Catalysis Research is an international peer-reviewed Open Access journal published quarterly online by LIDSEN Publishing Inc. This periodical is devoted to publishing high-quality papers that describe the most significant and cutting-edge research in all areas of catalysts and catalyzed reactions. Its aim is to provide timely, authoritative introductions to current thinking, developments and research in carefully selected topics.

Topics contain but are not limited to:

  • Photocatalysis
  • Electrocatalysis
  • Environmental catalysis
  • Biocatalysis, enzymes, enzyme catalysis
  • Catalysis for biomass conversion
  • Organocatalysis, catalysis in organic and polymer chemistry
  • Nanostructured Catalysts
  • Catalytic materials
  • Computational catalysis
  • Kinetics of catalytic reactions

It publishes a variety of article types: Original Research, Review, Communication, Opinion, Comment, Conference Report, Technical Note, Book Review, etc.

There is no restriction on paper length, provided that the text is concise and comprehensive. Authors should present their results in as much detail as possible, as reviewers are encouraged to emphasize scientific rigor and reproducibility.

Publication Speed (median values for papers published in 2023): Submission to First Decision: 4.1 weeks; Submission to Acceptance: 13.0 weeks; Acceptance to Publication: 9 days (1-2 days of FREE language polishing included)

Current Issue: 2024  Archive: 2023 2022 2021
Open Access Original Research

Fabrication and Catalytic Property of an Ordered Terpyridine Pd(II)/Ni(II) Catalytic Monolayer for Suzuki Coupling Reactions

Wei Wang 1, †, Sa Bi 1, †, Huanhuan Li 2, Tiesheng Li 1,*

  1. College of Chemistry, Zhengzhou University, Zhengzhou 450001, Henan Province, P.R. China

  2. Zhengzhou University of Technology, Zhengzhou 450044, Henan Province, P.R. China

† These authors contributed equally to this work.

Correspondence: Tiesheng Li

Academic Editor: Jacques Muzart

Special Issue: Advanced in Palladium-Catalyzed Reactions

Received: September 21, 2022 | Accepted: November 17, 2022 | Published: December 02, 2022

Catalysis Research 2022, Volume 2, Issue 4, doi:10.21926/cr.2204042

Recommended citation: Wang W, Bi S, Li H, Li T. Fabrication and Catalytic Property of an Ordered Terpyridine Pd(II)/Ni(II) Catalytic Monolayer for Suzuki Coupling Reactions. Catalysis Research 2022;2(4):25; doi:10.21926/cr.2204042.

© 2022 by the authors. This is an open access article distributed under the conditions of the Creative Commons by Attribution License, which permits unrestricted use, distribution, and reproduction in any medium or format, provided the original work is correctly cited.

Abstract

Fabrication, arrangement, and controllable composition of ordered organometallic films are critical for designing a highly active catalyst and investigating the catalytic mechanism. In this paper, an organometallic terpyridine Pd(II)/Ni(II) monolayer linked on the silicon substrate surface (denoted as Si-Tpy-Pd1/Ni1) was prepared and characterized using water contact angle, ultraviolet spectra, X-ray diffraction, scanning electron microscopy, and X-ray photoelectron microscopy. Si-Tpy-Pd1/Ni1 exhibited high catalytic activity, substrate applicability, and reusability after 5 runs. During recycling, the deactivation was induced by the aggregation of active Pd/Ni nanoparticles. The catalytic mechanism was heterogeneous and occurred on the Si-Tpy-Pd1/Ni1 monolayer surface; the mechanism was confirmed using hot filtrate, poison test, and a three-phase experiment. The real active center was Pdδ/Niδ+ and was formed in situ on the organometallic monolayer surface, which acted as a precursor with a synergistic effect between Pd and Ni. The electron density of Pd became more negative because of electron transfer from Ni to Pd, which facilitated the oxidative addition reaction.

Keywords

Palladacycle; self-assembled monolayer; Pd/Ni bimetallic catalyst; Suzuki coupling reaction; synergistic effect

1. Introduction

Palladium (Pd)-catalyzed cross-coupling is a useful tool in organic synthesis [1]. Hetero-multimetallic catalysts exhibit unique properties because of the synergistic effects between different metals [2]. Hetero-bimetallic nano-catalysts, including Pd-based or nickel-based catalysts, were widely applied in synthesis because of their high activity, stability, and recyclability [3,4,5,6]. However, the active sites easily aggregate, with the changes in active sites being difficult to detect during catalysis [7], leading to difficulty in elucidating the true mechanism underlying heterogeneous catalysis [8].

The ordered multi-organometallic monolayer has been widely developed in catalytic systems, retaining most of the performances as precursors [9,10]. The self-assembly method can offer desired design, control the orientation, and facilitate easy recycling in the forms of films or monolayers [11,12,13]. The surface properties of organometallic catalytic monolayers, such as Langmuir–Blodgett (LB) films [14,15,16] and self-assembled films supported on solid supports, can be easily characterized, identified, and evaluated using various scientific techniques [17]. Such studies can provide deep insights into catalytic behavior at the molecular level, such as ligands containing certain functional groups (e.g., Schiff bases) exhibiting excellent performance in heterogeneous catalysis.

Studies have demonstrated that the catalytic activity of the mono- or bi-organometallic catalytic monolayers can be improved by quantifying film morphology, tuning the composition of different metals, elucidating the relative distributions of electrons (which are related to the electrical characteristics of the ligand), and modifying the support and synergy of hetero-bimetals [18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35]. Appropriate supports, such as silicon with a greater flat surface, can result in a high dispersion and reproducibility of ordered organometallic catalytic monolayers [21,22,23,25,36,37,38]. Pd-based catalysts doped with other metals not only have less amount of Pd but also have high catalytic activity enhanced by the adjustment of the distance between Ni and Pd and the distribution and electronic structure of active sites [25,27,31]. Application of Pd catalysts exhibiting catalyst poisoning is restricted and is a major challenge [1]. Nickle (Ni) has the same chemical property as Pd [39]. Ni is non-expensive and more nucleophilic because of its smaller size compared with Pd, a characteristic critical in Suzuki coupling reaction; moreover, the catalytic performances of Ni and Pd vary [40,41]. Therefore, the combined use of Pd and Ni in the catalysis of the Suzuki cross-coupling reaction has attracted much attention [42,43,44]. For this, the development of a regular, cost-effective, and eco-friendly film to enhance the catalytic property and elucidate synergy is warranted [45,46,47].

In this paper, an organometallic compound containing a terpyridine group acting as the binding site for Pd and Ni and its terpyridine Pd/Ni catalytic monolayer grafted onto the surface of silicon substrate was fabricated through self-assembly (SA). The compound was characterized, and its catalytic property in the Suzuki coupling reaction, the structure of the active sites, and the synergistic effect were explored.

2. Materials and Methods

2.1 Reagent, Instruments, and General Procedure

Chemical reagents were obtained from chemical companies. Solvents were treated and distilled before use.

X-Ray diffraction was performed using a PAN analytical X-Pert PRO. The water contact angle (WCA) was recorded using a HARKE-SPCX1 (Beijing). UV–vis spectra were recorded using a Lambda 35 (UV–vis) spectrophotometer (Perkin Elmer Inc., USA). An SPM-9500J3 (Shimadzu Corporation, Japan) was employed for atomic force microscopy (AFM) measurements in the air under ambient conditions. Scanning electron microscopy images were measured using a Hitachi S-4800 instrument. X-ray photoelectron spectroscopy images were measured using an ESCALab220i-XL electron spectrometer (VG Scientific, 300 W Al Kα radiation).

