Next Article in Journal
Synthesis, Characterization of the Novel Heterojunction Photocatalyst Sm2NdSbO7/BiDyO3 for Efficient Photodegradation of Methyl Parathion
Previous Article in Journal
Ionic Liquid Modification of High-Pt-Loading Pt/C Electrocatalysts for Proton Exchange Membrane Fuel Cell Application
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Constructing Polyphosphazene Microsphere-Supported Pd Nanocatalysts for Efficient Hydrogenation of Quinolines under Mild Conditions

1
National & Local Joint Engineering Research Center for Textile Fiber Materials and Processing Technology, College of Material Science and Engineering, Zhejiang Sci-Tech University, Hangzhou 310018, China
2
Zhijiang College, Zhejiang University of Technology, Hangzhou 310024, China
3
Key Laboratory of Bio-Based Materials, Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences, Qingdao 266101, China
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(6), 345; https://doi.org/10.3390/catal14060345
Submission received: 18 April 2024 / Revised: 17 May 2024 / Accepted: 23 May 2024 / Published: 27 May 2024
(This article belongs to the Special Issue State-of-the-Art Polymerization Catalysis)

Abstract

:
The efficient hydrogenation of N-heterocycles with H2 under mild conditions remains a significant challenge. In this work, polyphosphazene (PZs) microspheres, novel organic–inorganic hybrid materials possessing unique –P=N– structural units and a diverse range of side groups, were used to serve as support for the design of a stable and efficient Pd nanocatalyst (Pd/PZs). The PZs microspheres were prepared by self-assembly induced by precipitation polymerization, and Pd nanoparticles were grown and loaded on the support by a chemical reduction process. Several characterization techniques, including XRD, FTIR, SEM, TEM, XPS, BET and TGA, were used to study the structural features of the nanocomposites. The results revealed that Pd nanoparticles were uniformly distributed on the PZs microspheres, with primary sizes ranging from 4 to 9 nm based on the abundance of functional P/N/O groups in PZs. Remarkably high catalytic activity and stability were observed for the hydrogenation of quinoline compounds using the Pd/PZs nanocatalyst under mild conditions. Rates of 98.9% quinoline conversion and 98.5% 1,2,3,4-tetrahydroquinoline selectivity could be achieved at a low H2 pressure (1.5 bar) and temperature (40 °C). A possible reaction mechanism for quinoline hydrogenation over Pd/PZs was proposed. This work presents an innovative approach utilizing a Pd-based nanocatalyst for highly efficient multifunctional hydrogenation.

Graphical Abstract

1. Introduction

Catalytic hydrogenation reactions are a crucial step in the production of various fine chemicals and pharmaceuticals in industry. Quinoline and its derivative compounds, which are a kind of N-heterocycles, are commonly found in natural products and pharmaceuticals [1,2,3]. The direct hydrogenation of quinolines using H2 is considered as the most convenient and atom-efficient strategy to obtain the corresponding reduction products [4]. Nevertheless, this hydrogenation process often requires harsh reaction conditions with significant safety risks [5]. Additionally, as illustrated in Scheme 1, the quinoline hydrogenation process may involve a competitive hydrogenation of both benzene rings and heterocyclic rings, leading to the formation of multiple hydrogenated products, e.g., 1,2,3,4-Tetrahydroquinoline (py-THQ), 5,6,7,8-tetrahydroquinoline (bz-THQ), and decahydroquinoline (DHQ) [6]. N-heterocycles, including quinolines and their hydrogenated forms, exhibit strong coordination ability with metal active nanoparticles, which can potentially poison metal catalysts. Therefore, achieving selective hydrogenation of N-containing aromatics remains a challenging task. From an environmental perspective and emphasizing safety considerations, it is imperative to develop an effective and highly chemoselective hydrogenation system that performs under mild conditions.
Metal catalysts play a pivotal role in industrial applications of hydrogenation reactions. The hydrogenation of N-heteroarenes can be achieved using a variety of homogeneous and heterogeneous transition metal catalysts [7,8,9,10,11,12]. Numerous homogenous metal catalysts (e.g., Pd, Ru, Rh) have demonstrated the ability to catalyze the conversion of quinoline into py-THQ. However, they typically require high H2 pressure, elevated temperatures (>100 °C), prolonged reaction times, or harsh additives (e.g., acid, I2, or PPh3) for efficient transformation [13,14,15,16]. Moreover, the challenging separation and recovery processes as well as the limited reusability further hinder their practical application. To overcome these limitations, substantial endeavors have been devoted to developing novel heterogeneous metal catalysts for quinoline hydrogenation [4,5,8,10,17]. A series of transition metal supported catalysts, such as M/Al2O3 (M = Pd, Ru or Rh) [18], Pd/hydroxyapatites [19], Rh/montmorillonite [20], polymer-supported Pd catalyst (Pd-pol) [21], and Co3Cu1Ox [22], have been explored for reducing quinolines. However, most of these catalytic systems still require drastic reaction conditions (60–130 °C, 1–4 MPa H2) or rely on expensive carbon nitride supports, which involve complicated synthesis steps and harsh conditions, to obtain a high conversion of quinoline towards py-THQ (Table S1). Among these active metals, Pd nanocatalysts have garnered considerable attention due to their excellent performance in the transformation [18,19,21]. Hence, there is a need to develop efficient and stable Pd-based nanocatalysts which are active in low-pressure H2 and low-temperature systems.
The catalytic activity of noble metal catalysts is strongly related to their particle size and dispersion. Noble metal nanocatalysts of small size and with high dispersion universally display outstanding catalytic hydrogenation performance for quantum size effects [23,24]. However, because of the high surface energy, small-size metallic nanoparticles are prone to aggregate during the preparation process, resulting in reduced catalytic activity and selectivity [25]. Therefore, a variety of supports, such as TiO2, ZnO, SiO2, carbon materials with different morphologies, graphitic carbon nitride, and organic polymers, have been applied to stabilize the metal nanoparticles [26,27,28,29,30,31,32]. Among them, as a novel organic–inorganic hybrid materials, polyphosphazenes (PZs) have unique –P=N– structural units with alternating P and N atoms in the backbone and a structural multiplicity of side groups, such as organic, organometallic or inorganic units [33]. The polymers have the characteristics of flexible molecular design, high thermal stability and biocompatibility, which makes them suitable candidates for immobilizing metal particles and biomedical carriers. For instance, Wang et al. and others showed that polyphosphazene nanotubes could be chemically anchored to Ag or Ag-Au nanoparticles via the phosphazene ring and –OH groups and displayed high catalytic performance in the hydrogenation of 4-nitrobenzene [34,35,36]. However, so far, superior applications of polyphosphazenes as catalyst supports for hydrogenation reactions are still in their infancy.
In this work, a novel Pd nanocatalyst was constructed by using poly-(cyclotriphosphazene-co-4,4′-sulfonyldiphenol) microspheres as a good scaffold for Pd nanoparticles. The PZs microspheres, as a typical crosslinked polymer, provided abundant phosphazene rings, S=O groups and hydroxyl groups to stably immobilize the Pd active sites by the chemical reduction method with NaBH4 as the reductant. The Pd nanocatalyst was applied in the direct catalytic hydrogenation of quinoline compounds under mild conditions and it exhibited outstanding catalytic activity and reusability at low temperatures of 30–50 °C and a low H2 pressure of 1.5 bar. The possible reaction mechanism for quinoline hydrogenation was proposed. This work constructed a high-performance Pd nanocatalyst for N-heterocycles hydrogenation by using a crosslinked polymer with facile synthesis and high chemical stability as the support, contributing to the development of green and sustainable chemistry.

