DOI:
10.1039/C6RA07560F
(Paper)
RSC Adv., 2016,
6, 43980-43984
Three-dimensional highly branched Pd3Cu alloy multipods as enhanced electrocatalysts for formic acid oxidation
Received
23rd March 2016
, Accepted 27th April 2016
First published on 28th April 2016
Abstract
Pd3Cu alloy multipod nanocrystals are synthesized via a facile polyol reduction method without addition of any other surfactants and additives. The Pd3Cu alloy multipods exhibit a three-dimensional (3D) highly branched morphology which are constructed by arthrogenous branches jointed together at the center. Owing to the superior 3D multipod nanostructure, the Pd3Cu alloy presents remarkably enhanced catalytic performance toward formic acid oxidation.
Introduction
Alloying noble metals such as Pt and Pd with other non-precious transition metals (e.g., Cu, Fe, Co, Ni) has been demonstrated to be an effective strategy towards the development of high-performance and low-cost fuel cells electrocatalysts.1–6 The introduction of the second metal compositions could not only decrease the amount of noble metal required in electrocatalysts but also modify the catalytic activity and stability of noble metal catalysts. The crystallographic structure, electronic state, and surface elemental distribution of alloy nanocrystals can be finely modulated via optimizing the composition, shape and morphology in various synthetic procedures.1–17 Recently, noble metal alloy nanocrystals with complex nanostructures, including nanoframes, branched structures, concave structures, and ultrathin structures, have been synthesized by using different synthetic methodologies.6,10 Particularly, noble metal alloy nanocrystals with highly branched morphologies such as nanodendrites and multipods have been demonstrated with enhanced electrocatalytic properties because of their high density of active sites derived from the atomic steps, ledges and kinks on surface.5,6,17,18
Although, several excellent works have been reported, studies that aimed at the synthesis of highly branched noble metal alloy with multipod nanostructures are still very limited.17,19–27 In this work, we report the synthesis of three-dimensional highly branched Pd3Cu alloy multipod nanocrystals and its catalytic properties towards electrooxidation of formic acid. The Pd3Cu multipods were synthesized via polyol reduction process, by using a mixed solvent of ethylene glycol and water. Particularly, no surfactants and additives were required in this synthetic method. As known, noble metal nanocrystals with multipod nanostructures possess high density of active sites on surface. In this case, the obtained Pd3Cu alloy multipods with three-dimensional highly branched structure could further ensure the maximum exposure of the catalytically active sites. In addition, the porous framework formed by Pd3Cu multipods could remarkably enhance the electron-transfer kinetics in electrocatalysis. Electrochemical experiments showed that the as-prepared Pd3Cu multipods catalysts possess remarkably enhanced catalytic activity toward formic acid oxidation (FAO) over the commercial Pd black catalysts.
Experimental section
Chemicals and reagents
CuCl2·2H2O and ethylene glycol were purchased from Sinopharm Chemical Reagent Co. Ltd. Palladium black (99.9%) was purchased from Alfa Aesar. Na2PdCl4 (98%) was purchased from Sigma-Aldrich. All the chemicals were used as received without further purification.
Synthesis of Pd3Cu multipods
In a typical procedure, 40 mg of Na2PdCl4 and 93 mg of CuCl2·2H2O were dissolved in a mixed solvent containing 10 mL of water and 20 mL of ethylene glycol. The flask containing the above mixture was heated to 150 °C in an oil bath and kept at this temperature for 2 h. The products were collected by centrifugation (10
000 rpm, 10 min), and were further purified by consecutive washing/centrifugation cycles with water and ethanol for several times. The product was re-dispersed in 5 mL of hexane before further characterization.
Structure characterizations
Scanning electron microscopy (SEM) images were recorded on a Hitachi S-4800 field-emission scanning electron microscope. Transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) images were collected on a FEI Tecnai G2 F20 electron microscope operated at 200 kV. High-angle annular dark-field scanning transmission electron microscopy/energy-dispersive X-ray spectroscopy (HAADF-STEM-EDS) analysis was carried out on FEI Tecnai G2 F20 electron microscope. X-ray diffraction (XRD) patterns were recorded on a Rigaku MiniFlex-600 X-ray diffractometer with Cu Kα radiation (λ = 1.5406 Å). X-ray photoelectron spectroscopy (XPS) measurements were performed with Thermo Scientific ESCALAB 250Xi X-ray Photoelectron Spectrometer. The exact composition of as-prepared alloy samples was determined by an Agilent 7500ce inductively coupled plasma-mass spectrometry (ICP-MS).