The Pd and Ni content in the catalysts before and after the cross-coupling reactions was measured using inductively coupled plasma atomic emission spectroscopy (ICP-AES) with an ICAP 6000 Series (Thermo Scientific).

1H NMR and 13C NMR spectra were recorded using a Bruker Avance III 400 MHz spectrometer in CDCl3 solution with tetramethyl silane as an internal standard. A VERTEX 70 V spectrometer (Bruker Optik, Ettlingen, Germany) at 293 K, with a spectral resolution of 4 cm–1 and a scanner velocity of 10 kHz, was used for monitoring the coupling reaction of 4-bromotoluene with phenylboronic acid.

2.2 Fabrication of Si-Tpy-Ni/Pd

This work combined two organometallics in a monolayer on supports by using self-assembly to improve the activity of active sites. The bi-organometallic catalytic monolayer was designed and prepared (Scheme 1).

Click to view original image

Scheme 1 Fabricating route of a catalytic monolayer on a silicon wafer (Si-Tpy-Pd/Ni).

Silicon and quartz substrates (3 cm × 1 cm × 0.1 cm) were sonicated in ethanol (30 mL) and deionized water (30 mL) for 15 min. Subsequently, the quartz slide or silicon wafer was immersed in H2SO4/HNO3 (1:1; v:v) solution at 90 °C or in “piranha” solution (H2O2/H2SO4, 1:2; v:v) at 80 °C for 1.5 h to remove any organic residues and to create silanol groups on the surfaces (Si-OH). The cleaned slides were rinsed extensively with water (50 mL), ethanol (30 mL), and acetone (30 mL), followed by drying in a stream of nitrogen.

The freshly prepared hydroxyl silicon or quartz (Si-OH) (3 cm × 1 cm × 0.1 cm) surfaces were placed in a solution containing 0.5 mL of 3-aminopropyl triethoxysilane (1.1 mmol) and 10 mL of toluene. After treatment at 80 °C for 24 h, the surfaces were washed with toluene (50 mL), methanol, water, and acetone, followed by drying in a nitrogen stream (Si-APTES). The amino-functionalized substrates were immersed in 10 mL of dried DMF, containing 4-([2,2':6',2''-Terpyridin]-4'-yl) benzoic acid (Tpy, 1.2 mmol), 2-dimethylaminopyridine (DMAP, 1.2 mmol), and 1-Ethyl-(3-dimethylaminopropyl) carbodiimide salt (EDC·HCl, 1.5 mmol) at 80 °C for 24 h. The resulting substrates were removed and rinsed with ethanol (50 mL), deionized water (50 mL), and acetone (50 mL) to obtain the modified surfaces (Si-Tpy).

Si-Tpy (3 cm × 1 cm × 0.1 cm) was immersed in 30 mL of EtOH (anhydrous) and sonicated for 0.5 h. Subsequently, a mixture with a certain mole ratio of Ni to Pd was added at 40 °C and incubated for 12 h. Si-Tpy-Pd/Ni was removed and washed with methanol (50 mL), ethyl acetate (50 mL), and ethanol (50 mL), followed by vacuum drying at 40 °C for 24 h.

The real atomic ratios in the prepared Si-Tpy-Pd/Ni (3 cm × 1 cm × 0.1 cm) were further determined using ICP–AES. For sample treatment, the sample was broken down through nitrolysis with concentrated nitric acid (5 mL). Subsequently, the solution was transferred into a 10 mL volumetric flask to fix its quantity in water.

2.3 Procedure for Suzuki Reaction and Recycling

Si@Typ-Pd/Ni (3 cm × 1 cm × 0.1 cm), a base (0.2 mmol), and a reactant (aryl halide derivatives, 0.10 mmol; arylboronic acid derivatives, 0.15 mmol) were added to a 10 mL round-bottom flask with 7 mL solvent (H2O: EtOH, v:v = 1:5). The reaction was performed in an oil bath at 80 °C for 12 h, with the subsequent separation of catalysts from the reaction mixture. The residue solution was concentrated, and the cross-coupling product was separated using silica gel chromatography with ethyl acetate as eluent. The product was then purified two times using chromatography.

The recycling experiments were performed under the aforementioned conditions. The used catalyst was recovered from the reaction mixture after each run. The catalyst was reused in sequential runs after washing with ethyl acetate (20 mL), methyl alcohol (20 mL), and water (30 mL) three times.

3. Results and Discussion

3.1 Characterization of Si@Tpy-Pd/Ni

The fabricating processes were characterized. First, the water contact angle (WCA) was used to measure the wettability of functionalized surfaces [48,49]. The silicon wafer surface was hydrophilic because of the hydroxyl groups formed (2.0°) after treatment with the “piranha” solution (Figure 1). The WCA increased to 38° after treatment with APTES and further increased to 54° when treated with terpyridine derivative because of the introduced aryl groups. The WCA of the modified silicon was 45° when Pd2LiCl4 and NiCl2 were added to Si-Typ. The significant changes in the WCAs of the modified surfaces varied during the fabrication step because of the surface configuration of the catalytic monolayer.

Click to view original image

Figure 1 Changes in the silicon surface WCA in the fabrication process (Si-OH, Si-APTES, Si-Typ, and Si-Typ-Pd/Ni).

UV spectra of the fabricating processes of Si-Typ-Pd/Ni monolayer modified on quartz were measured (Figure 2). Hydrophilic quartz (called Si-OH, black line) had no peak, and the absorption at approximately 200 nm increased gradually because of the introduced amine (Si-APTES, red line). An absorption at 290 nm for the terpyridine group appeared after modification with the terpyridine derivative (Si-Typ, blue line). Two new peaks at 223 and 237 nm appeared, indicating coordination with Li2PdCl4 and NiCl2, respectively (green line and pink line). After Si-Typ coordination with a mixture of Li2PdCl4 and NiCl2, the characteristic peak at 290 nm for terpyridine shifted to 282 nm (yellow-green line) because of the ordered monolayer. The vicinal terpyridine complex had a face-to-face configuration, indicating the interaction between the molecules and regular packing.

Click to view original image

Figure 2 The UV–vis spectra of the fabrication self-assembly processes of Si-Tpy-Pd/Ni (Si-OH, black line; Si-APTES, red line; Si-Typ, blue line; Si-Tpy-Pd, green line; Si-Tpy-Ni, pink line; and Si-Tpy-Pd/Ni, yellow-green line).

Low-angle X-ray diffractograms (LAXD) of the Si-Tpy-Pd/Ni preparation process on the silicon surface were measured (Figure 3). A characteristic peak at 1.02° for Si-OH was observed (Figure 3A), which shifted to 1.05° after APTES grafting (Figure 3B). The same peak slightly shifted to 1.03° after a terpyridine derivative was grafted (Figure 3C). The diffraction peak of the monolayer self-assembled with Pd and Ni was also observed approximately at the same position (Figure 3D), indicating coordination between terpyridine group metals to form Si-Tpy-Pd/Ni. Therefore, the ordered hetero-bimetallic self-assembled catalytic monolayer was fabricated on the silicon surface.

Click to view original image

Figure 3 Low X-ray diffraction patterns of Si-Tpy-Pd/Ni. (A) Si-OH, (B) Si-APTES, (C) Si-Tpy, (D) Si-Tpy-Pd/Ni in the assembly processes.

