalyst (MNP-16) with an average particle size of about 15 nm, and successfully used it in the Heck-type arylation of different substituted maleimides with aryl iodides to obtain the target products in good to excellent yields (77% – 88%), which could be stably reused for five times (Fig. 15)[37].
图式15 MNP-16的合成及其在Heck偶联反应中的应用[37]

Scheme 15 Synthesis of MNP-16 and its application in Heck coupling reactions[37]

2.3 P-Pd-P coordination bond catalytic system

In 2015, Zarnaghash et al. Prepared a heterogeneous catalyst MNP-17 with an average particle size of 10 nm by phosphine functionalization on the surface of MNPs synthesized by coprecipitation method and then loading palladium, and applied it to the Buchwald-Hartwig amination coupling reaction of aryl halides and amines using a three-step method shown in Figure 16[38]. Under solvent-free conditions, the efficient amination of aryl chlorides with secondary cyclic amines was achieved with 1. 2 mol% catalytic amount. The catalyst can be reused for only 5 times with retained catalytic activity. This may be due to the redox reaction between some phosphine ligands and active palladium species during the reaction, which makes Pd (0) species aggregate to form Pd nanoparticles to overflow the catalytic system and reduce the activity and recycling rate of the catalyst.
图式16 MNP-17的合成及其在Buchwald-Hartwig 胺化反应中的应用[38]

Scheme 16 Synthesis of MNP-17 and its application in Buchwald-Hartwig amination[38]

2.4 S-Pd-N coordination bond catalytic system

Sulfur ligands are viable alternatives due to their reduced sensitivity to air and moisture, strong electron donating properties, solubility in different solvents and stability in solution. In 2019, Sobhani et al. Synthesized a novel water-dispersed/magnetically recoverable palladium heterogeneous catalyst MNP-18 by surface modification of hydrophilic groups and used it for C-C coupling reaction in pure water system[39]. Due to the presence of hydrophilic TEG in the catalyst, MNP-18 can be effectively dispersed in the aqueous phase. Aryl cyanides were obtained in 99% yield by the reaction of halobenzenes with K4[Fe(CN)6]·3H2O at 80 ℃ for 6 H when the catalytic amount was 0.2 mol%. When the catalytic amount is 0.1 mol%, MNP-18 can effectively catalyze the Hiyama cross-coupling reaction of halogenated benzene and triethoxyphenylsilane under fluorine-free conditions; When the catalytic amount of MNP-18 was 0. 01 mol%, it could effectively catalyze the Suzuki cross-coupling reaction of different halogenated benzenes with phenylboronic acid, and the yield could reach 98%. Compared with other similar catalytic systems, the catalytic system has good versatility for the coupling reaction of aryl halides and even chlorides. In the corresponding reaction, the catalyst could be recycled nine times without a significant decrease in catalytic activity (Fig. 17).
图式17 MNP-18的合成及其在C-C偶联反应中的应用[39]

Scheme 17 Synthesis of MNP-18 and its application in C-C coupling reactions[39]

2.5 Se-Pd-N coordination bond catalytic system

The catalytic activity of organoselenium-containing ligands is not only comparable to that of the corresponding phosphorus analogues, but also shows excellent catalytic activity in many cases. In 2018, Nemati et al. First synthesized magnetic nano-palladium catalyst MNP-19 (Fig. 18) with organic selenium-nitrogen chelate coordination, and designed and prepared MNP-20 (Fig. 19) in the following year, which was used in "phosphorus-free" Heck-Mizoroki coupling reaction and Suzuki-Miyaura coupling reaction, respectively[40,41]. The results show that both catalysts have good thermal stability and catalytic activity, and can be stably recycled for 5 and 7 times in the template reaction, respectively.
图式18 MNP-19的合成及其在Heck偶联反应中的应用[40]

Scheme 18 Synthesis of MNP-19 and its application in Heck coupling reactions[40]

图式19 MNP-20的合成及其在Suzuki偶联反应中的应用[41]

Scheme 19 Synthesis of MNP-20 and its application in Suzuki coupling reactions[41]

3 Palladium catalytic system based on tridentate coordination mode

Sobhani et al. Synthesized MNP-supported NNN-Pd chelating catalyst MNP-21 through the synthetic route of Scheme 20 in 2015[42]. Under the optimized conditions, MNP-21 can effectively catalyze the C-C coupling of different aryl halides with a broad substrate range, good functional group tolerance, and moderate to excellent yields. In the Heck coupling reaction of iodobenzene with n-butyl acrylate, MNP-21 could be recycled 10 times without any significant decrease in catalytic activity. The catalyst after 10 cycles was characterized by TEM, and no obvious aggregation of Pd was observed.
图式20 MNP-21的合成及其在C-C偶联反应中的应用[42]

Scheme 20 Synthesis of MNP-21 and its application in C-C coupling reactions[42]