Electrochemical experiments
The electrochemical experiments were carried out with a CHI 660E electrochemical workstation (CH Instruments) at room temperature. A three-electrode cell was used that consists of a glassy carbon electrode as the working electrode, platinum wire as the counter electrode and a Ag/AgCl (3 M KCl) electrode as the reference electrode. For the preparation of catalysts inks, 5 mg of the catalysts powder was ultrasonically dispersed in 1 mL of water containing 10 μL of 5 wt% Nafion solution until a dark homogeneous dispersion was formed. Then 10 μL of the catalysts ink was dropped on a freshly polished glassy carbon electrode with a micropipette and dried at room temperature. For comparison, the electrochemical measurements of commercial Pd black (Johnson Matthey) were also carried out under the same conditions.
Results and discussion
Fig. 1a shows the representative SEM images of as-prepared Pd3Cu multipod nanocrystals. It can be seen that the Pd3Cu multipods are of high uniformity with size of ca. 500 nm, displaying a flower-like 3D nanostructure. The high magnification SEM image (Fig. 1b) reveals that each nanostructure is composed of many arthrogenous branches which point out in various directions from the center to form the 3D highly branched structure. The highly branched multipod nanostructure can be also clearly seen in TEM image (Fig. 1c and d). The diffraction spots of the SEAD pattern (insert of Fig. 1c) reveal the highly crystalline nature of the Pd3Cu alloy nanocrystals. A small size Pd3Cu multipod nanocrystal (Fig. 1e) shows that the branches are jointed together at the center. This suggests the formation of Pd3Cu multipods might involve the growth from a common point center.28 The HRTEM image (Fig. 1f) and the corresponding FFT pattern (insert of Fig. 1f) shows lattice fringes with an interplanar spacing of 0.22 nm, which was indexed to (111) planes of face-centered cubic (fcc) Pd–Cu alloy.29–31 Furthermore, the arthrogenous branches of Pd3Cu alloy nanocrystals exhibit a step or concave surface topology which is in rich of atomic steps, ledges and kinks.19 High-angle annular dark-field scanning transmission electron microscopy/energy-dispersive X-ray spectroscopy (HAADF-STEM-EDS) mapping was used to analyze the Pd and Cu distribution in Pd3Cu alloy nanocrystals. As shown in Fig. 2a–c, the spatial distributions of Pd and Cu are uniform and completely overlapped. The cross-sectional compositional line profiles (Fig. 2d) of a single Pd3Cu nanocrystal further confirmed the homogeneous distribution of the two elements. The EDX analysis shows that the atomic ratio of Pd and Cu is 76.3 and 23.7%, respectively, which was almost the same as ICP-MS data (Pd/Cu = 76
:
24).
 |
| Fig. 1 (a and b) SEM, (c–e) TEM and (f) HRTEM images of Pd3Cu. Insert of (c): corresponding SAED pattern. Insert of (f): corresponding FFT pattern. | |
 |
| Fig. 2 (a) HAADF-STEM image, (b and c) EDS mapping images, (d) the cross-sectional compositional line profiles and (e) EDX spectrum of Pd3Cu. | |
XRD patterns of as-prepared Pd3Cu multipods were depicted in Fig. 3, with the commercial Pd black as a comparison. The diffraction peaks of Pd3Cu at 2θ = 40.5, 47.0, 68.8, 82.9 and 87.6° can be indexed as the (111), (200), (220), (311) and (222) facets of face-centered cubic (fcc) Pd–Cu alloy, respectively.22,29–31 These diffraction peaks of Pd3Cu are located between those of pure fcc Pd (JCPDS 46-1043) and Cu (JCPDS 04-0836), indicating the formation of single-phase uniform alloy structure.32,33 In comparison with Pd black, the diffraction peaks of as-prepared Pd3Cu alloy upshift to higher angles with an increase in Cu content, which can be ascribed to the decreased lattice constant due to the alloying of Pd atom with smaller Cu atoms.32,33
 |
| Fig. 3 XRD patterns of Pd3Cu and Pd black. | |