SEM images of Si-OH, Si-APTES, Si-Tpy, and Si-Tpy-Pd/Ni were measured (Figure 4). The surface of the silicon substrate after hydrophilic treatment was smooth and without defects (Figure 4A). Uniform domains on the surface after silanization with APTES were observed (Figure 4B), and similar domains were observed in the image after terpyridine group grafting (Figure 4C). Finally, the uniform monolayer was fabricated after coordination with a mixture of Pd and Ni salts, indicating the formation of the ordered organometallic Pd/Ni monolayer (Figure 4D).

Click to view original image

Figure 4 SEM images of the assembly processes of Si-Tpy-Pd/Ni (A), Si-OH (B), Si-APTES (C), Si-Tpy, and Si-Tpy-Pd/Ni (D) (white bar: 200 nm).

The chemical elements on the surface in the fabrication processes of Si-Tpy-Pd/Ni were analyzed using X-ray photoelectron spectroscopy (Figure 5). The peak at 399.1 eV was attributed to N1s (NH2), and Si 2p (C-Si) was detected in Si-APTES (Figure 5A, black line). Peaks at 401.5 eV and 399.5 eV were attributed to N1s (CONH) and N1s (C = NH); the peaks were observed after terpyridine grafting in Si-Tpy (Figure 5A, red line). Peaks for Pd 3d and Ni 2p were detected after complexing Pd and Ni. Compared with Si-Tpy, in Si-Tpy-Pd/Ni, the HR-XPS of Ni 2p revealed characteristic peaks at 337.92 eV and 343.22 eV, which were attributed to the BE of Pd(II) (Figure 6A). The peaks at 855.67 eV, 861.57 eV (Sat peak), and 873.47 eV were attributed to the BE of Ni(II) (Figure 6B).

Click to view original image

Figure 5 XPS spectra of the self-assembly process of Si-Tpy-Pd/Ni (A), Si-APTES (B), Si-Tpy (C), and Si-Tpy-Pd/Ni.

Click to view original image

Figure 6 High resolution XPS spectra of (A) Pd 3d and (B) Ni 2p.

The contents of Pd and Ni in Si-Tpy-Pd/Ni as measured using ICP-AES were 1.12 × 10–6 mmol·cm–2 and 9.30 × 10–8 mmol·cm–2, respectively. In Si-Tpy-Pd, the content of Pd was 4.69 × 10–6 mmol/cm2, whereas, in Si-Tpy-Ni, the content of Ni was 3.82 × 10–6 mmol/cm2.

AFM images in fabrication processes were also measured (Figure 7). The surface of the silicon substrate exhibited a clear morphology, with arrays having a roughness average (Ra) of 0.49 nm (Figure 7A). Ordered surfaces with a Ra of 2.55 nm (Figure 7B) were observed after APTES grafting. Subsequently, the images exhibited a homogeneous surface with a Ra of 0.83 nm, except for a few defects after terpyridine group grafting (Figure 7C); this indicated that a monolayer may has formed. The surface image of the Si-Tpy-Pd monolayer (Figure 7D) exhibited high-density arrays, with a Ra of 1.19 nm, indicating that the coordination of Pd2+ with Si-Tpy resulted in morphology changes.

Click to view original image

Figure 7 AFM images of a hydrophilic silicon wafer (A) (Si-OH), salinization (B) (Si-APTE), terpyridine grafting (C) (Si-Tpy), terpyridine (D) (Si-Tpy-Pd), and terpyridine Pd (II) grafting.

Characterizations mentioned above confirmed the oriental self-assembly of the Si-Tpy-Pd/Ni monolayer on the silicon wafer surface.

3.2 Catalytic Property of Si-Tpy-Pd/Ni

3.2.1 Evolution of Si-Tpy-Pd/Ni Catalytic Property

The optimization of reaction conditions in the Suzuki cross-coupling reaction was initially explored, and the results are listed in Table 1. Lower yields were obtained with different solvents (entries 1–6), whereas higher yields were achieved using H2O/EtOH (v/v = 1:5) (entries 7–9). The effects of various bases were investigated, and the highest yield (95%) was obtained using K2CO3 (entries 9–13). With a decrease in temperature and time, yield decreased (entries, 14–19), and increased substrate loading resulted in low yields because of the limited amount of catalyst used. The optimized conditions were K2CO3, H2O/EtOH (v/v = 1:5), 4-bromotoluene (0.10 mmol), 80 °C, and 12 h; these conditions were used in further investigation.

Table 1 Optimization of Suzuki–Miyaura reaction conditionsa.

3.2.2 Effect of Composition on Si-Tpy-Pdx/Ni1-x Catalytic Activity

The influence of different ratios of Pd/Ni on Si-Tpy-Pdx/Ni1-x was explored under the optimized conditions, and the results are presented in Table 2. Among the six compositions used, GO-Tpy-Pd1/Ni1 provided the highest yield of up to 96%. The yield of GO-Tpy-Ni and GO-Tpy-Pd were up to 61% and 86%, respectively, indicating the high activity of Ni and Pd complexes. In the case of Pd doped with Ni at a proper ratio, the activity was significantly enhanced. This was because the proper ratio of Pd/Ni effectively supplied favorable active sites on the catalytic monolayer surface to enhance the catalytic activity through a “cooperative effect” [27,50]. Therefore, the novel structure of active sites efficiently came in contact with reactants, thereby ensuring a prompt reaction.

Table 2 Catalytic activity of Si-Tpy-Pdx/Ni1-x with different ratios of Pd/Ni in the Suzuki coupling reaction a.

3.2.3 Screening of Substrates Catalyzed by Si-Tpy-Pd1/Ni1

The Suzuki coupling reaction catalyzed by Si-Tpy-Pd1/Ni1 in the presence of aryl halides and arylboronic acid derivatives was conducted under optimized conditions, and the results are summarized in Table 3. High yields of coupling products were achieved using both electron-donating and electron-withdrawing substituted aryl bromide and arylboronic acids (Table 3, entries 1–8). However, low yields were obtained for aryl chlorides (entry 9, 10) because of the high binding energy of the C-Cl bond. Cross-coupling compounds were also at higher yields by using benzyl bromide with naphthylboronic acid (entry 11). However, lower yields were obtained in the reaction of thiophene derivative with arylboronic acids because of its poison property of thiophene (entry 12). The results confirmed the high activity of Si-Tpy-Pd1/Ni1 in the synthesis of aromatic compounds with a high turnover number (TON) and turnover frequency (TOF) values.

Table 3 Suzuki coupling reaction of aryl halides with different arylboronic acids a.

3.2.4 Effects of Supports and Fabrication

Control experiments were performed to investigate the effect of different supports, such as glass, gel, silicon, and homogeneous Pd/Ni, on their catalytic activities (Table 4).

Table 4 Effect of different supports on the catalytic properties in Suzuki coupling reaction.

