Next Article in Journal
Electron Mediation and Photocurrent Enhancement in Dunalliela salina Driven Bio-Photo Electrochemical Cells
Previous Article in Journal
The Roles of Precursor-Induced Metal–Support Interaction on the Selective Hydrogenation of Crotonaldehyde over Ir/TiO2 Catalysts
Previous Article in Special Issue
Novel Recyclable Pd/H-MOR Catalyst for Suzuki-Miyaura Coupling and Application in the Synthesis of Crizotinib
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetically Reusable Fe3O4@NC@Pt Catalyst for Selective Reduction of Nitroarenes

1
School of Materials Science and Engineering, Institute of Molecular Engineering and Applied Chemistry, Anhui University of Technology, Ma’anshan 243002, China
2
College of Biology and Environmental Engineering, Zhejiang Shuren University, Hangzhou 310015, China
3
College of Petroleum Chemical Industry, Changzhou University, Changzhou 213164, China
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(10), 1219; https://doi.org/10.3390/catal11101219
Submission received: 15 September 2021 / Revised: 5 October 2021 / Accepted: 6 October 2021 / Published: 9 October 2021
(This article belongs to the Special Issue Characterization Analysis of Heterogeneous Catalysts)

Abstract

:
A novel reusable Fe3O4@NC@Pt heterogeneous catalyst was synthesized by immobilizing platinum on nitrogen-doped carbon magnetic nanostructures. It was characterized by infrared analysis (FT-IR), X-ray diffraction (XRD), transmission electron microscopy (TEM), and vibrating sample magnetometer (VSM). The catalytic efficiency of Fe3O4@NC@Pt was investigated by reduction of nitro aromatic compounds. The catalyst showed good catalytic activity, wide range of substrates, and good chemical selectivity, especially for the substrates of compounds containing halide and carbonyl groups. The magnetically catalyst can readily be reused up to ten cycles without loss of catalytic activity. Moreover, the key pharmaceutical intermediate Lorlatini can be facilely achieved through this strategy.

1. Introduction

Anilines are one of the most important intermediates in fine chemicals, which can manufacture highly value-added chemicals such as natural products, pharmaceuticals, agrochemicals, surfactants, plastics, polymers, dyes, and pigments [1,2,3,4,5]. Traditionally, it can be synthesized by reduction of the corresponding nitroarenes using stoichiometric amounts of base metals like Fe/Zn and sulfides as the reducing reagents [6]. However, these reduction procedures may lead to vast amounts of solid wastes and hazardous by-products, causing serious environmental problems [7,8].
Catalytic reductions by heterogeneous transition metal catalysts have attracted much interest as a sustainable and efficient strategy because the catalysts can be easily separated and recovered from the reaction mixtures for reuse [9,10,11]. Magnetic heterogeneous catalysts have attracted remarkable attention by organic chemists due to their feasible removal and reusing by means of an external magnetic field [12,13,14,15]. A heterogeneous catalyst can be formed by loading active metals or molecular catalysts onto magnetic materials, which is not time-consuming and prevents losing of catalyst during the separation process. Supported noble metal catalysts like Pd, Au, Ag, Ru, and Rh have been widely investigated for the selective reduction of various substituted nitroarenes to the corresponding anilines due to their high activity [16,17,18,19,20]. However, these catalysts frequently generate some unwanted by-products, for example, dehalogenation of aromatic nitro compounds, which seriously lowers the product quality, limiting their large-scale industrial applications [21].
Biomass materials are widely used in industrial catalysis, carbon materials in particular, excellent catalytic activity was obtained by loading transition metals onto them [22,23,24,25], which are our longstanding interest [26,27,28,29,30]. Platinum has been demonstrated to exhibit much higher selectivity and dissociative capacity toward H2 in hydrogenation of the nitroarenes. Meanwhile, its lower price and lower dehalogenation trend than palladium means it has great potential in the chemoselective hydrogenation of halogenated nitroarenes. In this paper, we find a facile approach to prepare a magnetically-reusable, nitrogen-doped carbon core-shell Fe3O4@NC@Pt catalyst for the selective reduction of nitroarenes, especially aromatic nitro compounds containing halogen and carbonyl substituent. The catalytic activity was optimized, and the substrate was expanded. The catalyst has also been used to synthesize lorlatini, the key pharmaceutical intermediates with high efficiency and cyclability.

