Skip to content
BY 4.0 license Open Access Published by De Gruyter December 23, 2020

Preparation, characterization, and catalytic performance of Pd–Ni/AC bimetallic nano-catalysts

  • Yu Zhang , Yalong Liao EMAIL logo , Gongchu Shi , Wei Wang and Bowen Su

Abstract

Palladium–nickel (Pd–Ni) bimetallic nano-catalysts supported on activated carbon (Pd–Ni/AC) have been successfully prepared by impregnation method enhanced with ultrasonic. The prepared Pd–Ni/AC catalysts were used for the catalytic hydrodechlorination reaction of bleached shellac and characterized by Fourier-transform infrared spectroscopy, nitrogen adsorption, X-ray diffraction, X-ray photoelectron spectroscopy, field-emission scanning electron microscope, and high-resolution transmission electron microscopy. The results show that Pd–Ni bimetallic structures in catalytic particles with the diameter of 4 and 14 nm were distributed uniformly on AC support, and the lattice fringe spacing in catalytic particles was measured as 0.213 nm which is lying in the region between monometallic Pd (111; 0.225 nm) and Ni (111; 0.203 nm), and that Pd1–Ni1/AC catalyst exhibits the best catalytic hydrodechlorination performance with the dechlorination efficiency of 92.58 wt% while it is used for the hydrodechlorination of bleached shellac, and the optimum catalytic performance is related to the synergistic electronic effect and bimetallic structure of the Pd1–Ni1/AC sample.

1 Introduction

Bleached shellac, a natural product refined from seedlac by bleaching with sodium hypochlorite, has excellent chemical properties, making it widely applied in many industries [1,2,3,4,5]. However, in the bleaching process a part of chlorine from the bleaching agent (sodium hypochlorite), which is widely used, were combined with the double bond of terpenic acid of shellac by addition reaction, which deteriorates the storage performance and then affects the transportation, sales, and application of the product. In order to eliminate the effects of combined chlorine, it has been reported to use complex bleaches or chlorine-free bleaches in the bleaching process; but compared with sodium hypochlorite, these bleaching agents not only are at a disadvantage in terms of bleaching ability and price but also introduce other harmful substances.

The aspect of eliminating the combined chlorine in the bleached shellac by bleaching with sodium hypochlorite was investigated by Liao and Chai [6] based on the catalytic hydrodechlorination of bleached shellac with Ni–Fe bimetallic catalyst. The results showed that the prepared bimetallic catalyst can remove the combined chlorine produced in the bleaching process with a dechlorination efficiency of 82.7 wt% under optimum conditions. In order to reduce the chlorine content of the bleached shellac as much as possible to meet the increasing rigorous demand of the consumers for the product with lower chlorine content, it is necessary to explore catalysts with better hydrodechlorination performance. In recent years, the bimetallic catalyst, of which the unique synergistic effect of two different cations in the crystal lattice is believed to provide multiple oxidation states, enhances the electronic conductivity, and then greatly promotes catalytic performance [7,8,9], is favored by researchers. Especially, supported Pd-based catalysts have been widely used in liquid-phase hydrodechlorination of organochlorines due to their strong ability to adsorb and dissociate H2 [10,11]. However, expensive cost for manufacturing supported Pd-based catalysts has become a key factor limiting its development. Alloying of Pd with a second metallic component to form Pd-based bimetallic catalysts can improve the catalytic performance (activity, durability, and selectivity) while reducing the cost of the catalyst [12,13]. The valence state of the two components of the catalyst and the topological distribution of the nanoparticles are major factors affecting the catalytic activity [14]. Among Pd-based bimetallic systems, palladium–nickel (Pd–Ni) is one of the most widely investigated due to their excellent performance compared to the Pd monometallic catalysts in many industrially important reactions such as hydrodechlorination [15,16], hydrodeoxygenation [17,18,19], hydrogenation [20,21], oxygen reduction [22,23], and hydrogen evolution reactions [24,25]. However, in the field of hydrodechlorination, most of the existing research are on the hydrodechlorination of small molecular chlorine-containing compounds, and few research on hydrodechlorination of chlorine-containing organic compounds, especially chlorine-containing macromolecular natural products. It is also necessary and interesting to synthesize Pd–Ni bimetallic catalysts for the study of hydrodechlorination of bleached shellac.

Pd–Ni bimetallic catalysts supported on AC and/or carbon nanofibers have also been reported to have outstanding properties in the hydrodeoxygenation of aldehydes, cellobiose conversion, and the methanol oxidation reaction, etc. [26,27,28,29]. However, to the best of our knowledge, nano Pd–Ni/AC bimetallic catalyst has not been reported in literature on research for selective hydrodechlorination of bleached shellac. Herein, a series of nano Pd–Ni/AC bimetallic catalysts with different molar ratio of Pd to Ni, and the structure and morphology of the catalysts were characterized and their catalytic performance for catalytic hydrodechlorination of bleached shellac were investigated in the present article. And the reason that activated carbon (AC) was selected as the support because of its high specific surface area, low price, and excellent absorption capability. In addition, carbon can be produced directly from biomass resources, making it the most “sustainable” support material for supported catalysts [12,19].