Cross-coupling compounds were not detected when silicon, glass, or gel was used (entry 1). A yield of 10% was obtained after mixing Li2PdCl4 with NiCl2 (entry 2). When terpyridine derivative was mixed with Li2PdCl4/NiCl2 as a catalyst without supports, only 8% yield was obtained (entry 3). In the case of Si-Tpy-Ni, Gel-Tpy-Pd/Ni, and Glass-Tpy-Pd/Ni, low yields were obtained (entries 5, 7, 8), indicating their poor activity compared with those of Si-Tpy-Pd (entry 4), Si-Tpy-Ni (entry 5), and Si-Tpy-Pd1/Ni1 (entry 6). Si-Tpy-Pd1/Ni1 also exhibited a higher activity than Si-Tpy-Pd and Si-Tpy-Ni, although the content of Pd in Si-Tpy-Pd1/Ni1 was half of GO-Tpy-Pd. Therefore, the activity of bi-organometallic Pd/Ni was enhanced because of the ordered orientation of organometallic compounds in monolayer and synergy between Pd and Ni [51,52,53,54,55].

The catalytic properties of Si-Tpy-Pd1/Ni1 were also compared with those reported by studies (Table 5). The results indicated that supports selection, ligand structure, formation and composition of metals, and orientation could efficiently affect catalytic activity.

Table 5 Comparison of Si-Tpy-Pd1/Ni1 with Ni-based or Pd-based catalysts in the Suzuki reaction.

3.3 Stability and Recycling Ability

Recycling tests were performed to investigate the recyclability of Si-Tpy-Pd1/Ni1 (Figure 8). A significant loss in recyclability was not observed after the fourth run; however, a low yield was obtained after the fifth run (51%). When the reaction time was increased to 24 h at the seventh run, an yield of 60% was obtained, suggesting deactivation in the recycling process [27,28,29,30,31,32,33,34]. The contents of Pd and Ni reused at the fourth and fifth runs were also measured, with the values being 1.34 × 10–7 (fresh: 2.43 × 10–6) mmol/cm2 and 1.48 × 10–8 (fresh: 3.38 × 10–7) mmol/cm2, respectively. Therefore, the loss of Pd and Ni was responsible for the activity loss.

Click to view original image

Figure 8 Recycling of Si-Tpy-Pd1/Ni1 in the Suzuki coupling reaction.

The stability of Si-Tpy-Pd1/Ni1 was investigated and compared with that of Si-Tpy-Pd (Figure 9). When Si-Tpy-Pd1/Ni1 and Si-Tpy-Pd were exposed for 1 week at ambient conditions, 87% yield was obtained, and 61% yield of Si-Tpy-Pd was obtained in the model reaction. Subsequently, 90% and 48% yield was achieved by Si-Tpy-Pd1/Ni1 and Si-Tpy-Pd in the second run after exposure for 2 weeks, indicating that Si-Tpy-Pd1/Ni1 exhibited more stability than Si-Tpy-Pd because of Ni doping.

Click to view original image

Figure 9 Investigation of Si-Tpy-Pd1/Ni1 and Si-Tpy-Pd stability in the Suzuki coupling reaction.

The catalytic activity of Si-Tpy-Pd1/Ni1 was also compared with commercial Pd/C under optimized conditions because Pd/C is a promising catalyst in Suzuki cross-coupling reactions (Figure 10) [56,57]. A little cross-coupling product (yield of less than 3%) was detected using Pd/C as a catalyst with a lower TOF value (Figure 10a), indicating the higher catalytic activity of Si-Tpy-Pd1/Ni1. In recycling tests, Si-Tpy-Pd1/Ni1 had higher stability than commercial Pd/C even after reusing three times (Figure 10b). The effect of temperature on TOFs of Si-Tpy-Pd1/Ni1 and Pd/C was also investigated (Figure 10c), with the results indicating similar catalytic performance. In addition, catalytic performances of Si-Tpy-Pd1/Ni1 and Pd/C at different substrate concentrations were explored. The results indicated that Si-Tpy-Pd1/Ni1 exhibited higher activity with decreasing substrate concentration. Therefore, the orientation, arrangement, and composition of the catalyst, as well as the selected support, had a stronger effect on the catalytic performances and stability.

Click to view original image

Figure 10 Yields per Pd atom of Si-TpyPd1/Ni1 compared with commercial Pd/C in the Suzuki cross-coupling reaction. Reaction condition: (a) PhB(OH)2 (0.15 mmol), 4-bromotoluene (0.1 mmol), K2CO3 (0.2 mmol), Si-TpyPd1/Ni1 , solvent (H2O:EtOH = 1:5 (1 mL + 5 mL)) at 80 °C for 12 h. (b) PhB(OH)2 (0.15 mmol), 4-bromotoluene (0.10 mmol), K2CO3 (0.20 mmol), Si@TpyPd1/Ni1, solvent (H2O:EtOH = 1:5 (1 mL + 5 mL)) at 80 °C for 4 h.

3.4 Exploration of the Catalytic Mechanism

3.4.1 Hot Filtration Experiment

The differentiation of heterogeneous from homogeneous catalysts is crucial [56]. The results of the hot filtration experiment indicated that the yield gradually increased before 8 h, followed by almost no change after 12 h (Figure 11, black line). A yield of 96% was obtained after 12 h. To investigate whether Pd leaching occurred during the catalytic process, Si-Tpy-Pd1/Ni1 was separated from the catalysis system at 4 h. The yields remained constant (approximately 50%, Figure 11, red line), indicating Pd and Ni leaching did not occur during the catalytic process [26,29].

Click to view original image

Figure 11 Hot filtration experiment of Si-Tpy-Pd1/Ni1 in the Suzuki coupling reaction.

3.4.2 Poison Experiments

To investigate the active center on the catalytic monolayer surface, poison tests were performed (Table 6). A low yield of 17% was obtained by the addition of a drop of mercury into the reaction mixture before reaction initiation. The active center formed on the surface was partially coordinated with less Hg because of its poor dispersibility (entry 2). In the case of thiophene additives added with less than 1.0 aquiver, significant deactivation was observed because of its stronger coordination with active sites, such as Pd, resulting in a decrease in active sites [58]. Therefore, the catalytic reaction mainly occurred on the surface of Si-Tpy-Pd1/Ni1 (entry 4).

Table 6 Poison experiments of Si-Tpy-Pd1/Ni1 catalyst a.

3.4.3 Three-Phase Test

The three-phase test is an effective method to determine the presence of heterogeneous catalysis [30,55]. The experiment design is depicted in Scheme 2 and was performed as described in Figure 12.

Click to view original image

Scheme 2 Three-phase experiment.

Click to view original image

Figure 12 UV spectra of SiO2@PhBr in the three-phase test. Treatment: washing solvent (DCM, EtOH, and MeOH) and ultrasonic (naphthybronoic acid, black; SiO2@PhBr, red; SiO2@PhBr [add Si-Tpy-Pd1N1 without wash], blue; SiO2@PhBr [add washed Si-Tpy-Pd1N1], green; SiO2@PhBr [add washed and ultrasonicated Si-Tpy-Pd1N1], purple; SiO2@PhBr [no cat and unwashed], yellow; SiO2@PhBr [no cat, and washed] water).

For the three-phase test, 4-chlorobenzyl bromide was covalently anchored on a quartz surface as one phase as described in a previous report (denoted as SiO2@PhBr) [30] (Figure 12).

