Next Article in Journal
Insights over Titanium Modified FeMgOx Catalysts for Selective Catalytic Reduction of NOx with NH3: Influence of Precursors and Crystalline Structures
Next Article in Special Issue
Electrospun Nanofibers Embedding ZnO/Ag2CO3/Ag2O Heterojunction Photocatalyst with Enhanced Photocatalytic Activity
Previous Article in Journal
Oxygen Vacancy Enhanced Photoreduction Cr(VI) on Few-Layers BiOBr Nanosheets
Previous Article in Special Issue
Solar-Light-Driven Efficient ZnO–Single-Walled Carbon Nanotube Photocatalyst for the Degradation of a Persistent Water Pollutant Organic Dye
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of Palladium Nanoparticles Decorated Polyethyleneimine/Polycaprolactone Composite Fibers Constructed by Electrospinning with Highly Efficient and Recyclable Catalytic Performances

1
State Key Laboratory of Metastable Materials Science and Technology, Yanshan University, Qinhuangdao 066004, China
2
Hebei Key Laboratory of Applied Chemistry, School of Environmental and Chemical Engineering, Yanshan University, Qinhuangdao 066004, China
3
National Engineering Research Center for Equipment and Technology of Cold Strip Rolling, Yanshan University, Qinhuangdao 066004, China
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(6), 559; https://doi.org/10.3390/catal9060559
Submission received: 24 May 2019 / Revised: 20 June 2019 / Accepted: 21 June 2019 / Published: 22 June 2019
(This article belongs to the Special Issue New Trends in the Photocatalytic Removal of Organic Dyes)

Abstract

:
Nano-sized palladium nanoparticles showed high catalytic activity with severe limitations in catalytic field due to the tendency to aggregate. A solid substrate with large specific surface area is an ideal carrier for palladium nanoparticles. In present work, polyethyleneimine/polycaprolactone/Pd nanoparticles (PEI/PCL@PdNPs) composite catalysts were successfully designed and prepared by electrospinning and reduction methods using PEI/PCL elexctrospun fiber as carrier. The added PEI component effectively regulated the microscopic morphology of the PEI/PCL fibers, following a large number of pit structures which increased the specific surface area of the electrospun fibers and provided active sites for loading of the palladium particles. The obtained PEI/PCL@PdNPs catalysts for reductions of 4-nitrophenol (4-NP) and 2-nitroaniline (2-NA) exhibited extremely efficient, stable, and reusable catalytic performance. It was worth mentioning that the reaction rate constant of catalytic reduction of 4-NP was as high as 0.16597 s−1. Therefore, we have developed a highly efficient catalyst with potential applications in the field of catalysis and water treatment.

1. Introduction

In recent years, there has been an increasing shortage of water sources available in the context of rapid technological and economic development, and widespread attention has been paid to the protection and purification of water sources. Harmful organic waste is a main source of water pollution. That conversion of harmful organic chemicals into harmless or low toxicity compounds under mild conditions has become an extremely important area of research. It is well known that 4-nitrophenol (4-NP) is widely used in the synthesis of dyes and medicines, and is one of the common chemical products used in the chemical industry [1]. However, it is prone to contaminate soil and water [2]. Methods for purifying such compounds are photocatalytic degradation [3], adsorption [4], microbial degradation [5], and nitro reduction [6], etc. Among these methods, the nitro reduction is one of the most commonly used methods and has the characteristics of simple process, mild conditions and environmental protection [7]. Therefore, it is important to develop effective catalysts to convert hazardous organics into reusable products.
Precious metal particles with excellent catalytic activity, such as palladium, have received extensive attention [8,9,10]. However, the palladium nanoparticles with small particle size and high surface energy are unstable and tend to agglomerate, inhibiting their catalytic activity and catalytic efficiency [11]. An effective way to solve the aggregation of metal particles is to load the particles on a carrier substrate [12,13,14,15,16]. Recently, several different palladium-based catalysts have been reported. Lebaschi et al. synthesized [email protected] NPs catalyst which used environmentally friendly natural black tea extract as a reducing agent, was simple in preparation and excellent in catalytic performance [17]. Wu et al. synthesized CN-supported PdNPs nanocomposites using cellulose nanocrystals (CN) that acted as both supporting substrate and reducing agent, which exhibited higher catalytic activity in catalytic reduction of methylene blue and 4-NP [18]. Chen et al. prepared bimetallic AuPdNPs/GNs catalysts by one-pot method utilizing graphene nanosheets (GNs) with large specific surface area, which showed higher catalytic ability than single metal catalysts [19]. Le et al. reported the preparation of fibrous nano-silica supported palladium nanocatalyst (Pd/KCC-1), and the unique dendritic fiber morphology of the support (KCC-1) resulted in poor aggregation of palladium nanoparticles [20].
Electrospinning is a simple and effective method for the direct and continuous preparation of polymer nanofibers [21,22]. The prepared electrospun fibers have unique characteristics, such as specific surface area and high porosity [23,24], which can be widely used in the fields of adsorption [25,26], catalysis and filtration [27,28,29,30]. Therefore, electrospun fibers can be used as an ideal carrier for metal nanoparticles. Polycaprolactone (PCL) containing abundant polar methylene and ester groups has good biodegradability, biocompatibility and mechanical properties, which is one of the most easily used polymers for electrospinning [31,32,33]. However, the hydrophobicity of polycaprolactone fibers limits their application in water soluble systems. Polyethyleneimine (PEI) contains many amino functional groups, which are combined with PCL to improve the hydrophilicity of the fibers. In present work, based on the characteristic of large surface area of electrospinning fibers, PEI/PCL composite fibers were prepared as a carrier for supporting palladium nanoparticles, which provide active sites. The surface modification of the PEI/PCL fibers was achieved by changing the content of the PEI component, and the appearance of the pit structures further increased the specific surface area of the fibers and the active sites for supporting palladium particles. The catalytic reduction of 4-nitrophenol and 2-nitroaniline by PEI/PCL@PdNPs composite catalyst showed extremely high catalytic activity and stability, which meant potential application in water purification and catalysis.