4 Palladium catalytic system based on tetradentate coordination mode

In recent years, Schiff bases, which are easy to synthesize and have high stability, have played a key role in coordination chemistry and catalysis as chelating multidentate ligands. Studies of Schiff base – transition metal complexes immobilized on various supports have been widely reported. Due to the high chemical stability and catalytic activity of tetradentate chelate coordination compounds, in 2014, Jaridi et al. And their colleagues successively synthesized superparamagnetic nanocatalysts MNP-22 and MNP-23 (Fig. 21 and 22) with average particle size of about 30 nm and 38 nm by loading Pd (II) Schiff base complexes on the surface of functionalized Fe3O4, and applied them to the C-C bond coupling reaction of aryl halides[43][43,44]. When the loading of MNP-22 was 0.7 mol%, iodobenzene reacted with phenylacetylene at 90 ℃ for 0.5 H to give the target product in 94% yield. Secondly, in the Heck coupling reaction, only 0.3 mol% of MNP-23 was needed, and bromobenzene and n-butyl acrylate reacted at 110 ℃ for 3 H to obtain the coupling product in 90% yield. In the Suzuki coupling reaction of bromobenzene and phenylboronic acid, the product yield was 88% at 100 ℃ for 2. 5 H when the mole fraction of MNP-23 was 0. 3 mol%. In the corresponding reaction, MNP-22 and MNP-23 can be reused many times, while the catalytic activity remains basically unchanged, and the Pd leaching rate is negligible (Figure 21).
图式21 MNP-22的合成及其在Sonogashira偶联反应中的应用[43]

Scheme 21 Synthesis of MNP-22 and its application in Sonogashira coupling reactions[43]

图式22 MNP-23的合成及其在C-C偶联反应中的应用[44]

Scheme 22 Synthesis of MNP-23 and its application in C-C coupling reactions[44]

Sardarian et al. Synthesized MNPs supported Schiff base coordinated palladium polymer catalyst MNP-24 in 2019, with an average particle size of 30 nm[45]. The catalyst has good catalytic activity for Heck coupling reaction and Sonogashira coupling reaction in the presence of aryl iodides. In the corresponding reaction, MNP-24 can be stably recycled for 8 times with good stability (Figure 23).
图式23 MNP-24的合成及其在C-C偶联反应中的应用[45]

Scheme 23 Synthesis of MNP-24 and its application in C-C coupling reactions[45]

In May of the same year, Mahdavi et al. Prepared a new catalyst MNP-25 of (pyridine-2-methyl) dithiocarbamate (PDTC-Pd) complex supported on γ-Fe2O3@SiO2, and investigated its catalytic performance in the Heck/Sonogashira coupling reaction of various aryl halides with various olefins/phenylacetylene in water[46]. The catalyst showed the advantages of efficient recyclability, high turnover rate (TON) and high turnover frequency (TOF) even at low catalytic amount (5 mg, less than 0.1 mol%). There was no significant loss of activity in 10 consecutive runs of the Heck and Sonogashira reactions (Scheme 24).
图式24 MNP-25的合成及其在C-C偶联反应中的应用[46]

Scheme 24 Synthesis of MNP-25 and its application in C-C coupling reactions[46]

Iron, nickel and their alloys are widely concerned as typical magnetic materials[47]. In 2020, Sadeghzadeh et al. Prepared mesoporous silica-coated magnetic core-shell FeNi3/DFNS MNPs with dendritic silica fiber morphology by hydrothermal method and microemulsion method, and supported Schiff base complex of Pd (II), and finally added melamine for heat treatment to synthesize Schiff base coordinated magnetic nanocatalyst MNP-26[48]. The newly synthesized MNP-26 was applied to the preparation of cyclic carbonates from CO2, which showed good catalytic activity and could be stably recycled for 10 times (Fig. 25).
图式25 MNP-26的合成及其在CO2环碳酸酯反应中的应用[48]

Scheme 25 Synthesis of MNP-26 and its application in CO2 cyclocarbonate reaction[48]

In order to maximize the availability of metal catalysts and prevent the leaching of metal species, Rahman et al. Prepared catalyst MNP-27 by modifying the structure of phenanthroline Schiff base and successfully applied it to Suzuki-Miyara coupling[49]. The catalyst showed good reactivity for both electron-deficient and electron-rich aromatic halides in the presence of a variety of organoboronic acids. MNP-27 can be simply recovered from the reaction system by using an external strong magnet and can be stably reused for 7 times, showing good catalytic activity and stability (Scheme 26).
图式26 MNP-27的合成及其在Suzuki偶联反应中的应用[49]

Scheme 26 Synthesis of MNP-27 and its application in Suzuki coupling reactions[49]

5 Palladium catalytic system based on multidentate coordination mode

Diethylenetriaminepentaacetic acid (DTPA) is widely used as a strong complexing agent for many metals because it contains three N atoms and four free carboxyl groups. If DTPA is bound to Fe3O4@SiO2, palladium will be in the core position of three nitrogen atoms, while the overflowing Pd or Pd in solution can be captured again by the blocked four carboxyl groups, achieving efficient loading and stability of palladium. Li et al. First reported a DTPA-modified Fe3O4@SiO2 magnetic nanoparticle-supported Pd nanocatalyst, MNP-28[50]. With a high loading of Pd2+/Pd0 (0.535 mmol/G), the catalyst showed efficient catalytic activity for the Suzuki-Miyaura coupling reaction of aryl halides with phenylboronic acid, and could be stably reused for 20 times, showing great potential for industrial practical applications (Fig. 27).
图式27 MNP-28的合成及其在Suzuki偶联反应中的应用[50]

Scheme 27 Synthesis of MNP-28 and its application in Suzuki coupling reactions[50]