Another soluble reagent (naphthyl boronic acid) was dissolved in a mixture solvent as the second phase to detect even a low amount of the naphthyl group with UV. The mixture was subjected to a reaction in the presence of a catalyst immobilized on the silicon as the third phase (Si-Tpy-Pd1Ni1). UV spectra of SiO2@PhBr and naphthyl boronic acid in ethanol were measured (Figure 12, red and black line), and the characteristic peak at 210 nm for SiO2@PhBr and at 225, 271, 280, and 291 nm for naphthyl boronic acid were observed. Because of the adsorption of naphthylboronic acid on one phase (SiO2@PhBr), designs were also presented under the reaction conditions. Reactions proceeded under the optimized conditions without Si-Tpy-Pd1Ni1. After the removal of SiO2@PhBr without washing, UV was measured (Figure 12, yellow line). The characteristic peaks of naphthylboronic acid at 236, 296, and 300 nm were observed, indicating the adsorption of naphthylboronic acid. However, UV spectra of SiO2@PhBr washed with DCM, EtOH, and MeOH or after ultrasonic treatment revealed no characteristic peaks (Figure 12, pink and green crimson line), indicating that adsorbed naphthylboronic acid was removed from SiO2@PhBr surface. If the active Pd species from Si-Tpy-Pd1Ni1 leached into the solution, the anchored reactant (one phase) may be converted to coupling compounds with naphthylboronic acid. Subsequently, SiO2@PhBr was carefully treated after reaction with Si-Tpy-Pd1Ni1 and analyzed under UV. A stronger broad peak for naphthylboronic acid (Figure 12, bright green line) and a gradual decrease after treatment with solvent and ultrasonication were observed (Figure 12, crimson and light blue line), indicating that the naphthyl boronic acid was absorbed on SiO2@PhBr surface through intermolecular interaction. No coupling reaction occurred on SiO2@PhBr surface. Therefore, the leaching of Pd active species into solution from Si-Tpy-Pd1Ni1 did not occur. Moreover, the leaching of Pd active species into solution from Si-Tpy-Pd1Ni1 did not occur. These results confirmed the presence of heterogeneous catalysis.

3.4.4 WCA Analysis

Wettability, a crucial property of solid materials, is governed by both surface chemical composition and the geometrical structure, and optimal wettability can enhance the catalytic activity [48,49]. Figure 13 depicts the WCA changes in Si-Tpy-Pd/Ni during catalysis. The WCAs of Si-Tpy-Pd/Ni were 44° before catalysis, which changed from 44° (0 h) to 48° (1 h), 50° (2 h), 54° (4 h), 46° (8 h), and 43° (12 h), demonstrating that catalysis occurred on the Si-Tpy-Pd/Ni monolayer surface; moreover, substrate absorption, intermediate formation, and desorption of product occurred because the Si-Tpy-Pd/Ni monolayer surface almost recovered after catalysis. This phenomenon was investigated using XPS analysis.

Click to view original image

Figure 13 Changes of WCA in Suzuki coupling reaction catalyzed by Si-Tpy-Pd1/Ni1 at 0, 1, 2, 4, 8, and 12 h.

3.4.5 Low X-ray Diffraction (LXRD) of Si-Tpy-Pd1/Ni1 During Catalysis.

The LXRD patterns of Si-Tpy-Pd1/Ni1 used for catalysis at different times were measured (Figure 14). The LXRD pattern at 1.1° for Si-Tpy-Pd1/Ni1 proved its regular structure, and the diffraction peak shifted from a high to a low angle. Moreover, the pattern type remained unchanged during catalysis (from 1 to 12 h), indicating that the thickness of the catalyst monolayer decreased and increased and finally retained its original thickness. Therefore, the structure of the catalyst was not damaged before and after catalysis, demonstrating that the hetero-bimetallic catalyst was durable and could maintain its orientation under the optimized conditions.

Click to view original image

Figure 14 LXRD of Si-Tpy-Pd1/Ni1 in Suzuki coupling reaction at 0, 1, 2, 4, 8, and 12 h.

3.4.6 UV Spectra of Si-Tpy-Pd1/Ni1 During Catalysis.

Changes in Si-Tpy-Pd1/Ni1 UV spectra were investigated during catalysis, and the results are displayed in Figure 15. The peak at 275 nm for Si-Tpy-Pd1/Ni1 shifted to 218 nm after 1 h, indicating that substrate adsorption led to the high density of monolayer. The adsorption became instant at 2, 4, and 8 h, thereby demonstrating the equivalence. Finally, it recovered at 12 h, demonstrating the stability of the catalytic monolayer.

Click to view original image

Figure 15 UV spectra of Si-Tpy-Pd1/Ni1 during catalysis (0 h, black line; 1 h, red line; 2 h, blue line; 8 h, purple line; 12 h, yellow line).

3.4.7 SEM Images of SiTpy-Pd1/Ni1 During Catalysis

To investigate the changes on the surfaces, SEM images at different reaction times were obtained (Figure 16). SEM revealed a regular surface before catalysis (Figure 16A), and the morphology of the catalytic surface changed at 1 h (Figure 16B), indicating that the adsorption of substrates occurred on the catalytic monolayer surface. Extreme changes occurred on the catalyst surfaces at 2 and 4 h (Figures 16C and 16D), indicating the automatic agglomeration of coupling compounds because of the higher adsorption rate, which was the key factor in the catalytic process. As observed in Figure 16E, the image became smooth at 8 h. This indicated that a violent reaction occurred, which was accompanied by a sequential process, including the adsorption of substrate and formation of intermediates and compounds on the catalytic surface. The morphology of the catalytic surface was more ordered at 12 h than at 2, 4, and 8 h (Figure 16F). The significant changes in the catalytic surface may be because of the in situ formation of Pd on the catalytic surface. The catalytic process included adsorption, synergism, and desorption. More time was required for catalysis because of the covered active sites, adsorption, and desorption rate of substrate or product affecting the catalytic process.

Click to view original image

Figure 16 SEM images of Si-Tpy-Pd1/Ni1 at (A) 0 h, (B) 1 h, (C) 2 h, (D) 4 h, (E) 8 h, and (F) 12 h (white bar: 200 nm).

The Si-Tpy-Pd1/Ni1 surface morphology became random with some particles after the 1st, 2nd, 3rd, 4th, 5th, and 6th run (Figure 17). After the 1st run, some larger particles were observed, which might be the substrates or coupling products adsorbed on the catalytic monolayer surface (Figure 17A). As the recycling proceeded, much more large particles were obtained, especially after five or six runs (Figures 17B, C, D, E, and F). This may be the main reason for deactivation attributable to the aggregation of the active species forming Pd/Ni clusters through the migration of active atoms and coupling with each other [58]. Therefore, the prepared Si-Tpy-Pd1/Ni1 was stable during catalysis.

Click to view original image

Figure 17 SEM images of Si@Tpy-Pd1/Ni1 reused. (A) 1st run, (B) 2nd run, (C) 3rd run. (D) 4th run, (E) 5th run, (F) 6th run. (white bar: 200 nm).

3.4.8 XPS Analysis of GO-Tpy-Pd1/Ni1 During Catalysis

The HR-XPS of Pd 3d, Br 3d, and N 1s during catalysis was measured (Figure 18). During the catalytic process, the peaks at 399.89 eV for N 1s (Figure 18A) shifted to 399.99 eV (1 h), 400.14 eV (2 h), 399.99 eV (4 h), 400.04 eV (8 h), and 400.19 eV (12 h). The peak position of N 1s changed during the catalytic process because of the electron transfer to or electron acceptance from the different valence states of Pd and Ni coordinated with the ligand. The peaks at 340.54 eV and 335.24 eV for Pd 3d3/2 and Pd 3d5/2 were assigned to Pd(0) and appeared at 1 h (Figure 18B). The intensity of the two peaks of Pd(0) gradually increased with increasing catalysis time, and the intensity of the peaks for Pd(II) gradually decreased with time, indicating that the real active site Pd(0) was formed in situ by Pd(II) reduction. The peak of Br 3d was observed at 8 and 12 h (Figure 18C). Therefore, the basic elements of the catalyst and substrates appeared and changed during catalysis, implying the oxidation of benzyl bromide with metal to produce an oxidative intermediate on the surface. The oxidative intermediate reacted with phenylboronic acid to form a transmetalation intermediate, which yielded cross-coupling compounds through elimination reduction, and finally desorbed from the surface.