2. Results and Discussion

2.1. Characterization of PEI/PCL@PdNPs Composites

Figure 1 illustrated the experimental process of electrospinning, chemical modification and catalytic reduction in the preparation and application of PEI/PCL@PdNPs composite catalyst. Firstly, the granular PCL and the liquid PEI were dissolved in a good organic solvent, chloroform, to obtain a homogeneous electrospinning precursor solution which was magnetically stirred at room temperature overnight. The precursor solution, placed in a 10 mL syringe with a stainless steel needle (20G), was electrospun under a high voltage driving force to obtain a PEI/PCL composite fiber. Secondly, NaBH4 as a good reducing agent reduced aqueous PdCl2 to palladium to obtain a solution containing palladium nanoparticles. The prepared PEI/PCL fibers were immersed in the palladium nanoparticles solution for one hour to complete chemical modification and obtained PEI/PCL@PdNPs composites. Finally, the prepared PEI/PCL@PdNPs was used as catalyst to catalytically reduce 4-NP and 2-NA to explore its catalytic activity. Due to the large number of functional groups present in PEI, it was possible that some NH2 groups reacted with the carbonyl ester bond, and other functional amine groups remained free on the surface, allowing them to bind to other molecules such as palladium [34].
Figure 2 depicted the scanning electron microscope (SEM) surface morphologies of PEI/PCL fibers (Samples 1–6) at different PEI contents. It was clear that the structures of the PEI/PCL electrospun fibers were related to the mixing ratio of PEI and PCL. PEI/PCL fibers of different mixing ratios exhibited inhomogeneous thickness structure. As the PEI content increased, the smooth PEI/PCL composite fibers gradually appeared pits on the fiber surface. The larger the mixing ratio of PEI to PCL, the more and deeper pits on the fiber surface. It was well known that parameters such as polymer viscosity and concentration were key factors influencing electrospinning [22]. The pit structure of the PEI/PCL electrospun fiber was attributed to the difference of PEI component contained in the polymer fibers. As shown in Table 1, the specific surface area of PEI/PCL fibers improved from 8.34 to 12.93 m2g−1 with the increment of PEI component in composite fibers. The electrospun fibers with pit-like surface structures further increased the specific surface area of the fibers, which increased the particles attachment sites and helpful to next modification of palladium nanoparticles on surface of the synthesized composite fiber.
In order to further observe the structure of as-prepared PEI/PCL@PdNPs composite, the transmission electron microscopy (TEM) morphologies of the obtained fibers were examined, as shown in Figure 3. It was obvious that the surface of PEI/PCL fiber (sample 6, w/w, 35:65) was uneven, as shown in Figure 3b. The rugged structure increased the specific surface area helpful to next chemical modification of the fibers. It could be seen in Figure 3c that the PdNPs modified on the surface of PEI/PCL fibers showed the mean size range of 5–10 nm. As shown in Figure 3d, the interplanar spacing of the palladium nanoparticles was 0.2243 nm, which well matched the (111) crystal plane of Pd [35]. Furthermore, the Pd element mapping shown in Figure 3e was good evidence of the presence and distribution of palladium nanoparticles on the surface of the prepared PEI/PCL fibers. It was reasonably speculated that hydrophilic PdNPs were successfully anchored on the surface of PEI/PCL composite fibers via weak intermolecular interactions, then the composites could be expected good catalytic activity in next catalytic process [36,37].
In addition, the morphology of PEI/PCL@PdNPs was characterized by scanning electron microscope (SEM), as shown in Figure 4. Figure 4a depicted the SEM structure of PEI/PCL@PdNPs, and it was apparent that deposits were present on the surface of the PEI/PCL fibers. The energy dispersive spectroscopy (EDS) image of the obtained PEI/PCL@PdNPs shown in Figure 4b–d clearly showed the presence of carbon, oxygen, nitrogen, and palladium elements. Due to the shielding effect, only a part of the nitrogen element in the PEI could be displayed, as shown in Figure 4d. The presence of small amounts of gold and copper was due to the coated-gold treatment and copper substrate used during the measurement. From EDS data, the detailed atomic percentages were found to be C = 75.37%, O = 12.61%, N = 7.88%, and Pd = 1.23%. The above results indicated that the palladium nanoparticles were successfully modified on the surface of the PEI/PCL fibers.
The prepared PEI/PCL fibers and PEI/PCL@PdNPs composite were further characterized by X-ray diffraction spectroscopy, as shown in Figure 5. It was obvious from curves (a) and (b) that both the PEI/PCL composite fibers with different weight ratios showed characteristic diffraction peaks at 2θ = 21.4° and 23.8°. The PEI/PCL fiber with low PEI component showed strong characteristic peaks, while the binary fiber with high PEI component showed the weaker intensity of the characteristic peak. The difference in the intensity of the characteristic peaks of PEI/PCL fibers with different blend ratios demonstrated that the higher the PEI component, the lower the crystallinity of the composite fibers. From the X-ray diffraction spectrum of the PEI/PCL@PdNPs composite, it could be seen that a characteristic peak appeared at 2θ = 40.1°, corresponding to the (111) crystal plane of the palladium particles, which further proved that the palladium nanoparticles successfully modified the PEI/PCL composite fibers [38,39]. In addition, the characteristic peak intensity of the PEI/PCL@PdNPs composite increased as the palladium particles were successfully modified on the surface of the PEI/PCL composite fibers.
To investigate the thermal stability of the prepared PEI/PCL@PdNPs composite, the thermogravimetric characterization of the obtained PCL fibers, PEI/PCL fibers, and PEI/PCL@PdNPs composite were measured under argon gas atmosphere, as shown in Figure 6. TG analysis of the three composite fiber materials showed that the weight loss was slower when the temperature was lower than 120 °C due to the volatilization of residual moisture in the samples [40]. In addition, when the temperature was higher than 500 °C, the quality of the three species of samples tended to be stable. Between 120 °C and 450 °C, the weight of above-mentioned three samples was firstly slow loss, and then the weight of these samples was sharply lost at 400 °C, which was due to the thermal decomposition of various chemical groups and alkyl chains contained in these samples at high temperature [41]. The tempertature points with mass loss 5 wt.% showed the values of 353, 311, and 236 °C for PCL fibers, PEI/PCL fibers, and PEI/PCL@PdNPs composite fibers, respectively, indicating that the addition of the PEI component and PdNPs reduced the thermal stability of the composite fibers. The above significant difference in the thermal stability could reasonably be attributed to the inclusion of palladium nanoparticles on the surface of the composite fibers.
Next, X-ray photoelectron spectroscopy (XPS) was measured to investigate the interface element composition and the reduction level of Pd2+ of the obtained PEI/PCL@PdNPs composite. The results of XPS spectroscopy of PEI/PCL fiber and PEI/PCL@PdNPs composite were shown in Figure 7a. The signals of C1s, N1s and O1s appeared in the XPS full spectrum of PEI/PCL fibers, while the new Pd3d and Si2p signals appeared in the XPS full spectrum of the prepared PEI/PCL@PdNPs. The Si2p signal is due to the contact of the sample with the silicon plate during the examining process. From surface analysis data, the detailed atomic percentages were found to be C = 71.44%, O = 16.08%, N = 6.63%, and Pd = 0.88%. Figure 7b was the deconvolution of the Pd3d peak in the prepared PEI/PCL@PdNPs composite. The XPS spectrum of Pd3d showed double peaks with binding energies at 334.5 eV and 339.7 eV, corresponding to the Pd3d5/2 and Pd3d3/2 components of the metal Pd(0) state, respectively, which confirmed the presence of metal Pd [20,42,43,44]. The above result indicated that excess NaBH4 reduced all Pd2+ in PdCl2 to Pd0 without any residual Pd2+ ions. The mentioned XPS results further demonstrated the successful preparation of PEI/PCL@PdNPs composites.