2. Results and Discussion

2.1. Fabrication Process of the Pd/PZs Catalyst

Scheme 2 shows the fabrication procedure of the Pd/PZs nanocatalysts, involving the precipitation polymerization of hexachlorocyclotriphosphazene (HCCP) and 4,4′-sulfonyldiphenol (BPS), with triethylamine (TEA) as the acid acceptor and the chemical reduction of the Pd precursor with NaBH4 as the reductant. In the first step, HCCP and BPS were used as comonomers and dissolved in acetonitrile. The PZs microspheres were obtained through self-assembly induced by precipitation polymerization between HCCP and BPS after adding TEA to neutralize acid ions. In the second step, PdCl2 was dissolved in a dilute hydrochloric acid solution to form H2PdCl4 before use. The PZs microspheres were mixed with H2PdCl4 solution, and PdCl42− ions were adsorbed on the surface of the PZs microspheres. Then, the Pd nanoparticles were grown and loaded on the surface of PZs microspheres by in situ reduction of PdCl42−, with NaBH4 as the reductant, to yield the Pd/PZs nanocatalyst. The rich phosphazene rings and O-anchoring sites would help the adsorption and confinement of Pd nanoparticles on the PZs support [34]. Unless otherwise specified, the initial Pd loading in the Pd/PZs was 5% in this work.

2.2. Composition and Structure

The phase structure of the PZs and 5%Pd/PZs samples were studied by XRD. Figure 1 shows that there was one broad peak at around 15° in the PZs polymer, attributable to the amorphous PZs matrix. The intensity of the PZs peak in the 5%Pd/PZs was much weaker than that in the PZs, which might be caused by the partial coverage of Pd particles on the surface of PZs. Apart from the PZs peak, several diffraction peaks at 39.4°, 45.2°, and 66.4°were also found in 5%Pd/PZs, corresponding to the (111), (200), and (220) lattice planes of the face-centered cubic crystalline structure of metallic Pd [37]. Based on the Scherrer equation, the crystallite sizes of Pd0 particles in the 5%Pd/PZs were about 4.5 nm. The Pd content in the 5%Pd/PZs, determined by ICP, was about 3.8%. This result implies that PZs can act as an excellent scaffold to anchor Pd nanoparticles of small size.
The FTIR spectra in Figure 2 confirmed the presence of abundant functional groups on the surface of the PZs and Pd/PZs. Similar adsorption bands appeared in the spectra of the PZs and Pd/PZs. The broad band at 3427 cm−1 was assigned to −OH groups. The bands at 1323, 1290, and 1150 cm−1 belonged to the characteristic O=S=O absorption in the sulfonyldiphenol units [34]. The band at 1200 cm−1 corresponded to P=N bonds in the cyclotriphosphazene structure, while the band at 1014 cm−1 was ascribed to the P−O bond [33,38]. The band at 1373 cm−1 might be related to C-O-P groups. The presence of O=S=O and P−O bands confirmed successful condensation between HCCP and BPS to form PZs. The band at 1587 and 1488 cm−1 could be ascribed to −C=C− in the benzene ring. The bands at 3420, 1405 and 1105 cm−1 indicated the existence of abundant C−OH groups in the PZs. The FTIR spectra of PZs and Pd/PZs were in agreement with the characteristic absorption bands of the PZs microspheres.
The pore structure of 5%Pd/PZs was measured by N2 adsorption/desorption measurement. As presented in Figure S1, the Pd/PZs displayed a type-III adsorption curve without an obvious hysteresis ring, indicating little pore structure in the PZs microspheres. A certain amount of N2 adsorption was observed in the relative pressure range of p/p0 > 0.8, suggesting the existence of an external surface area. The Brunauer–Emmett–Teller (BET) surface area and pore volume of the Pd/PZs were 7.1 m2 g−1 and 0.033 cm3 g−1, respectively. The morphological features of PZs and 5%Pd/PZs were examined by SEM and TEM. The typical SEM images in the Figure 3a,b show the uniform spherical shapes with a smooth surface and a 500–1000 nm diameter that could be observed in the PZs. The TEM images in Figure 3c–e reveal that most of the Pd nanoparticles with a primary diameter of about 4–9 nm were evenly deposited on the surface of the PZs microspheres without obvious aggregation. The HRTEM picture (Figure 3f) shows the presence of a Pd (111) plane d-spacing of 0.23 nm, supporting the formation of metallic Pd nanoparticles on the surface of the PZs microspheres.
X-ray photoelectron spectroscopy (XPS) measurements were used to further examine the chemical composition and chemical states of the 5%Pd/PZs nanocomposites. Figure 4a shows the XPS survey spectrum of the 5%Pd/PZs catalyst, supporting that Pd/PZs was mainly composed of C, N, O, P, S, and Pd elements. The high resolution XPS spectra of Pd 3d in Figure 4b shows that there were four peaks in the Pd/PZs catalyst. The two peaks at 336.3 and 341.5 eV belonged to Pd0, and the other two peaks at 337.9 and 343.1 eV were assigned to Pd2+, respectively [24]. The atomic ratios of Pd0 and Pd2+ in the total Pd species are 34.3% and 65.7%, respectively. The Pd2+ species might be anchored on PZs by Pd–N or Pd–O bonds.
Figure 4c displays the C 1s XPS spectra of the PZs and Pd/PZs samples. The C 1s spectra of both PZs and Pd/PZs possessed four similar peaks at 284.6, 285.0, 285.7, and 286.6 eV, attributable to C–C, C–O–P, C–N, and C=C groups, respectively [38]. In the N 1s spectra in Figure 4d, two peaks at 298.3 and 397.5 eV could be fitted in the PZs sample, which were assigned to N–P/N=P groups in the cyclotriphosphazene structure and N–C groups, respectively [38]. Compared to the PZs sample, Pd/PZs had two similar peaks for the N–P/N=P and N–C groups, while the N–P/N=P peak shifted 0.1 eV to a higher binding energy, implying the formation of strong ligand bonds between PZs and Pd nanoparticles. Thus, Pd might be anchored on the PZs support by the coordination of Pd−N bonding. In the P 2p spectra in Figure 4e, two similar peaks at 133.7 and 134.2 eV, belonging to P–N/P=N and P–O groups, were observed in both the PZs and Pd/PZs samples [38]. Figure 4f reflects that the PZs and Pd/PZs samples had three similar oxygen species at 531.9, 533.4 and 534.8 eV, which were related to the C–O group, surface hydroxyl species (OH), and S=O group, respectively [38,39]. Compared to the PZs support, the binding energy of the OH group and S=O group peaks in the Pd/PZs sample shifted positively about 0.1–0.2 eV, suggesting that the OH and S=O groups also contribute to the coordination behavior of Pd nanoparticles and PZs microspheres. Thus, the phosphazene rings, OH groups and S=O groups might serve as the primary anchoring sites to bond Pd species.
The thermal behaviors of Pd/PZs and PZs were studied by using thermo-gravimetric analysis (TGA). Figure 5 shows the TGA curves of the PZs polymer and 5%Pd/PZs under N2 atmosphere. A dramatic weight loss was observed in the pure PZs in the temperature range of 470–580 °C, attributable to the decomposition of the polymer framework [38]. The loading of Pd did not change the initial decomposition temperature of the polymer, suggesting the high thermal stability of Pd/PZs. It should be noted that the 5%Pd/PZs nanocomposites had a weight loss of 9.5% at 70–200 °C, which might have arisen from the adsorbed water. These results indicate that the Pd nanoparticles loading enhanced the water absorbance of PZs. This is consistent with a similar phenomenon reported previously [38].

