Rahul Krishnaa,
Diana M. Fernandesb,
Valdemar F. Domingosc,
Edivagner S. Ribeirod,
João Campos Gild,
Catarina Diase,
João Venturae,
Cristina Freireb and
Elby Titus*a
aCentre for Mechanical Technology and Automation (TEMA), Department of Mechanical Engineering, University of Aveiro, 3810-193, Portugal. E-mail: elby@ua.pt; Fax: +351-234370953; Tel: +351-234370830
bREQUIMTE/LAQV, Department of Chemistry and Biochemistry, Faculty of Sciences, University of Porto, 4169-007 Porto, Portugal
cLIP-Coimbra, University of Coimbra, Coimbra 3004-516, Portugal
dCFisUC, Department of Physics, University of Coimbra, Coimbra 3004-516, Portugal
eIFIMUP, IN – Institute of Nanoscience and Nanotechnology, Department of Physics and Astronomy, Faculty of Sciences, University of Porto, 4169-007 Porto, Portugal
First published on 7th July 2015
A nanocomposite catalyst containing palladium–nickel boride–silica and reduced graphene oxide (Pd@NixB–SiO2/RGO, abbreviated as Pd@NSG) was successfully fabricated and its enhanced hydrogen spillover mechanism and high catalytic performance towards reduction of 4-nitrophenol (4-NP) to 4-aminophenol (4-AP) is discussed. The structure, composition and morphology of the Pd@NSG nanocomposite were characterized by various techniques. The H2 adsorption experiment directly reveals the spillover effect on the Pd@NSG nanocomposite and its enhanced H2 uptake capacity (0.7 wt%) compared to SiO2/RGO (0.05 wt%) under 50 bar pressure at RT. 4-NP reduction reaction shows remarkably high activity (120 s) of Pd@NSG compared to NixB–SiO2/RGO (7200 s) with excellent stability up to 5 cycles. Both the experiments showed the facile H2 dissociation on the Pd (active sites) activator and subsequent transportation of hydrogen atoms on receptor sites.
Attractive results were also obtained by Wu et al. for methylene blue (MB) dye reduction using the Pt–Fe3O4/RGO nanocomposite.12 Qin et al. also successfully demonstrated the recyclable catalytic performance of the PdNi@Pd/RGO nanocomposite for hydrogen generation via formic acid decomposition.13 On the other hand, 4-nitrophenol (4-NP) is a pollutant and an effluent in many chemicals and drug industries and therefore its efficient quick degradation is required.14 Nevertheless, its reduction is only possible in the presence of a catalyst.15 For this hydrogenation process, several kinds of catalysts are still being investigated and several examples can be found in the scientific literature.14,16–19 A recent work demonstrated the successful reduction of 4-NP using Pd/RGO nanocomposite in a very short time (120 s).20 Ji et al. also reported the reduction of 4-NP by Ni/RGO.21 Although, the reduction process of 4-NP with Ni/RGO was too slow compared to Pd/RGO. The fast reduction in Pd was evident due to the high spillover capability and its noble behaviour.13,20,22 Considering the high cost of Pd and slow reduction process using Ni/RGO, it is crucial to develop new materials for this catalysis reaction with retention of catalytic activity and low price.