2.2. Catalytic Performance of PEI/PCL@PdNPs Composites

Free palladium nanoparticles exhibit high catalytic activity, while the tendency to aggregate during the catalytic process reduced their catalytic performance. Various palladium composite catalysts reported previously exhibited different catalytic performance at 25 °C, as shown in Table 2 [17,18,19,20,45,46,47]. Therefore, the prepared PEI/PCL@PdNPs composite catalyst was applied to catalytic reduction of 4-NP and 2-NA to explore its catalytic performance. The catalytic activity of PEI/PCL@PdNPs containing active functional groups such as C=O and -NH2 was attributed to the interaction of PdNPs with the 4-NP substrate adjacent to the catalytic site to promote the reduction of the nitro group. It could be reasonably speculated that the PEI/PCL@PdNPs played a leading role during the catalytic reduction process. When the PEI/PCL@PdNPs catalyst was added to the solution system of the oxidant 4-NP and the reducing agent NaBH4, the electrons were transferred from the BH4 donor to the 4-NP receptor, thereby producing amino derivative [17]. The 4-nitrophenolate and BH4 were adsorbed on the surface of the PEI/PCL@PdNPs catalyst by weak electrostatic interaction. When electrons transferred to Pd NPs, the active hydrogen groups formed and reduced 4-nitrophenol molecules [17]. This electron transfer induced 4-NP hydrogenation spontaneously occurred on the surface of the metal catalyst. Finally, 4-nitroaniline (4-AP) was desorbed from the catalyst surface.
The catalytic performance of PEI/PCL@PdNPs catalyst for 4-NP and 2-NA was observed by UV-visible spectroscopy at room temperature. As shown in Figure 8a, it was clear that the maximum absorption peak at 317 nm of 4-NP aqueous solution shifted to 400 nm owing to the addition of freshly prepared reducing agent NaBH4, indicating the formation of 4-nitrophenol ions [48]. At the same time, the color of the mixed solution system changed from light yellow to bright yellow. In addition, 10 mg of PEI/PCL@PdNPs catalyst (PEI:PCL, w/w, 35:65) was added to the system of 4-NP (10 mL, 0.005 M) and NaBH4 (20 mL, 0.1 M). As shown in Figure 8b, after several seconds, the intensity of the absorption peak at 400 nm significantly decreased and a new absorption peak appeared at 300 nm. Referring to previous literature, the emerging absorption peak corresponded to 4-AP [20]. After 20 seconds, the intensity of the absorption peak at 400 nm of the 4-NP and NaBH4 mixed system decreased to zero, and the color of the solution became transparent (inset photo in Figure 8b), indicating that all of the 4-NP was reduced to 4-AP. The PEI/PCL@PdNPs catalyst showed high catalytic activity for the reduction of 4-NP, which is attributed to the active sites of PEI/PCL substrate fibers for palladium nanoparticles. In addition, the prepared PEI/PCL@PdNPs composite with weight ratios (w/w, 10:90) catalyst was applied to catalytic reduction of 4-NP to explore its catalytic performance, as shown Figure S1. The PEI/PCL@PdNPs catalyst at 10:90 weight ratio completely reduced 4-NP to 4-AP in 100 s, and the rate constant k of catalytic reduction reaction was 0.00061 s−1. At the same time, The loading of Pd in the catalyst was found to be Pd = 0.63% in the detailed atomic percentages of PEI/PCL@PdNPs (w/w, 10:90) material. Compared with the PEI/PCL@PdNPs composite at 35:65 weight ratio, it could be reasonably speculated that the improved catalytic performance was owing to more palladium nanoparticles anchored on the surface of PEI/PCL ((w/w), 35:65) fibers with the increment of PEI component in fiber structures. For comparison, the synthesized palladium nanoparticle in solution catalytically reduced the same solution system of 4-NP (10 mL, 0.005 M) and NaBH4 (20 mL, 0.1 M), and completely reduced 4-NP after 10 minutes, The kinetic constant k of the reduction reaction was 0.006 s−1 as shown in Figure S2. The catalytic performance of PEI/PCL@PdNPs catalyst was significantly better than that of palladium nanoparticles in solution, because the PEI/PCL substrate improved the dispersibility of palladium nanoparticles. At the same time, 10 mg of PEI/PCL fibers without Pd nanoparticles was added to the system of 4-NP (10 mL, 0.005 M) and NaBH4 (20 mL, 0.1 M). After 20 min, no color changes in the solution system had occurred. Moreover, the catalytic properties of PEI/PCL@PdNPs catalyst for the reduction of 2-NA were investigated by the same method. As shown in Figure 8d,e, it was worth mentioning that there was no significant change in the absorption peak of the mixed solution after adding fresh NaBH4 to the 2-NA solution. After 9 min of addition of the PEI/PCL@PdNPs catalyst, the 2-NA was completely reduced to o-phenylenediamine (OPD), as reported in the literature [49]. The prepared PEI/PCL@PdNPs exhibited different catalytic capabilities for 4-NP and 2-NA. Apparently, PEI/PCL@PdNPs exhibited higher catalytic efficiency for 4-NP. Furthermore, the component amont of palladium nanoparticles on fibers surface depended on the modification times. Compared with appropriate and enough modification time like 1 hour, palladium nanoparticles were hardly found on the surface of PEI/PCL composite fibers at 10 minutes modification with poor catalytic performance.
Since the concentration of excess NaBH4 in the catalytic reaction was much higher than that of 4-NP or 2-NA, the concentration of NaBH4 in the mixed solution system could be regarded as a constant. Therefore, the concentration of 4-NP or 2-NA was a physical quantity that changed with time t, and the catalytic reduction reaction was a pseudo-first-order reaction [50,51,52,53,54,55,56,57]. The rate of the catalytic reaction (containing adsorption process) was evaluated according to Lambert-Beer’s law and pseudo-first-order reaction kinetics (Equation (1)), where A0, At and C0, Ct were the absorbance and concentration of the mixed solution at the initial and time t, respectively. Figure 8c and f were the linear relationships between the reaction time t and ln(Ct/C0) of catalytic reduction of 4-NP and 2-NA by PEI/PCL@PdNPs, respectively, reflecting the rate of catalytic reduction of 4-NP and 2-NA by PEI/PCL@PdNPs. The kinetic constant k of the above catalytic reduction reaction was calculated to be 0.16597s−1 and 0.00401 s−1, respectively. The kinetic constant k could intuitively reflect the speed of the catalytic reduction reaction [58,59,60,61,62,63,64,65]. Obviously, PEI/PCL@PdNPs catalytically reduced 4-NP faster.
dCt/dt = −kt or ln(Ct/C0) = ln(At/A0) = −kt
The stability and recyclability of the catalyst was another aspect of evaluating the catalyst. Therefore, it was important and necessary to explore the repeated catalytic ability of PEI/PCL@PdNPs catalyst for 4-NP and 2-NA, as shown in Figure 9. The catalytic efficiency of the PEI/PCL@PdNPs catalyst after 8 cycles of repeated catalytic reduction of fresh 4-NP and 2-NA systems were 95% and 92%, respectively. The above result indicated the high catalytic activity, stability and recyclability of the PEI/PCL@PdNPs catalyst. Referring to previous literature, the few decrement in catalytic efficiency was reanonably attributed to the attachment of organics on the surface of catalyst and the loss of palladium particles on the catalyst surface during the washing of PEI/PCL@PdNPs with ethanol and ultrapure water [36].

3. Materials and Methods

3.1. Materials

Polyethylenimine (PEI, M.W. 600.99%) was purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China), and the same with Palladium chloride (PdCl2, Pd 59–60%). Polycaprolactone (PCL, average Mn 80,000) was obtained from Sigma-Aldrich (Shanghai) Trading Co. Ltd. (China). Trichloromethane (CHCl3) was obtained from Jindong Tianzheng Co. Ltd. In addition, Sodium borohydride (NaBH4), 2-nitroaniline (2-NA) and 4-nitrophenol (4-NP) were obtained from Alfa Aesar (Beijing, China). All chemicals were not further purified. A Milli-Q Millipore Filtration System (Millipore Co., Bedford, MA, USA) was used to prepare ultrapure water.