Click to view original image

Figure 18 HR-XPS spectra of Pd 3d, N 1s, and Br 3d at 0, 1, 2, 4, 8, and 12 h.

In addition, the changes in the bonding energy of Ni 2p during catalysis were analyzed (Figure 19). Peaks at 855.67, 861.57, and 873.47 eV were attributed to N 2p3/2 and N 2p1/2, indicating the existence of Ni2+ (Figure 19A). The binding energy of Ni 2p increased to 856.14 eV after 4 h, indicating the loss of electrons in Ni (Figure 19B); it decreased to 854.24 eV after 8 h, indicating electron acceptance by Ni (Figure 19C). Finally, the bonding energy gradually returned to the original state at 855.77 eV at 12 h, indicating that Ni(II) was reduced to Ni(0) (Figure 19D). The HR-XPS of Pd revealed that Pd(0) quantity increased with increasing catalytic time (Figure 18B). This may be attributed to the electron transfer from Ni to Pd, which facilitated the oxidation step because of more negative charge in the Pd active center (Niδ+/Pdδ–/Pd(0)) formed during the catalysis process (Scheme 3). Niδ+/Pdδ–/Pd(0) surrounded by appropriate Pd(II)/Ni(II) played a crucial role in catalysis because of the synergistic effect between Pd and Ni linked as a “guest” to a “host” in the entity, thereby modifying the electronic properties of its active center [45,59,60,61] and increasing the activity. The catalytic film was regarded as an infinite “pool” of organometallic compounds containing a few active species on its surface [62]. This indicated the importance of catalysis surface modification by organometallic compounds to tune the catalytic activity.

Click to view original image

Figure 19 High resolution XPS of Ni 2p at different times: (A) 0 h, (B) 4 h, (C) 8 h, and (D) 12 h.

Click to view original image

Scheme 3 Proposal of active center in Ni(II)/Pd(0) formed in situ during catalysis.

On the basis of the aforementioned results, a hypothetic catalytic mechanism was proposed (Scheme 4). An “induced step” was the adsorption of one or more substrates on the organometallic monolayer surface during the catalytic process. Pd(0) adsorbed p-bromotoluene to form the oxidative insertion intermediate. Pd(II) or Ni(II) in the vicinity of active Pd(0) adsorbed phenyl boronic acid, which transmetalated with the oxidative insertion intermediate, followed by the formation of a coupling compound through elimination reduction. Finally, coupling compounds diffused from the catalytic monolayer surface.

Click to view original image

Scheme 4 Supposed catalytic mechanism and synergy on the self-assembled organometallic framework film (OMFF) surface.

Bi-organometallic catalytic monolayer exhibits higher activity than the mono-organometallic catalytic monolayer because of the synergistic effect existing on the catalytic surface, such as the electron distribution of active species. Moreover, active metals or nonactive metals in active sites simultaneously joined in the catalytic process, as described in Scheme 3 [63,64,65,66]. The results indicated that the morphology, catalyst orientation, ligand design, support selection, the rational combination of different metals, and the ratio of multi-metals were crucial factors for achieving a highly active and stable catalyst.

4. Conclusions

A new organometallic terpyridine Ni/Pd self-assembled catalytic monolayer supported on a silicon wafer surface (denoted as Si-Tpy-Ni1/Pd1) was fabricated and characterized. The monolayer exhibited high catalytic activity, substrate applicability, and recyclability after 5 runs, and the deactivation was induced by Pd aggregation during recycling. The catalytic mechanism on the surface of Si-Tpy-Pd1/Ni1 was heterogeneous. The synergistic mechanism between Ni and Pd was investigated in detail, and the real active species was Pd/Ni, which was formed in situ. The electron density of Pd increased because of electron transfer from Ni to Pd, which was the crucial factor for the initial oxidation step. These results indicated that the high activity of a catalyst could be achieved through optimal ligand designing, support selection, catalyst arrangement, and a suitable combination of bi-organometallics.

Acknowledgments

The authors gratefully acknowledge the Henan Natural Science Foundation of China (192102210046) for their financial support. Thank Prof. Luyuan Mao, Zhengzhou University for AFM measurement.

Author Contributions

Prof. Tiesheng Li designed, analysis and writing; Wei Wang and Huanhuan Li for synthesis and measurement; Sa Bi for synthesis.

Competing Interests

The authors have declared that no competing interests exist.

