Feng Zhao,
Weixiao Kong,
Zonggao Hu,
Jindun Liu,
Yafei Zhao* and
Bing Zhang*
School of Chemical Engineering and Energy, Zhengzhou University, Zhengzhou 450001, P. R. China. E-mail: zhaoyafei007@126.com; zhangb@zzu.edu.cn
First published on 15th August 2016
An environmentally benign and economic reaction system with an effective catalyst for 4-nitrophenol reduction is highly desirable. Here, we synthesized reduced graphene oxide (RGO) supported Pt–Ni alloy catalysts with different atomic ratios of Pt and Ni, investigated their morphology, size, dispersity, structure and elemental valence, and studied their catalytic activity in order to tune their performance for 4-nitrophenol reduction. It is worth pointing out that the RGO support can efficiently avoid the aggregation of Pt–Ni alloy nanoparticles, and the most dispersed and smallest Pt–Ni particles on RGO can be obtained when the atomic ratio of Pt to Ni is 1:
9. The Pt–Ni/RGO (1
:
9) nanocatalyst also shows a higher catalytic activity toward the conversion of 4-NP to 4-AP with a catalytic rate constant of 0.3700 min−1 than Pt–Ni/RGO (1
:
3) and Pt–Ni/RGO (1
:
25), and much higher than that of Pt/RGO, Ni/RGO and bare Pt–Ni owing to the well-defined composition, small particle size (10 nm), good dispersion, synergistic effect between Pt and Ni, and electron transfer between RGO and Pt–Ni alloy nanoparticles. In addition, the catalyst possesses good stability and recyclability for the catalytic reduction reaction. The Pt–Ni/RGO nanocatalyst, with well-defined composition, small particle size, uniform dispersity, high catalytic rate, and recyclability, should be an ideal catalyst for specific applications in liquid phase reactions.
Graphene, consisting of a single atomic layer of sp2 carbon atoms,16 has many unique properties such as high specific surface area,17 charge carrier mobility,18 good electrical conductivity,19 high mechanical strength and special chemical properties,20 making it an ideal support for growth of functional nanoparticles, which shows great potential for diverse applications in catalysis,21,22 energy storage devices23,24 and sensing.25 As graphene could not only stabilize metals but also facilitate electron transfer between graphene and metals in catalytic reaction process, the catalytic performance of the nanocatalysts could be highly improved.26 However, graphene sheets are unstable with respect to scrolling or restacking and tend to form agglomerations through van der Waals interactions,27 and their poor dispersity in water solution originated from hydrophobic surface leads to weak interaction between metals and graphene.28 As is well known, graphene sheets can be produced by solution-based chemical reduction of graphene oxide (GO), known as reduced graphene oxide (RGO). Unless graphene, the hydrophilicity of GO allows it to be readily uniformly dispersed in water and GO possesses abundant oxygen-containing functional groups and a large surface area,29 which provides chemically active sites available for nanoparticles anchoring and dispersion.27,30 Nevertheless, GO exhibits a significant loss of conductivity, which limits electron transfer between metal nanoparticle and support and thus leads to low catalytic activity. It needs to be reduced to restore the sp2 hybrid network (RGO) to reintroduce the conductive property.31 Although intensive effort has focused on graphene- or graphene oxide-based materials, the goal of direct growth and anchoring bimetals on their surfaces with high dispersity, uniform distribution, improved catalytic activity and good recyclability is still full of challenges.32,33
Herein, we report a facile and simple in situ co-reduction approach for one-pot synthesis of Pt–Ni bimetals supported on RGO with different atomic ratios using GO as intermediate support, in which metal ions and GO were reduced simultaneously by NaBH4, for catalytic hydrogenation of 4-NP. The oxygen functional groups on GO are in favor of metal ions anchoring through electrostatic adsorption, and after co-reduction, RGO avoids the aggregation of Pt–Ni nanoparticles and facilitates the reactant adsorption and electron transfer to the well dispersed Pt–Ni nanoparticles catalyst.32 The as-synthesized Pt–Ni/RGO nanocatalyst with atomic ratio of Pt and Ni 1:
9 displays better catalyst activity and good stability for the reduction of 4-nitrophenol than other Pt–Ni/RGO nanocatalysts, monometallic Pt/RGO, Ni/RGO and bare Pt–Ni, owing to the well-defined composition, small particle size, good dispersion, synergistic effect between Pt and Ni, and electron transfer between RGO and Pt–Ni alloy nanoparticles.