2.3. Catalytic Performance of the Pd/PZs Catalyst

To determine the catalytic behavior of the Pd/PZs catalyst, the catalytic performance for quinoline hydrogenation towards py-THQ was conducted in ethanol. The reactions were first carried out at 40 °C and 1.5 bar H2 for 2 h. As presented in Table 1, little quinoline was converted with PZs support, reflecting that PZs support had no catalytic hydrogenation activity for quinoline hydrogenation under mild conditions. Under the same conditions, 1%Pd/PZs could convert 6.4% of quinoline, with 97.2% py-THQ selectivity. When the Pd loading increased to 5%, the conversion increased to 67.3% for 2 h and maintained high selectivity towards py-THQ (>98%). Extending the reaction time to 4 h, the quinoline was almost completely transformed into py-THQ, with 98.9% conversion and 98.5% py-THQ selectivity. These results indicate that Pd/PZs could effectively catalyze the hydrogenation reaction of quinoline towards py-THQ. In comparison, the 5%Pd/C catalyst had a higher catalytic activity than 5%Pd/PZs, with 99% conversion for 2 h, but a lower selectivity of py-THQ (80.2%). The by-products bz-THQ (2.1%) and DHQ (17.3%) were also formed via the benzene ring hydrogenation. These results reveal that the properties of the support strongly impact the product distribution of Pd catalysts. There might be a strong synergetic effect of Pd and PZs in the chemselective hydrogenation of quinoline towards py-THQ.
To establish the optimal reaction conditions, the effects of reaction parameters, including reaction temperature, reaction time, and solvent, on the catalytic activity of 5%Pd/PZs were investigated in detail. The reaction rate was sensitive to reaction temperature. When varying the reaction temperature from 30 °C to 40 °C, the conversion of quinoline improved significantly from 41.8% to 67.9%, along with a high py-THQ selectivity above 98%. When increasing the temperature to 50 °C, the conversion increased to 97.2%, but the py-THQ selectivity reduced to 96.9%. Therefore, 40 °C was chosen in most reactions in this work. Figure 6 shows the results of quinoline hydrogenation for different times over a 5%Pd/PZs catalyst. At the beginning of the reaction, quinoline was converted to py-THQ by the selective hydrogenation of the heterocyclic ring, while little bz-THQ or DHQ was formed by hydrogenation of the benzene ring. When the reaction continued for 0.5 h, 33.1% of the quinoline was mostly converted into py-THQ. As the reaction continued, the conversion gradually increased from 33.1% to 98.9% at 4 h, while maintaining the high selectivity of py-THQ (>98%). Further extending the time to 5 h, py-THQ selectivity was still maintained at >98%. These results suggest that Pd/PZs has a highly catalytic activity and chemselectivity towards py-THQ under mild conditions.
As reported before, fewer polar substrates preferred to hydrogenate in more polar solvents in the heterogeneous catalytic system [40]. A series of solvents with different polarities (ethanol, water, tetrahydrofuran (THF), toluene, hexane) were used to investigate the influence of solvent on the reaction rates. As listed in Table 2, the polarity of solvents was in the following order: water > ethanol > THF > toluene > hexane. The reaction rates of Pd/PZs were influenced by the choice of different solvents, but little difference in the py-THQ selectivity (97.9–99.2%) was observed among the five solvents. The conversion of quinoline in the different solvents decreased as follows: toluene > ethanol > hexane > water > THF. Through careful analysis, it was shown that the reaction rates in the various solvents did not correlate with the polarity, which might be due to the higher multipole moments of some solvents [28]. Considering the toxicity of toluene, ethanol was selected for most catalytic reactions in this work.
The recyclability of 5%Pd/PZs catalyst is a crucial factor for industrial application. The catalytic reaction was performed in ethanol at 40 °C and 1.5 bar H2 for 4 h. After each cycle, the catalyst was recovered, washed with deionized water and ethanol, dried, and then used in the next run. As shown in Figure 7, the conversion did not reduce obviously during the five runs, while py-THQ selectivity kept almost constant (>98%), indicating that the Pd/PZs catalyst was highly stable for the hydrogenation of quinoline in ethanol. The structural properties of the reused Pd/PZs catalyst were characterized by XRD. As shown in Figure S2, no obvious change was observed in the structure of Pd/PZs after the reaction. The ICP analysis showed that the quantity of Pd leaching into the reaction solution in each run was negligible (<0.02%). These results reveal the good stability of the novel Pd nanocatalyst.