3.2. Preparation of PEI/PCL@PdNPs Composites

11% (w/v) polycaprolactone polymer solution was prepared by dissolving PCL in chloroform referring to previous literature [66,67]. Secondly, according to the different blend ratios of PEI:PCL (10:90, 15:85, 20:80, 25:75, 30:70, 35:65 (w/w)), PEI were added to the abovementioned 11% polycaprolactone solution to obtained different concentration gradients of PEI/PCL mixed solutions. The PEI/PCL mixed solutions of different blend ratios were magnetically stirred overnight at room temperature to obtain uniform electrospinning precursor solutions. The precursor solutions were sequentially electrospun under the parameters of 15 kV voltage, 1 mL/h flow rate, and 25 cm distance between the needle and the receiving plate to obtain PEI/PCL electrospun composite fibers (samples 1–6) of different blend ratios.
0.003 M PdCl2 aqueous solution and 0.01 M NaBH4 solution were disposed in 50 mL volumetric bottles respectively. The above-mentioned PdCl2 aqueous solution (20 mL) in a clean beaker was stirred vigorously and subsequently added 0.01 M NaBH4 solution drop by drop until the color of the solution changed from khaki to dark brown, indicating the reduction of Pd2+ to Pd0 and the formation of Pd particles [68]. Then two species of 10 mg PEI/PCL composite fibers (sample 1 and 6) were added to the above-mentioned Pd particles solution and continued stirring vigorously to avoid the Pd particles agglomeration. After an hour, the PEI/PCL fibers were taken out and washed with deionized water three times before drying. Thus, the PEI/PCL@PdNPs composite materials (sample 7 and 8) were obtained and completed the chemical modification of Pd particles.

3.3. Catalytic Performance of PEI/PCL@PdNPs Composite

According to the previous literature, the prepared PEI/PCL@PdNPs composite was used as a catalyst to reduce 4-NP and 2-NA. The entire catalytic experiment was carried out at 25 °C. Fresh NaBH4 solution (20 mL, 0.1 M) was thoroughly mixed with 4-NP solution (10 mL, 0.005 M), and the UV–vis spectrum of the mixed solution was measured as the absorbance value at the initial time (0 s). Subsequently, 10 mg of the prepared PEI/PCL@PdNPs composite was quickly added to the mixed solution. The absorbance of the mixed solution was measured by UV–vis spectroscopy at intervals (5 s, 15 s, 20 s) until it was reduced to zero, and the catalytic curve of PEI/PCL@PdNPs catalyst for 4-NP was obtained. The same method was used to investigate the catalytic performance of PEI/PCL@PdNPs catalyst for 2-NA solution. NaBH4 acted as a reducing agent during the catalytic reduction process. At the same time, according to the above experimental method, PEI/PCL@PdNPs catalyst was carried out repeated catalytic experiments on fresh 4-NP and NaBH4 mixed solutions. After the catalytic reduction reaction is completed, the PEI/PCL@PdNPs catalyst was removed and the washed several times with ethanol and ultrapure water until the color of the washing liquid did not change. The washed PEI/PCL@PdNPs were then dried for the next catalytic experiment. Catalytic reduction of fresh 4-NP and 2-NA solutions was carried out 8 times in succession.

3.4. Characterization

Scanning electron microscope (SEM, FEI Corporate, Hillsboro, OR, USA) with a field emission gun FEI QUANTA FEG 250 were measured to obtain the morphologies of prepared electrospun fibers. Before measurement, the samples were coated with AuNPs by sputtering in vacuum for 30 seconds. In addition, energy dispersive spectroscopy (EDS) were utilized to acquire the chemical elements analysis of the as-obtained fiber samples. The morphology of the samples was further characterized via Transmission electron microscopy (TEM, HT7700, High-Technologies Corp., Ibaraki, Japan).The microstructure was observed by High-resolution transmission electron microscopy (HRTEM, Tecnai-G2 F30 S-TWIN, Philips, Netherlands). X-ray diffraction patterns were conducted on an X-ray diffractometer (SMART LAB, Rigaku, Akishima, Japan) with a CuKα X-ray radiation source and a Bragg diffraction apparatus. Thermogravimetry-differential scanning calorimetry (TG-DSC) analysis was carried out by NETZSCH STA 409 PC Luxxsi thermal analyzer (Netzsch Instruments Manufacturing Co., Ltd., Seligenstadt, Germany) under argon gas atmosphere. X-ray photoelectron spectroscopy (XPS) was investigated on Thermo Scientific ESCALab 250Xi (Netzsch Instrument Manufacturing GmbH, Seligenstadt, Germany) equipped with 200 W of monochromatic AlKα radiation. UV-vis absorption spectra were measured with UV-visible spectroscopy (UV-2550, Shimadzu Corporation, Tokyo, Japan). BET measurements (NOVA 4200-P, Quantachrome Instruments, Boynton Beach, FL, USA) were obtained to characterize the specific surface areas and pore diameter distribution. The surface areas were calculated by the Brunauer–Emmett–Teller (BET) method.

4. Conclusions

In summary, we have successfully designed and prepared PEI/PCL@PdNPs composite fiber catalyst with remarkable catalytic activity and stability. After the microscopic morphology of the PEI/PCL fibers was regulated, the plenty of pits structures appeared on the surface of the fibers increased the specific surface area of solid substrate fibers and provided abundant active attachment sites and sufficient space for the chemical modification of the palladium nanoparticles, which favored the adhesion of the palladium particles. The catalytic reduction of 4-nitrophenol and 2-nitroaniline by PEI/PCL@PdNPs catalyst exhibited fast, stable and reusable catalytic performance. This work provided new research clues for the preparation of composites loaded with PdNPs and ideal metal particle-based carriers.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/6/559/s1, Figure S1. UV–vis absorption curves of catalytic reduction of 4-NP with (PEI:PCL, w/w, 10:90) PEI/PCL@PdNPs composite (a) and linear relationship of the reduction process (b). Figure S2. UV–vis absorption curves of catalytic reduction of 4-NP via PdNPs solution (a), PEI/PCL fibers (b), and linear relationship of the reduction process (c).

Author Contributions

T.J. and Q.P. conceived and designed the experiments; C.W. and J.Y. performed the experiments; S.H., J.Z., and L.Z. analyzed the data; Z.B. and Q.P. contributed reagents/materials/analysis tools; C.W. and T.J. wrote the paper.