2. Results and Discussion

From Scheme 1, the preparation process of Fe3O4@NC@Pt included three steps. Firstly, the Fe3O4 microspheres were prepared via a facile solvothermal reaction using FeCl3·6H2O as the raw material and ethylene glycol as the solvent. Secondly, a nitrogen-doped carbon layer was wrapped on the outside of Fe3O4 microspheres, the glucose served as a carbon source, and ethylenediamine (EDA) served as a nitrogen source. Finally, Pt(0) nanoparticles, obtained by reduction of H2PtCl6·6H2O by sodium borohydride, were supported on Fe3O4@NC@Pt catalysts surface. To investigate the effects of nitrogen-doped carbon on the Fe3O4@NC@Pt catalytic activity of reduction of nitroarenes, Fe3O4@C@Pt was prepared in the same way mentioned above without adding ethylenediamine.
Figure 1 exhibits the FT-IR spectra of the Fe3O4, Fe3O4@NC, and Fe3O4@NC@Pt. In the spectrum (Figure 1a), the peaks were observed at around 583 cm−1(Fe-O stretching), 1626 cm−1 (C=O bound stretching), and 1425 cm−1 (pyrrole ring framework stretching), respectively. The characteristic peaks of Fe3O4@NC appeared at 579, 1689, and 3428 cm−1 in Figure 1b, they were related to Fe-O, C=O band stretching, and the -OH stretching vibrations of carboxylic acids, respectively, indicating the surface functionalization on the surface of the Fe3O4@NC. It is very important for Pt immobilization with the help of these functional groups. As shown in Figure 1c, the absorption peaks of Fe3O4@NC@Pt at 581 and 1705 cm−1 related to Fe3O4@NC. Meanwhile, it can be seen that the peak intensity of Fe-O, C=O, and -OH was obviously weaker compared with Fe3O4@NC. This may suggest that the interactions exist between platinum and Fe3O4@NC. A similar conclusion was reached in previous reports [31].
The catalyst composition and formulation were characterized by X-ray diffraction (XRD). In Figure 2a–c, the peaks at 2θ values of 18.2°, 30.1°, 35.4°, 37.1°, 43.1°, 53.4°, 53.4°, 56.9°, and 62.5° are well in agreement with the pure Fe3O4 (JCPDS 74-0748) [32,33]. In Figure 2c, the peaks at 2θ values of 40.0°, 46.5°, and 67.8° are attributed to the Pt (JCPDS 01-1194) [14]. All XRD patterns have no other impurities peaks, the sharp and strong diffraction peaks also confirm the well crystallization of the products. Meanwhile, in Figure 2b,c, the broad peak at 2θ = 21.6° is the characteristic reflection of amorphous nitrogen-doped carbon [34].
The specific surface area and pore structure of Fe3O4@NC@Pt were examined using low-temperature nitrogen adsorption–desorption measurements, the specific surface area was calculated by BET method [35], and the pore size distribution was calculated by Barrett–Joyner–Halenda (BJH) method [36]. From Figure 3, it can be seen that the isotherm of Fe3O4@NC@Pt is type IV compared with various type isotherms classified by the IUPAC [37]. The isotherms showed a small hysteresis loop which is the characteristic of mesoporosity. The specific surface area is 5.8781 m2/g and the total pore volume is 0.019498 cm3/g, respectively. As shown in the BJH pore size distribution curves (Figure 3, inset), the catalyst pores are mainly in the range of 5–80 nm, the average diameter of the catalyst is 19.8 nm.
Vibrating sample magnetometer (VSM) was used to test the magnetic properties of the Fe3O4@NC@Pt in fields ranging from +30 to −30 kOe at room temperature (Figure 4). The Fe3O4@NC@Pt shows ferromagnetic properties from the hysteresis loop, whose saturation magnetization value was 60.8 emu/g. With the help of an external magnetic field, the Fe3O4@NC@Pt could be quickly separated from the reaction system and reused for the next time.
Transmission Electron Microscopy (TEM) was used to further confirm the micromorphology and structure of materials. The prepared catalysts have obvious core structure (Fe3O4 Core), and small particles can be found on the surface of the core, the catalyst particles are uniform, and the size is ca. 3–4 μm (Figure 5a). In addition, the EDS mapping verified the presence of C, N, O, Fe, and Pt in Figure 5b and the dispersion and confinement of Pt on Fe3O4@NC@Pt. The EDS mapping of Fe3O4@NC@Pt showed (Figure 5c) that the content of the platinum is 16.36% in semi-quantified.
The magnetic catalyst Fe3O4@NC@Pt was evaluated in the reduction reaction. 1-chloro-4-nitrobenzene 1a was used as the model compound to explore the catalytic activity of Fe3O4@NC@Pt. As we all know, the catalysts and reducing agents are two important factors for reduction reaction. For this purpose, a series of controlled experiments were studied, toluene was used as solvent and hydrazine hydrate was reducing agent, and the reaction temperature was 70 °C. The reduction reaction did not occur when Fe3O4@NC was used in the reaction, nevertheless, when replaced by Fe3O4@C@Pt, it showed much better catalytic activity but was less than Fe3O4@NC@Pt under the same conditions. Meanwhile, the reaction cannot occur without adding any catalyst (Table 1, entries 1–4). Therefore, two conclusions can be drawn: firstly, platinum (Pt) particles on the surface of the catalyst can effectively catalyze the reduction of nitro compounds to amine; secondly, N doped Fe3O4@NC@Pt catalyst is further improved in the catalytic activity, indicating N doping promotes the electron transfer on the catalyst surface during the reaction [38]. The reducing agent plays an important role in the reduction of aromatic nitro compounds, so the investigation of reducing agent is particularly important. First, the effect of different reducing agents on the yield of the reaction was studied (Table 1, entries 4–7). It can be seen that hydrazine hydrate can reduce the 1-chloro-4-nitrobenzene very well, while hydrogen, sodium borohydride, and ammonium formate are not very effective, and hydrazine hydrate is also a commonly used reducing agent which is green, therefore, hydrazine hydrate was chosen as the optimized reducing agent. Furthermore, as a contrast, available comparative literature data obtained in similar conditions are listed in Table 1, entries 8–10, where it can be seen that these works are excellent.