The synthesis of finely divided Ni (without surfactant) is critical due to its magnetic behaviour (agglomeration of nanoparticles) and the requirement of harsh reaction conditions.23 For example, to synthesise the Ni nanoparticles, the occasional involvement of toxic chemicals (hydrazine hydrate) and high-temperature reaction conditions (to decompose the organometallic precursors) disfavour the process easiness.24–27 Furthermore, NixB can be easily synthesised by the simple reduction of Ni2+ ions by sodium borohydride (NaBH4).28 NixB is a very active hydrogenation catalyst and a hydrogen generator (from aqueous solution of NaBH4) and this type of materials is attractive for the H2 spillover mechanism.29–31
Similarly, H2 is the most promising energy fuel for automobiles and small portable devices (e.g. mobile phones and laptops) due to its light weight, high energy density and clean combustion.13,32 Yet, significant challenges hinder its widespread application as a choice of fuel due to the lack of a safe and easy method of its storage.33–35 Very recently, Li et al. reported the H2 storage capacity of Pd crystal and Pd loaded HKUST at 303 K.36 However, specific synthesis protocol of porous material HKUST and overall dependency on precious metals was a major drawback. Therefore, it is crucial to develop new materials for efficient H2 storage as well. Considering the increasing need for efficient H2 storage (Scheme 1) and the fact that Pd and NixB are well-known materials for hydrogen spillover, we report a stepwise synthesis of Pd@NSG nanocomposite and systematic investigation of hydrogen spillover on Pd@NSG nanocomposite. For this, we have synthesized (i) RGO–SiO2 nanocomposite then (ii) NixB–SiO2/RGO and finally, (iii) Pd@NixB–SiO2/RGO (Pd@NSG). Moreover, to synthesize the NixB species on SiO2/RGO nanocomposite, these steps are crucial due to high redox potential and hydrogen spillover capability of Pd compared to Ni, the entity NixB can only form in the absence of Pd. Whereas, in one step synthesis there was chance of formation of Pd–Ni alloy instead of Pd@NixB entity. This is an economically viable method since it uses, as major components, low cost materials such as NixB, SiO2 and RGO and a small percentage of Pd as an activator. Moreover, in this work, the catalytic activity of Pd@NSG nanocomposite was investigated using the reduction of 4-NP in an aqueous medium in the presence of NaBH4. This is a standard hydrogenation reaction to evaluate the catalytic activity of nanocatalysts, owing to its ease of execution.14 The H2 spillover effect was confirmed through the high volumetric adsorption of the gaseous H2 in Pd@NSG nanocomposite and also the disappearance of the phenolate absorbance peak at λ ≈ 400 nm during the progress of 4-NP reduction to 4-AP.
![]() | ||
Scheme 1 Schematic illustration of H2 spillover on Pd and NixB@SiO2 (left) and adsorption of H2 in Pd@NSG nanocomposite (right). |
The XRD spectra of Pd@NSG shows some additional peaks that are also related to the various crystallographic diffraction planes of Pd/PdO phase. The major diffraction peaks at 39.9° and 46.4° can be assigned as diffraction from (111) and (200), respectively, planes of face-centered cubic (fcc) crystal lattice structure of Pd.39 The other two peaks at 34.29° and 59.9° indicates the reflection from PdO (due to surface oxidation of Pd).40 In addition, in order to elucidate the structure of Pd@NSG nanocomposite and reaction mechanism in more detail, we have also carried out the FTIR spectroscopy. Fig. 1(b) shows the FTIR spectra of GO, NixB–SiO2/RGO and Pd@NSG nanocomposite (in the range of 4000–800 cm−1). In FTIR spectra of GO, a strong broad band was observed in the high frequency area (3400–3200 cm−1) which is assigned to the vibration stretching mode of –OH groups due to the surface adsorbed water molecules. After the reduction, the intensity of this band was continuously decreased from NixB–SiO2/RGO to Pd@NSG nanocomposites which suggests the subsequent removal of surface adsorbed water molecules during the reduction process.41 Moreover, the peak related to the vibration stretching mode of carbonyl functionality also became deprived in NixB–SiO2/RGO to Pd@NSG nanocomposites indicating the elimination of edge related –CO groups and formation of GO to RGO.14 Finally, the absorption peaks at 1385 cm−1 (stretching vibration of C–O of carboxylic acid) and 1110 cm−1 (C–OH of alcohol) were also efficiently reduced in both samples compared to the GO.41 However, in NixB–SiO2/RGO and Pd@NSG nanocomposites a new peak was observed at 1008 cm−1 which indicates the incorporation of SiO2 nanoparticles.42 Then, to ascertain the change in carbon system Raman spectroscopy was also performed. This technique is very useful for assigning the corresponding changes of graphene material on the basis of peak position and intensity. Fig. 1(c) shows the Raman spectra of GO, NixB–SiO2/RGO and Pd@NSG nanocomposite in the range of 1050–1900 cm−1. The Raman spectrum of GO displays the two characteristic D and G bands at 1353 and 1598 cm−1, respectively, with an ID/IG ratio of 0.96. It has