2.4. Possible Reaction Pathway

Based on the results mentioned above, a possible reaction pathway for the quinoline hydrogenation over Pd/PZs was proposed. It is considered that the quinoline hydrogenation process catalyzed by Pd/PZs catalyst under H2 might involve the competitive hydrogenation of benzene rings and heterocyclic rings into bz-THQ and py-THQ, as well as a further hydrogenation reaction to form DHQ [28]. In our reaction system, py-THQ was almost the only product in the hydrogenation reaction catalyzed by Pd/PZs, implying that the quinoline hydrogenation reaction is occurring through the exclusive adsorption of the N-heterocycle rather than the benzene ring. As displayed in Scheme 3, initially, the catalytic reaction begins with the adsorption of the quinoline substance on the surface of the PZs support via an interaction between the N-heterocycle of quinoline and –OH groups. At the same time, molecular H2 is adsorbed on the Pd nanoparticles and is activated and decomposed into H atoms. Then the N-heterocycle is attacked by active H atoms to form a hydrogenated N-heterocycle, thus producing py-THQ [27,28]. The N-heterocycle would be more easily hydrogenated by active H atoms at the Pd–PZs interface. Meanwhile, the dissociated hydrogen also might transfer to the surface of the PZs near the Pd–PZs interface to reduce the adsorbed N-heterocycle. Subsequently, py-THQ leaves the surface of the Pd/PZs catalyst quickly with the assistance of the solvent, and a new quinoline molecule fills the vacant space. Hence, the Pd/PZs nanocatalyst is highly active and stable in the controlled transformation of quinolines into py-THQ under low pressure of H2 and low temperature.
The catalytic activity of Pd/PZs might be affected by three factors as follows. Firstly, PZs microspheres have a unique spherical structure, rich phosphazene rings and O-anchoring sites, offering abundant anchoring sites to fix Pd nanoparticles of small size. Secondly, the rich hydroxyl groups on the surface of PZs would favor adsorption of N-heterocycle of quinoline via the strong interaction between the pyridine ring of quinoline and –OH groups, which might be the primary reason for the high selectivity for py-THQ. Thirdly, the Pd nanocatalyst is effective in the activation of H2 and the hydrogenation of quinoline to produce py-THQ. In a word, the metal–support synergetic effects between the PZs polymer and Pd nanoparticles contribute to the superior catalytic activity and high py-THQ selectivity of the Pd/PZs catalyst.

2.5. Hydrogenation of Other N/O/S-Heterocycles Compounds

The Pd/PZs catalyst was not only effective in catalyzing the hydrogenation of quinoline under low pressure of H2, but also the hydrogenation of various other quinoline compounds, O-heterocycles and S-heterocycles compounds. As presented in Table 3, 6-methylquinoline and 8-methylquinoline were almost completely converted into 6-methyl-1,6,7,8-tetrahydroquinoline and 8-methyl-1,6,7,8-tetrahydroquinoline over Pd/PZs under 40 °C and 1.5 bar H2 for 6 h, with a high yield above 98%. Pd/PZs gave moderate catalytic activity to transform 2-methylquinoline into 2-methyl-1,6,7,8-tetrahydroquinoline, with 58.3% conversion, and >99% conversion of 2-methyl-1,6,7,8-tetrahydroquinoline, under 60 °C and 1.5 bar H2 for 4 h. Prolonging the reaction time would enhance the conversions. Pd/PZs was also active in the transformation of 2,3-benzofuran into 2,3-dihydrobenzofuran, with >99% yield under 40 °C and 1.5 bar H2 for 4 h. When 2,3-benzothiophene was used as substance, a low conversion (2.6%) was obtained 40 °C and 1.5 bar H2 for 6 h, but it maintained a high selectivity of the 2,3-dihydrobenzothiophene above 99%. These results confirm that the Pd/PZs nanocatalyst was highly selective in the hydrogenation of N/O/S-heterocycles under mild conditions.

3. Materials and Methods

3.1. Materials

All reagents were purchased from Aladdin and used without further purification. The commercial carbon support was purchased from Shanghai Macklin Biochemical Technology Co., Ltd. (Activated charcoal) (Shanghai, China).

3.2. Preparation of the Poly-[Cyclotriphosphazene-co-(4,4′-Sulfonyldiphenol)] Microspheres

PZs microspheres were prepared by precipitation polymerization of HCCP and BPS according to procedures reported in the literature. In brief, 3.48 g of HCCP and 7.51 g of BPS was dissolved in 500 mL of acetonitrile. Then, 15 mL of TEA was added to the solution. The obtained mixtures were stirred in an ultrasonic bath at 40 °C for 1 h. Subsequently, the resulting solid was separated from the solution by centrifugation and washed with ethanol and water several times. The solid was dried under a vacuum, giving PZs microspheres in the form of a white powder.

3.3. Preparation of the Pd/PZs Catalyst

The Pd/PZs catalyst was prepared via an ultrasound-assisted reduction method. First, 0.2 g of PZs microspheres was mixed with 20 mL of water. A certain amount of PdCl2 aqueous solution (e.g., 0.1 g Pd /mL, 0.1 mL for 5%Pd/PZs) was added to the aqueous PZs microspheres dispersion and was stirred at room temperature for 30 min. After that, 10 mL of NaBH4 aqueous solution (0.1 mol/L) was added the above mixture dropwise under ultrasonic treatment, resulting in a grayish suspension. The reaction system was treated in an ultrasonic bath for 30 min. The resulting solids were separated by centrifugation and washing with water and then dried at 80 °C in a vacuum oven. The 5%Pd/C was also prepared via a similar process with commercial carbon as support. The size of the Pd nanoparticles, calculated by the Scherrer equation, was about 14.3 nm (Figure S3).