As is known, the solvent also has an important effect on the reaction, and various solvents were used to track the reaction (Table 2, entries 1–8). Surprisingly, when using water as solvent, the yield of the product almost reached 99%, compared with the others for this reaction. The result was exciting because water as a solvent is green and nontoxic, which is more sustainable as a solvent for this reaction. As far as the effect of temperature on the reaction is concerned, interestingly, the reaction could also occur at room temperature, but in lower yields it can be seen (Table 2, entries 9–10). The reaction time was also one of the most important factors affecting the yield of the amine-based compounds. The results showed that with the decrease of reaction time, the yield of the product decreased, and the yield was only 65% for 2 h under the optimal conditions compared with 4 h (Table 2, entries 1 and 11). It can be concluded that the final optimal reaction conditions were confirmed: catalyst (Fe3O4@NC@Pt), reducting agent (hydrazine hydrate), solvent (H2O), temperature (70 °C), and reaction time (4 h).
To demonstrate the general applicability, the selective hydrogenation of the Fe3O4@NC@Pt catalyst was further evaluated for 23 substituted nitroarenes in Table 3. Each substituted nitroarene could be transformed to corresponding aromatic amine in the presence of Fe3O4@NC@Pt catalyst, which showed excellent activity and selectivity. It also can be seen from Table 3 that the yields were all above 93%. From the experimental results, it can be found that the electronic effects of aryl nitro compounds with different substituents hardly affects the yield of the product. First, it can be found that all the halogenated nitroarenes compound could be favorably reduced to desired haloanilines almost without any by-product, however, 4-iodo nitrobenzene exhibited much lower reaction rates (Table 3, 2a2f), the yield of 4-iodine aniline was lower than the others halogenated nitroarenes, as we know it is a tendency to dehalogenate for iodoaromatic nitro compounds more easily during reduction. Compounds with carbonyl groups were also reduced when the nitro groups was reduced in the process (Table 3, 2g2j). After 8 h of reduction, 4-nitrobenzaldehyde was reduced to 4-amino benzaldehyde (yield 99%), when the aldehyde group was replaced by other reductive groups, such as -COCH3 and -COOH, the yield of the product could also be obtained according to the same strategy. In addition, other nitroarene compound containing sulfo, alkyl, alkoxy, amino, and biphenyl, could be reduced to corresponding anilines in good yields (Table 3, 2k2v). Finally, a heterocyclic aromatic nitro compound was selected as a substrate, which could also be reduced to amino, and the yield was 95% (Table 3, 2w). All 1H NMR spectrum of products (2a2w) (See supporting information Figures S1–S23).
For industrial applications of catalytic systems, the Fe3O4@NC@Pt catalytic hydrogenation strategy was used to obtain the important pharmaceutical intermediates of commercial drug Lorlatini (Scheme 2). Compound 4 had great significance, which can be served as precursors for the synthesis of antitumor drugs Lorlatini. Typically, 4 mmol compounds 3 was input to 6 mL distilled water and then 0.080 g ultrasonically dispersed Fe3O4@NC@Pt catalyst in water (8 mL) was introduced to this solution. Next, 10 mmol N2H4·H2O was added, and the mixture was stirred in 70 °C for 4 h. The reaction system was detected by TLC. Conversion from compound 3 to 4 with excellent results in the presence of Fe3O4@NC@Pt catalyst in water at 70 °C for 4 h. The target product 4 was achieved with high isolated yield (up to 99% yield). After the reaction, Pt in the reaction solvent was determined by ICP-OES, however, no platinum contamination was detected (Pt content < 10 ppm). The 1H NMR spectrum of products 4 (See supporting information Figure S24).
For heterogeneous catalysts, the stability, ease of separation, and reusability are the most important features. First, hot leaching experiment was conducted when reacted for 1 h, the catalyst Fe3O4@NC@Pt was recovered with a magnet, and then the system without a catalyst continued to react at 70 °C for 3 h. 1-chloro-4-nitrobenzene conversion maintained the same after removal of the catalyst. This further confirmed that Pt particles were immobilized onto the catalyst surface and not leached into the reaction solution, the dispersion state photographs of the magnetically separable Fe3O4@NC@Pt (Figure 6a) and magnetic separation of Fe3O4@NC@Pt after reaction (Figure 6b). The recyclability of Fe3O4@NC@Pt was tested for the hydrogenation of 1-chloro-4-nitrobenzene in Figure 6c. In cycle performance test, the yield of 4-chloroaniline was still good for the sixth run. However, the catalytic activity was slightly reduced, and the yield of aniline remained at 92% after ten cycles. All these results suggested that the Fe3O4@NC@Pt catalyst could maintain activity without a significant efficiency reduction.