Funding

This research was funded by the National Natural Science Foundation of China (No. 21872119), the Support Program for the Top Young Talents of Hebei Province, China Postdoctoral Science Foundation (No. 2015M580214), and the Research Program of the College Science & Technology of Hebei Province (No. ZD2018091).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Amezquita-Garcia, H.J.; Razo-Flores, E.; Cervantes, F.J.; Rangel-Mendez, J.R. Activated carbon fibers as redox mediators for the increased reduction of nitroaromatics. Carbon 2013, 55, 276–284. [Google Scholar] [CrossRef]
  2. El-Sheikh, S.M.; Ismail, A.A.; Al-Sharab, J.F. Catalytic reduction of p-nitrophenol over precious metals/highly ordered mesoporous silica. New J. Chem. 2013, 37, 2399–2407. [Google Scholar] [CrossRef]
  3. Dieckmann, M.S.; Gray, K.A. A comparison of the degradation of 4-nitrophenol via direct and sensitized photocatalysis in TiO2 slurries. Water Res. 1996, 30, 1169–1183. [Google Scholar] [CrossRef]
  4. Marais, E.; Nyokong, T. Adsorption of 4-nitrophenol onto Amberlite® IRA-900 modified with metallophthalocyanines. J. Hazard. Mater. 2008, 152, 293–301. [Google Scholar] [CrossRef]
  5. Liu, N.; Li, H.; Ding, F.; Xiu, Z.; Liu, P.; Yu, Y. Analysis of biodegradation by-products of nitrobenzene and aniline mixture by a cold-tolerant microbial consortium. J. Hazard. Mater. 2013, 260, 323–329. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, L.; Zheng, S.; Kang, D.E.; Shin, J.Y.; Suh, H.; Kim, I. Synthesis of multi-amine functionalized hydrogel for preparation of noble metal nanoparticles: Utilization as highly active and recyclable catalysts in reduction of nitroaromatics. RSC Adv. 2013, 3, 4692–4703. [Google Scholar] [CrossRef]
  7. Boothroyd, S.R.; Kerr, M.A. A mild and efficient method for the preparation of anilines from nitroarenes. Tetrahedron Lett. 1995, 36, 2411–2414. [Google Scholar] [CrossRef]
  8. Pérez-Lorenzo, M. Palladium nanoparticles as efficient catalysts for Suzuki cross-coupling reactions. J. Phys. Chem. Lett. 2012, 3, 167–174. [Google Scholar] [CrossRef]
  9. Hildebrand, H.; Mackenzie, K.; Kopinke, F.D. Highly active Pd-on-magnetite nanocatalysts for aqueous phase hydrodechlorination reactions. Environ. Sci. Technol. 2009, 43, 3254–3259. [Google Scholar] [CrossRef]
  10. Esumi, K.; Isono, R.; Yoshimura, T. Preparation of PAMAM− and PPI−Metal (silver, platinum, and palladium) nanocomposites and their catalytic activities for reduction of 4-nitrophenol. Langmuir 2004, 20, 237–243. [Google Scholar] [CrossRef]
  11. Lin, Y.; Qiao, Y.; Wang, Y.; Yan, Y.; Huang, J. Self-assembled laminated nanoribbon-directed synthesis of noble metallic nanoparticle-decorated silica nanotubes and their catalytic applications. J. Mater. Chem. 2012, 22, 18314–18320. [Google Scholar] [CrossRef]
  12. Okaa, S.; Kamegawa, T.; Mori, K.; Yamashita, H. An electroless deposition technique for the synthesis of highly active and nano-sized Pd particles on silica nanosphere. Catal. Today 2012, 185, 109–112. [Google Scholar] [CrossRef]
  13. Li, Y.; Fan, X.; Qi, J.; Ji, J.; Wang, S.; Zhang, G.; Zhang, F. Palladium nanoparticle-graphene hybrids as active catalysts for the Suzuki reaction. Nano Res. 2010, 3, 429–437. [Google Scholar] [CrossRef] [Green Version]
  14. Hermans, S.; Bruyr, V.; Devillers, M. A templating effect of carbon nanomaterials on the synthesis of Pd nanoparticles by covalent grafting onto surface O-groups. J. Mater. Chem. 2012, 22, 14479–14486. [Google Scholar] [CrossRef]
  15. Liu, J.; Li, Y.Q.; Zheng, W.J. Synthesis of immobilized nanopalladium on polymer-supported Schiff base, and study of its catalytic activity in the Suzuki–Miyaura reaction. Monatsh. Chem. 2009, 140, 1425. [Google Scholar] [CrossRef]
  16. Domènech, B.; Muñoz, M.; Muraviev, D.N.; Macanás, J. Polymer-stabilized palladium nanoparticles for catalytic membranes: Ad hoc polymer fabrication. Nanoscale Res. Lett. 2011, 6, 406. [Google Scholar] [CrossRef]
  17. Lebaschi, S.; Hekmati, M.; Veisi, H. Green synthesis of palladium nanoparticles mediated by black tea leaves (Camellia sinensis) extract: Catalytic activity in the reduction of 4-nitrophenol and Suzuki-Miyaura coupling reaction under ligand-free conditions. J. Colloid Interf. Sci. 2017, 485, 223–231. [Google Scholar] [CrossRef]
  18. Wu, X.; Lu, C.; Zhang, W.; Yuan, G.; Xiong, R.; Zhang, X. A novel reagentless approach for synthesizing cellulose nanocrystal-supported palladium nanoparticles with enhanced catalytic performance. J. Mater. Chem. A 2013, 1, 8645–8652. [Google Scholar] [CrossRef]
  19. Chen, X.; Cai, Z.; Chen, X.; Oyama, M. AuPd bimetallic nanoparticles decorated on graphene nanosheets: Their green synthesis, growth mechanism and high catalytic ability in 4-nitrophenol reduction. J. Mater. Chem. A 2014, 2, 5668–5674. [Google Scholar] [CrossRef]
  20. Le, X.; Dong, Z.; Li, X.; Zhang, W.; Le, M.; Ma, J. Fibrous nano-silica supported palladium nanoparticles: An efficient catalyst for the reduction of 4-nitrophenol and hydrodechlorination of 4-chlorophenol under mild conditions. Catal. Commun. 2015, 59, 21–25. [Google Scholar] [CrossRef]
  21. Bhardwaj, N.; Kundu, S.C. Electrospinning: A fascinating fiber fabrication technique. Biotechnol. Adv. 2010, 28, 325–347. [Google Scholar] [CrossRef] [PubMed]
  22. Huang, Z.M.; Zhang, Y.Z.; Kotaki, M.; Ramakrishna, S. A review on polymer nanofibers by electrospinning and their applications in nanocomposites. Compos. Sci. Technol. 2003, 63, 2223–2253. [Google Scholar] [CrossRef]
  23. Shariful, M.I.; Sharif, S.B.; Lee, J.J.L.; Habiba, U.; Ang, B.C.; Amalina, M.A. Adsorption of divalent heavy metal ion by mesoporous-high surface area chitosan/poly (ethylene oxide) nanofibrous membrane. Carbohyd. Polym. 2017, 157, 57–64. [Google Scholar] [CrossRef] [PubMed]
  24. Su, C.; Li, Y.; Dai, Y.; Gao, F.; Tang, K.; Cao, H. Fabrication of three-dimensional superhydrophobic membranes with high porosity via simultaneous electrospraying and electrospinning. Mater. Lett. 2016, 170, 67–71. [Google Scholar] [CrossRef]
  25. Habiba, U.; Afifi, A.M.; Salleh, A.; Ang, B.C. Chitosan/(polyvinyl alcohol)/zeolite electrospun composite nanofibrous membrane for adsorption of Cr6+, Fe3+ and Ni2+. J. Hazard. Mater. 2017, 322, 182–194. [Google Scholar] [CrossRef] [PubMed]
  26. Zhao, R.; Wang, Y.; Li, X.; Sun, B.; Wang, C. Synthesis of β-cyclodextrin-based electrospun nanofiber membranes for highly efficient adsorption and separation of methylene blue. ACS Appl. Mater. Inter. 2015, 7, 26649–26657. [Google Scholar] [CrossRef] [PubMed]
  27. Shang, C.; Li, M.; Wang, Z.; Wu, S.; Lu, Z. Electrospun nitrogen-doped carbon nanofibers encapsulating cobalt nanoparticles as efficient oxygen reduction reaction catalysts. ChemElectroChem 2016, 3, 1437–1445. [Google Scholar] [CrossRef]
  28. Ye, J.S.; Liu, Z.T.; Lai, C.C.; Lo, C.T.; Lee, C.L. Diameter effect of electrospun carbon fiber support for the catalysis of Pt nanoparticles in glucose oxidation. Chem. Eng. J. 2016, 283, 304–312. [Google Scholar] [CrossRef]
  29. Zhu, M.M.; Han, J.Q.; Wang, F.; Shao, W.; Xiong, R.H.; Zhang, Q.L.; Pan, H.; Yang, Y.; Samal, S.K.; Zhang, F.; et al. Electrospun nanofibers membranes for effective air filtration. Macromol. Mater. Eng. 2017, 302, 1600353. [Google Scholar] [CrossRef]
  30. Jiang, S.; Hou, H.; Agarwal, S.; Greiner, A. Polyimide nanofibers by “Green” electrospinning via aqueous solution for filtration applications. ACS Sustain. Chem. Eng. 2016, 4, 4797–4804. [Google Scholar] [CrossRef]
  31. Yoshimoto, H.; Shin, Y.M.; Terai, H.; Vacanti, J.P. A biodegradable nanofiber scaffold by electrospinning and its potential for bone tissue engineering. Biomaterials 2003, 24, 2077–2082. [Google Scholar] [CrossRef]