6 Palladium catalytic system based on Pd-C covalent bond

N-heterocyclic carbene (NHC) -Pd complexes combine the catalytic activity of noble metals with the strong σ-electron donating ability, low toxicity, stability and easy modification of N-heterocyclic carbene, and become one of the most promising catalytic systems. In 2014, Esmaeilpou et al. Prepared bis-NHC palladium catalyst MNP-29 supported on poly (N-vinylpyrrolidone) grafted MNP[51]. Under the optimal conditions, MNP-29 can effectively catalyze the Heck reaction of aryl halides with n-butyl acrylate or styrene, and the yield can reach 74% -98%; When the supported palladium nanoparticles were used for the Suzuki reaction of aryl halides with organoboronic acids, the yields were up to 77% – 97%. In Suzuki and Heck coupling reactions, the catalyst could be used continuously for 6 times without significant decrease in catalytic activity. In addition, ICP analysis confirmed that the stability of the catalytic system was good when the leaching amount of palladium was low after 6 cycles, 6% and 4%, respectively (Figure 28).
图式28 MNP-29的合成及其在C-C偶联反应中的应用[51]

Scheme 28 Synthesis of MNP-29 and its application in C-C coupling reactions[51]

Although magnetic silica core-shell nanocatalysts are well dispersed in many solvents, most of the catalyst centers are hydrophobic environments, and when water is used as a solvent, a large amount of polyethylene glycol (PEG) and ionic liquid (IL) or n-Bu4NBr(TBAB) are needed as solvents or phase transfer reagents. In 2015, Mahdavi et al. Reported a highly water dispersible/magnetically separable NHC-Pd catalyst MNP-30 based on PEG substituted imidazolium phosphite ionic liquid[52]. In the aqueous Heck and Sonogashira coupling reactions, the catalytic activity of the catalyst was significantly improved, and a small amount (< 1 mg (0.0087 mol%)) of Pd catalyst was used to reduce 4-nitrophenol to 4-aminophenol. Under the template reaction condition, the Pd catalyst can be stably recycled for 10 times, and the leaching amount of Pd is less than 1 ppm. The unique water solubility and negligible Pd leaching rate indicate the high efficiency, versatility, environmental protection and economy of the catalyst (Fig. 29).
图式29 MNP-30的合成及其在C-C偶联反应中的应用[52]

Scheme 29 Synthesis of MNP-30 and its application in C-C coupling reactions[52]

In 2016, Andr Andrés et al. Prepared catalysts MNP-31 and MNP-32 by grafting pre-synthesized single NHC-Pd complexes and double NHC-Pd complexes onto the surface of γ-Fe2O3@SiO2, and compared their performance differences through Suzuki coupling reaction for the first time (Fig. 30)[53]. The results showed that MNP-31 and MNP-32 had good catalytic activity and recycling rate compared with homogeneous Pd complexes without MNPs. Compared with MNP-31, MNP-32 has better stability, higher catalytic activity, mild reaction conditions, and even catalytic activity in the coupling of aryl chlorides with phenylboronic acid (MNP-31 can not catalyze well). In terms of recycling performance, both of them can be stably recycled for 12 times, and the leaching rate of palladium is very low (≤ 10 ppm).
图式30 MNP-31和MNP-32的结构示意图[53]

Scheme 30 Schematic representation of the structures of MNP-31 and MNP-32[53]

In the same year, Singh et al. Designed and prepared a uniform magnetic nanoparticle-supported oxime complex catalyst MNP-33 (Fig. 31) with an average particle size in the range of 5.90 ~ 14.39 nm by anchoring carbon heterocycle on the surface of magnetic Fe3O4@SiO2, and proved that it could effectively promote the Heck, Suzuki and Sonogashira coupling reactions of aryl iodides, with high reaction yield and stable recycling up to 7 times[54]. Although the catalytic activity of MNP-33 is not outstanding, the stability of the catalyst is greatly improved due to the formation of Pd-C bond.
图式31 MNP-33的合成及其在C-C偶联反应中的应用[54]

Scheme 31 Synthesis of MNP-33 and its application in C-C coupling reactions[54]

In 2017, Tu et al., for the first time, synthesized a novel molecular acenaphthylimidazole-type NHC-palladacycle complex catalyst MNP-34 with a well-defined structure through the strategy of enlarging the conjugated system of NHC[55]. The catalyst has a uniform morphology, an average particle size of 150 nm (TEM), and good thermal stability (DSC-TGA), and can be stored in the environment for many years. MNP-34 exhibits unprecedented catalytic activity of heterogeneous palladium catalysts: efficient Suzuki-Miyaura coupling reactions of inert heteroaryl chlorides with various boronic acids were achieved to give the corresponding functional acridine derivatives at 0.5 mol% catalyst dosage. The catalyst can be stably recycled for 5 times without obvious activity loss. The leaching rate of Pd in the filtrate tested by ICP is negligible. According to TEM, the morphology of the catalyst before and after recovery does not change significantly. The sixth decrease in catalytic activity may be due to the poisoning of the catalyst by the substrate (Figure 32).
图式32 MNP-34的合成及其在Suzuki偶联反应中的应用[55]

Scheme 32 Synthesis of MNP-34 and its application in Suzuki coupling reactions[55]