To investigate the morphology and microstructure of the as-synthesized nanocatalysts, Pt–Ni/RGO nanocatalysts and bare Pt–Ni were characterized by TEM and HRTEM (Fig. 2). As seen from Fig. 2a–c, the RGO nanosheets are clearly distinguished from the background and are transparent monolayer with a few wrinkles, which is a characteristic feature of RGO sheets. Meanwhile, the Pt–Ni nanoparticles are homogeneously and uniformly anchored on the surface of RGO nanosheets and well separated from each other although some of them tend to stack together. Notably, the sizes of the Pt–Ni nanoparticles on RGO nanosheets strongly depend on atomic ratios of Pt and Ni (inset in Fig. 2a–c). Pt–Ni/RGO (1:
9) has the smallest average particle size of about 10 nm and narrower size distribution than Pt–Ni/RGO (1
:
3) of ∼21 nm and Pt–Ni/RGO (1
:
25) of ∼16 nm, which means that Pt–Ni particle size decreases at first but increases later with increasing of Ni content. Similar result was also reported by Zhang et al.36 Particle size plays an important role on the catalytic activity of the catalysts. A decrease of metal particle sizes leads to an increase in their surface area, edges on their surface and corner atoms, which can improve the catalytic properties. For comparison, TEM image of the bare Pt–Ni (1
:
9) prepared in the absence of GO is shown in Fig. 2d. The bare Pt–Ni (1
:
9) nanoparticles without graphene are aggregated severely and assembled into chain-like particles with an average particle size of about 120 nm, which is much bigger than the Pt–Ni nanoparticles in Pt–Ni/RGO (1
:
9) nanocatalysts (Fig. 2b). This result indicates that GO nanosheets have significant influence on the size of the Pt–Ni nanoparticles. GO can serve as an effective dispersing agent which can anchor Pt–Ni nanoparticles and thus control their sizes and distribution on graphene during the synthetic process. In addition, the lattice spacing of the particle of the Pt–Ni/RGO (1
:
9) nanocatalyst was measured to be 0.215 nm from a high resolution of TEM image (the inset in the top of Fig. 2b). The value is smaller than that of the (111) plane of Pt (0.227 nm, JCPDS-04-0802) but larger than that of the (111) plane of Ni (0.203 nm, JCPDS-46-0850), indicating that the Pt–Ni nanoparticles are in crystalline alloy state.30 The molar ratio and actual total loading content of Pt and Ni in the Pt–Ni/RGO nanocatalysts were further determined by ICP-MS, as shown in Table 1. The results are in good agreement with the theoretical molar ratio and total amount of Pt and Ni (40%), meaning that most of the Pt and Ni were loaded on the surface of RGO.