  32. Xu, T.; Liang, Z.; Ding, B.; Feng, Q.; Fong, H. Polymer blend nanofibers containing polycaprolactone as biocompatible and biodegradable binding agent to fabricate electrospun three-dimensional scaffolds/structures. Polymer 2018, 151, 299–306. [Google Scholar] [CrossRef]
  33. Baker, S.R.; Banerjee, S.; Bonin, K.; Guthold, M. Determining the mechanical properties of electrospun poly-ε-caprolactone (PCL) nanofibers using AFM and a novel fiber anchoring technique. Mater. Sci. Eng. C 2016, 59, 203–212. [Google Scholar] [CrossRef] [PubMed]
  34. Delyanee, M.; Solouk, A.; Akbari, S.; Seyedjafari, E. Modification of electrospun poly(L-lactic acid)/polyethylenimine nanofibrous scaffolds for biomedical application. Inter. J. Polym. Mater. 2018, 67, 247–257. [Google Scholar] [CrossRef]
  35. Guo, Y.; Xu, Y.T.; Gao, G.H.; Wang, T.; Zhao, B.; Fu, X.Z.; Sun, R.; Wong, C.P. Electro-oxidation of formaldehyde and methanol over hollow porous palladium nanoparticles with enhanced catalytic activity. Catal. Commun. 2015, 58, 40–45. [Google Scholar] [CrossRef]
  36. Guo, R.; Jiao, T.F.; Xing, R.R.; Chen, Y.; Guo, W.C.; Zhou, J.X.; Zhang, L.X.; Peng, Q.M. Hierarchical AuNPs-loaded Fe3O4/polymers nanocomposites constructed by electrospinning with enhanced and magnetically recyclable catalytic capacities. Nanomaterials 2017, 7, 317. [Google Scholar] [CrossRef]
  37. Zhang, Y.W.; Peng, H.H.; Huang, W.; Zhou, Y.F.; Zhang, X.H.; Yan, D.Y. Hyperbranched poly (amidoamine) as the stabilizer and reductant to prepare colloid silver nanoparticles in situ and their antibacterial activity. J. Phys. Chem. C 2008, 112, 2330–2336. [Google Scholar] [CrossRef]
  38. Li, C.L.; Su, Y.; Lv, X.Y.; Shi, H.J.; Yang, X.G.; Wang, Y.J. Enhanced ethanol electrooxidation of hollow Pd nanospheres prepared by galvanic exchange reactions. Mater. Lett. 2012, 69, 92–95. [Google Scholar] [CrossRef]
  39. Yang, L.; Hu, C.G.; Wang, J.L.; Yang, Z.X.; Guo, Y.M.; Bai, Z.Y.; Wang, K. Facile synthesis of hollow palladium/copper alloyed nanocubes for formic acid oxidation. Chem. Commun. 2011, 47, 8581–8583. [Google Scholar] [CrossRef]
  40. Ji, Y.; Ghosh, K.; Shu, X.Z.; Li, B.Q.; Sokolov, J.C.; Prestwich, G.D.; Clark, R.A.F.; Rafailovich, M.H. Electrospun three-dimensional hyaluronic acid nanofibrous scaffolds. Biomaterials 2006, 27, 3782–3792. [Google Scholar] [CrossRef]
  41. Konwer, S.; Boruah, R.; Dolui, S.K. Studies on conducting polypyrrole/graphene oxide composites as supercapacitor electrode. J. Electron. Mater. 2011, 40, 2248. [Google Scholar] [CrossRef]
  42. Yang, S.D.; Dong, J.; Yao, Z.H.; Shen, C.M.; Shi, X.Z.; Tian, Y.; Lin, S.X.; Zhang, X.G. One-pot synthesis of graphene-supported monodisperse Pd nanoparticles as catalyst for formic acid electro-oxidation. Sci. Rep. 2014, 4, 4501. [Google Scholar] [CrossRef] [PubMed]
  43. Zhu, Q.L.; Song, F.Z.; Wang, Q.J.; Tsumori, N.; Himeda, Y.; Autrey, T.; Xu, Q. A solvent-switched in situ confinement approach for immobilizing highly-active ultrafine palladium nanoparticles: Boosting catalytic hydrogen evolution. J. Mater. Chem. A 2018, 6, 5544–5549. [Google Scholar] [CrossRef]
  44. Jin, Z.; Nackashi, D.; Lu, W.; Kittrell, C.; Tour, J.M. Decoration, migration, and aggregation of palladium nanoparticles on graphene sheets. Chem. Mater. 2010, 22, 5695–5699. [Google Scholar] [CrossRef]
  45. Dong, Z.P.; Le, X.D.; Liu, Y.S.; Dong, C.X.; Ma, J.T. Metal organic framework derived magnetic porous carbon composite supported gold and palladium nanoparticles as highly efficient and recyclable catalysts for reduction of 4-nitrophenol and hydrodechlorination of 4-chlorophenol. J. Mater. Chem. A 2014, 2, 18775–18785. [Google Scholar] [CrossRef]
  46. Li, H.; Han, L.; Cooper-White, J.; Kim, I. Palladium nanoparticles decorated carbon nanotubes: Facile synthesis and their applications as highly efficient catalysts for the reduction of 4-nitrophenol. Green Chem. 2012, 14, 586–591. [Google Scholar] [CrossRef]
  47. Le, X.; Dong, Z.; Liu, Y.; Jin, Z.; Huy, T.D.; Le, M.; Ma, J. Palladium nanoparticles immobilized on core–shell magnetic fibers as a highly efficient and recyclable heterogeneous catalyst for the reduction of 4-nitrophenol and Suzuki coupling reactions. J. Mater. Chem. A 2014, 2, 19696–19706. [Google Scholar] [CrossRef]
  48. Kim, K.; Kim, K.L.; Shin, K.S. Co-reduced Ag/Pd bimetallic nanoparticles: Surface enrichment of Pd revealed by Raman spectroscopy. J. Phys. Chem. C 2011, 115, 14844–14851. [Google Scholar] [CrossRef]
  49. Abay, A.K.; Kuo, D.H.; Chen, X.; Saragih, A.D. A new V-doped Bi2(O,S)3 oxysulfide catalyst for highly efficient catalytic reduction of 2-nitroaniline and organic dyes. Chemosphere 2017, 189, 21–31. [Google Scholar] [CrossRef]
  50. Jana, S.; Ghosh, S.K.; Nath, S.; Pande, S.; Praharaj, S.; Panigrahi, S.; Basu, S.; Endo, T.; Pal, T. Synthesis of silver nanoshell-coated cationic polystyrene beads: A solid phase catalyst for the reduction of 4-nitrophenol. Appl. Catal. A-Gen. 2006, 313, 41–48. [Google Scholar] [CrossRef]
  51. Jiao, T.F.; Zhao, H.; Zhou, J.X.; Zhang, Q.R.; Luo, X.N.; Hu, J.; Peng, Q.M.; Yan, X.H. Self-Assembly Reduced Graphene Oxide Nanosheet Hydrogel Fabrication by Anchorage of Chitosan/Silver and Its Potential Efficient Application toward Dyes Degradation for Wastewater Treatments. ACS Sustain. Chem. Eng. 2015, 3, 3130–3139. [Google Scholar] [CrossRef]
  52. Guo, R.; Jiao, T.F.; Li, R.F.; Chen, Y.; Guo, W.C.; Zhang, L.X.; Zhou, J.X.; Zhang, Q.R.; Peng, Q.M. Sandwiched Fe3O4/carboxylate graphene oxide nanostructures constructed by layer-by-layer assembly for highly efficient and magnetically recyclable dye removal. ACS Sustain. Chem. Eng. 2018, 6, 1279–1288. [Google Scholar] [CrossRef]
  53. Wang, C.R.; Sun, S.X.; Zhang, L.X.; Yin, J.J.; Jiao, T.F.; Zhang, L.; Xu, Y.L.; Zhou, J.X.; Peng, Q.M. Facile preparation and catalytic performance characterization of AuNPs-loaded hierarchical electrospun composite fibers by solvent vapor annealing treatment. Colloid Surf. A 2019, 561, 283–291. [Google Scholar] [CrossRef]
  54. He, Y.; Wang, R.; Jiao, T.F.; Yan, X.Y.; Wang, M.L.; Zhang, L.X.; Bai, Z.H.; Zhang, Q.; Peng, Q.M. Facile Preparation of Self-Assembled Layered Double Hydroxide-Based Composite Dye Films as New Chemical Gas Sensors. ACS Sustain. Chem. Eng. 2019, 7, 10888–10899. [Google Scholar] [CrossRef]
  55. Feng, Y.; Jiao, T.F.; Yin, J.J.; Zhang, L.; Zhang, L.X.; Zhou, J.X.; Peng, Q.M. Facile Preparation of Carbon Nanotube-Cu2O Nanocomposites as New Catalyst Materials for Reduction of p-Nitrophenol. Nanoscale Res. Lett. 2019, 14, 78. [Google Scholar] [CrossRef]
  56. Chen, K.Y.; Jiao, T.F.; Li, J.K.; Han, D.X.; Wang, R.; Tian, G.J.; Peng, Q.M. Chiral Nanostructured Composite Films via Solvent-Tuned Self-Assembly and Their Enantioselective Performances. Langmuir 2019, 35, 3337–3345. [Google Scholar] [CrossRef] [PubMed]
  57. Chen, K.Y.; Yan, X.Y.; Li, J.K.; Jiao, T.F.; Cai, C.; Zou, G.D.; Wang, R.; Wang., M.L.; Zhang, L.X.; Peng, Q.M. Preparation of Self-Assembled Composite Films Constructed by Chemically-Modified MXene and Dyes with Surface-enhanced Raman Scattering Characterization. Nanomaterials 2019, 9, 284. [Google Scholar] [CrossRef]
  58. Sun, S.X.; Wang, C.R.; Han, S.Q.; Jiao, T.F.; Wang, R.; Yin, J.J.; Li, Q.; Wang, Y.Q.; Geng, L.J.; Yu, X.D.; et al. Interfacial nanostructures and acidichromism behaviors in self-assembled terpyridine derivatives Langmuir-Blodgett films. Colloid Surf. A 2019, 564, 1–9. [Google Scholar] [CrossRef]
  59. Huang, X.X.; Jiao, T.F.; Liu, Q.Q.; Zhang, L.X.; Zhou, J.X.; Li, B.B.; Peng, Q.M. Hierarchical electrospun nanofibers treated by solvent vapor annealing as air filtration mat for high-efficiency PM2.5 capture. Sci. China Mater. 2019, 62, 423–436. [Google Scholar] [CrossRef]