In 2017, Patil et al. Synthesized catalysts MNP-35 and MNP-36 and applied them to Suzuki coupling and Heck coupling reactions[56,57]. The results showed that both MNP-35 and MNP-36 exhibited excellent catalytic activity for Suzuki coupling reaction of different substrates under mild conditions, with yields up to 95%. In the Suzuki cross-coupling reaction of bromobenzene and phenylboronic acid, MNP-36 could be recycled for 7 times without significant activity loss, and the product could still be obtained in a considerable yield of 81% even after 12 times of recycling. In contrast, MNP-35 could be recycled for 5 cycles, and a decrease in catalytic activity was observed when the sixth cycle was started. This may be due to the change of catalytic activity caused by the change of steric hindrance effect caused by the addition of nitro group on the phenyl ring in the N-ligand. At the same time, under the optimized conditions, in the Heck cross-coupling reaction of bromobenzene and styrene as a model reaction, there was no significant decrease in the activity of MNP-36 after five cycles. However, further recycling led to a decrease in the yield of the cross-coupled product, which was still 74% after 12 recycles. One of the reasons for the reduced recovery efficiency of the Heck cross-coupling compared with the Suzuki cross-coupling reaction may be the difference in the reaction conditions required for the two coupling reactions (Scheme 33).
图式33 MNP-35和MNP-36的合成及其在Suzuki偶联反应中的应用[56,57]

Scheme 33 Synthesis of MNP-35 and MNP-36 and their application in Suzuki coupling reactions[56,57]

In addition, vitamin B1 is a water-soluble vitamin and a degradable green thiazolyl nucleophilic heterocyclic carbon, which can be used as a biodegradable and non-toxic NHC ligand to replace the toxic and poorly biodegradable imidazolyl NHC ligand. In 2018, Rafiee et al. Synthesized catalyst MNP-37 by covalently linking thiamine hydrochloride (VB1) with Fe3O4@SiO2 to form a biodegradable complex with palladium, and applied it to Suzuki coupling reaction[58]. Under the optimal conditions, only 0. 001 G of catalyst (0. 022 mmol%, Pd) can effectively catalyze the coupling reaction of various aryl halides with substituted phenylboronic acids, and the corresponding products can be obtained in a short time with a yield of 98%. However, the catalytic system has low reactivity for -Cl-based aromatic substituents, and the catalyst can be recycled five times (Figure 34).
图式34 MNP-37的合成及其在Suzuki偶联反应中的应用[58]

Scheme 34 Synthesis of MNP-37 and its application in Suzuki coupling reactions[58]

Pincer pyridine NHCs have attracted much attention because of their trinity rigid structure, which makes the metal complexes have high stability against air and moisture. Rafiee et al. Prepared a magnetic chitosan-supported Pd-CNC-type pincer complex catalyst (MNP-38) by grafting modified chitosan onto MNP, and applied it to Suzuki coupling reaction.Under the optimal conditions, the catalytic system can effectively catalyze the coupling reaction of various aryl halides and arylboronic acids, and the corresponding biphenyls can be obtained in high yields except for the coupling reaction involving aryl chlorides[59]. The catalyst can be stably recycled for 13 times. In this pincer composite, the metal atom is more tightly bound to the ligand, which is particularly stable and durable compared to traditional pincer complexes containing heteroatoms in the side arm (Scheme 35).
图式35 MNP-38的合成及其在Suzuki偶联反应中的应用[59]

Scheme 35 Synthesis of MNP-38 and its application in Suzuki coupling reactions[59]

In 2018, Sardarian et al. Reported theophylline-NHC-Pd (II) catalyst MNP-39 with Pd (II) coordination compound supported on MNPs, and used it in Suzuki and Sonogashira coupling reactions[60]. MNP-39 was able to efficiently catalyze the Suzuki and Sonogashira cross-coupling reactions of various aryl halides with phenylboronic acids, terminal aromatics, and aliphatic alkynes in good yields. Compared with other supported catalysts, the reaction conditions are mild and the catalyst content is less (only 0.37 ~ 0.5 mol% under the optimal conditions). As a stable phosphine-free Pd catalyst, MNP-39 showed good stability, could be recovered by magnetic separation and repeated for 8 reaction cycles, no significant activity loss was observed, and the leaching rate of palladium was less than 0.18 ppm (Scheme 36).
图式36 MNP-39的合成及其在C-C偶联反应中的应用[60]

Scheme 36 Synthesis of MNP-39 and its application in C-C coupling reactions[60]

In 2019, Tahmasbi et al. Formed magnetic MCM-41 nanoparticles by doping Fe3O4 nanoparticles functionalized by silica layer on the MCM-41 framework, and then immobilized the N-heterocyclic carbene-coordinated Pd (II) complex on the surface of magnetic MCM- 41 by APTES or CPTMS to prepare magnetic nanocatalysts MNP-40-A and B, respectively (Fig. 37)[61][62]. It has been shown that the two catalysts are widely suitable for the coupling of various aryl halides (Suzuki reaction and Heck reaction). Under the optimized experimental conditions, the surface structure and morphology of catalyst MNP-40-a did not change when it was recycled, and it could be reused for 8 times with an average separation rate of 97.5%. In addition, the AAS results showed that the recycled catalyst MNP-40-B could be reused for 7 times, and the average separation rate could reach 93. 14%, and the activity of both was not reduced.
图式37 MNP-40的合成及其在C-C偶联反应中的应用[61,62]

Scheme 37 Synthesis of MNP-40 and its application in C-C coupling reactions[61,62]