2 Experimental section

2.1 Materials and chemicals

AC (16 mesh) was obtained from Yaqi Environmental Protection Technology Co., Ltd. (Jiangsu, China). Palladium dichloride (palladium content ≥59.5 wt%) was purchased from Kaida Chemical Co., Ltd. (Shanxi, China). Nickel nitrate hexahydrate (nickel nitrate hexahydrate content ≥98.0 wt%) was purchased from Fengchuan Chemical Reagent Technology Co., Ltd. (Tianjin, China). Polyethylene glycol 6000 (PEG-6000) was purchased from Kemiou Chemical Reagent Co., Ltd. (Tianjin, China). Bleached shellac was purchased from Anning DECCO Fine Chemical Co., Ltd. (Kunming, China). In addition, all other reagents used in the present study were of analytical grade. Double-distilled water was used to carry out the experiment.

2.2 Preparation of catalysts

The commercial AC obtained was pretreated in 20 wt% nitric acid for 30 min at 40°C with magnetically stirring at 300 rpm to remove any residual metal particles thoroughly to gain better absorptivity for metallic ions, followed by filtering out the solid sample and washing it with deionized water to a final pH of 7.0. The products from the pretreated process was purified by sonicating at 720 W with ultrasonic cell grinder in deionized water for 10 min to reduce ash content inside the sample and remove surface unstable oxygen-containing groups. Then the modified AC was obtained after the purified product was filtered and dried at 60°C overnight. Bimetallic Pdx–Niy/AC nano-catalysts (Pd:Ni = x:y, molar radio) were synthesized by ultrasonic-enhanced impregnation method, and the mass fraction of the loaded palladium in catalyst was 1.0 wt%. First, after the calculated amounts of PdCl2 and Ni(NO3)2·6H2O according to the molar radio of Pd and Ni were dissolved into 0.06 mol/L hydrochloric acid under the stirring speed of 300 rpm with magnetic stirrer, 5 g of the modified AC was mixed into the solution, and then 1.5 g of polyethylene glycol 6000 (PEG-6000) was added into the mixture. Second the mixtures got were dispersed for 30 min with ultrasonic cell crushing apparatus (UH-1200E; Tianjin Auto Science Instrument Co., Ltd., China), followed by impregnated for 30 min at ambient temperature, and then dried in oven at 80°C overnight for getting precursors. Finally, the precursors were reduced by H2/N2 mixture gas (45 mL/min) at 300°C (the ramping rate was 10°C/min) for 2 h to obtain the catalysts. The principle flow chart for preparation of the catalysts is shown in Figure 1.

Figure 1 The principle flowchart for the preparation of catalysts.
Figure 1

The principle flowchart for the preparation of catalysts.

2.3 Hydrodechlorination of bleached shellac with catalysts

In order to evaluate the catalytic performance of the catalysts with different Pd–Ni ratio, catalytic hydrodechlorination experiments of bleached shellac were performed. The results of these experiments were evaluated in terms of dechlorination efficiency. First, the bleached shellac (100 g) was dissolved in a sodium carbonate solution (700 mL, 10 wt%) at room temperature, heated to 80°C with a magnetic stirring speed of 400 rpm, 1.8 g of catalyst was added, and 0.1 L/min H2 was introduced for the removal of the chlorine combined in the bleached shellac. After allowing it to react for 90 min, the reactant was cooled to room temperature, and the catalyst used in the reaction was separated by filtration and dried in oven at 65°C overnight. Then a certain amount of diluted sulfuric acid solution was added to the filtrate to get the dechlorination products of shellac. Finally, according to the Chinese national standard [30], the chlorine content of the obtained shellac is measured, and the dechlorination efficiency is calculated by Eq. 1:

(1)Da=XaX×100%

where Da is dechlorination efficiency, Xa represents the chlorine content of the obtained shellac, and X is the chlorine content of the bleached shellac.

2.4 Characterization of catalysts

The phase and crystallinity of prepared catalysts were characterized by powder X-ray diffraction (XRD) patterns with a PANalytical Xpert3 powder diffractometer using Cu Kα radiation (λ = 1.54056 Å) operating at 40 kV and 40 mA in the 2θ range of 5–90°.

The X-ray photoelectron spectroscopy (XPS) measurement was performed on a photoelectron spectrometer (PHI5000 Versa probe-II; ULVAC-PHI, Japan) with Al-Kα radiation source operating at 15 kV and 50 W to characterize the binding form of Pd and Ni in the catalysts. High-resolution scans of elemental lines were recorded at 46.95 eV pass energy. All binding energies were referenced to the C 1s peak at 284.8 eV to compensate for the effect of surface charging. The processing and curve fitting of the high-resolution spectra was performed using MultiPak software.

The specific surface area of the prepared catalyst was determined with Brunauer–Emmett–Teller (BET) method, and pore diameter and pore volume were determined according to Barrett–Joyner–Halenda (BJH) method by the Quantachrome nitrogen adsorption instrument (Quadrasorb-Evo, USA). Structure and surface morphology of bimetallic Pd1–Ni1/AC nano-catalysts was observed by Fourier-transform infrared spectroscopy (FTIR; Bruker Tensor 27, Germany) and field-emission scanning electron microscope (FESEM) with an FEI Nova NanoSEM 450 electron microscope with an accelerating voltage of 10 kV and a beam current of 10 mA. High-resolution transmission electron microscopy (HRTEM) was used to measure the particle size and lattice fringe spacing of catalyst using JEOL JEM-2100 electron microscope with an accelerating voltage of 200 kV. Sample for the analysis was ultrasonically dispersed in anhydrous ethanol and then 5 μL of the dispersed sample solution was pipetted on a Duplex Copper Ring for HRTEM measurement. The particle size and lattice fringe spacing were obtained by ImageJ software.