3. Experimental Materials

The reagent was commercially available Ferric chloride hexahydrate (FeCl3·6H2O), Glucose, Sodium acetateanhydrous, ethylene glycol, Vinylpyrrolidinone polymer (PVP), and ethylenediamine (EDA), which were purchased from Shanghai Aladdin Reagent Co. Ltd (Shanghai, China). Chloroplatinic acid hexahydrate (H2PtCl6·6H2O, ≥37.50%) was purchased from J&K Scientific Ltd (Beijing, China). All reagents are analytically pure and used without further purification.

3.1. Characterization

Fourier-transform infrared (FTIR) spectra were recorded with a Bruker Tensor 27 using KBr pellets(4000–400 cm−1). Transmission electron microscopy (TEM) images were performed on a JEOL 2100 TEM microscope operating at 200 kV. The X-ray diffraction (XRD) ware tested on D/max 2500 PC X-ray diffraction (XRD, Rigaku, Tokyo, Japan) with Cu-Kα radiation (λ= 0.154 nm) operated at 40 kV and 40 mA. The N2 adsorption and desorption isotherms were recorded on a Micromeritics ASAP 2020 sorptometer at −196 °C, and the samples were degassed for 10 h at 180 °C before analysis. The magnetic properties of the catalysts were tested on a vibrating sample magnetometer (VSM) at room temperature from +30 to −30 kOe. 1H and 13C NMR spectra were carried out with a Bruker Advance 400 spectrometer by using TMS as the internal standard and DMSO-d6 or CDCl3 as solvents. The Pt content in the catalyst was carried out with a Perkin-Elmer Optima 2100 DV.