In addition, our group successfully prepared a core-shell structure of magnetic mesoporous nanoparticle-supported NHC-palladacycle catalyst MNP-41 in 2022, which exhibited broad substrate adaptability in Suzuki cross-coupling of arylboronic acids and arylhalides[63]. Moreover, MNP-41 has high catalytic activity and reusability, and its catalytic performance does not decrease significantly even after 12 cycles (Figure 38).
图式38 MNP-41的合成及其在Suzuki偶联反应中的应用[63]

Scheme 38 Synthesis of MNP-41 and its application in Suzuki coupling reactions[63]

7 Conclusion and prospect

The catalytic performance of heterogeneous catalysts is not only related to the inherent characteristics of the support and the link mode between the dominant catalyst and the support, but also closely related to the specific structure of the dominant catalyst fragment, that is, the coordination mode of the active species itself. In recent years, magnetic nanoparticles (MNPs) have become an attractive carrier due to their dual advantages of magnetic properties and nanosize effect. MNPs @ L-Pd has also attracted much attention because of its simple synthesis, easy structural characterization, easy separation, and high activity. MNPs @ L-Pd has not only been developed with rich support types and loading strategies, which has the advantages of easy in situ magnetic separation and recycling, but also designed a variety of superior catalyst fragments through the multidentate coordination of C, N, O, P, S, Se and other coordination atoms.It has high catalytic activity and stability in C-C bond formation reactions with aryl halides and heterocyclic halides as substrates, and even has certain catalytic activity and universality for aryl chlorides with low reactivity (Table 1-3), showing good catalytic activity and stability, and has certain industrial application prospects.
表1 不同磁性纳米颗粒负载的催化剂对Suzuki-Miyaura反应的性能

Table 1 Performance of catalysts supported by different magnetic nanoparticles for Suzuka-Miyaura reaction

Pd catalyst (mol%) Standard conditions Yield(%) Reusability TOF(h-1) ref
MNP-2 (0.14 mol% Pd) K2CO3, EtOH/H2O(1∶1), 80 ℃, 50 min 95% 8 814.28 20a
MNP-3 (0.01 mol% Pd) K2CO3, EtOH/H2O(1∶1), 70 ℃, 30 min 96% 12 19 200 21a
MNP-4 (0.1 mol% Pd) Et3N or Na2CO3, EtOH/H2O(1∶1), 80 ℃,
30 min
99% 7 1980 22b
MNP-10 (0.009 mol% Pd) K2CO3, EtOH/H2O(1∶1), 75 ℃, 1.0 h 97% 5 10 778 29a
MNP-12 (0.47 mol% Pd) K2CO3, H2O, 60~90 ℃, 20 min 95% 6 612 31a
MNP-13 (0.1 mol% Pd) K2CO3, H2O, 90 ℃, 1.0 h 95% 6 950 32a
MNP-14 (0.09 mol% Pd) K2CO3, H2O, reflux, 1.0 h 93% 6 1033.3 33a
MNP-18 (0.01 mol% Pd) Et3N, H2O, 80 ℃, 45 min 91% 9 12 133.3 39a
MNP-20 (0.04 mol% Pd ) K2CO3, EtOH/H2O(2∶1), 60 ℃, 1.5 h 95% 7 1583.3 41a
MNP-21 (0.5 mol% Pd) Et3N, DMF, 100 ℃, 3 h 92% 10 61.3 42a
MNP-23 (0.3 mol% Pd) K2CO3, NMP, 100 ℃, 2.5 h 88% 8 117 44a
MNP-27 (0.017 mol% Pd) K2CO3, EtOH/H2O(1∶1), 80 ℃, 3.0 h 86% 7 1686.3 49a
MNP-28 (0.34 mol% Pd) K2CO3, EtOH/H2O(2∶1), 60 ℃, 3.0 h 92% 20 6903 50a
MNP-29 (0.825 mol% Pd) K2CO3, NMP, 90 ℃, 1.0 h 88% 6 107 51a
MNP-34 (0.5 mol%) K3PO4, Toluene, 100 ℃, 24.0 h 99% 7 8.25 55c
MNP-35 (0.15 mol% Pd) K2CO3, EtOH/H2O(1∶1), 70 ℃, 1.0 h 95% 5 633.3 56a
MNP-36 (0.15 mol% Pd) K2CO3, EtOH/H2O(1∶1), R.T, 2.0 h 95% 7 316.7 57a
MNP-37 (0.022 mmol% Pd ) Na2CO3, EtOH, 60 ℃, 20 min 95% 5 12 954.5 58a
MNP-38 (0.021 mmol% Pd) NaHCO3, EtOH/H2O(1∶1), 70 ℃, 10 min 98% 13 2940×104 59a
MNP-39 (0.37 mol% Pd ) K2CO3, H2O, 60 ℃, 3.0 h 96% 8 86.5 60a
MNP-40-A (1.5 mol% ) Na2CO3, PEG-400, 80 ℃, 100 min 88% 8 35.2 61a
MNP-40-B (0.83 mol% ) Na2CO3, PEG-400, 80 ℃, 3.0 h 93% 7 37.3 62a
MNP-41 (0.5 mol%) K2CO3, EtOH/H2O(2∶1), 60 ℃, 12.0 h 93% 12 15.5 63d

a. Suzuki-Miyaura reaction of bromobenzene with phenylboronic acid.; b. Suzuki-Miyaura reaction of bromobenzene with 4-tolylboronic acid.; c. Suzuki-Miyaura reaction of 9-chloroacridine with phenylboronic acid.; d. Suzuki-Miyaura reaction of 4-chloroanisole with phenylboronic acid.