3 Results and discussion

3.1 Catalytic performance of catalysts

As we can see from Figure 2 that all catalysts exhibit different catalytic hydrodechlorination activities; and with an increase in nickel content, the catalytic performance of bimetallic catalyst gradually improves, but the addition of excessive nickel will lead to a sharp decrease. Apparently the Pd1–Ni1/AC catalyst displays the best catalytic activities with the dechlorination efficiency of 92.58 wt%; and it is used for hydrodechlorination of bleached shellac, which is consistent with the study result of gas-phase hydrogenation of p-chloronitrobenzene over Pd–Ni/Al2O3 reported by Cárdenas-Lizana et al. [31]. The reason for this was that catalytic reaction performance was sensitive to the structure and the activated component’s ratio of the catalysts, and that Pd-Ni = 1:1 delivered the highest reaction yield, attributing catalytic response to bimetallic particle formation. Additionally, the dechlorination efficiency of the Pd1–Ni1/AC catalyst prepared in the present work is better than that of Pd1–Ni1/γ-Al2O3 reported in literature [32], indicating that the support of catalysts plays a vital influence on its catalytic performance.

Figure 2 Dechlorination efficiency of the synthesized Ni/AC, Pd1-Ni2/AC, Pd1–Ni1/AC, Pd2-Ni1/AC, and Pd/AC samples.
Figure 2

Dechlorination efficiency of the synthesized Ni/AC, Pd1-Ni2/AC, Pd1–Ni1/AC, Pd2-Ni1/AC, and Pd/AC samples.

The stability of all the catalysts Pd1-Ni2/AC, Pd1–Ni1/AC, and Pd2-Ni1/AC with good performance (as shown in Figure 2) of hydrodechlorination of bleached shellac was evaluated by continuously running tests for five times, where the catalyst was used directly for the next reaction without any treatment. The results obtained are shown in Figure 3. As observed, the activity of all catalysts decreases gradually after continuous running, but the decreasing rates are obviously different, and the stability of bimetallic catalyst is obviously better than that of monometallic catalyst. The better stability may be due to the stronger interaction between the bimetal and the carbon support, which is effective in preventing exfoliation and agglomeration of the active Pd component. It is worth noting that the stability of Pd1–Ni1/AC catalyst is the best, wherein the dechlorination efficiency can reach 90.03 wt% in the fifth recycling use. However, the stability of Pd1-Ni2/AC catalyst is not as good as that of Pd1–Ni1/AC catalyst, indicating that the additional amount of nickel will affect the structure of the catalyst, thus affecting the durability of the catalyst, and excessive nickel is not conducive to the combination of bimetallic and support. Babu et al. [16] reported that the high activity and stability of optimal catalyst for hydrodechlorination of chlorobenzene was due to the formation of very active Pd–Ni interfaces embedded with Pdδ+ species. In order to explore the mechanism of the highest catalytic performance of Pd1–Ni1/AC catalyst for bleached shellac, a series of characterization methods are required to further investigate the structure and surface morphology of Pd1–Ni1/AC catalyst.

Figure 3 Dechlorination efficiency of the synthesized Pd1–Ni1/AC, Pd2-Ni1/AC, and Pd/AC samples measured in runs 1–5.
Figure 3

Dechlorination efficiency of the synthesized Pd1–Ni1/AC, Pd2-Ni1/AC, and Pd/AC samples measured in runs 1–5.

3.2 Structural features of the catalysts by nitrogen adsorption

The N2 adsorption analysis showed that the specific surface area (BET) of the support (modified AC) and the catalyst (Pd1–Ni1/AC) was 681.226 and 642.86 m2/g, respectively; the pore diameter of the support and the catalyst was 3.598 and 3.600 nm, respectively; and the pore volume of the support and the catalyst was 0.055 and 0.057 mL/g, respectively, indicating the loading of Pd and Ni bimetal on the modified AC enhanced by ultrasonic impregnation method affects little the specific surface area, pore diameter, and pore volume of the support. The nitrogen absorption–desorption equilibrium isotherm (Figure 4) is in accordance with the standard classification of the International Union of Pure and Applied Chemistry, which conforms to the class IV adsorption–desorption isotherm, indicating the presence of slit-like pores not only in the support but in the catalyst as well, which is consistent with the specification of active carbon.

Figure 4 N2 adsorption–desorption isotherm curve of the support (a) and the catalyst (b).
Figure 4

N2 adsorption–desorption isotherm curve of the support (a) and the catalyst (b).

3.3 Structural features of the catalysts by FTIR

Adsorption properties of AC is strongly dependent on its surface chemical structure. However, after its modification by nitric acid, the AC surface is oxidizes to carbonyl, which further oxidizes to other groups such as carboxyl, phenolic hydroxyl, lactone group, etc. [33,34]. The IR spectra of AC modified by nitric acid and AC supported by Pd–Ni are shown in Figure 5. In Figure 5, the peak at 3412.526 cm−1 is classified as the stretching vibration absorption peak of –OH. The peak at 1621.057 cm−1 is the stretching vibration absorption peak of C═C in AC skeleton. The peaks at 1390.5 or 1401.95 cm−1 are the O–H bending vibration peak in the carboxyl group. The spectral peak around 1092.638 cm−1 was attributed to the asymmetric stretching vibration of C–O–C, and the spectral peak at 796.208 cm−1 was attributed to the stretching vibration absorption peak of cis-epoxy ether bond. It can be seen from Figure 5 that the O–H bending vibration peak in the carboxyl group after loading Pd–Ni moves toward the high-frequency region, from 1390.5 to 1401.95 cm−1.