already been reported that the G band is an intrinsic feature of graphene and closely related to the vibrations in all sp2 carbon materials.14 The D band becomes prominent when defects are introduced in graphene and in GO it is activated due to the reduction in size of the in-plane sp2 domains due to the attachment of various functionalities in edge and basal plane sites.41 In NixB–SiO2/RGO and Pd@NSG nanocomposite spectra, these two prominent bands (D and G) were shifted to lower wave numbers and are located at 1346 and 1596 and 1343 and 1594 cm−1, respectively. In both spectra ID/IG ratio was increased compared to GO. The continuous increment of ID/IG ratio from GO to NixB–SiO2/RGO and Pd@NSG nanocomposite can be clearly observed in Fig. 1(d) suggested the clear change in carbon system due to the incorporation of some extra defects in graphene.38
We have also performed the XPS analysis to identify the degree of reduction from GO to RGO and investigated the oxidation states of Pd, Ni, B and Si entities in Pd@NSG nanocomposite. The C 1s XPS spectrum of GO shows two large broad peaks that are deconvoluted into four peaks at approximately 284.3, 285.2, 287.2 and 288.9 eV (see Fig. S1 ESI†). The peaks at 284.3 and 285.2 eV are attributed to the sp2 CC and sp3 C–C bonding, respectively.14 The peak at around 287.2 eV is assigned to the binding energies of carbon in C–O and C
O and that at 288.9 eV to carbon in COOH groups.43
However, the C 1s core level spectrum of Pd@NSG (Fig. 2(a)) shows one intense peak and another less intense at higher binding energies. These peaks are deconvoluted into five peaks at approximately 285.0, 286.3, 287.6, 289.2 and 290.3 eV. The peak at 285.0 is attributed to sp2 C–C,44 that at 286.3 eV to C–N, the ones at 287.6 and 289.2 eV to binding energies of carbon in C–O (hydroxyl and epoxy) and CO, respectively. Similarly the peak at around 290.3 eV corresponds to carbon in COOH groups.43,45,46 Normally, the peak intensity due to C–O and C
O is very high in GO. But our results show the less intense peak, which is proportionate to the reduction of GO to RGO during the metal NPs (Ni and Pd) deposition.43,47 During this process there was an easy spillover of hydrogen gas on metal NPs which further reduces the oxygen functionalities of GO. The deconvoluted spectra of O 1s, B 1s, N 1s, Si 2p and Ni 2p are shown in Fig. S2 (see ESI†). In Table S1 (see ESI†) are summarized the XPS data obtained for Pd@NSG. The XPS spectrum of Pd 3d is shown in Fig. 2(b). The binding energies of Pd 3d can be resolved into 3d5/2 and 3d3/2 doublets caused by spin-orbital coupling. Upon deconvolution of the spectra, the curves are fitted with two pairs of binding energies for Pd0 and PdII at 335.8 eV, 341.1 eV and 337.4 eV, 342.7 eV, respectively.48 The percentage of PdII relatively to Pd0 is 12.5%. The O 1s core level spectrum of Pd@NSG (Fig. S2a, ESI†) is deconvoluted into 2 main peaks approximately at 531.8 and 533.5 eV assigned to O
C–OH and O–C or C–OH, respectively.46,49 The Ni 2p spectrum is also deconvoluted into 2p3/2 and 2p1/2 doublets caused by spin-orbital coupling. The peak of metallic nickel is observed at approximately 853.5 eV, while those at 857.0 and 874.4 eV and 862.2 and 880.6 eV are attributed to NiO and NiOOH, respectively (see Fig. S2e, ESI†).50,51
The existence of B species is confirmed by the presence of sharp peak at 193 eV in B 1s XPS core spectrum.52,53 Systematic microscopic investigations of initial material GO, intermediate composite material RGO–SiO2 (after the loading of synthesized SiO2 nanoparticles within the RGO matrix), NixB–SiO2/RGO nanocomposite and final product Pd@NSG nanocomposite were also carried out. Fig. 3(a) and (b) show the SEM images of GO (after the 2 h exfoliation in methanol at RT) at higher and lower magnifications, respectively. Images of GO clearly exhibit the presence of few layers of graphitic carbon with typical wrinkle behaviour.38 Fig. 3(c) and (d) display the initial morphology and size of the SiO2 nanoparticles with graphene sheet and show that all particles were small sized and with spherical shape. Moreover, images shows that SiO2 nanoparticles are covered with carbon sheets and well separated without any specific agglomerations.54 Similar behaviour was observed after the intercalation of NixB nanoparticles on SiO2/RGO as observed in Fig. 3(e) and (f). Additionally, these two images show higher density of nanoparticles compared to previous images of SiO2/RGO indicating the successful formation of NixB–SiO2/RGO nanocomposite. Fig. 4(a)–(c) shows the SEM images of the final product Pd@NSG at different magnifications. Fig. 4(a) and (b), at lower magnification, show the intercalation of tiny Pd nanoparticles on NixB–SiO2/RGO nanocomposite at a superficial position and Fig. 4(c) depicts the arrangement of the nanoparticles in a cone-type structure. This arrangement may be due to the grafting of small Pd nanoparticles on high defect sites of NixB–SiO2/RGO nanocomposite and the filling of cavities by them. Moreover, to confirm the presence of Ni and Pd elements, we also have carried out the EDX analysis.