3.2. Preparation of Fe3O4 Microspheres

Fe3O4 microspheres were obtained according to a previous report [40]. In a typical experiment, 3 g FeCl3·6H2O, 4 g NaAc, and 2 g PVP were dispersed in 60 mL ethylene glycol under magnetic stirring. Next, the mixture was transferred to a Teflon-lined autoclave and maintained at 200 °C for 12 h. After that, the desired product was separated using an external magnet and washed with water and ethanol for several times, respectively. Finally, the black Fe3O4 microspheres were dried in a vacuum oven at 60 °C for 24 h.

3.3. Preparation of Fe3O4@C Microspheres

Fe3O4@C was fabricated by a simple carbonization of glucose on Fe3O4 surface under hydrothermal conditions. 200 mg Fe3O4 microspheres were dispersed in 10 mL water of glucose (3.2 g) aqueous solution and ultrasonicated for 0.5 h. After that, they were added into a 100mL Teflon lined autoclave heated at 180 °C for 10 h. After cooling to room temperature, the final Fe3O4@C microspheres were obtained by an external magnet and washed with ethanol, followed by water. Finally, the product with black color was dried in a vacuum oven for 24 h.

3.4. Preparation of Fe3O4@NC Microspheres

Fe3O4@NC was synthesized according to the previous report described elsewhere [41]. 100 mg Fe3O4 microspheres were dispersed into 10 mL water of ethylenediamine (EDA) (0.2 mL) and 1.6 g glucose solution, then ultrasonicated for 0.5 h. Subsequently, the mixture was transferred into a Teflon-lined autoclave and treated at 180 °C for 10 h. After cooling to room temperature, the obtained black product Fe3O4@NC microspheres were washed with ethanol and water, respectively. Lastly, they were dried in a vacuum oven for 24 h.

3.5. Preparation of the Fe3O4@C@Pt and Fe3O4@NC@Pt Catalyst

The Fe3O4@C@Pt and Fe3O4@NC@Pt catalyst were prepared according to the previous report, with some modifications [42]. In a typical procedure, take the Fe3O4@NC preparation process for example, 400 mg Fe3O4@NC were dispersed in 40 mL ethanol by ultrasonic treatment for 0.5 h. Next, 3 mL H2PtCl6·6H2O (35 mg) of ethanol solution were added into Fe3O4@NC suspension solution and continuously ultrasonicated for 1 h, the last, 8 mL sodium borohydride (1 g) of ethanol solution was dropped into the above mixture with vigorous stirring under 60 °C. After 2 h of reduction, the products were separated and washed several times with water. The products were dried in a vacuum oven to obtain Fe3O4@NC@Pt. The same prepared process was undertaken for Fe3O4@C@Pt.

3.6. Procedure for the Selective Hydrogenation of Nitroarenes Reactions

1 mmol nitro compounds was input to 2 mL distilled water, and then 0.020 g ultrasonically dispersed Fe3O4@NC@Pt catalyst in water (2 mL) was introduced to this solution. Next, 4 mmol N2H4·H2O was added, and the mixture was stirred in 70 °C. The reaction system was detected by TLC. After that, the catalyst was removed with external magnetic field and washed several times with ethanol and used after drying in subsequent reactions, then the residual solvent was evaporated under vacuum to obtain the pure amines. The conversions were determined by the gas chromatography (GC) analysis. All the synthesized amines were characterized by comparison of NMR spectral data with the reported values in literatures.

3.7. Procedure for Catalyst Reused

1.0 mmol 1-chloro-4-nitrobenzene, 4.0 mmol N2H4·H2O, and Fe3O4@NC@Pt were mixed in 3 mL H2O. The mixture was stirred at 70 °C. After the reaction, the catalyst was separated by an external magnet and washed with water for three times (2 mL Water/time) and ethanol for three times (2 mL ethanol/time), then dried in vacuum and directly used in the next run.

4. Conclusions

A novel magnetically recyclable Fe3O4@NC@Pt catalyst was prepared. The results show that the catalyst has a good catalytic activity and especially high selectivity for the hydrogenation of nitroarenes with halogen and carbonyl groups under mild reaction conditions. These magnetic catalysts could readily be reused at least ten times without significant decrease of catalytic activity. Moreover, the key pharmaceutical intermediates Lorlatini could be easily synthesized using this catalyst. Therefore, this new catalyst system should be a very useful, sustainable, and environmentally friendly solution in industrial applications.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11101219/s1, Figure S1: 1H NMR spectrum of 2a, recorded in Chloroform-d at 25 °C; Figure S2: 1H NMR spectrum of 2b, recorded in Chloroform-d at 25 °C; Figure S3: 1H NMR spectrum of 2c, recorded in DMSO-d6 at 25 °C; Figure S4: 1H NMR spectrum of 2d, recorded in Chloroform-d at 25 °C; Figure S5: 1H NMR spectrum of 2e, recorded in Chloroform-d at 25 °C; Figure S6: 1H NMR spectrum of 2f, recorded in DMSO-d6 at 25 °C; Figure S7: 1H NMR spectrum of 2g, recorded in Chloroform-d at 25 °C; Figure S8: 1H NMR spectrum of 2h, recorded in Chloroform-d at 25 °C; Figure S9: 1H NMR spectrum of 2i, recorded in Chloroform-d at 25 °C; Figure S10: 1H NMR spectrum of 2j, recorded in Chloroform-d at 25 °C; Figure S11: 1H NMR spectrum of 2k, recorded in Chloroform-d at 25 °C; Figure S12: 1H NMR spectrum of 2l, recorded in Chloroform-d at 25 °C; Figure S13: 1H NMR spectrum of 2m, recorded in Chloroform-d at 25 °C; Figure S14: 1H NMR spectrum of 2n, recorded in Chloroform-d at 25 °C; Figure S15: 1H NMR spectrum of 2o, recorded in Chloroform-d at 25 °C; Figure S16: 1H NMR spectrum of 2p, recorded in Chloroform-d at 25 °C; Figure S17: 1H NMR spectrum of 2q, recorded in Chloroform-d at 25 °C; Figure S18: 1H NMR spectrum of 2r, recorded in DMSO-d6 at 25 °C; Figure S19: 1H NMR spectrum of 2s, recorded in Chloroform-d at 25 °C; Figure S20: 1H NMR spectrum of 2t, recorded in DMSO-d6 at 25 °C; Figure S21: 1H NMR spectrum of 2u, recorded in Chloroform-d at 25 °C; Figure S22: 1H NMR spectrum of 2v, recorded in Chloroform-d at 25 °C; Figure S23: 1H NMR spectrum of 2w, recorded in Chloroform-d at 25 °C; Figure S24: 1H NMR spectrum of 4, recorded in DMSO-d6 at 25 °C.