Figure 5 The FTIR pattern of the support and the catalyst Pd1–Ni1/AC.
Figure 5

The FTIR pattern of the support and the catalyst Pd1–Ni1/AC.

3.4 Structural features of the catalysts by XRD

As shown in Figure 6a, the XRD profiles of Pd1–Ni1/AC catalyst display four typical diffraction peaks of face-centered cubic (fcc) crystalline of Pd. The Pd (111) peak was employed to calculate the crystallite size from the Scherrer equation [35,36]:

(2)D=kλβ2θcosθ

where D means the average crystallite size (nm), K represents a constant of the equipment, λ is the wavelength of the X-ray radiation (0.154056 nm), β2θ is the full width at half maximum (rad) of the identification peak, and θ is the diffraction angle. In addition, the lattice parameter (ahkl) values were computed according to the Bragg’s formula [37,38,39]:

(3)sinθ=λh2+k2+l22aforacubicstructure

where h, k, and l are the Miller indices and λ is the interplanar distance for the (111) plane of the catalyst.

Figure 6 Powder XRD patterns of the synthesized: Pd1–Ni1/AC (a), Pd/AC (b), and AC (c) samples.
Figure 6

Powder XRD patterns of the synthesized: Pd1–Ni1/AC (a), Pd/AC (b), and AC (c) samples.

The obtained crystallite size and the lattice parameter for nanoparticles of Pd1–Ni1/AC catalyst are 11.47 nm and 3.96 Å, respectively; the lattice parameter is 0.02 Å larger than the value of pure palladium (3.94 Å) reported in the literature [40].

As shown in Figure 6b, the 2θ value of the Pd (111) plane of Pd/AC monometallic catalyst is 40.15°, which shifted toward higher 2θ value in comparison to the pure Pd (ICDD Card No. 87-0637), due to the interaction between the active component Pd and the support during the preparation process. After adding the second component Ni, the peak of the Pd (111) plane shifts to 39.7°, resulting in an increase of the lattice parameter, probably due to the lattice distortion and defects, demonstrating the strong electronic interaction between Pd and Ni during the formation of bimetallic nanostructures. However, due to the relatively strong signal of palladium, no obvious nickel diffraction peak can be observed, which is also related to the amorphous structure of the material [24,37].

3.5 XPS

XPS was further employed to examine the oxidation state of active species on the surface of catalyst and the extent of metal alloy formation of the Pd1–Ni1/AC catalyst, which was shown in Figure 7. The peak of Pd 3d5/2 can be deconvolved into two peaks, 335.8 and 338.2 eV. The lower peak belongs to the Pd0, and the higher peak does not correspond to the value of the Pd2+ reported in the literature but belongs to the existence of electron-deficient Pd species (Pdn+) [41,42,43]. Similar to palladium, Ni 2p3/2 also consists of two peaks of values 856.1 and 862.0 eV, respectively. The former is the peak of Ni0, and the latter is still not corresponding to Ni2+ but belongs to the existence of electron-deficient Ni species (Nin+). Interestingly, however, compared with the corresponding standard values [44,45], the peaks of both Pd0 and Ni0 in the Pd1–Ni1/AC catalyst shifted toward higher binding energy, indicating the interaction between the active ingredient and the support, which explains the better durability of the catalyst. The formation of Pdn+ and Nin+ confirms a strong electronic interaction between Pd–Ni bimetallic nanoparticles, which increases the catalytic activity of the catalyst, and is in excellent agreement with the result of XRD.

Figure 7 The survey XPS spectra of Pd 3d (a) and Ni 2p (b) in Pd1–Ni1/AC.
Figure 7

The survey XPS spectra of Pd 3d (a) and Ni 2p (b) in Pd1–Ni1/AC.

In addition, there exists strong electro heterogeneity on the surface of bimetallic nanoparticles as a result of the electronegativity discrepancy between the active components [46], which would speed up the heterolysis process of hydrogen [47,48,49] and enhance the interaction between the adsorbed reactant molecules and the active components [50]; therefore, greatly accelerate the hydrogenolysis process.

3.6 FESEM

The FESEM images for the synthesized Pd1–Ni1/AC bimetallic catalyst are presented in Figure 8. These lighting dots in Figure 8 are relevant to the spherical metal nanoparticles which have been well dispersed and stably supported on the inner wall of the pores and the surface of the AC, and most of them have fine particle size. No obvious agglomeration was observed between the nanoparticles. However, a small amount of large granular metal is also supported on the surface of the AC. During the impregnation process, when the concentration of metal ions on the surface of the AC is relatively high, a small number of atoms have sufficient chances to selectively adsorb on certain crystal faces of the formed small crystals and then preferentially grow along specific crystal directions, resulting in the formation of large crystals with regular shapes [51].

Figure 8 The FESEM images of Pd1–Ni1/AC catalyst.
Figure 8

The FESEM images of Pd1–Ni1/AC catalyst.