3.4. Catalytic Performance Evaluation

The hydrogenation of quinolines was conducted in a 50 mL stainless-steel autoclave. The experimental procedure was as follows: 0.5 mmol of quinoline, 65 mg of catalyst, and 10 mL of ethanol were injected into the autoclave. The autoclave was flushed with H2 five times to remove the air, stirred at 400 rpm under 1.5 bar H2 and then heated to the desired temperature. After the reaction, the substrate and products were quantified by a gas chromatograph (GC, Shimadzu GC-2010, Kyoto, Japan) with an Agilent capillary column DB-5, and the products were also identified by GC coupled with a mass spectrometer (GC-MS, Shimadzu GCMS-QP2010, Kyoto, Japan). The quantification of products was identified by an external standard method.
The calculation of conversion was as follows:
c o n v e r s i o n % = n 0 n 0 n 0 × 100 %
where n0 and n0 are the initial mol of quinoline added to the reactor and the residual mol of quinoline after the reaction, respectively.
The selectivity of products was calculated by the mol of products and the mol of converted quinoline.
S e l e c t i v i t y   o f   p r o d u c t = n ( p r o d u c t ) n 0 n 0 × 100 %

4. Conclusions

We have reported a simple strategy for the fabrication of Pd/PZs nanocomposites by precipitation polymerization and a chemical reduction process. The PZs support offered rich N/O-anchoring sites to anchor stably Pd nanoparticles. Various characterization results revealed that the small-sized Pd nanoparticles were well dispersed on the surface of the PZs microspheres. The novel catalyst was found to effectively catalyze quinoline hydrogenation towards py-THQ through the exclusive hydrogenation of the N-heterocycle rather than the benzene ring under mild conditions. The reaction rates could be impacted by many factors, including Pd content, reaction time, reaction temperature, and solvents. An optimal activity with 98.9% conversion and 98.5% py-THQ selectivity was obtained over 5%Pd/PZs under 1.5 bar H2 and 40 °C for 4 h. The metal–support synergistic effects of metallic Pd and PZs support co-contributed to the outstanding catalytic activity of Pd/PZs in the N-heterocycles hydrogenation reactions. This work provides novel efficient and stable catalysts for the hydrogenation reactions of N-heterocycles under mild conditions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14060345/s1. Characterizations; Figure S1: (a) Nitrogen sorption isotherms of 5%Pd/PZs, (b) pore-size distribution curves of 5%Pd/PZs; Figure S2: The XRD pattern of 5%Pd/PZs before and after reaction; Figure S3: XRD pattern of 5%Pd/C sample; Table S1: The comparison of chemoselective hydrogenation of quinoline towards py-THQ with various catalysts [5,8,21,22,28,41,42,43,44,45,46].

Author Contributions

X.C. was responsible for original draft preparation, review, editing, and funding acquisition. Q.X. and Y.Y. were responsible for original draft preparation, investigation, and validation. B.D. and Z.Z. were responsible for investigation, data curation and validation. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Baima Lake Laboratory Joint Funds of the Zhejiang Provincial Natural Science Foundation of China (LBMHY24E060003), the Scientific Research Foundation of Zhejiang Sci-Tech University (19212450-Y), and the Fundamental Research Funds of Zhejiang Sci-Tech University (23212112-Y).