Catalysts | Pt loading (wt%) | Ni loading (wt%) | Pt![]() ![]() |
PtNi loading (wt%) |
---|---|---|---|---|
Pt–Ni/RGO (1![]() ![]() |
20.23 | 17.36 | 1![]() ![]() |
37.59 |
Pt–Ni/RGO (1![]() ![]() |
10.45 | 27.48 | 1![]() ![]() |
37.93 |
Pt–Ni/RGO (1![]() ![]() |
4.55 | 33.54 | 1![]() ![]() |
38.09 |
Fig. 3 shows the FTIR spectra of the as-synthesized GO and Pt–Ni/RGO (1:
9) nanocatalyst. In the spectrum of GO, the broad and intense band around 3410 cm−1 and the band at 1623 cm−1 can be attributed to the –OH stretching vibration due to the surface adsorbed water molecules, while the peak at 1733 cm−1 is assigned to the C
O stretching vibration of the carboxylic groups. The peaks at 1224 cm−1 and 1053 cm−1 correspond to the stretching vibration peaks of C–O (epoxy) and C–O (alkoxy), respectively. The results clearly confirm that oxygen-containing groups were successfully bound to the edges of the graphene nanoplates through overoxidation. After the reduction, the intensity of the band at 3410 cm−1 for Pt–Ni/RGO (1
:
9) composite decreased dramatically, suggesting the subsequent removal of surface adsorbed water molecules during the reduction process.37 Additionally, all the other peaks at 1733, 1623, 1224 and 1053 cm−1 corresponding to the oxygen functionalities disappeared and new peaks at 1636, 1560, 1390, 1128 cm−1 occurred, confirming the formation of the Pt–Ni nanoparticles and RGO from GO. The results also demonstrate that the nanoparticles attach to the graphene support via the oxygen functional groups, and the interactions were strong enough to ensure the nanoparticles remained attached even after chemical cleaning and ultrasonication.38
The reduction of GO to RGO in the Pt–Ni/RGO nanocatalyst is further verified by Raman spectroscopy. As shown in Fig. 4, both the GO and Pt–Ni/RGO (1:
9) display two prominent peaks of the D and G bands, which are associated with disorder features induced by lattice defect and vibrations of the graphite sp2 carbon domains respectively. The intensity ratio (ID/IG) of the D to the G band correlates with the average size of sp2 domains.30 The D and G bands of GO are located at 1348 and 1603 cm−1 with ID/IG ratio of 1.02. While in the case of Pt–Ni/RGO (1
:
9), the two bands move to lower wavenumber of 1345 and 1589 cm−1. Besides, the ID/IG ratio increases from 1.02 for GO to 1.38 for Pt–Ni/RGO (1
:
9) because the average size of in-plane sp2 domains become smaller when GO is effectively reduced to RGO.39 The results indicate that the GO has been well deoxygenated and reduced in Pt–Ni/RGO (1
:
9) nanocatalyst.
In order to further ascertain the formation of Pt–Ni/RGO nanocatalysts, the XRD of the as-synthesized GO, Pt/RGO and Pt–Ni/RGO was performed as shown in Fig. 5. The characteristic diffraction peak of GO at around 2θ = 10.2° corresponding to (001) reflection disappeared and no new peak occurred at 2θ = 22.5° corresponds with the (002) reflection of graphene in the Pt/RGO and Pt–Ni/RGO nanocatalysts, indicating that the GO nanosheets were reduced and the restacking and aggregation of RGO sheets has been prevented by the supported metal alloy nanoparticles.35 The Pt/RGO has four characteristic diffraction peaks at 40.09°, 46.59°, 67.87° and 81.51° corresponding to the (111), (200), (220) and (311) planes of face-centered-cubic crystalline Pt nanoparticles (JCPDS 04-0802). In the case of Pt–Ni/RGO nanocatalysts, the diffraction peaks continuously shift to higher angles with elevated loading of Ni content in contrast with Pt/RGO, but no characteristic peaks of Ni or its oxides were detected. This shift reveals that Ni atoms have entered into Pt lattice and the formation of Pt–Ni alloy, which is consistent with the observation result of TEM.40 Similar shift can also be observed by Xu et al. on SiO2 supported Pt–Ni alloy and Pt–Ni/C nanocrystallites.41,42 The diffraction peak intensity decreases with the increasing of Ni content and even disappears when the atomic ratio of Pt to Ni is 1:
25, indicating that excess Ni in the Pt–Ni alloys will lead to weaker crystallinity.35
![]() | ||
Fig. 5 XRD patterns of GO, Pt/RGO and Pt–Ni/RGO nanocatalysts with Pt–Ni atomic ratios of 1![]() ![]() ![]() ![]() ![]() ![]() |
The surface composition and the electronic properties of Pt–Ni on Pt–Ni/RGO nanocatalyst were further analyzed by X-ray photoelectron spectroscopy (XPS). Fig. 6a and b shows the Pt 4f and Ni 2p XPS spectra of the Pt–Ni/RGO (1:
9) nanocatalyst, respectively. In Fig. 6a, the diffraction peak at the binding energy of 68.4 eV can be assigned to the XPS peak position of Ni 3p according to the report by Ma et al.43 The two intense peaks located at 71.37 eV (Pt 4f7/2) and 74.79 eV (Pt 4f5/2) could be assigned to metallic Pt0 species, and two more weak peaks detected at 72.91 eV and 76.83 eV are attributed to Pt2+ species in PtO and Pt(OH)2. A comparison of the relative intensities of the peaks due to Pt0 and Pt2+ indicated that Pt in Pt–Ni/RGO was predominately metallic although there exists Pt2+ species in Pt–Ni/RGO nanocatalyst.44 In Fig. 6b, the peaks located at 855.80 eV and 873.50 eV are assigned to elemental Ni2+ 2p3/2 and 2p1/2 of Ni(OH)2, respectively. The other two peaks at 856.90 eV and 874.80 eV could be assigned to NiOOH. This indicates that Ni0 on the surface of the nanocatalyst is low irrespective, which because bulk Ni is easily oxidized in the atmosphere. Similar results can also be observed in previous report.30 Additionally, there exists a characteristic intense shake up satellite signal adjacent to the main peaks at 861.42 eV and 879.99 eV, which is due to the multielectron excitation.44 Comparison of Pt–Ni/RGO and Pt/RGO nanocatalysts as shown in Fig. 6c, it can be observed that a lower binding energy shift in Pt0 4f7/2 peak of the Pt–Ni/RGO nanocatalysts in comparison to that of Pt/RGO, which signifies the electronic modification of Pt in the Pt–Ni alloy. The electrons transfer from Ni to Pt due to the electronegative difference between Ni (1.91) and Pt (2.28) has lowered the density of state on the Fermi level and thus modify the electronic properties of Pt, which is expected to enhance the catalytic activity.32 For instance, it has been reported by Park et al. that a negative binding energy shift is responsible for an electron donation of Ni to Pt, resulting in a change in the electronic properties of the Pt 4f peaks and improving the electrocatalytic activity of Pt–Ni alloys.45
![]() | ||
Fig. 6 XPS spectra of (a) Pt 4f, (b) Ni 2p of the Pt–Ni/RGO (1![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
![]() | ||
Fig. 7 Time-dependent UV-vis absorption spectra of the reduction of 4-nitrophenol by NaBH4 catalyzed by Pt–Ni/RGO (1![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
To further compare the catalytic properties of different nanocatalysts and investigate the mechanism, a pseudo-first-order kinetics equation is applied to calculate the rate constant (k) since excess NaBH4 was used in this study, which is given as follows:32
kt = ln(Ct/C0) = ln(At/A0) |