  60. Wang, C.R.; Yin, J.J.; Wang, R.; Jiao, T.F.; Huang, H.M.; Zhou, J.X.; Zhang, L.X.; Peng, Q.M. Facile preparation of self-assembled polydopamine-modified electrospun fibers for highly effective removal of organic dyes. Nanomaterials 2019, 9, 116. [Google Scholar] [CrossRef]
  61. Guo, R.; Wang, R.; Yin, J.J.; Jiao, T.F.; Huang, H.M.; Zhao, X.M.; Zhang, L.X.; Li, Q.; Zhou, J.X.; Peng, Q.M. Fabrication and highly efficient dye removal characterization of beta-cyclodextrin-based composite polymer fibers by electrospinning. Nanomaterials 2019, 9, 127. [Google Scholar] [CrossRef] [PubMed]
  62. Huang, X.X.; Wang, R.; Jiao, T.F.; Zou, G.D.; Zhan, F.K.; Yin, J.J.; Zhang, L.X.; Zhou, J.X.; Peng, Q.M. Facile Preparation of Hierarchical AgNP-Loaded MXene/Fe3O4/Polymer Nanocomposites by Electrospinning with Enhanced Catalytic Performance for Wastewater Treatment. ACS Omega 2019, 4, 1897–1906. [Google Scholar] [CrossRef]
  63. Yin, Y.R.; Ma, N.; Xue, J.; Wang, G.Q.; Liu, S.B.; Li, H.L.; Guo, P.Z. Insights into the Role of Poly(vinylpyrrolidone) in the Synthesis of Palladium Nanoparticles and Their Electrocatalytic Properties. Langmuir 2019, 35, 787–795. [Google Scholar] [CrossRef] [PubMed]
  64. Liu, K.; Yuan, C.Q.; Zou, Q.L.; Xie, Z.C.; Yan, X.H. Self-Assembled Zinc/Cystine-Based Chloroplast Mimics Capable of Photoenzymatic Reactions for Sustainable Fuel Synthesis. Angew. Chem. Int. Ed. 2017, 56, 7876–7880. [Google Scholar] [CrossRef] [PubMed]
  65. Yan, X.Y.; Wang, M.L.; Sun, X.; Wang, Y.H.; Shi, G.H.; Ma, W.L.; Hou, P. Sandwich-like Ag@Cu@CW SERS substrate with tunable nanogaps and component based on the Plasmonic nanonodule structures for sensitive detection crystal violet and 4-aminothiophenol. Appl. Surf. Sci. 2019, 479, 879–886. [Google Scholar] [CrossRef]
  66. Ma, K.; Chen, W.; Jiao, T.; Jin, X.; Sang, Y.; Yang, D.; Zhou, J.; Liu, M.; Duan, P. Boosting Circularly Polarized Luminescence of Small Organic Molecules via Multi-dimensional Morphology Control. Chem. Sci. 2019. [Google Scholar] [CrossRef]
  67. Liu, J.Z.; Bauer, A.J.P.; Li, B.B. Solvent vapor annealing: An efficient approach for inscribing secondary nanostructures onto electrospun fibers. Macromol. Rapid Comm. 2014, 35, 1503–1508. [Google Scholar] [CrossRef]
  68. Nasrollahzadeh, M.; Sajadi, S.M.; Honarmand, E.; Maham, M. Preparation of palladium nanoparticles using Euphorbia thymifolia L. leaf extract and evaluation of catalytic activity in the ligand-free Stille and Hiyama cross-coupling reactions in water. New J. Chem. 2015, 39, 4745–4752. [Google Scholar] [CrossRef]
Figure 1. Preparation process of polyethyleneimine/polycaprolactone/Pd nanoparticles (PEI/PCL@PdNPs) composite and its catalytic performance.
Figure 1. Preparation process of polyethyleneimine/polycaprolactone/Pd nanoparticles (PEI/PCL@PdNPs) composite and its catalytic performance.
Catalysts 09 00559 g001
Figure 2. SEM images of AuNPs coated PEI/PCL electrospun fibers with different weight ratios ((w/w) (a), 10:90; (b), 15:85; (c), 20:80; (d), 25:75; (e), 30:70; (f), 35:65).
Figure 2. SEM images of AuNPs coated PEI/PCL electrospun fibers with different weight ratios ((w/w) (a), 10:90; (b), 15:85; (c), 20:80; (d), 25:75; (e), 30:70; (f), 35:65).
Catalysts 09 00559 g002
Figure 3. TEM images of the synthesized PdNPs (a), PEI/PCL fibers at 35:65 weight ratio (b) and modified with PdNPs (c), high resolution of PdNPs (d), and C/N/O/Pd elemental mapping (e) of PEI/PCL@PdNPs composites. The scale size of elemental mapping is 1.5 μm.
Figure 3. TEM images of the synthesized PdNPs (a), PEI/PCL fibers at 35:65 weight ratio (b) and modified with PdNPs (c), high resolution of PdNPs (d), and C/N/O/Pd elemental mapping (e) of PEI/PCL@PdNPs composites. The scale size of elemental mapping is 1.5 μm.
Catalysts 09 00559 g003
Figure 4. Scanning electron microscope (SEM) and EDS images of as-prepared PEI/PCL@PdNPs composite (a,b) and C/O/N/Pd elemental mapping (cf). The scale bar in iamge a represent 1 μm.
Figure 4. Scanning electron microscope (SEM) and EDS images of as-prepared PEI/PCL@PdNPs composite (a,b) and C/O/N/Pd elemental mapping (cf). The scale bar in iamge a represent 1 μm.
Catalysts 09 00559 g004
Figure 5. X-ray diffraction spectroscopy (XRD) pattern (A) with zoom-in region (B) of PEI/PCL fibers with the weight ratios of 10:90 (a) and 35:65 (b) and PEI/PCL@PdNPs composite (c).
Figure 5. X-ray diffraction spectroscopy (XRD) pattern (A) with zoom-in region (B) of PEI/PCL fibers with the weight ratios of 10:90 (a) and 35:65 (b) and PEI/PCL@PdNPs composite (c).
Catalysts 09 00559 g005
Figure 6. Thermogravimetry (TG) curves of the prepared PCL fibers and PEI/PCL fibers with the weight ratios of 35:65 and PEI/PCL@PdNPs composite.
Figure 6. Thermogravimetry (TG) curves of the prepared PCL fibers and PEI/PCL fibers with the weight ratios of 35:65 and PEI/PCL@PdNPs composite.
Catalysts 09 00559 g006
Figure 7. Survey X-ray photoelectron spectroscopy (XPS) spectra of the obtained PEI/PCL fiber and PEI/PCL@PdNPs (a) and deconvolutions of XPS peaks of the Pd3d region (b).
Figure 7. Survey X-ray photoelectron spectroscopy (XPS) spectra of the obtained PEI/PCL fiber and PEI/PCL@PdNPs (a) and deconvolutions of XPS peaks of the Pd3d region (b).
Catalysts 09 00559 g007
Figure 8. UV absorption curves of 4-NP and 2-NA before and after adding NaBH4 aqueous solution (a,d); catalytic reduction of 4-NP and 2-NA with PEI/PCL@PdNPs composite (PEI:PCL, w/w, 35:65) and photos of 4-NP and 2-NA solution after reduction (b,e); linear relationship of the reduction process (c,f).
Figure 8. UV absorption curves of 4-NP and 2-NA before and after adding NaBH4 aqueous solution (a,d); catalytic reduction of 4-NP and 2-NA with PEI/PCL@PdNPs composite (PEI:PCL, w/w, 35:65) and photos of 4-NP and 2-NA solution after reduction (b,e); linear relationship of the reduction process (c,f).
Catalysts 09 00559 g008
Figure 9. Recyclability test of PEI/PCL@PdNPs catalyst for the reduction of 4-NP (a) and 2-NA (b).
Figure 9. Recyclability test of PEI/PCL@PdNPs catalyst for the reduction of 4-NP (a) and 2-NA (b).
Catalysts 09 00559 g009
Table 1. Physical data of as-prepared different fiber composites.
Table 1. Physical data of as-prepared different fiber composites.
PEI/PCL Fibers Samples of Different Weight RatiosSpecific Surface Area (m2g−1)Average Pore Diameter (nm)Pore Volume (cm3g−1)
10:908.3423.830.00867
15:858.8425.310.00919
20:809.7327.860.01018
25:7511.1631.920.01144
30:7011.8233.560.01231
35:6512.9336.850.01322
Table 2. Comparative characteristics and catalytic performance of PdNPs-based materials.
Table 2. Comparative characteristics and catalytic performance of PdNPs-based materials.
NO.MaterialsKinetic Constant k/ s−1CharacteristicsRefs
1[email protected] NPs0.00059Natural bioactive black tea extract reducing agent17
2CN-supported PdNPs nanohybrids0.00570Cellulose nanocrystals as support matrix and reducing agent18
3AuPdNPs/ graphene nanosheets (GNs)0.01445Bimetallic nanoparticles, monodisperse method, graphene nanosheet carrier19
4nano-silica supported palladium nanocatalyst (Pd/KCC-1)0.00800Unique dendritic fibrous morphology led to poor aggregation of PdNPs20
5Pd/ magnetic porous carbon (MPC)0.01200Porous carbon composite catalyst carrier, magnetic separation45
6carbon nanotubes/ hyperbranched polymers/Pd (CNT/PiHP/Pd)0.02630Carbon nanotube assisted with hyperbranched polymer support46
7Pd/Fe3O4@SiO2@KCC-1(silica nanospheres) 0.01960Core-shell magnetic fibrous nanocatalyst47
8Polyethyleneimine(PEI) /Polycaprolactone (PCL)@PdNPs0.16597Electrospun fibers carrier, simple methodPresent work