表2 不同磁性纳米颗粒负载的催化剂对Heck反应的性能

Table 2 Performance of catalysts supported by different magnetic nanoparticles for Heck reaction

Pd catalyst (mol%) Standard conditions Yield(%) Reusability TOF(h-1) ref
MNP-5 (3.58 mol%) K2CO3, DMF, 100 ℃, 8 h 93% 8 3.25 23a
MNP-6 (1.24 mol% Pd) Et3N, DMSO, 100 ℃, 1.5 h 92% 6 49 24a
MNP-10 (0.009 mol% Pd) Et3N, Solvent-free, 120 ℃, 40 min 96% 5 16000 29b
MNP-12 (0.61mol% Pd) K2CO3, H2O, 80~90 ℃, 50 min 95% 6 187 31c
MNP-13 (0.1 mol% Pd) K2CO3, H2O/DMF(1∶1), reflux, 8.0 h 95% 6 122.5 32a
MNP-14 (0.09 mol% Pd) K2CO3, H2O, reflux, 12 h 96% 6 88.8 33d
MNP-15 (0.71 mol% Pd) Et3N, DMF, TBAB, 120 ℃, 20 min 93% 10 393 34b
MNP-19 (0.15 mol% Pd ) Et3N, DMF, 120 ℃, 45 min 98% 5 871.1 40b
MNP-21 (0.5 mol% Pd) Et3N, DMF, 100 ℃, 0.25 h 98% 10 784 42c
MNP-23 (0.3 mol% Pd) K2CO3, DMF, 110 ℃, 0.75 h 97% 8 430 44c
MNP-24 (0.2 mol% Pd) K2CO3, H2O/DMF(2∶1), 90 ℃, 0.5 h 95% 8 950 45c
MNP-25 (0.08 mol% Pd) NaOAc, H2O, R.T., 1.0 h 98% 10 1225 46a
MNP-29 (0.99 mol% Pd ) K2CO3, NMP, 110 ℃, 0.5 h 96% 6 194 51c
MNP-30 (0.0435 mol% Pd) NaOAc, H2O, R.T., 1.0 h 95% 10 2183.9 52a
MNP-36 (1.0 mol% Pd) Na3PO4·12H2O, MeCN, 80 ℃, 4.0 h 96% 7 24 57e
MNP-40-A(2.27 mol% ) Na2CO3, PEG-400, 120 ℃, 80 min 96% 7 31.7 61c
MNP-40-B(1.46 mol% ) Na2CO3, PEG-400, 120 ℃, 1.0 h 98% 7 67.1 62c

a. R=Ph; b. R=COOMe; c. R=COOBun; d. R=COOH; e. R=COOtBu

表3 不同磁性纳米颗粒负载的催化剂对其他C-C偶联反应的性能

Table 3 Performance of catalysts supported by different magnetic nanoparticles for other C-C coupling reactions

Pd catalyst (mol%) Standard conditions Yield(%) Reusability TOF(h-1) ref
MNP-21(0.5 mol% Pd) Et3N, DMF, 100 ℃, 1.5 h 96% 10 128 42a
MNP-18(0.1 mol% Pd) Et3N, H2O, 80 ℃, 0.5 h 90% 9 1800 42b
MNP-18(0.2 mol% Pd) Et3N, H2O, 80 ℃, 6.0 h 99% 9 82.5 42c
MNP-22 (0.7 mol% Pd) Et3N, DMF, 90 ℃, 0.5 h 94% 6 268.6 43a
MNP-24 (0.2 mol% Pd) K2CO3, H2O/DMF(2∶1), 90 ℃, 1.0 h 96% 8 480 45a
MNP-25 (0.08 mol%) NaOAc, H2O, R.T., 1.0 h 94% 10 1175 46a
MNP-30(0.0435 mol% Pd) NaOAc, H2O, R.T., 1.0 h 95% 10 2184 52a
MNP-39(0.43 mol% Pd ) Piperidine, Solvent-free, 90 ℃, 1.5 h 95% 8 147.3 60a

a. Sonogashira reaction of Iodobenzene with phenylacetylene; b. Cyanation Reaction of Iodobenzene with K4[Fe(CN)6]·3H2O;c. Hiyama Cross-Coupling Reaction of Halobenzenes with Triethoxyphenylsilane.

Although the development and application of MNPs @ L-Pd have made great progress in recent years, there are still many challenges. (1) According to the existing work progress and development trend in this field, it is not difficult to find that the magnetic nanocatalyst supported by organic palladium complex still has the defects of poor thermal stability, short service life of the catalyst, and is not conducive to continuous production. Therefore, the design of MNPs @ L-Pd catalysts with higher support stability and coordination stability is one of the goals pursued by researchers. (2) The existing catalytic system can catalyze a single type of reaction, and the catalytic amount in the reaction is high, the reaction temperature is not mild enough, the applicability of the substrate needs to be expanded, and the catalyst activity needs to be improved. It is urgent to develop diversified new catalytic materials with high loading and efficient catalytic activity through the careful design of supports and superior catalyst ligands. (3) There is also much room for research on the preparation process of MNPs supported palladium catalyst and its catalytic mechanism in the coupling reaction. (4) MNPs supported chiral catalysts have good application prospects in the field of asymmetric catalysis, but they are still in the initial stage and are worthy of further exploration. (5) With the development of organic technology, more and more practical needs have been put forward, and in the past few years, the combination of photocatalysis, electrocatalysis and organocatalysis has made remarkable progress in modern chemical synthesis[64~66][67]. At present, it has been proved that compared with heat source, the rational application of light and electricity can achieve catalysis under mild conditions, which is conducive to the recycling of catalysts[68]. Therefore, the cross-domain synergistic catalysis of organometallic catalysts needs to be studied and expanded urgently.
[1]
Zhao G, Xu X J. Nanoscale, 2021, 13 (24): 10649.