![]() | ||
Fig. 3 SEM images of: GO (a) and (b), SiO2/RGO (c) and (d) and NixB–SiO2/RGO nanocomposites (e) and (f) in higher and lower magnifications, respectively. |
![]() | ||
Fig. 4 SEM images of Pd@NSG nanocomposite at different magnifications: 6000× (a) 15000× (b) and 40000× (c); (d) EDX spectra of Pd@NSG. |
Fig. 4(d) displays the EDX spectrum of Pd@NSG nanocomposite and results show the presence of all elements: Pd, Ni, Si, C and O with the exception of the light weight elements B and H. TEM analysis was also performed to confirm the shape, morphology and internal structure of Pd@NSG nanocomposite for detailed investigation of final product. Fig. 5(a) and (b) show the TEM images at lower and higher magnification, respectively. The images clearly displays the homogeneous distribution of nanoparticles on the graphene sheets; the higher magnification image shows that all particles have spherical shape without any specific kind of agglomeration of nanoparticles.55 Moreover, to find the crystallinity in Pd@NSG nanocomposite we have performed the selected area electron diffraction (SAED) analysis (see Fig. S3 ESI†) which clearly shows the (111) lattice fringe of Pd NPs along with bright spots of graphene layers.55 Furthermore, to establish the detailed mechanism of reduction of Pd2+ ions to Pd(0) in presence of anhydrous methanol (CH3OH),56 we have also carried out the HRTEM analysis of bare Pd NPs which clearly shows the formation of monodisperse small spherical Pd NPs in the range of (2.5–4 nm) without any agglomeration (see Fig. S4 ESI† for detail). Here, we used Pd(OAc)2 precursor and methanol as solvent and reducing agent. At moderate temperature (45 °C) and under sonication condition methanol provided the reducing species H2 as per the following eqn (1):
2CH3OH = 2HCHO + H2 | (1) |
![]() | ||
Fig. 5 TEM images (a) and (b) of Pd@NSG nanocomposite at higher and lower magnifications, respectively. |
Also, the produced H2 reduces the Pd(OAc)2 to Pd(0) as shown in eqn (2).
Pd(OAc)2 + H2 = Pd(0) + 2CH3COOH | (2) |
Finally, to investigate the spillover effect we have performed the physical H2 storage measurement using Sievert's instrument and isotherms were recorded under pressure from 1–50 bar. Fig. 6 shows the H2 uptake characteristic of SiO2/RGO, NixB–SiO2/RGO, Pd–SiO2/RGO and Pd@NSG nanocomposite at RT. As displayed in the graph, the H2 storage increases in the materials with increasing pressure.
![]() | ||
Fig. 6 H2 adsorption isotherm of SiO2/RGO, NixB–SiO2/RGO, Pd–SiO2/RGO and Pd@NSG nanocomposite up to 50 bar pressure at RT. |
At 50 bar, the maximum H2 storage of SiO2/RGO is 0.25 mmol while, after insertion of nanoparticles in SiO2/RGO matrix the H2 uptake capacity was dramatically increased. For NixB–SiO2/RGO and Pd–SiO2/RGO nanocomposites the maximum H2 uptake was estimated to be 0.76 and 1.55 mmol, respectively. At similar adsorption isotherm conditions the major changes were obtained for Pd@NSG nanocomposite which shows a 3.5 mmol H2 uptake. This value was approximately 14 times higher than SiO2/RGO nanocomposite. The corresponding wt% values were 0.05 and 0.7, respectively. This shows the enhancement of H2 uptake in Pd@NSG sample which was higher than SiO2/RGO in the whole range of pressures tested.
Scheme 1 illustrates the H2 spillover mechanism and subsequent diffusion in Pd@NSG nanocomposite. Here, the metal nanoparticles (Pd and NixB@SiO2) act as spillover centre for H2 molecules and dissociates the molecular H2 into H˙ radicals.56 The generated H˙ radicals then migrates from the catalyst centre to the storage material and easily diffuse into the graphene layers.57 Especially, they migrate to the defect sites of graphene sheets such as the edges locations and saturate the hexagonal sp2 hybridized (–CC) network and make the sp3 hybridized (–C–H) structure.38 This phenomenon can be explained through the formation of “bridge” built structure on catalyst centre where H2 molecules easily form the dangling bonds with catalyst centre and dissociate into H˙ radicals.30 In detail, the Pd nanoparticles work as a source, the NixB nanoparticles act as an activator to dissociate the H2 molecules and finally, RGO and SiO2 play the role of receptor.