Fig. 8 shows the plots of ln(At/A0) vs. reaction time of the as-synthesized nanocatalysts. It can be clearly seen that all the plots fit well with the pseudo-first-order kinetics model with R2 values larger than 0.99. The turnover frequency (TOF) of the catalyst, which is the number of reactant molecules that 1 g of catalyst can convert into products with one second, is used to compare catalyst efficiencies. The rate constant and TOF of different catalysts are listed in Table 2. It is noted that the rate constants and TOF values for Pt–Ni/RGO nanocatalysts are higher than those of monometallic Pt/RGO and Ni/RGO, indicative of higher catalytic activity of Pt–Ni/RGO nanocatalysts. This may be due to the synergetic chemical coupling effects of Pt–Ni alloy. According to the traditional theory about the catalytic reduction of 4-NP to 4-AP, BH4− would adsorb onto the catalyst surface and donate electrons to noble metal nanoparticles, and then transfer to the 4-NP adhered to catalysts surfaces via chemical adsorption. As a result, when Ni atoms enter into Pt lattice in Pt–Ni alloy, it can enhance and transfer electrons to Pt, facilitating the transfer process of electrons to the substrate and thus enhancing the catalytic effect.32 In view of this, the rate constant can be improved by increasing the content of Ni (Pt–Ni/RGO (1:
9) > Pt–Ni/RGO (1
:
3)). However, it doesn't continually increase (Pt–Ni/RGO (1
:
25) < Pt–Ni/RGO (1
:
9)), because fewer metallic Pt atoms are available on the catalysts surface for the reduction of 4-NP.12 The apparent activation energy and pre-exponential factor of the Pt–Ni/RGO nanocatalysts were also estimated based on the linear correlation between ln
k and the 1/T (as shown in Fig. S1†) which follows the Arrhenius equation: ln
k = ln
A − Ea/RT.46 And their values are listed in Table S1.† The apparent activation energy of the Pt–Ni/RGO (1
:
9) was measured to be 29.1 kJ mol−1, which is lower compared to Pt–Ni/RGO (1
:
3) (34.1 kJ mol−1) and Pt–Ni/RGO (1
:
25) (44.3 kJ mol−1). The results are in well agreement with the order of the rate constants of the Pt–Ni/RGO nanocatalysts, which further indicates the optimum molar ratio of Pt–Ni is 1
:
9. In addition, researchers also found that nanoparticles with the smaller size and better dispersion may have greatly enhanced catalytic properties owing to their higher surface area, more edges and corner atoms, since heterogeneous catalysis takes place only on the surface of the nanoparticles.47 Based on the above, the Pt–Ni/RGO (1
:
9) nanocatalyst which has the optimum Pt–Ni atomic ratio and the smallest particle size as illustrated in TEM image (Fig. 2) shows the best catalytic performance. Moreover, its rate constant of 0.3700 min−1 is about 70 times higher than that of bare Pt–Ni (1
:
9) of 0.0054 min−1, which indicates that the enlarged surface area and high conductivity of RGO can not only facilitate the reactant adsorption but also make the Pt–Ni nanoparticles well dispersed and transfer electron to them, thereby dramatically improving the catalytic effect.30
![]() | ||
Fig. 8 Plot of ln(At/A0) versus time for the reduction of 4-NP catalyzed by (a) Pt–Ni/RGO (1![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
Catalysts | Pt–Ni/RGO (1![]() ![]() |
Pt–Ni/RGO (1![]() ![]() |
Pt–Ni/RGO (1![]() ![]() |
Pt/RGO | Ni/RGO | Bare Pt–Ni (1![]() ![]() |
k (min−1) | 0.1890 | 0.3700 | 0.0850 | 0.0575 | 0.0729 | 0.0054 |
R2 | 0.9973 | 0.9907 | 0.9926 | 0.9903 | 0.9905 | 0.9908 |
TOF(×1017) s−1 | 49.8 | 110.9 | 28.8 | 17.5 | 22.2 | 4.4 |
For the consideration of practical application, the recyclability of Pt–Ni/RGO (1:
9) nanocatalyst was examined for four times. As shown in Fig. 9, the conversion of 4-NP to 4-AP maintains higher than 86.27%, showing that there is no obvious loss of catalytic activity after each recycling. This result also indicates that the as-prepared Pt–Ni/RGO (1
:
9) nanocatalyst has a high stability in the reduction of 4-NP. Therefore, the recyclability of Pt–Ni/RGO (1
:
9) makes it a promising candidate for the catalytic reduction of 4-AP to 4-NP.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra16045j |
This journal is © The Royal Society of Chemistry 2016 |