Author Contributions

The experimental work was conceived and designed by J.Q. and Q.Z.; T.W., K.Z., and E.Z. performed the experiments; J.Q., C.S. and A.J. analyzed the data; J.Q. and Q.Z. drafted the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Zhejiang Shuren University Basic Scientific Research Special Funds, grant number 2020XZ011 and 2021 Scientific and Technological Innovation Activity Plan for College Students in Zhejiang Province (New Talent Plan), grant number 2021R421009.

Data Availability Statement

There is no additional data available.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jagadeesh, R.V.; Surkus, A.E.; Junge, H.; Pohl, M.M.; Radnik, J.; Rabeah, J.; Huan, H.M.; Schunemann, V.; Brückner, A.; Beller, M. Nanoscale Fe2O3-based catalysts for selective hydrogenation of nitroarenes to anilines. Science 2013, 342, 1073–1076. [Google Scholar] [CrossRef]
  2. Pang, H.B.; Gallou, F.; Sohn, H.; Camacho-Bunquin, J.; Delferro, M.; Lipshutz, B.H. Synergistic effects in Fe nanoparticles doped with ppm levels of (Pd+Ni). A new catalyst for sustainable nitro group reductions. Green Chem. 2018, 20, 130–135. [Google Scholar] [CrossRef]
  3. MacDonald, L.; Rausch, B.; Symes, M.D.; Cronin, L. Selective hydrogenation of nitroarenes using an electrogenerated polyoxometalate redox mediator. Chem. Commun. 2018, 54, 1093–1096. [Google Scholar] [CrossRef] [Green Version]
  4. Tian, M.; Cui, X.L.; Yuan, M.; Yang, J.; Ma, J.T.; Dong, Z.P. Efficient chemoselective hydrogenation of halogenated nitrobenzenes over an easily prepared γ-Fe2O3-modified mesoporous carbon catalyst. Green Chem. 2017, 19, 1548–1554. [Google Scholar] [CrossRef]
  5. Lian, C.; Liu, H.Q.; Xiao, C.; Yang, W.; Zhang, K.; Liu, Y.; Wang, Y. Solvent-free selective hydrogenation of chloronitrobenzene to chloroaniline over a robust Pt/Fe3O4 catalyst. Chem. Commun. 2012, 48, 3124–3126. [Google Scholar] [CrossRef]
  6. Wei, H.S.; Liu, X.Y.; Wang, A.Q.; Zhang, L.L.; Qiao, B.T.; Yang, X.F.; Huang, Y.Q.; Miao, S.; Liu, J.Y.; Zhang, T. FeOx-supported platinum single-atom and pseudo-single-atom catalysts for chemoselective hydrogenation of functionalized nitroarenes. Nat. Commun. 2014, 5, 5634–5641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Kelly, S.M.; Lipshutz, B.H. Chemoselective reductions of nitroaromatics in water at room temperature. Org. Lett. 2014, 16, 98–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Junge, K.; Wendt, B.; Shaikh, N.; Beller, M. Iron-catalyzed selective reduction of nitroarenes to anilines using organosilanes. Chem. Commun. 2010, 46, 1769–1771. [Google Scholar] [CrossRef]
  9. Liu, L.C.; Corma, A. Metal catalysts for heterogeneous catalysis: From single atoms to nanoclusters and nanoparticles. Chem. Rev. 2018, 118, 4981–5079. [Google Scholar] [CrossRef] [Green Version]
  10. Li, Z.; Ji, S.F.; Liu, Y.W.; Cao, X.; Tian, S.B.; Chen, Y.J.; Niu, Z.Q.; Li, Y.D. Well-defined materials for heterogeneous catalysis: From nanoparticles to isolated single-atom sites. Chem. Rev. 2020, 120, 623–682. [Google Scholar] [CrossRef] [PubMed]
  11. Robert, A.W.; Johnstone, A.H.; Wilby, A.H.; Entwistle, I.D. Heterogeneous catalytic transfer hydrogenation and its relation to other methods for reduction of organic compounds. Chem. Rev. 1985, 85, 129–170. [Google Scholar]
  12. Sharma, R.K.; Monga, Y.; Puri, A. Magnetically separable silica@Fe3O4 core-shell supported nano-structured copper(II) composites as a versatile catalyst for the reduction of nitroarenes in aqueous medium at room temperature. J. Mol. Catal. A-Chem. 2014, 393, 84–95. [Google Scholar] [CrossRef]
  13. Zhang, B.; Guo, X.W.; Liang, H.J.; Ge, H.B.; Gu, X.M.; Chen, S.; Yang, H.M.; Qin, Y. Tailoring Pt-Fe2O3 interfaces for selective reductive coupling reaction to synthesize imine. ACS Catal. 2016, 6, 6560–6566. [Google Scholar] [CrossRef]
  14. Byun, S.; Song, Y.M.; Kim, B.M. Heterogenized bimetallic Pd–Pt–Fe3O4 nanoflakes as extremely robust, magnetically recyclable catalysts for chemoselective nitroarene reduction. ACS Appl. Mater. Interfaces 2016, 8, 14637–14647. [Google Scholar] [CrossRef] [PubMed]
  15. Liu, Y.; Chung, J.; Jang, Y.; Mao, S.; Kim, B.M.; Wang, Y.; Guo, X. Magnetically recoverable nanoflake-shaped iron oxide/Pt heterogeneous catalysts and their excellent catalytic performance in the hydrogenation reaction. ACS Appl. Mater. Interfaces 2014, 6, 1887–1892. [Google Scholar] [CrossRef]
  16. Shokouhimehr, M.; Kim, T.; Jun, S.W.; Shin, K.; Jang, Y.J.; Kim, B.H.; Kim, J.; Hyeon, T. Magnetically separable carbon nanocomposite catalysts for efficient nitroarene reduction and Suzuki reactions. Appl. Catal. A-Gen. 2014, 476, 133–139. [Google Scholar] [CrossRef]
  17. Li, Y.; Li, N.; Jiang, W.; Ma, G.Y.; Zangene, M.M. In situ decorated Au NPs on pectin-modified Fe3O4 NPs as a novel magnetic nanocomposite (Fe3O4/Pectin/Au) for catalytic reduction of nitroarenes and investigation of its anti-human lung cancer activitiesInternational. Int. J. Biol. Macromol. 2020, 163, 2162–2171. [Google Scholar] [CrossRef] [PubMed]
  18. Rahimi, J.; Taheri-Ledari, R.; Niksefat, M.; Maleki, A. Enhanced reduction of nitrobenzene derivatives: Effective strategy executed by Fe3O4/PVA-10%Ag as a versatile hybrid nanocatalyst. Catal. Commun. 2020, 134, 105850. [Google Scholar] [CrossRef]
  19. Dayan, S.; Arslan, F.; Ozpozan, N.K. Ru(II) impregnated Al2O3, Fe3O4, SiO2 and N-coordinate ruthenium(II) arene complexes: Multifunctional catalysts in the hydrogenation of nitroarenes and the transfer hydrogenation of aryl ketones. Appl. Catal. B Environ. 2015, 164, 305–315. [Google Scholar] [CrossRef]