Data Availability Statement

The authors can confirm that all relevant data are included in the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chakraborty, S.; Brennessel, W.W.; Jones, W.D. A Molecular Iron Catalyst for the Acceptorless Dehydrogenation and Hydrogenation of N-heterocycles. J. Am. Chem. Soc. 2014, 136, 8564–8567. [Google Scholar] [CrossRef]
  2. Wang, D.S.; Chen, Q.A.; Lu, S.M.; Zhou, Y.G. Asymmetric Hydrogenation of Heteroarenes and Arenes. Chem. Rev. 2012, 112, 2557–2590. [Google Scholar] [CrossRef] [PubMed]
  3. Hegedűs, L.; Nguyen, T.T.T.; Lévay, K.; László, K.; Sáfrán, G.; Beck, A. Poisoning and Reuse of Supported Precious Metal Catalysts in the Hydrogenation of N-Heterocycles, Part II: Hydrogenation of 1-Methylpyrrole over Rhodium. Catalysts 2022, 12, 730. [Google Scholar] [CrossRef]
  4. Zhang, Y.J.; Pang, S.F.; Wei, Z.H.; Jiao, H.J.; Dai, X.C.; Wang, H.L.; Shi, F. Synthesis of a Molecularly Defined Single-active Site Heterogeneous Catalyst for Selective Oxidation of N-heterocycles. Nat. Commun. 2018, 9, 1465. [Google Scholar] [CrossRef]
  5. Wei, Z.Z.; Chen, Y.Q.; Wang, J.; Su, D.F.; Tang, M.H.; Mao, S.J.; Wang, Y. Cobalt Encapsulated in N-Doped Graphene Layers: An Efficient and Stable Catalyst for Hydrogenation of Quinoline Compounds. ACS Catal. 2016, 6, 5816–5822. [Google Scholar] [CrossRef]
  6. Zhu, Q.Z.; Yin, X.G.; Tan, Y.J.; Wei, D.D.; Li, Y.; Pei, X.L. Highly Dispersed Palladium Nanocatalyst Anchored on N-doped Nanoporous Carbon Microspheres Derived from Chitosan for Efficient and Stable Hydrogenation of Quinoline. Int. J. Biol. Macromol. 2024, 254, 127949. [Google Scholar] [CrossRef]
  7. Fish, R.H.; Thormodsen, A.D.; Cremer, G.A. Homogeneous Catalytic Hydrogenation. 1. Regiospecific Reductions of Polynuclear Aromatic and Polynuclear Heteroaromatic Nitrogen Compounds Catalyzed by Transition Metal Carbonyl Hydrides. J. Am. Chem. Soc. 1982, 104, 5234–5237. [Google Scholar] [CrossRef]
  8. Zhang, S.; Xia, Z.M.; Ni, T.; Zhang, H.; Wu, C.; Qu, Y.Q. Tuning Chemical Compositions of Bimetallic AuPd Catalysts for Selective Catalytic Hydrogenation of Halogenated Quinolines. J. Mater. Chem. A 2017, 5, 3260–3266. [Google Scholar] [CrossRef]
  9. Lyons, T.W.; Leibler, I.N.; He, C.Q.; Gadamsetty, S.; Estrada, G.J.; Doyle, A.G. Broad Survey of Selectivity in the Heterogeneous Hydrogenation of Heterocycles. J. Org. Chem. 2024, 89, 1438–1445. [Google Scholar] [CrossRef]
  10. Chen, F.; Surkus, A.E.; He, L.; Pohl, M.M.; Radnik, J.; Topf, C.; Junge, K.; Beller, M. Selective Catalytic Hydrogenation of Heteroarenes with N-Graphene-Modified Cobalt Nanoparticles (Co3O4–Co/NGr@α-Al2O3). J. Am. Chem. Soc. 2015, 137, 11718–11724. [Google Scholar] [CrossRef]
  11. Colliere, V.; Verelst, M.; Lecante, P.; Axet, M.R. Colloidal Ruthenium Catalysts for Selective Quinaldine Hydrogenation: Ligand and Solvent Effects. Chem.-Eur. J. 2024, 30, e202302131. [Google Scholar] [CrossRef] [PubMed]
  12. Chen, W.M.; Che, Y.X.; Xia, J.; Zheng, L.R.; Lv, H.F.; Zhang, J.; Liang, H.W.; Meng, X.M.; Ma, D.; Song, W.G.; et al. Metal–Sulfur Interfaces as the Primary Active Sites for Catalytic Hydrogenations. J. Am. Chem. Soc. 2024, 146, 11542–11552. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, Z.; Du, H. Enantioselective Metal-free Hydrogenations of Disubstituted Quinolines. Org. Lett. 2015, 17, 6266–6269. [Google Scholar] [CrossRef] [PubMed]
  14. Yamaguchi, R.; Ikeda, C.; Takahashi, Y.; Fujita, K. Homogeneous Catalytic System for Reversible Dehydrogenation—Hydrogenation Reactions of Nitrogen Heterocycles with Reversible Interconversion of Catalytic Species. J. Am. Chem. Soc. 2009, 131, 8410–8412. [Google Scholar] [CrossRef] [PubMed]
  15. Rosales, M.; Bastidas, L.J.; Gonzalez, B.; Vallejo, R.; Baricelli, P.J. Kinetics and Mechanisms of Homogeneous Catalytic Reactions. Part 11. Regioselective Hydrogenation of Quinoline Catalyzed by Rhodium Systems Containing 1,2-Bis(diphenylphosphino)ethane. Catal. Lett. 2011, 141, 1305–1310. [Google Scholar] [CrossRef]
  16. Facchetti, G.; Christodoulou, M.S.; Binda, E.; Fusè, M.; Rimoldi, I. Asymmetric Hydrogenation of 1-aryl substituted-3,4-Dihydroisoquinolines with Iridium Catalysts Bearing Different Phosphorus-Based Ligands. Catalysts 2020, 10, 914. [Google Scholar] [CrossRef]
  17. Wu, J.G.; Li, X.; Fu, K.; Cao, D.; Cheng, D.J. Constructing Fully Exposed Pt Atomically Dispersed Catalysts for Enhanced Multifunctional Selective Hydrogenation Reactions. Chem. Eng. J. 2024, 481, 148706. [Google Scholar] [CrossRef]
  18. Campanati, M.; Vaccari, A.; Piccolo, O. Mild Hydrogenation of Quinoline: 1. Role of Reaction Parameters. J. Mol. Catal. A Chem. 2002, 179, 287–292. [Google Scholar] [CrossRef]
  19. Hashimoto, N.; Takahashi, Y.; Hara, T.; Shimazu, S.; Mitsudome, T.; Mizugaki, T.; Jitsukawa, K.; Kaneda, K. Fine Tuning of Pd0 Nanoparticle Formation on Hydroxyapatite and Its Application for Regioselective Quinoline Hydrogenation. Chem. Lett. 2010, 39, 832–834. [Google Scholar] [CrossRef]
  20. Campanati, M.; Casagrande, M.; Fagiolino, I.; Lenarda, M.; Storaro, L.; Battagliarin, M.; Vaccari, A. Mild Hydrogenation of Quinoline: 2. A Novel Rh-containing Pillared Layered Clay Catalyst. J. Mol. Catal. A Chem. 2002, 184, 267–272. [Google Scholar] [CrossRef]
  21. Dell’Anna, M.M.; Capodiferro, V.F.; Mali, M.; Manno, D.; Cotugno, P.; Monopoli, A.; Mastrorilli, P. Highly Selective Hydrogenation of Quinolines Promoted by Recyclable Polymer Supported Palladium Nanoparticles under Mild Conditions in Aqueous Medium. Appl. Catal. A Gen. 2014, 481, 89–95. [Google Scholar] [CrossRef]