References

  1. Isley NA, Gallou F, Lipshutz BH. Transforming Suzuki–Miyaura cross-couplings of MIDA boronates into a green technology: No organic solvents. J Am Chem Soc. 2013; 135: 17707-17710. [CrossRef]
  2. Xi P, Cao Y, Yang F, Ma C, Chen F, Yu S, et al. Facile synthesis of Pd-based bimetallic nanocrystals and their application as catalysts for methanol oxidation reaction. Nanoscale. 2013; 5: 6124-6130. [CrossRef]
  3. Xiang J, Li P, Chong H, Feng L, Fu F, Wang Z, et al. Bimetallic Pd-Ni core-shell nanoparticles as effective catalysts for the Suzuki reaction. Nano Res. 2014; 7: 1337-1343. [CrossRef]
  4. Rai RK, Gupta K, Behrens S, Li J, Xu Q, Singh SK. Highly active bimetallic nickel–palladium alloy nanoparticle catalyzed Suzuki–Miyaura reactions. ChemCatChem. 2015; 7: 1806-1812. [CrossRef]
  5. Son SU, Jang Y, Park J, Na HB, Park HM, Yun HJ, et al. Designed synthesis of atom-economical Pd/Ni bimetallic nanoparticle-based catalysts for sonogashira coupling reactions. J Am Chem Soc. 2004; 126: 5026-5027. [CrossRef]
  6. Kutubi MS, Sato K, Wada K, Yamamoto T, Matsumura S, Kusada K, et al. Dual Lewis acidic/basic Pd0.5Ru0.5–poly(N-vinyl-2-pyrrolidone) alloyed nanoparticle: Outstanding catalytic activity and selectivity in Suzuki–Miyaura cross-coupling reaction. ChemCatChem. 2015; 7: 3887-3894. [CrossRef]
  7. Yang XF, Wang A, Qiao B, Li J, Liu J, Zhang T. Single-atom catalysts: A new frontier in heterogeneous catalysis. Acc Chem Res. 2013; 46: 1740-1748. [CrossRef]
  8. Xue L, Lin Z. Theoretical aspects of palladium-catalysed carbon–carbon cross-coupling reactions. Chem Soc Rev. 2010; 39: 1692-1705. [CrossRef]
  9. Sancheti SP, Urvashi, Shah MP, Patil NT. Ternary catalysis: A stepping stone toward multicatalysis. ACS Catal. 2020; 10: 3462-3489. [CrossRef]
  10. Zhou X, Dai J, Wu K. Steering on-surface reactions with self-assembly strategy. Phys Chem Chem Phys. 2017; 19: 31531-31539. [CrossRef]
  11. Qiu Y, Chen P, Liu M. Evolution of various porphyrin nanostructures via an oil/aqueous medium: Controlled self-assembly, further organization, and supramolecular chirality. J Am Chem Soc. 2010; 132: 9644-9652. [CrossRef]
  12. Yuan J, Liu M. Chiral molecular assemblies from a novel achiral amphiphilic 2-(heptadecyl) naphtha[2, 3]imidazole through interfacial coordination. J Am Chem Soc. 2003; 125: 5051-5056. [CrossRef]
  13. Shen Z, Jiang Y, Wang T, Liu M. Symmetry breaking in the supramolecular gels of an achiral gelator exclusively driven by π–π stacking. J Am Chem Soc. 2015; 137: 16109-16115. [CrossRef]
  14. Kakkar AK. Nano-organometallics: Heterogenizing homogeneous catalysts via thin film methodology. Chem Rev. 2002; 102: 3579-3588. [CrossRef]
  15. Huang X, Li C, Jiang S, Wang X, Zhang B, Liu M. Self-assembled spiral nanoarchitecture and supramolecular chirality in Langmuir−Blodgett films of an achiral amphiphilic barbituric acid. J Am Chem Soc. 2004; 126: 1322-1323. [CrossRef]
  16. Mu B, Li T, Li C, Liu P, Shang W, Wu Y. Langmuir–Blodgett films of cyclopalladated ferrocenylimine: Preparation, characterization, and application in Suzuki coupling reaction. Tetrahedron. 2009; 65: 2599-2604. [CrossRef]
  17. Marshall ST, O’Brien M, Oetter B, Corpuz A, Richards RM, Schwartz DK, et al. Controlled selectivity for palladium catalysts using self-assembled monolayers. Nat Mater. 2010; 9: 853-858. [CrossRef]
  18. Xiao M, Feng L, Zhu J, Liu C, Xing W. Rapid synthesis of a PtRu nano-sponge with different surface compositions and performance evaluation for methanol electrooxidation. Nanoscale. 2015; 7: 9467-9471. [CrossRef]
  19. Wang J, Mu B, Fu Z, Wang L, Li T, Wu Y. Cyclopalladated ferrocenylimines with ester groups for Heck and Suzuki coupling reactions. Chin J Catal. 2014; 35: 1059-1067. [CrossRef]
  20. Key RJ, Tengco JMM, Smith MD, Vannucci AK. A molecular/heterogeneous nickel catalyst for Suzuki–Miyaura coupling. Organometallics. 2019; 38: 2007-2014. [CrossRef]
  21. Fu Z, Zhang N, Liu J, Li T, Xu W, Wang F, et al. N-hydroxymethyl acrylamide polymer brush and its application in catalyzing coupling reaction. J Colloid Interface Sci. 2013; 394: 409-418. [CrossRef]
  22. Zhao N, Li T, Zhai Z, Qiu J, Xu W, Liu H, et al. Cyclopalladated arylimine self-assembly films for Suzuki reaction. ChemCatChem. 2013; 5: 1481-1489. [CrossRef]
  23. Fu Z, Li T, He X, Liu J, Xu W, Wu Y. Cyclopalladated ferrocenylimine functionalized polymer brushes film and its mechanism investigation of heterogeneous catalysis. J Mol Catal A Chem. 2014; 395: 293-299. [CrossRef]
  24. Xue Z, Huang P, Li T, Qin P, Xiao D, Liu M, et al. A novel “tunnel-like” cyclopalladated arylimine catalyst immobilized on graphene oxide nano-sheet. Nanoscale. 2017; 9: 781-791. [CrossRef]
  25. Gu X, Bi S, Guo L, Zhao Y, Li T, Liu M, et al. Facile fabrication of ordered component-tunable heterobimetallic self-assembly nanosheet for catalyzing “click” reaction. ACS Omega. 2017; 2: 5415-5433. [CrossRef]
  26. Wang L, Huang P, Yang J, Li T, Mao L, Liu M, et al. Fabrication and catalytic properties of ordered cyclopalladated diimine monolayer: Investigation on catalytic mechanism. RSC Adv. 2018; 8: 31860-31867. [CrossRef]
  27. Huang P, Xue Z, Li T, Liu Z, Wei D, Liu M, et al. Investigation on electron distribution and synergetic to enhance catalytic activity in bimetallic Ni(II)/Pd(II) molecular monolayer. ChemCatChem. 2018; 10: 5141-5153. [CrossRef]
  28. Huang P, Song E, Sun Y, Li T, Wei D, Liu M, et al. Schiff-based Pd(II)/Fe(III) bimetallic self-assembly monolayer---Preparation, structure, catalytic dynamic and synergistic. Mol Catal. 2019; 469: 75-86. [CrossRef]
  29. Shang W, Zeng X, Li T, Xu W, Wei D, Liu M, et al. Controlled distribution of active centre to enhance catalytic activity of ordered Pd/Co catalytic nano-monolayer. J Catal. 2019; 376: 228-237. [CrossRef]
  30. Song E, Wang J, Li T, Zhao W, Liu M, Wu Y. Novel ordered cyclopalladated aryl imine monolayers—Structure Designing for Enhancing Catalytic Performance. Mol Catal. 2020; 482: 110671. [CrossRef]
  31. Ren R, Bi S, Wang L, Zhao W, Wei D, Li T, et al. Terpyridine-based Pd(II)/Ni(II) organometallic framework nano-sheets supported on graphene oxide—Investigating the fabrication, tuning of catalytic properties and synergetic effects. RSC Adv. 2020; 10: 23080-23090. [CrossRef]
  32. Ren R, Huang P, Zhao W, Li T, Liu M, Wu Y. A new ternary organometallic Pd(II)/Fe(III)/Ru(III) self-assembly monolayer: The essential ensemble synergistic for improving catalytic activity. RSC Adv. 2021; 11: 1250-1260. [CrossRef]
  33. Li Z, Song E, Ren R, Zhao W, Li T, Liu M, et al. Pd–Pd/PdO as active sites on intercalated graphene oxide modified by diaminobenzene: Fabrication, catalysis properties, synergistic effects, and catalytic mechanism. RSC Adv. 2022; 12: 8600-8610. [CrossRef]