  20. Hu, Z.N.; Zhou, J.J.; Ai, Y.J.; Liu, L.; Qi, L.; Jiang, R.H.; Bao, H.J.; Wang, J.T.; Hu, J.S.; Sun, H.B.; et al. Two dimensional Rh/Fe3O4/g-C3N4-N enabled hydrazine mediated catalytic transfer hydrogenation of nitroaromatics: A predictable catalyst model with adjoining Rh. J. Catal. 2018, 368, 20–30. [Google Scholar] [CrossRef]
  21. Sheng, Y.; Wang, X.G.; Xing, Z.K.; Chen, X.B.; Zou, X.J.; Lu, X.G. Highly active and chemoselective reduction of halogenated nitroarenes catalysed by ordered mesoporous carbon supported platinum nanoparticles. ACS Sustain. Chem. Eng. 2019, 7, 8908–8916. [Google Scholar] [CrossRef]
  22. Antony, R.; Marimuthu, R.; Murugavel, R. Bimetallic nanoparticles anchored on core-shell support as an easily recoverable and reusable catalytic system for efficient nitroarene reduction. ACS Omega 2019, 4, 9241–9250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Choi, Y.; Bae, H.S.; Seo, E.Y.; Jang, S.W.; Park, K.H.; Kim, B.S. Hybrid gold nanoparticle-reduced graphene oxide nanosheets as active catalysts for highly efficient reduction of nitroarenes. J. Mater. Chem. 2011, 21, 15431. [Google Scholar] [CrossRef]
  24. Ombaka, L.M.; Ndungu, P.G.; Kibet, J.; Nyamori, V.O. The effect of pyridinic- and pyrrolic-nitrogen in nitrogen-doped carbon nanotubes used as support for Pd-catalyzed nitroarene reduction: An experimental and theoretical study. J. Mater. Sci. 2017, 52, 10751–10765. [Google Scholar] [CrossRef]
  25. Nie, R.F.; Wang, J.H.; Wang, L.N.; Qin, Y.; Chen, P.; Hou, Z.Y. Platinum supported on reduced graphene oxide as a catalyst for hydrogenation of nitroarenes. Carbon 2012, 50, 586–596. [Google Scholar] [CrossRef]
  26. Shen, C.; Xu, J.; Yu, W.; Zhang, P. ChemInform abstract: A highly active and easily recoverable Chitosan@Copper catalyst for the C-S coupling and its application in the synthesis of zolimidine. Green Chem. 2014, 16, 3007–3012. [Google Scholar] [CrossRef]
  27. Shen, C.; Shen, H.; Yang, M.; Xia, C.; Zhang, P. Novel D-glucosamine-derived pyridyl-triazole@palladium catalyst for solvent-free Mizoroki-Heck reactions and its application in the synthesis of Axitinib. Green Chem. 2014, 17, 225–230. [Google Scholar] [CrossRef]
  28. Shen, H.; Shen, C.; Chen, C.; Wang, A.; Zhang, P. Novel glycosyl pyridyl-triazole@palladium nanoparticles: Efficient and recoverable catalysts for C–C cross-coupling reactions. Catal. Sci. Technol. 2015, 5, 2065–2071. [Google Scholar] [CrossRef]
  29. Shen, C.; Zhang, P.; Sun, Q.; Bai, S.; Andy Hor, T.A.; Liu, X. Recent advances in C-S bond formation via C-H bond functionalization and decarboxylation. Chem. Soc. Rev. 2015, 46, 291–314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Zhao, L.W.; Zheng, K.; Tong, J.Y.; Jin, J.Z.; Shen, C. Novel biomass-derived Fe3O4@Pd NPs as efcient and sustainablenanocatalyst for nitroarene reduction in aqueous media. Catal. Lett. 2019, 149, 2607–2613. [Google Scholar] [CrossRef]
  31. Zhu, M.Y.; Diao, G.W. Magnetically recyclable Pd nanoparticles immobilized on magnetic Fe3O4@C nanocomposites: Preparation, characterization, and their catalytic activity toward Suzuki and heck coupling reactions. J. Phys. Chem. C 2011, 115, 24743–24749. [Google Scholar] [CrossRef]
  32. Li, Y.; Wang, Z.; Liu, R.J. Superparamagnetic α-Fe2O3/Fe3O4 heterogeneous nanoparticles with enhanced biocompatibility. Nanomaterials 2021, 11, 834. [Google Scholar] [CrossRef]
  33. Serga, V.; Burve, R.; Maiorov, M.; Krumina, A.; Skaudžius, R.; Zarkov, A.; Kareiva, A.; Popov, A.I. Impact of Gadolinium on the structure and magnetic properties of nanocrystalline powders of iron oxides produced by the extraction-pyrolytic method. Materials 2020, 13, 4147. [Google Scholar] [CrossRef]
  34. Alzahrani, F.M.; Alsaiari, N.S.; Katubi, K.M.; Amari, A.; Elkhaleefa, A.M.; Rebah, F.B.; Tahoon, M.A. Magnetic nitrogen-doped porous carbon nanocomposite for Pb(II) adsorption from aqueous solution. Molecules 2021, 26, 4809. [Google Scholar] [CrossRef]
  35. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of gases in multimolecular layers. J. Am. Chem. Soc. 1938, 60, 309–319. [Google Scholar] [CrossRef]
  36. Barrett, E.P.; Joyner, L.G.; Halend, P.P. The determination of pore volume and area distributions in porous substances. I. Computations from nitrogen isotherms. J. Am. Chem. Soc. 1951, 73, 373–380. [Google Scholar] [CrossRef]
  37. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reino, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef] [Green Version]
  38. Chen, P.; Chew, L.M.; Kostka, A.; Muhlera, M.; Xia, W. The structural and electronic promoting effect of nitrogendoped carbon nanotubes on supported Pd nanoparticles for selective olefin hydrogenation. Catal. Sci. Technol. 2013, 3, 1964–1971. [Google Scholar] [CrossRef]
  39. Xu, S.L.; Shen, S.C.; Zhao, S.; Ding, Y.W.; Chu, S.Q.; Chen, P.; Lin, Y.; Liang, H.W. Synthesis of carbon-supported sub-2 nanometer bimetallic catalysts by strong metal–sulfur interaction. Chem. Sci. 2020, 11, 7933–7939. [Google Scholar] [CrossRef]
  40. Zheng, K.; Shen, C.; Qiao, J.; Tong, J.Y.; Jin, J.Z.; Zhang, P.F. Novel magnetically-recyclable, nitrogen-doped Fe3O4@Pd NPs for Suzuki–Miyaura coupling and their application in the synthesis of crizotinib. Catalysts 2018, 8, 443. [Google Scholar] [CrossRef] [Green Version]