[2]
Vogel P, Lam Y H, Simon A, Houk K N. Catalysts, 2016, 6 (9): 128.

[3]
Gabriele B. Molecules, 2022, 27 (4): 1227.

[4]
Niu B, Yang K, Lawrence B, Ge H. ChemSusChem, 2019, 12 (13): 2955.

[5]
Wang Q J, Su Y J, Li L X, Huang H M. Chem.Soc.Rev., 2016, 45 (5): 1257.

[6]
Xia Y, Qiu D, Wang J B. Chem.Rev., 2017, 117 (23): 13810.

[7]
Sain Shalu, Jain S, Srivastava M, Vishwakarma R, Dwivedi J. Curr. Org. Synth., 2019, 16 (8): 1105.

[8]
Wu X F, Anbarasan P, Neumann H, Beller M. Angew. Chem. Int. Ed., 2010, 49 (48): 9047.

[9]
Beller M, Zapf A, Mägerlein W. Chem. Eng. Technol., 2001, 24 (6): 575.

[10]
Dobrounig P, Trobe M, Breinbauer R. Monatsh. Chem., 2017, 148 (1): 3.

[11]
Cole-Hamilton D J, Tooze R P. Catalyst Separation, Recovery and Recycling. New York: Springer, 2006.

[12]
Mallesham B, Raikwar D, Shee D. Advanced Functional Solid Catalysts for Biomass Valorization. Eds.: Mustansar H C, Sudarsanam P. Amsterdam: Elsevier, 2020.1.

[13]
Liu J, Zhu Q R, Du J, Zhang X L. Chin. J. Org. Chem., 2015, 35(1): 15.

( (刘杰, 朱庆仁, 杜娟, 张袖丽. 有机化学, 2015, 35(1): 15.)

[14]
Franzen R G. Top. Catal., 2016, 59 (13): 1143.

[15]
Qiao N N. Mater Dissertation of Guangdong Pharmaceutical University, 2019.

(乔妮娜. 广东药科大学硕士论文, 2019.).

[16]
Yan T L, Zhang Y, Zhu X Q, Zhang X H, Liu Z T, Dai J T, Xu G D. Journal of Yangzhou University Natural Science Edition, 2022, 25(1): 19.

( 颜廷良, 张玉, 朱啸庆, 张雪华, 刘总堂, 戴兢陶, 徐国栋. 扬州大学学报(自然科学版), 2022, 25(1): 19.)

[17]
Mu J L, Hu T W, Shan S Y, Jiang L H, Wang Y M, Jia Q M. Polymeric Materials Science and Engineering, 2013, 29(7): 179.

母佳利, 胡庭维, 陕绍云, 蒋利红, 王亚明, 贾庆明. 高分子材料科学与工程, 2013, 29(7): 179.).

[18]
Wang H B. Mater Dissertation of Lanzhou University, 2013.

(王海博. 兰州大学硕士论文, 2013.).

[19]
Bui N T, Dang T B, Le H V, Phan N T S. Chin. J. Catal., 2011, 32 (11/12): 1667.

[20]
Beygzadeh M, Alizadeh A, Khodaei M M, Kordestani D. Catal. Commun., 2013, 32: 86.

[21]
Yang P B, Ma R, Bian F L. ChemCatChem, 2016, 8 (24): 3746.

[22]
Nehlig E, Waggeh B, Millot N, Lalatonne Y, Motte L, GuÉnin E. Dalton Trans., 2015, 44 (2): 501.

[23]
Bodaghifard M A. J. Organomet. Chem., 2019, 886: 57.

[24]
Aryanasab F, Shabanian M, Laoutid F, Vahabi H. Appl. Organomet. Chem., 2021, 35 (5): e6198.

[25]
Dehbanipour Z, Zarnegaryan A. Inorg. Chem. Commun., 2022, 141.

[26]
Xu X Q, Wang W Q, Lu L, Zhang J Z, Luo J. Catal. Lett., 2022, 152 (10): 3031.

[27]
Cai Y, Yuan H L, Gao Q, Wu L L, Xue L J, Feng N J, Sun Y. Catal. Lett., 2023, 153(2): 460.

[28]
Bauer R E, Grimsdale A C, Müllen K. Functional Molecular Nanostructures. Ed.: Schlüter A D. Berlin, Heidelberg: Springer, 2005. 253.

[29]
Ma R, Yang P B, Bian F L. New J. Chem., 2018, 42 (6): 4748.

[30]
Sorokina S A, Kuchkina N V, Lawson B P, Krasnova I Y, Nemygina N A, Nikoshvili L Z, Talanova V N, Stein B D, Pink M, Morgan D G, Sulman E M, Bronstein L M, Shifrina Z B. Appl. Surf. Sci., 2019, 488: 865.

[31]
Sheikh S, Nasseri M A, Chahkandi M, Reiser O, Allahresani A. RSC Adv., 2022, 12 (15): 8833.