Share and Cite

MDPI and ACS Style

Wang, C.; Yin, J.; Han, S.; Jiao, T.; Bai, Z.; Zhou, J.; Zhang, L.; Peng, Q. Preparation of Palladium Nanoparticles Decorated Polyethyleneimine/Polycaprolactone Composite Fibers Constructed by Electrospinning with Highly Efficient and Recyclable Catalytic Performances. Catalysts 2019, 9, 559. https://doi.org/10.3390/catal9060559

AMA Style

Wang C, Yin J, Han S, Jiao T, Bai Z, Zhou J, Zhang L, Peng Q. Preparation of Palladium Nanoparticles Decorated Polyethyleneimine/Polycaprolactone Composite Fibers Constructed by Electrospinning with Highly Efficient and Recyclable Catalytic Performances. Catalysts. 2019; 9(6):559. https://doi.org/10.3390/catal9060559

Chicago/Turabian Style

Wang, Cuiru, Juanjuan Yin, Shiqi Han, Tifeng Jiao, Zhenhua Bai, Jingxin Zhou, Lexin Zhang, and Qiuming Peng. 2019. "Preparation of Palladium Nanoparticles Decorated Polyethyleneimine/Polycaprolactone Composite Fibers Constructed by Electrospinning with Highly Efficient and Recyclable Catalytic Performances" Catalysts 9, no. 6: 559. https://doi.org/10.3390/catal9060559

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