3.7 HRTEM

The morphology of Pd1–Ni1/AC catalyst was further investigated by HRTEM and the image is given in Figure 9. It could be observed that nanoparticles of Pd1–Ni1/AC catalyst were distributed uniformly on AC support and the size of the nanoparticles mainly distribute among 4 and 14 nm. The ultrafine nanoparticles can provide larger contact area between the active site and the reactants, which can result in high catalytic efficiency.

Figure 9 TEM image of Pd1–Ni1/AC (a and b); high-resolution TEM image of Pd1–Ni1/AC (c); size distribution of Pd1–Ni1/AC nanoparticles (d).
Figure 9

TEM image of Pd1–Ni1/AC (a and b); high-resolution TEM image of Pd1–Ni1/AC (c); size distribution of Pd1–Ni1/AC nanoparticles (d).

The surface-averaged diameter, ds, was calculated by applying the following equation [52,53]:

(4)ds=i=1nnidi3i=1nnidi2

where ni represents the number of particles of diameter di and the ds value was found to be 9.8 nm, and it is close to the value of the XRD report.

As shown in Figure 9, the lattice fringe spacing in catalytic particles was measured as 0.213 nm, which was located within the values of monometallic Pd (111) (0.225 nm) and Ni (111) (0.203 nm) [20,54], representing the formation of Pd–Ni bimetallic structures related to the changed lattice fringe spacing [12].

4 Conclusions

In general, a novel bimetallic nanoparticle catalyst for the hydrodechlorination of bleached shellac was prepared by impregnation method enhanced with ultrasonic, and the nanoparticles of Pd1–Ni1 bimetallic-activated components of the catalyst were distributed uniformly on AC support with the size of the nanoparticles mainly distribute among 4 and 14 nm.

The lattice fringe spacing in catalytic particles was measured as 0.213 nm which is laying in the region between monometallic Pd (111) (0.225 nm) and Ni (111) (0.203 nm), representing the formation of Pd–Ni bimetallic structures.

The Pd1–Ni1Pd1–Ni1/AC catalyst exhibited the highest catalytic activity and good stability for hydrodechlorination of the bleached shellac. The dechlorination efficiency reached 92.58 wt% while it was used to eliminate the chlorine combined in the bleached shellac and can reach 90.03 wt% in the fifth recycling use. The excellent catalytic performance of Pd1–Ni1/AC is mainly due to the formation of a Pd–Ni bimetallic structure in catalysis. The intrinsic synergy between the bimetallic particles accelerates the hydrogenolysis process of H2 on the surface of bimetallic nanoparticles and ultimately improves the catalytic activity of the catalyst for the hydrodechlorination of the bleached shellac. And the highest stability of Pd1–Ni1/AC can be explained by the fact that a reasonable Pd/Ni molar ratio can promote the interaction between bimetal and support to the greatest extent.

The excellent catalytic performance of nano Pd–Ni bimetallic catalysts in the present research demonstrates the natural advantages of bimetallic particles, and it is confirmed that Pd–Ni bimetallic nanoparticles may have potential prospects for catalytic hydrodechlorination of chlorine-containing macromolecular organics.


tel: +86-871-6519-8154

Acknowledgments

This research was supported by National Natural Science Foundation of China (Project No. 21978122, 21566017, and 21266011), and the National College Student Innovation and Entrepreneurship Training Program of China (201710674161) as well as Kunming University of Science and Technology Test Fund (2018M20172202027).

  1. Conflict of interest: The authors declare no conflict of interest.

References

[1] Liao YL, Peng JH, Liu ZH. National and international seedlac processing development and its trend. Scientla Silvae Sinicae. 2007;7:93–100. 10.3321/j.issn:1001-7488.2007.07.016.Search in Google Scholar

[2] Jo WS, Song HY, Song NB, Lee JH, Min SC, Song KB. Quality and microbial safety of ‘Fuji’ apples coated with carnauba-shellac wax containing lemongrass oil. LWT Food Sci Technol. 2014;55:490–14. 10.1016/j.lwt.2013.10.034.Search in Google Scholar

[3] Remadevi OK, Siddiqui MZ, Nagaveni HC, Rao MV, Shiny KS, Ramani R. Efficacy of shellac-based varnishes for protection of wood against termite, borer and fungal attack. J Indian Acad Wood Sci. 2015;12:9–14. 10.1007/s13196-015-0138-2.Search in Google Scholar

[4] Al-Gousous J, Penning M, Langguth P. Molecular insights into shellac film coats from different aqueous shellac salt solutions and effect on disintegration of enteric-coated soft gelatin capsules. Int J Pharm. 2015;484:283–91. 10.1016/j.ijpharm.2014.12.060.Search in Google Scholar PubMed

[5] Wang X, Yu DG, Li XY, Bligh SW, Williams GR. Electrospun medicated shellac nanofibers for colon-targeted drug delivery. Int J Pharm. 2015;490:384–90. 10.1016/j.ijpharm.2015.05.077.Search in Google Scholar PubMed

[6] Liao YL, Chai XJ. Preparation of bleached low-chlorine shellac by catalytic hydrogenation and its structural characterization. Chem Ind For Products. 2008;28:100–4. 10.3321/j.issn:0253-2417.2008.05.020.Search in Google Scholar