[32]
Yang J H, Wang D F, Liu W D, Zhang X, Bian F L, Yu W. Green Chem., 2013, 15 (12): 3429.

[33]
Wang D F, Liu W D, Bian F L, Yu W. New J. Chem., 2015, 39 (3): 2052.

[34]
Barazandehdoust M, Mamaghani M, Kefayati H. React. Kinet., Mech. Catal., 2020, 129 (2): 1007.

[35]
Khoobi MM, Ma?mani L, Rezazadeh F, Zareie Z, Foroumadi A, Ramazani A, Shafiee A. J. Mol. Catal. A: Chem., 2012, 359: 74.

[36]
Ma?mani L, Sheykhan M, Heydari A, Faraji M, Yamini Y. Appl. Catal., A, 2010, 377 (1): 64.

[37]
Mohammadnia M, Poormirzaei N. Nanoscale Adv., 2021, 3 (7): 1917.

[38]
Zarnaghash N, Panahi F, Khalafi-Nezhad A. J. Iran. Chem. Soc., 2015, 12 (11): 2057.

[39]
Sobhani S, Habibollahi A, Zeraatkar Z. Org. Process Res. Dev., 2019, 23 (7): 1321.

[40]
Rangraz Y, Nemati F, Elhampour A. New J. Chem., 2018, 42 (18): 15361.

[41]
Rangraz Y, Nemati F, Elhampour A. J. Phys. Chem. Solids, 2020, 138: 109251.

[42]
Sobhani S, Zeraatkar Z, Zarifi F. New J. Chem., 2015, 39 (9): 7076.

[43]
Esmaeilpour M, Sardarian A R, Javidi J. J. Organomet. Chem., 2014, 749: 233.

[44]
Esmaeilpour M, Javidi J. J. Chin. Chem. Soc., 2015, 62 (7): 614.

[45]
Sardarian A R, Kazemnejadi M, Esmaeilpour M. Dalton Trans., 2019, 48 (9): 3132.

[46]
Tashrifi Z, Bahadorikhalili S, Lijan H, Ansari S, Hamedifar H, Mahdavi M. New J. Chem., 2019, 43 (23): 8930.

[47]
Liao Q L, Tannenbaum R, Wang Z L. J. Phys. Chem. B, 2006, 110 (29): 14262.

[48]
Qiu J P, Yu L, Ni J G, Fei Z X, Li W D, Sadeghzadeh S M. New J. Chem., 2020, 44 (4): 1269.

[49]
Rahman M, L Sarjadi M S, Akhter M S, Hannan J J, Sarkar S M. Arabian J. Chem., 2022, 15 (8): 103983.

[50]
Ren F F, Li S, Zheng W Q, Song Q Y, Jia W H, Nan Y Q, Jia H J, Liu J, Bao J J, Li Y X. Colloids Surf., A, 2022, 642: 128611.

[51]
Esmaeilpour M, Javidi J, Dodeji F N, Hassannezhad H. J. Iran. Chem. Soc., 2014, 11 (6): 1703.

[52]
Bahadorikhalili S, Ma?mani L, Mahdavi H, Shafiee A. RSC Adv., 2015, 5 (87): 71297.

[53]
Martinez-Olid F, AndrÉs R, Flores J C, GÓmez-Sal P, HeuzÉ K, Vellutini L. Dalton Trans., 2016, 45 (29): 11633.

[54]
Kaur A, Singh V. Chem. Lett., 2016, 45 (1): 83.

[55]
Deng Q Y, Shen Y J, Zhu H B, Tu T. Chem. Commun. (Camb), 2017, 53 (97): 13063.

[56]
Vishal K, Fahlman B D, Sasidhar B S, Patil S A, Patil S A. Catal. Lett., 2017, 147 (4): 900.

[57]
Kandathil V, Fahlman B D, Sasidhar B S, Patil S A, Patil S A. New J. Chem., 2017, 41 (17): 9531.

[58]
Rafiee F, Mehdizadeh N. Catal. Lett., 2018, 148 (5): 1345.

[59]
Rafiee F, Hosseini S A. Appl. Organomet. Chem., 2018, 32 (11): e4519.

[60]
Esmaeilpour M, Sardarian A R, Firouzabadi H. J. Organomet. Chem., 2018, 873: 22.

[61]
Tahmasbi B, Ghorbani-Choghamarani A. New J. Chem., 2019, 43 (36): 14485.

[62]
Ghorbani-Choghamarani A, Tahmasbi B, Hudson R H E, Heidari A. Micropor. Mesopor. Mater., 2019, 284: 366.

[63]
Zuo B, Shao H, Zheng Y, Ma Y H, Li W F, Huang M X, Deng Q Y. Asian J. Org. Chem., 2022, 11 (5): e202200018.

[64]
Lakhdar S. Synthesis, 2020, 52 (22): 3492.

[65]
Liu Y Y, Liu J, Lu L Q, Xiao W J. Top Curr. Chem. (Cham), 2019, 377 (6): 37.

[66]
Zhou W J, Cao G M, Zhang Z P, Yu D G. Chem. Lett., 2019, 48 (3): 181.

[67]
Sauermann N, Meyer T H, Qiu Y A, Ackermann L. ACS Catal., 2018, 8 (8):7086.

[68]
Luscombe C K, Yadav P, Velmurugan N. Synthesis, 2023, 55(01): 1.

Outlines

/