  22. He, Z.H.; Li, N.; Wang, K.; Wang, W.T.; Liu, Z.T. Selective Hydrogenation of Quinolines over a CoCu Bimetallic Catalyst at Low Temperature. Mol. Catal. 2019, 470, 120–126. [Google Scholar] [CrossRef]
  23. Liu, X.; Zhang, B.; Fei, B.; Chen, X.; Zhang, J.; Mu, X. Tunable and Selective Hydrogenation of Furfural to Furfuryl alcohol and Cyclopentanone over Pt Supported on Biomass-derived Porous Heteroatom Doped Carbon. Faraday Discuss. 2017, 202, 79–98. [Google Scholar] [CrossRef] [PubMed]
  24. Ji, G.J.; Duan, Y.N.; Zhang, S.C.; Fei, B.H.; Chen, X.F.; Yang, Y. Selective Semihydrogenation of Alkynes Catalyzed by Pd Nanoparticles Immobilized on Heteroatom-doped Hierarchical Porous Carbon Derived from Bamboo Shoots. ChemSusChem 2017, 10, 3427–3434. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, X.; Cheng, L.L.; Yang, Y.G.; Chen, X.F.; Chen, F.T.; Lu, W.Y. Construction of High-Density Fe Clusters Embedded in a Porous Carbon Nitride Catalyst with Effectively Selective Transformation of Benzene. ACS Sus. Chem. Eng. 2023, 11, 1518–1526. [Google Scholar] [CrossRef]
  26. Cheng, L.; Sun, S.; Chen, X.; Chen, F.; Chen, X.; Lu, W. Convenient Fabrication of Ultrafine VOx Decorated on Porous g-C3N4 for Boosting Photocatalytic Degradation of Pharmaceuticals with Peroxymonosulfate. Surf. Interfaces 2023, 42, 103300. [Google Scholar] [CrossRef]
  27. Zhang, L.; Cheng, L.; Hu, Y.; Xiao, Q.; Chen, X.; Lu, W. Robust Co3O4 Nanocatalysts Supported on Biomass-derived Porous N-doped Carbon Toward Low-Pressure Hydrogenation of Furfural. Front. Mater. Sci. 2023, 17, 230645. [Google Scholar] [CrossRef]
  28. Gong, Y.T.; Zhang, P.F.; Xu, X.; Li, Y.; Li, H.R.; Wang, Y. A Novel Catalyst Pd@ompg-C3N4 for Highly Chemoselective Hydrogenation of Quinoline under Mild Conditions. J. Catal. 2013, 297, 272–280. [Google Scholar] [CrossRef]
  29. Spitaleri, A.; Pertici, P.; Scalera, N.; Vitulli, G.; Hoang, M.; Turney, T.W.; Gleria, M. Supported Ruthenium Nanoparticles on Polyorganophosphazenes: Preparation, Structural and Catalytic Studies. Inorg. Chim. Acta 2003, 352, 61–71. [Google Scholar] [CrossRef]
  30. Sun, S.S.; Peng, X.Y.; Guo, X.C.; Chen, X.F.; Liu, D. Boosting Solvent-Free Aerobic Oxidation of Benzylic Compounds into Ketones over Au-Pd Nanoparticles Supported by Porous Carbon. Catalysts 2024, 14, 158. [Google Scholar] [CrossRef]
  31. Ge, N.Q.; Hu, X.M.; Pan, Z.J.; Cai, S.J.; Fu, F.Y.; Wang, Z.Q.; Yao, J.M.; Liu, X.D. Sustainable Fabrication of Cellulose Aerogel Embedded with ZnO@noble Metal (Ag, Au, Ag-Au) NPs for Sensitive and Reusable SERS Application. Colloid. Surface. A 2023, 671, 131650. [Google Scholar] [CrossRef]
  32. Shi, M.; Hu, N.N.; Liu, H.M.; Qian, C.; Lv, C.; Wang, S. Controlled Synthesis of Pt-loaded Yolk-shell TiO2@SiO2 Nanoreactors as Effective Photocatalysts for Hydrogen Generation. Front. Mater. Sci. 2022, 16, 220591. [Google Scholar] [CrossRef]
  33. Zhao, Z.Z.; Xu, Z.; Chen, J.Y.; Zhong, M.Q.; Wang, J.H.; Chew, J.W. A Review on Functional Applications of Polyphosphazenes as Multipurpose Material for Lithium-ion Batteries. J. Energy Storage 2024, 85, 111049. [Google Scholar] [CrossRef]
  34. Wang, M.H.; Fu, J.W.; Huang, D.D.; Zhang, C.; Xu, Q. Silver Nanoparticles-decorated Polyphosphazene Nanotubes: Synthesis and Applications. Nanoscale 2013, 5, 7913–7919. [Google Scholar] [CrossRef]
  35. Wang, M.H.; Fu, J.W.; Chen, Z.H.; Wang, X.Z.; Xu, Q. In Situ Growth of Gold Nanoparticles onto Polyphosphazene Microspheres with Amino-groups for Alcohol Oxidation in Aqueous Solutions. Mater. Lett. 2015, 143, 201–204. [Google Scholar] [CrossRef]
  36. Ahmad, M.; Nawaz, T.; Assiri, M.A.; Hussain, R.; Hussain, I.; Imran, M.; Ali, S.; Wu, Z.P. Fabrication of Bimetallic Cu-Ag Nanoparticle-Decorated Poly(cyclotriphosphazene-co-4,4’-sulfonyldiphenol) and Its Enhanced Catalytic Activity for the Reduction of 4-Nitrophenol. ACS Omega 2022, 7, 7096–7102. [Google Scholar] [CrossRef]
  37. Huang, M.H.; Li, Z.; Du, L.L.; Jin, Z.K.; Li, R.H. CuPd/MgO for Efficient Catalytic Hydrogen Production from Formaldehyde Solution at Room Temperature. Chinese J. Inorg. Chem. 2022, 38, 2452–2458. [Google Scholar]
  38. Chen, S.S.; Li, Z.W.; Yuan, W.B.; Duan, W.S.; Qiao, C.D.; Yao, J.S.; Zhang, C.B.; Zhao, H.; Li, M.; Yang, G.H. Polyphosphazene-Functionalized Microspheres as Efficient Catalysts for the Knoevenagel Reaction under Mild Conditions. ChemPlusChem 2022, 87, e202200249. [Google Scholar] [CrossRef] [PubMed]
  39. Zhang, X.F.; Lin, X.Y.; Huang, X.Y.; Chen, Y.Q.; Lin, S.; Huang, X.; Xie, Z.L. The Role Identification of Nitrogen Dopant in Nanocarboncatalysis. Carbon Future 2024, 1, 9200008. [Google Scholar]
  40. Mukherjee, S.; Vannice, M.A. Solvent Effects in Liquid-phase Reactions: I. Activity and Selectivity during Citral Hydrogenation on Pt/SiO2 and Evaluation of Mass Transfer Effects. J. Catal. 2006, 243, 108–130. [Google Scholar] [CrossRef]
  41. Ren, Y.S.; Wang, Y.X.; Li, X.; Zhang, Z.H.; Chi, Q. Selective Hydrogenation of Quinolines into 1,2,3,4-Tetrahydroquinolines over a Nitrogen-doped Carbon-supported Pd catalyst. New. J. Chem. 2018, 42, 16694–16702. [Google Scholar] [CrossRef]
  42. Ren, D.; He, L.; Yu, L.; Ding, R.S.; Liu, Y.M.; Cao, Y.; He, H.Y.; Fan, K.N. An Unusual Chemoselective Hydrogenation of Quinoline Compounds Using Supported Gold Catalysts. J. Am. Chem. Soc. 2012, 134, 17592–17598. [Google Scholar] [CrossRef] [PubMed]