[7] Wei W, Ye W, Wang J, Huang C, Xiong JB, Qiao H, et al. Hydrangea-like alpha-Ni1/3Co2/3(OH)2 reinforced by ethyl carbamate “rivet” for all-solid-state supercapacitors with outstanding comprehensive performance. ACS Appl Mater Inter. 2019;11:32269–81. 10.1021/acsami.9b09555.Search in Google Scholar PubMed

[8] Almeida CVS, Tremiliosi-Filho G, Eguiluz KIB, Salazar-Band GR. Improved ethanol electro-oxidation at Ni@Pd/C and Ni@PdRh/C core–shell catalysts. J Catal. 2020;391:175–89. 10.1016/j.jcat.2020.08.024.Search in Google Scholar

[9] Wei W, Wu J, Cui S, Zhao Y, Chen W, Mi L. alpha-Ni(OH)2/NiS1.97 heterojunction composites with excellent ion and electron transport properties for advanced supercapacitors. Nanoscale. 2019;11:6243–53. 10.1039/c9nr00962k.Search in Google Scholar PubMed

[10] Chen H, Xu ZY, Wan HQ, Zheng JZ, Yin DQ, Zheng SR. Aqueous bromate reduction by catalytic hydrogenation over Pd/Al2O3 catalysts. Appl Catal B Environ. 2010;96:307–13. 10.1016/j.apcatb.2010.02.021.Search in Google Scholar

[11] Molina CB, Pizarro AH, Casas JA, Rodriguez JJ. Aqueous-phase hydrodechlorination of chlorophenols with pillared clays-supported Pt, Pd and Rh catalysts. Appl Catal B Environ. 2014;148–149:330–8. 10.1016/j.apcatb.2013.11.010.Search in Google Scholar

[12] Zhang JW, Sun KK, Li DD, Deng T, Lu GP, Cai C. Pd-Ni bimetallic nanoparticles supported on active carbon as an efficient catalyst for hydrodeoxygenation of aldehydes. Appl Catal A Gen. 2019;569:190–5. 10.1016/j.apcata.2018.10.038.Search in Google Scholar

[13] Gu J, Zhang YW, Tao FF. Shape control of bimetallic nanocatalysts through well-designed colloidal chemistry approaches. Chem Soc Rev. 2012;41:8050–65. 10.1039/c2cs35184f.Search in Google Scholar PubMed

[14] Śrębowata A, Juszczyk W, Kaszkur Z, Karpiński Z. Hydrodechlorination of 1,2-dichloroethane on active carbon supported palladium–nickel catalysts. Catal Today. 2007;124:28–35. 10.1016/j.cattod.2007.02.010.Search in Google Scholar

[15] Wu Y, Gan L, Zhang S, Song H, Lu C, Li W, et al. Carbon-nanotube-doped Pd-Ni bimetallic three-dimensional electrode for electrocatalytic hydrodechlorination of 4-chlorophenol: Enhanced activity and stability. J Hazard Mater. 2018;356:17–25. 10.1016/j.jhazmat.2018.05.034.Search in Google Scholar PubMed

[16] Seshu Babu N, Lingaiah N, Sai Prasad PS. Characterization and reactivity of Al2O3 supported Pd-Ni bimetallic catalysts for hydrodechlorination of chlorobenzene. Appl Catal B Environ. 2012;111–2:309–16. 10.1016/j.apcatb.2011.10.013.Search in Google Scholar

[17] Vijayakumar G, Pandurangan A. Up-gradation of α-tetralone to jet-fuel range hydrocarbons by vapour phase hydrodeoxygenation over Pd Ni/SBA-16 catalysts. Energy. 2017;140:1158–72. 10.1016/j.energy.2017.09.038.Search in Google Scholar

[18] Lai QH, Zhang C, Holles JH. Hydrodeoxygenation of guaiacol over Ni@Pd and Ni@Pt bimetallic overlayer catalysts. Appl Catal A Gen. 2016;528:1–12. 10.1016/j.apcata.2016.09.009.Search in Google Scholar

[19] Xu X, Li Y, Gong Y, Zhang P, Li H, Wang Y. Synthesis of palladium nanoparticles supported on mesoporous N-doped carbon and their catalytic ability for biofuel upgrade. J Am Chem Soc. 2012;134:16987–90. 10.1021/ja308139s.Search in Google Scholar PubMed

[20] Liu CL, Nan CS, Yang L, Li F. Facile synthesis and synergistically acting catalytic performance of supported bimetallic Pd-Ni nanoparticle catalysts for selective hydrogenation of citral. Mol Catal. 2017;436:237–47. 10.1016/j.mcat.2017.04.025.Search in Google Scholar

[21] Hou RJ, Yu WT, Porosoff MD, Chen JGG, Wang TF. Selective hydrogenation of 1,3-butadiene on Pd-Ni bimetallic catalyst: From model surfaces to supported catalysts. J Catal. 2014;316:1–9. 10.1016/j.jcat.2014.04.015.Search in Google Scholar

[22] Gunji T, Wakabayashi RH, Noh SH, Han B, Matsumoto F, DiSalvo FJ, et al. The effect of alloying of transition metals (M = Fe, Co, Ni) with palladium catalysts on the electrocatalytic activity for the oxygen reduction reaction in alkaline media. Electrochim Acta. 2018;283:1045–52. 10.1016/j.electacta.2018.06.051.Search in Google Scholar

[23] Xu C, Liu YQ, Hao Q, Duan HM. Nanoporous Pd-Ni alloys as highly active and methanol-tolerant electrocatalysts towards oxygen reduction reaction. J Mater Chem A. 2013;1(43):13542–8. 10.1039/c3ta12765f.Search in Google Scholar