  41. Liu, Z.Y.; Zhang, C.L.; Luo, L.; Chang, Z.J.; Sun, X.M. One-pot synthesis and catalyst support application of mesoporous N-doped carbonaceous materials. J. Mater. Chem. 2012, 22, 12149–12154. [Google Scholar] [CrossRef]
  42. Kumar, B.S.; Amali, A.J.; Pitchumani, K. Mesoporous microcapsules through d-Glucose promoted hydrothermal self-assembly of colloidal silica: Reusable catalytic containers for palladium catalyzed hydrogenation reactions. Appl. Mater. Interfaces 2015, 7, 22907–22917. [Google Scholar]
Scheme 1. Synthesis of Fe3O4@NC@Pt catalyst.
Scheme 1. Synthesis of Fe3O4@NC@Pt catalyst.
Catalysts 11 01219 sch001
Figure 1. FT-IR spectra of (a) Fe3O4; (b) Fe3O4@NC; (c) Fe3O4@NC@Pt.
Figure 1. FT-IR spectra of (a) Fe3O4; (b) Fe3O4@NC; (c) Fe3O4@NC@Pt.
Catalysts 11 01219 g001
Figure 2. XRD patterns of (a) Fe3O4; (b) Fe3O4@NC; (c) Fe3O4@NC@Pt.
Figure 2. XRD patterns of (a) Fe3O4; (b) Fe3O4@NC; (c) Fe3O4@NC@Pt.
Catalysts 11 01219 g002
Figure 3. N2 adsorption/desorption isotherms of Fe3O4@NC@Pt and BJH pore size distribution (inset).
Figure 3. N2 adsorption/desorption isotherms of Fe3O4@NC@Pt and BJH pore size distribution (inset).
Catalysts 11 01219 g003
Figure 4. Vibrating Sample Magnetometer (VSM) data of Fe3O4 @NC@Pt.
Figure 4. Vibrating Sample Magnetometer (VSM) data of Fe3O4 @NC@Pt.
Catalysts 11 01219 g004
Figure 5. TEM image of Fe3O4@NC@Pt (a) and EDS elemental mappings of C, N, O, Fe, and Pt for Fe3O4@NC@Pt (b), EDS map sum spectrum pattern of Fe3O4@NC@Pt (c).
Figure 5. TEM image of Fe3O4@NC@Pt (a) and EDS elemental mappings of C, N, O, Fe, and Pt for Fe3O4@NC@Pt (b), EDS map sum spectrum pattern of Fe3O4@NC@Pt (c).
Catalysts 11 01219 g005
Scheme 2. Application of the catalyst in the synthesis of intermediate of commercial drug Lorlatini.
Scheme 2. Application of the catalyst in the synthesis of intermediate of commercial drug Lorlatini.
Catalysts 11 01219 sch002
Figure 6. Photographs of the magnetically separable Fe3O4@NC@Pt (a) dispersion state of Fe3O4@NC@Pt; (b) magnetic separation of Fe3O4 @NC@Pt after reaction; (c) recyclability of the Fe3O4@NC@Pt catalyst.
Figure 6. Photographs of the magnetically separable Fe3O4@NC@Pt (a) dispersion state of Fe3O4@NC@Pt; (b) magnetic separation of Fe3O4 @NC@Pt after reaction; (c) recyclability of the Fe3O4@NC@Pt catalyst.
Catalysts 11 01219 g006
Table 1. Optimization of catalyst and reducing agent a.
Table 1. Optimization of catalyst and reducing agent a.
Catalysts 11 01219 i001
EntryCatalystReducing AgentTemp (°C)Time(h)Yield b (%)Ref.
1-hydrazine hydrate704-this work
2Fe3O4@NChydrazine hydrate704-this work
3Fe3O4@C@Pthydrazine hydrate70470this work
4Fe3O4@NC@Pthydrazine hydrate70495this work
5Fe3O4@NC@PtHydrogen70485this work
6Fe3O4@NC@PtSodium borohydride70470this work
7Fe3O4@NC@PtAmmonium formate70430this work
8Pt-Co/S-C2 bar H240299[39]
9Pd–Pt–Fe3O4NH3BH3RT0.8392[14]
10Pt/RGO-EG1 bar H240296[25]
a The reaction conditions: 1-Chloro-4-nitrobenzene 1a (1 mmol), 4 mmol reducing agent, 0.5 mol% catalyst, and Toluene (3 mL), under air. b Isolated yield.
Table 2. Optimization of reaction conditions a.
Table 2. Optimization of reaction conditions a.
Catalysts 11 01219 i002
EntryCatalystSolventTemp(°C)Time(h)Yield b (%)
1Fe3O4@NC@PtH2O70499
2Fe3O4@NC@PtDMSO70430
3Fe3O4@NC@PtToluene70495
4Fe3O4@NC@PtEtOH70440
5Fe3O4@NC@PtIsopropanol70450
6Fe3O4@NC@PtAcOEt70410
7Fe3O4@NC@PtTHF70420
8Fe3O4@NC@PtHexane70450
9Fe3O4@NC@PtH2O50470
10Fe3O4@NC@PtH2Ort440
11Fe3O4@NC@PtH2O70265
12Pt/CMK-3–HQEtOH80199 c
a the reaction conditions: 1-chloro-4-nitrobenzene 1a (1 mmol), 4 mmol N2H4·H2O, 0.5 mol% catalyst and solvent (3 mL) under air, all reaction temperature was 70 °C and the reaction time was 4 h, unless otherwise specified. b Isolated yield. c the work was done by the Xionggang Lu research group [21], as a comparison.
Table 3. Fe3O4@NC@Pt catalyzed hydrogenation of different nitroarenes a.
Table 3. Fe3O4@NC@Pt catalyzed hydrogenation of different nitroarenes a.
Catalysts 11 01219 i003
Catalysts 11 01219 i004 Catalysts 11 01219 i005 Catalysts 11 01219 i006 Catalysts 11 01219 i007 Catalysts 11 01219 i008
Catalysts 11 01219 i009 Catalysts 11 01219 i010 Catalysts 11 01219 i011 Catalysts 11 01219 i012 Catalysts 11 01219 i013
Catalysts 11 01219 i014 Catalysts 11 01219 i015 Catalysts 11 01219 i016 Catalysts 11 01219 i017 Catalysts 11 01219 i018
Catalysts 11 01219 i019 Catalysts 11 01219 i020 Catalysts 11 01219 i021 Catalysts 11 01219 i022 Catalysts 11 01219 i023
Catalysts 11 01219 i024 Catalysts 11 01219 i025 Catalysts 11 01219 i026
a The reaction conditions: nitroarenes 1a (1 mmol), N2H4·H2O (4 mmol), 0.5 mol% catalysts, 4 h and H2O (3 mL) under air, isolated yield; b 8 h; c 16 h.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Qiao, J.; Wang, T.; Zheng, K.; Zhou, E.; Shen, C.; Jia, A.; Zhang, Q. Magnetically Reusable Fe3O4@NC@Pt Catalyst for Selective Reduction of Nitroarenes. Catalysts 2021, 11, 1219. https://doi.org/10.3390/catal11101219

AMA Style

Qiao J, Wang T, Zheng K, Zhou E, Shen C, Jia A, Zhang Q. Magnetically Reusable Fe3O4@NC@Pt Catalyst for Selective Reduction of Nitroarenes. Catalysts. 2021; 11(10):1219. https://doi.org/10.3390/catal11101219

Chicago/Turabian Style

Qiao, Jun, Tian Wang, Kai Zheng, Enmu Zhou, Chao Shen, Aiquan Jia, and Qianfeng Zhang. 2021. "Magnetically Reusable Fe3O4@NC@Pt Catalyst for Selective Reduction of Nitroarenes" Catalysts 11, no. 10: 1219. https://doi.org/10.3390/catal11101219

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