  43. Niu, M.M.; Wang, Y.H.; Chen, P.; Du, D.J.; Jiang, J.Y.; Jin, Z.L. Highly Efficient and Recyclable Rhodium Nanoparticle Catalysts for Hydrogenation of Quinoline and Its Derivatives. Catal. Sci. Technol. 2015, 5, 4746–4749. [Google Scholar] [CrossRef]
  44. Chen, Y.G.; Yu, Z.J.; Chen, Z.A.; Shen, R.; Wang, Y.; Cao, X.; Peng, Q.; Li, Y.D. Controlled One-pot Synthesis of RuCu Nanocages and Cu@Ru Nanocrystals for the Regioselective Hydrogenation of Quinoline. Nano. Res. 2016, 9, 2632–2640. [Google Scholar] [CrossRef]
  45. Sahoo, B.; Kreyenschulte, C.; Agostini, G.; Lund, H.; Bachmann, S.; Scalone, M.; Junge, K.; Beller, M. A Robust Iron Catalyst for the Selective Hydrogenation of Substituted (iso) Quinolones. Chem. Sci. 2018, 9, 8134–8141. [Google Scholar] [CrossRef]
  46. Shaikh, M.N.; Abdelnaby, M.M.; Hakeem, A.S.; Nasser, G.A.; Yamani, Z.H. Co3O4/Nitrogen-doped Graphitic Carbon/Fe3O4 Nanocomposites as Reusable Catalysts for Hydrogenation of Quinoline, Cinnamaldehyde, and Nitroarenes. ACS Appl. Nano Mater. 2021, 4, 3508–3518. [Google Scholar] [CrossRef]
Scheme 1. Possible reaction pathways for quinolone hydrogenation.
Scheme 1. Possible reaction pathways for quinolone hydrogenation.
Catalysts 14 00345 sch001
Scheme 2. The fabrication procedure of the Pd/PZs composites (top). The polycondensation of the comonomers HCCP and BPS and the crosslinked structure of the PZs (bottom).
Scheme 2. The fabrication procedure of the Pd/PZs composites (top). The polycondensation of the comonomers HCCP and BPS and the crosslinked structure of the PZs (bottom).
Catalysts 14 00345 sch002
Figure 1. XRD patterns of 5%Pd/PZs and PZs samples.
Figure 1. XRD patterns of 5%Pd/PZs and PZs samples.
Catalysts 14 00345 g001
Figure 2. FTIR spectra of the 5%Pd/PZs and PZs samples.
Figure 2. FTIR spectra of the 5%Pd/PZs and PZs samples.
Catalysts 14 00345 g002
Figure 3. (a,b) Typical SEM images of PZs; (ce) TEM images; and (f) HRTEM images of 5%Pd/PZs.
Figure 3. (a,b) Typical SEM images of PZs; (ce) TEM images; and (f) HRTEM images of 5%Pd/PZs.
Catalysts 14 00345 g003
Figure 4. (a) XPS survey spectrum of 5%Pt/PZs catalyst; high resolution XPS spectra of (b) Pd 3d, (c) C 1s, (d) N 1s, (e) P 2p, and (f) O 1s.
Figure 4. (a) XPS survey spectrum of 5%Pt/PZs catalyst; high resolution XPS spectra of (b) Pd 3d, (c) C 1s, (d) N 1s, (e) P 2p, and (f) O 1s.
Catalysts 14 00345 g004
Figure 5. TGA curve of PZs and 5%Pd/PZs under N2 atmosphere.
Figure 5. TGA curve of PZs and 5%Pd/PZs under N2 atmosphere.
Catalysts 14 00345 g005
Figure 6. Conversion and py-THQ selectivity as a function of time at 40 °C using 10 mL ethanol as a solvent over 65 mg 5%Pd/PZs.
Figure 6. Conversion and py-THQ selectivity as a function of time at 40 °C using 10 mL ethanol as a solvent over 65 mg 5%Pd/PZs.
Catalysts 14 00345 g006
Figure 7. Recycling experiments of the 5%Pd/PZs catalyst under 40 °C and 1.5 bar H2 for 4 h.
Figure 7. Recycling experiments of the 5%Pd/PZs catalyst under 40 °C and 1.5 bar H2 for 4 h.
Catalysts 14 00345 g007
Scheme 3. Possible reaction pathway of quinoline hydrogenation over Pd/PZs.
Scheme 3. Possible reaction pathway of quinoline hydrogenation over Pd/PZs.
Catalysts 14 00345 sch003
Table 1. Catalytic activity of quinoline hydrogenation over various catalysts.
Table 1. Catalytic activity of quinoline hydrogenation over various catalysts.
EntrySamplesT (°C)Time (h)Con. (%)Sel. (%)
1PZs402<1<1
21%Pd/PZs4026.497.2
35%Pd/PZs40267.398.2
45%Pd/PZs40498.998.5
55%Pd/C40299.080.2
65%Pd/PZs30241.898.5
75%Pd/PZs40267.398.2
85%Pd/PZs50292.796.9
Reaction conditions: 0.5 mmol quinoline, 65 mg catalyst, 1.5 bar H2, 10 mL ethanol.
Table 2. Catalytic activity of quinoline hydrogenation over 5%Pd/PZs in different solvents.
Table 2. Catalytic activity of quinoline hydrogenation over 5%Pd/PZs in different solvents.
EntrySolventCon. (%)Sel. (%)Solvent Polarity
1ethanol67.398.24.3
2water46.299.210.2
3THF40.398.14.2
4toluene90.797.92.4
5hexane51.698.50.06
Reaction conditions: 0.5 mmol quinoline, 65 mg Pd/PZs, 40 °C, 2 h, 1.5 bar H2, 10 mL solvent.
Table 3. Catalytic activity of the hydrogenation of other N/O/S-heterocycles compounds.
Table 3. Catalytic activity of the hydrogenation of other N/O/S-heterocycles compounds.
EntrySubstratesProductsT (°C)t (h)Con. (%)Sel. (%)
1Catalysts 14 00345 i001Catalysts 14 00345 i00240698.1>99
2Catalysts 14 00345 i003Catalysts 14 00345 i00440699.3>99
3Catalysts 14 00345 i005Catalysts 14 00345 i00660458.3>99
4Catalysts 14 00345 i007Catalysts 14 00345 i008406100>99
5Catalysts 14 00345 i009Catalysts 14 00345 i0104062.6>99
Reaction conditions: 0.5 mmol substance, 65 mg Pd/PZs, 1.5 bar H2, 10 mL ethanol.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, X.; Xiao, Q.; Yang, Y.; Dong, B.; Zhao, Z. Constructing Polyphosphazene Microsphere-Supported Pd Nanocatalysts for Efficient Hydrogenation of Quinolines under Mild Conditions. Catalysts 2024, 14, 345. https://doi.org/10.3390/catal14060345

AMA Style

Chen X, Xiao Q, Yang Y, Dong B, Zhao Z. Constructing Polyphosphazene Microsphere-Supported Pd Nanocatalysts for Efficient Hydrogenation of Quinolines under Mild Conditions. Catalysts. 2024; 14(6):345. https://doi.org/10.3390/catal14060345

Chicago/Turabian Style

Chen, Xiufang, Qingguang Xiao, Yiguo Yang, Bo Dong, and Zhengping Zhao. 2024. "Constructing Polyphosphazene Microsphere-Supported Pd Nanocatalysts for Efficient Hydrogenation of Quinolines under Mild Conditions" Catalysts 14, no. 6: 345. https://doi.org/10.3390/catal14060345

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

Article Metrics

Back to TopTop