[24] Zhou YH, Zhang ZY, Wang SQ, Williams N, Cheng Y, Luo SZ, et al. rGO supported PdNi-CeO2 nanocomposite as an efficient catalyst for hydrogen evolution from the hydrolysis of NH3BH3. Int J Hydrog Energ. 2018;43:18745–53. 10.1016/j.ijhydene.2018.08.053.Search in Google Scholar

[25] Li J, Zhou PP, Li F, Ren R, Liu Y, Niu JR, et al. Ni@Pd/PEI–rGO stack structures with controllable Pd shell thickness as advanced electrodes for efficient hydrogen evolution. J Mater Chem A. 2015;3:11261–8. 10.1039/c5ta01805f.Search in Google Scholar

[26] Zhang JW, Sun KK, Li DD, Deng T, Lu GP, Cai C. Pd-Ni bimetallic nanoparticles supported on active carbon as an efficient catalyst for hydrodeoxygenation of aldehydes. Appl Catal A Gen. 2019;569:190–5. 10.1016/j.apcata.2018.10.038.Search in Google Scholar

[27] Frecha E, Torres D, Pueyo A, Suelves I, Pinilla JL. Scanning different Ni-noble metal (Pt, Pd, Ru) bimetallic nanoparticles supported on carbon nanofibers for one-pot cellobiose conversion. Appl Catal A Gen. 2019;585:117182. 10.1016/j.apcata.2019.117182.Search in Google Scholar

[28] González-Vera D, Bustamante TM, Díaz de León JN, Dinamarca R, Morales R, Osorio-Vargas PA, et al. Chemoselective nitroarene hydrogenation over Ni-Pd alloy supported on TiO2 prepared from ilmenite-type PdxNi1−xTiO3. Mater Today Commun. 2020;24:101091. 10.1016/j.mtcomm.2020.101091.Search in Google Scholar

[29] Araujo RB, Martín-Yerga D, Campos dos Santos E, Cornell A, Pettersson LGM. Elucidating the role of Ni to enhance the methanol oxidation reaction on Pd electrocatalysts. Electrochim Acta. 2020;360:136954. 10.1016/j.electacta.2020.136954.Search in Google Scholar

[30] Standardization Administration of China. Methods of testing lac products – Determination of chlorine content. Chinese national standard GB/T 8143-2008, Beijing; 2008.Search in Google Scholar

[31] Cárdenas-Lizana F, Gómez-Quero S, Amorim C, Keane MA. Gas phase hydrogenation of p-chloronitrobenzene over Pd–Ni/Al2O3. Appl Catal A Gen. 2014;473:41–50. 10.1016/j.apcata.2014.01.001.Search in Google Scholar

[32] Zhang Y, Wang YY, liao YL, Guo MY, Shi GC. Preparation, characterization and dechlorination property of nano Pd-Ni/γ-Al2O3 bimetallic catalyst. Funct Mater Lett. 2019;12(6):1951003. 10.1142/S1793604719510032.Search in Google Scholar

[33] Cao YH, Gu Y, Wang KL, Wang XM, Gu ZG, Ambrico T, et al. Adsorption of creatinine on active carbons with nitric acid hydrothermal modification. J Taiwan Inst Chem Eng. 2016;66:347–56. 10.1016/j.jtice.2016.06.008.Search in Google Scholar

[34] You FT, Yu GW, Xing ZJ, Li J, Xie SY, Li CX, et al. Enhancement of NO catalytic oxidation on activated carbon at room temperature by nitric acid hydrothermal treatment. Appl Surf Sci. 2019;471:633–44. 10.1016/j.apsusc.2018.12.066.Search in Google Scholar

[35] Cheng HH, Chen G, Zhang Y, Zhu YF, Li LQ. Boosting low-temperature de/re-hydrogenation performances of MgH2 with Pd-Ni bimetallic nanoparticles supported by mesoporous carbon. Int J Hydrog Energy. 2019;44:10777–87. 10.1016/j.ijhydene.2019.02.218.Search in Google Scholar

[36] Zafari R, Abdouss M, Zamani Y. Application of response surface methodology for the optimization of light olefins production from CO hydrogenation using an efficient catalyst. Fuel. 2019;237:1262–73. 10.1016/j.fuel.2018.10.074.Search in Google Scholar

[37] Koskun Y, Savk A, Sen B, Sen F. Highly sensitive glucose sensor based on monodisperse palladium nickel/activated carbon nanocomposites. Anal Chim Acta. 2018;1010:37–43. 10.1016/j.aca.2018.01.035.Search in Google Scholar PubMed

[38] Sen B, Kuzu S, Demir E, Akocak S, Sen F. Monodisperse palladium–nickel alloy nanoparticles assembled on graphene oxide with the high catalytic activity and reusability in the dehydrogenation of dimethylamine–borane. Int J Hydrog Energ. 2017;42:23276–83. 10.1016/j.ijhydene.2017.05.113.Search in Google Scholar

[39] Şen B, Aygün A, Okyay TO, Şavk A, Kartop R, Şen F. Monodisperse palladium nanoparticles assembled on graphene oxide with the high catalytic activity and reusability in the dehydrogenation of dimethylamine-borane. Int J Hydrog Energ. 2018;43:20176–82. 10.1016/j.ijhydene.2018.03.175.Search in Google Scholar