  34. Ren R, Li T, Liu M, Wu Y. Fabrication and catalytic properties of “cage like” aryl imine Pd(II)/Cu(II)-bimetallic catalytic monolayer supported on graphene oxide for Suzuki coupling reaction. Chem Eng Sci. 2022; 253: 117604. [CrossRef]
  35. Huang L, Liu H, Xue X, Zhao W, Li T. Controllable cyclopalladated polythiophene imine monolayer by self-assembly, hybrid doping and electrochemical polymerization: A simple way to enhance activity and stability. Catal Res. 2022; 2: 15. [CrossRef]
  36. Li T, Wang M, Liu J, Li J. A new chiral cyclopalladated ferrocenylimine self-assembly polymeric film supported on silicon for heck and Suzuki-coupling reactions. Catal Res. 2022; 2: 9. [CrossRef]
  37. Zou K, Liu H, Li T, Chen P, Liu M, Wu Y. An electrochemically polymerized and assembled cyclopalladated bi-thiophene imine for catalyzing coupling reactions: A modern strategy to enhance catalytic activity. RSC Adv. 2015; 5: 16654-16663. [CrossRef]
  38. Liu H, Li T, Xue X, Xu W, Wu Y. The mechanism of a self-assembled Pd(ferrocenylimine)–Si compound-catalysed Suzuki coupling reaction. Catal Sci Technol. 2016; 6: 1667-1676. [CrossRef]
  39. Tasker SZ, Standley EA, Jamison TF. Recent advances in homogeneous nickel catalysis. Nature. 2014; 509: 299-309. [CrossRef]
  40. Rosen BM, Quasdorf KW, Wilson DA, Zhang N, Resmerita AM, Garg NK, et al. Nickel-catalyzed cross-couplings involving carbon-oxygen bonds. Chem Rev. 2011; 111: 1346-1416. [CrossRef]
  41. Singh S, Sunoj RB. Mechanism and origin of enantioselectivity in nickel-catalyzed alkyl–alkyl Suzuki coupling reaction. J Phys Chem A. 2019; 123: 6701-6710. [CrossRef]
  42. Payard PA, Perego LA, Ciofini I, Grimaud L. Taming nickel-catalyzed Suzuki-Miyaura coupling: A mechanistic focus on boron-to-nickel transmetalation. ACS Catal. 2018; 8: 4812-4823. [CrossRef]
  43. Cooper AK, Leonard DK, Bajo S, Burton PM, Nelson DJ. Aldehydes and ketones influence reactivity and selectivity in nickel-catalysed Suzuki–Miyaura reactions. Chem Sci. 2020; 11: 1905-1911. [CrossRef]
  44. Wu Y, Wang D, Zhao P, Niu Z, Peng Q, Li Y. Monodispersed Pd−Ni nanoparticles: Composition control synthesis and catalytic properties in the Miyaura−Suzuki reaction. Inorg Chem. 2011; 50: 2046-2048. [CrossRef]
  45. Bao G, Bai J, Li C. Synergistic effect of the Pd–Ni bimetal/carbon nanofiber composite catalyst in Suzuki coupling reaction. Org Chem Front. 2019; 6: 352-361. [CrossRef]
  46. Sultanova ED, Samigullina AI, Nastapova NV, Nizameev IR, Kholin KV, Morozov VI, et al. Highly active Pd–Ni nanocatalysts supported on multicharged polymer matrix. Catal Sci Technol. 2017; 7: 5914-5919. [CrossRef]
  47. Rai RK, Gupta K, Tyagi D, Mahata A, Behrens S, Yang X, et al. Access to highly active Ni–Pd bimetallic nanoparticle catalysts for C–C coupling reactions. Catal Sci Technol. 2016; 6: 5567-5579. [CrossRef]
  48. Denysenko D, Werner T, Grzywa M, Puls A, Hagen V, Eickerling G, et al. Reversible gas-phase redox processes catalyzed by Co-exchanged MFU-4l(arge). Chem Commun. 2012; 48: 1236-1238. [CrossRef]
  49. Sharp JS, Farmer DJ, Kelly J. Contact angle dependence of the resonant frequency of sessile water droplets. Langmuir. 2011; 27: 9367-9371. [CrossRef]
  50. Herrmann WA, Cornils B. Organometallic homogeneous catalysis—Quo vadis? Angew Chem Int Ed Engl. 1997; 36: 1048-1067. [CrossRef]
  51. Metin Ö, Ho SF, Alp C, Can H, Mankin MN, Gültekin MS, et al. Ni/Pd core/shell nanoparticles supported on graphene as a highly active and reusable catalyst for Suzuki-Miyaura cross-coupling reaction. Nano Res. 2013; 6: 10-18. [CrossRef]
  52. Xia J, Fu Y, He G, Sun X, Wang X. Core-shell-like Ni-Pd nanoparticles supported on carbon black as a magnetically separable catalyst for green Suzuki-Miyaura coupling reactions. Appl Catal B Environ. 2017; 200: 39-46. [CrossRef]
  53. Jang S, Kim T, Park KH. Fabrication of crumpled ball-like nickel doped palladium-iron oxide hybrid nanoparticles with controlled morphology as effective catalyst for Suzuki–Miyaura coupling reaction. Catalysts. 2017; 7: 247. [CrossRef]
  54. Ohtaka A, Sansano JM, Nájera C, Miguel-García I, Berenguer-Murcia Á, Cazorla-Amorós D. Palladium and bimetallic palladium–nickel nanoparticles supported on multiwalled carbon nanotubes: Application to carbon—carbon bond-forming reactions in water. ChemCatChem. 2015; 7: 1841-1847. [CrossRef]
  55. Wang F, Li C, Sun LD, Xu CH, Wang J, Yu JC, et al. Porous single-crystalline palladium nanoparticles with high catalytic activities. Angew Chem Int Ed. 2012; 51: 4872-4876. [CrossRef]
  56. Bligaard T, Bullock RM, Campbell CT, Chen JG, Gates BC, Gorte RJ, et al. Toward benchmarking in catalysis science: Best practices, challenges, and opportunities. Acs Catal. 2016; 6: 2590-2602. [CrossRef]
  57. Phan NT, Van Der Sluys M, Jones CW. On the nature of the active species in palladium catalyzed Mizoroki–Heck and Suzuki–Miyaura couplings–homogeneous or heterogeneous catalysis, a critical review. Adv Synth Catal. 2006; 348: 609-679. [CrossRef]
  58. Bartholomew CH. Mechanisms of catalyst deactivation. Appl Catal A Gen. 2001; 212: 17-60. [CrossRef]
  59. Gu J, Zhang YW, Tao FF. Shape control of bimetallic nanocatalysts through well-designed colloidal chemistry approaches. Chem Soc Rev. 2012; 41: 8050-8065. [CrossRef]
  60. Saha J, Bhowmik K, Das I, De G. Pd–Ni alloy nanoparticle doped mesoporous SiO2 film: The sacrificial role of Ni to resist Pd-oxidation in the C–C coupling reaction. Dalton Trans. 2014; 43: 13325-13332. [CrossRef]
  61. Hooshmand SE, Heidari B, Sedghi R, Varma RS. Recent advances in the Suzuki–Miyaura cross-coupling reaction using efficient catalysts in eco-friendly media. Green Chem. 2019; 21: 381-405. [CrossRef]
  62. Green M. A new approach to the formal classification of covalent compounds of the elements. J Organomet Chem. 1995; 500: 127-148. [CrossRef]
  63. Ponec V. Alloy catalysts: The concepts. Appl Catal A Gen. 2001; 222: 31-45. [CrossRef]
  64. Voss GJ, Fløystad JB, Voronov A, Rønning M. The state of nickel as promotor in cobalt Fischer–Tropsch synthesis catalysts. Top Catal. 2015; 58: 896-904. [CrossRef]
  65. Samantaray MK, D’Elia V, Pump E, Falivene L, Harb M, Ould Chikh S, et al. The comparison between single atom catalysis and surface organometallic catalysis. Chem Rev. 2019; 120: 734-813. [CrossRef]
  66. Zhu S, Li Z, Ren R, Zhao W, Li T, Liu M, et al. Pd/Cu2O/CuO as active sites on the cyclometalated Pd(II)/Cu(II) nanosheet: Active centre formation, synergistic and catalytic mechanism. ChemistrySelect. 2022; 7: e202200340. [CrossRef]
Newsletter
Download PDF Download Citation
0 0

TOP