Due to this process an adduct species on catalyst surface is formed by the dissociation of H2 molecule and subsequently, π–π bonding (–CC) of graphene gets saturated and form sp3 (C–H) bonding.38 Finally, RGO works as a primary receptor site for preferential storage of H˙ radicals via chemisorption and SiO2 nanoparticles act as a secondary receptor to adsorbs the H2 molecule by physisorption.30 Due to these facts, the obtained result of H2 uptake for Pd@NSG nanocomposite at RT was remarkably higher than previously reported works.35,58 Huang et al. reported 0.15 wt% of H2 storage in Pd–Gr nanocomposite at similar conditions (298 K and 60 bar).59 Although, they have noticed that after loading of Pd metal on the graphene sheet, the H2 uptake was doubled. It is obviously due to the spillover capability of Pd to dissociate the H2 molecules and subsequent migration of H atoms on graphene sheet. The results reported here are also better than those obtained by Ansón et al.60 They reported a 0.16 wt% H2 storage at RT in Pd nanoparticles intercalated single walled carbon nanotube (SWNT) after applying 90 bar. However, in our work we applied only a maximum pressure of 50 bar and a 4 times higher H2 storage was observed. This may be attributed due to the porous structure of SiO2, edge defects and large surface area of graphene along with the presence of two spillover centres. Our results of storage capacity of Pd@NSG are higher compared to Latroche et al.61 in the giant-pore MOF MIL-101 (∼0.43 wt% at 80 bar), and Campesi et al.62 in Pd nanoparticles loaded porous carbon template composites at 298 K (0.08 wt% H2 storage).
The reduction of 4-NP by NaBH4 is a standard reaction in which the presence of the catalyst is the essential part for final accomplishment of the reaction. This reaction can be easily monitored by UV-vis spectroscopy under ambient conditions. In absence of the catalyst, NaBH4 only produces the 4-nitrophenolate intermediate ion (peak at 400 nm) in aqueous medium due to the increment of neutral aqueous solution alkalinity which is maintained for several hours (Fig. 7(a)). After the addition of Pd@NSG catalyst, the peak related to 4-NP was drastically decreased and almost disappeared within 2 min (120 s) as depicted in Fig. 7(b). At the same time, a new peak appears at around 300 nm which suggests the successful reduction process and the formation of 4-AP confirming the high catalytic activity of Pd@NSG nanocomposite compared to previously reported works.63,64 This higher catalytic activity of Pd@NSG nanocomposite can be attributed to the presence of bi-metallic condition of two well-known hydrogenation reaction catalyst (Pd and NixB) as shown in Scheme 2. In this aspect we also investigated the role of individual components viz. SiO2/RGO matrix, NixB–SiO2/RGO, Pd–SiO2/RGO without altering reaction conditions and protocol. SiO2/RGO matrix shows the negligible extent of reduction whereas, NixB–SiO2/RGO and Pd–SiO2/RGO reduces the 4-NP, but the rate of reaction was too slow compared to Pd@NSG nanocomposite and estimated times were 22 (1320 s) and 120 min (7200 s), respectively (see Fig. S5 ESI†).
![]() | ||
Scheme 2 Schematic representation of 4-NP molecule reduction on Pd@NSG nanocomposite in presence of NaBH4 in aqueous medium. |
Further, we have performed the kinetic measurement of Pd@NSG nanocomposite as shown in Fig. 7(c). The kinetics of the catalytic reaction measured as a function of time on the basis of absorbance at 400 nm.16 In this reaction, the concentration of NaBH4 was in excess compared with 4-NP and can be regarded as constant throughout the reaction and pseudo-first-order kinetics can be applied with respect to 4-NP.21 A linear relationship was obtained between ln(At/A0) and reaction time (t) which directly inferred the concentration of 4-AP during the reaction as ln(Ct/C0) as given in eqn (3).
k = t−1![]() ![]() | (3) |
Finally, we have performed the recycling test of Pd@NSG nanocomposite also for reduction of 4-NP organic molecule by the insertion of additional aliquots (25 μL) of reagents in same reaction cell. After each addition, the UV was recorded and the catalysts exhibited well stability towards the 4-NP reduction and corresponding time was increased up to 240 s for 5th cycle as shown in Fig. 7(d), suggests the robustness of catalyst at least for 5 consecutive cycles with an efficiency of 88%.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra05523g |
This journal is © The Royal Society of Chemistry 2015 |