[40] Guo YY, Dai CN, Lei ZG. Hydrogenation of 2-ethylanthraquinone on Pd-La/SiO2/cordierite and Pd-Zn/SiO2/cordierite bimetallic monolithic catalysts. Chem Eng Process. 2019;136:211–25. 10.1016/j.cep.2018.11.006.Search in Google Scholar

[41] Tan L, Li T, Zhou J, Chen H, Jiang F. Liquid-phase hydrogenation of N-nitrosodimethylamine over Pd-Ni supported on CeO2-TiO2: The role of oxygen vacancies. Colloids Surf A Physicochem Eng Asp. 2018;558:211–8. 10.1016/j.colsurfa.2018.08.066.Search in Google Scholar

[42] Jiang T, Huai Q, Geng T, Ying WY, Xiao TC, Cao FH. Catalytic performance of Pd–Ni bimetallic catalyst for glycerol hydrogenolysis. Biomass Bioenerg. 2015;78:71–9. 10.1016/j.biombioe.2015.04.017.Search in Google Scholar

[43] Chen H, Li T, Jiang F, Wang Z. Enhanced catalytic reduction of N-nitrosodimethylamine over bimetallic Pd-Ni catalysts. J Mol Catal A Chem. 2016;421:167–77. 10.1016/j.molcata.2016.05.026.Search in Google Scholar

[44] Singha S, Sahoo M, Parida KM. Highly active Pd nanoparticles dispersed on amine functionalized layered double hydroxide for Suzuki coupling reaction. Dalton Trans. 2011;40:7130–2. 10.1039/c1dt10697j.Search in Google Scholar PubMed

[45] Zhao JW, Shao MF, Yan DP, Zhang ST, Lu ZZ, Li ZX, et al. A hierarchical heterostructure based on Pd nanoparticles/layered double hydroxide nanowalls for enhanced ethanol electrooxidation. J Mater Chem A. 2013;1(19):5840–6. 10.1039/c3ta10588a.Search in Google Scholar

[46] Sarina S, Zhu H, Jaatinen E, Xiao Q, Liu H, Jia J, et al. Enhancing catalytic performance of palladium in gold and palladium alloy nanoparticles for organic synthesis reactions through visible light irradiation at ambient temperatures. J Am Chem Soc. 2013;135:5793–801. 10.1021/ja400527a.Search in Google Scholar PubMed

[47] Zhang JW, Lu GP, Cai C. Self-hydrogen transfer hydrogenolysis of β-O-4 linkages in lignin catalyzed by MIL-100(Fe) supported Pd–Ni BMNPs. Green Chem. 2017;19:4538–43. 10.1039/c7gc02087b.Search in Google Scholar

[48] Zhang JW, Cai Y, Lu GP, Cai C. Facile and selective hydrogenolysis of β-O-4 linkages in lignin catalyzed by Pd–Ni bimetallic nanoparticles supported on ZrO2. Green Chem. 2016;18:6229–35. 10.1039/c6gc02265k.Search in Google Scholar

[49] Zhang JG, Teo J, Chen X, Asakura H, Tanaka T, Teramura K, et al. A series of NiM (M = Ru, Rh, and Pd) bimetallic catalysts for effective lignin hydrogenolysis in water. ACS Catal. 2014;4:1574–83. 10.1021/cs401199f.Search in Google Scholar

[50] Xiao Q, Liu Z, Bo A, Zavahir S, Sarina S, Bottle S, et al. Catalytic transformation of aliphatic alcohols to corresponding esters in O2 under neutral conditions using visible-light irradiation. J Am Chem Soc. 2015;137:1956–66. 10.1021/ja511619c.Search in Google Scholar PubMed

[51] Qiu CC, Shang R, Xie YF, Bu YR, Li CY, Ma HY. Electrocatalytic activity of bimetallic Pd–Ni thin films towards the oxidation of methanol and ethanol. Mater Chem Phys. 2010;120:323–30. 10.1016/j.matchemphys.2009.11.014.Search in Google Scholar

[52] Ding E, Jujjuri S, Sturgeon M, Shore SG, Keane MA. Novel one step preparation of silica supported Pd/Sr and Pd/Ba catalysts via an organometallic precursor: Application in hydrodechlorination and hydrogenation. J Mol Catal A Chem. 2008;294:51–60. 10.1016/j.molcata.2008.07.020.Search in Google Scholar

[53] Osorio-Vargas P, Flores-González NA, Navarro RM, Fierro JLG, Campos CH, Reyes P. Improved stability of Ni/Al2O3 catalysts by effect of promoters (La2O3, CeO2) for ethanol steam-reforming reaction. Catal Today. 2016;259:27–38. 10.1016/j.cattod.2015.04.037.Search in Google Scholar

[54] Wang YW, Zeng YY, Wu XC, Mu MM, Chen LG. A novel Pd-Ni bimetallic synergistic catalyst on ZIF-8 for Sonogashira coupling reaction. Mater Lett. 2018;220:321–4. 10.1016/j.matlet.2018.03.006.Search in Google Scholar

Received: 2020-08-28
Revised: 2020-11-24
Accepted: 2020-12-01
Published Online: 2020-12-23

© 2020 Yu Zhang et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 29.3.2024 from https://www.degruyter.com/document/doi/10.1515/gps-2020-0071/html
Scroll to top button