The surface composition and chemical state of Pd3Cu multipods were analyzed by XPS. XPS spectra of Pd 3d for the Pd3Cu are shown in Fig. 4a. The Pd 3d spectra of Pd3Cu shows the double peaks with binding energy of 340.6 and 335.4 eV, representing the Pd 3d3/2 and 3d5/2 electrons of metal Pd, respectively.29–31 The XPS spectra of Pd 3d could be deconvoluted into two pairs of doublets. The most intense doublet of peaks is due to metallic Pd(0), while the weaker doublet of peaks (341.4 and 336.1 eV) could be assigned to the Pd(II) species such as PdO.29–31 It can be seen that Pd in Pd3Cu multipods is mostly at the chemical bonding state of Pd(0). Moreover, the Pd 3d binding energy of as-prepared Pd3Cu multipods exhibits a small positive shift to higher binding energy with respect to that of pure polycrystalline Pd (Pd 3d5/2 at 334.8 and Pd 3d3/2 at 340.2 eV). The upshift of Pd 3d binding energy with increasing Cu content in Pd–Cu alloy has been commonly observed that was ascribed to the electron transfer of Pd valance electrons to Cu.34 The Cu 2p spectra can be also fitted by two pairs of doublets. The principle doublet at 932.2 and 951.9 eV correspond to metallic Cu(0), suggesting that most of the Cu species were at zero-valence (Fig. 4b). The minor doublet at 933.7 and 954.1 eV and the presence of satellite signals indicates the presence of a small amount CuO species adsorbed on the alloy surfaces, due to the oxidation of metallic Cu under ambient conditions.29–31,35
 |
| Fig. 4 XPS spectra of the deconvoluted (a) Pd 3d and (b) Cu 2p peaks for Pd3Cu. | |
The electrochemistry of Pd3Cu multipods was first investigated by cyclic voltammograms (CVs) in N2-saturated 0.5 M H2SO4 solution. Fig. 5a shows the CV profiles of Pd3Cu multipods during the initial 50 cycles. The typical hydrogen adsorption/desorption (−0.2 to 0.1 V), Pd oxidation (>0.5 V), and reduction of Pd oxide (at ∼0.5 V) peaks can be clearly observed in the CV curves.36 Additionally, a small peak at 0.35 V can be ascribed to the anodic Cu stripping due to the dissolution of exposed Cu atoms from the Pd3Cu alloy surface.35–38 During the initial 50 cycles, the peaks in hydrogen adsorption/desorption region and Pd oxidation region sharpened and increased while the Cu stripping peak slightly decreased, suggesting a structural re-arrangement of Pd atoms at the surface.37,39 As for the Pt(Pd)–Cu alloy catalysts, a surface dealloying process were commonly observed which could generate a noble metal rich surface to enhance the electrocatalytic activity.37,40,41 The stable CV curve of Pd3Cu multipods was depicted in Fig. 5b, with that of Pd black for comparison. The electrochemically active surface area (ECSA) of the catalysts was determined by integrating the coulombic charge of Pd oxide stripping peak, by assuming a charge of 424 μC cm−2 for reduction of the PdO monolayer.42 The calculated ECSA value was 9.8 and 11.5 m2 gPd−1 for Pd3Cu multipods and Pd black, respectively. The ECSA of the samples was also calculated from the hydrogen adsorption/desorption peaks (210 μC cm−2 for monolayer adsorption of H), which was 13.1 and 20.6 m2 gPd−1 for Pd3Cu multipods and Pd black, respectively.43 The lower ECSA for the Pd3Cu multipods is mostly likely due to its larger particle size compared with that of Pd particles (∼3 nm) in commercial Pd black. This phenomenon has been commonly observed in Pt and Pd nanostructures with large particle size.21,44,45
 |
| Fig. 5 (a) CV curves of Pd3Cu recorded during the initial 50 cycles; (b) the stable CV curves for ECSA calculation. Scan rate: 50 mV s−1. | |
The FAO tests were conducted by cyclic voltammetry in 0.5 M H2SO4 + 0.5 M HCOOH solution with a sweep rate of 50 mV s−1. Fig. 6a shows the CV curves of formic acid oxidation on Pd3Cu multipods and Pd black catalysts. The Pd3Cu multipods catalyst displays lower onset potential (−0.12 V for Pd3Cu and −0.06 V for Pd black) and higher mass activity (1.6 times) than Pd black. Fig. 6b shows the CV profiles with current density normalized to the value of ECSA. The area-specific current density of formic acid oxidation on Pd3Cu multipods is about 1.9 times larger than that on Pd black, indicating enhanced intrinsic activity of Pd3Cu. The improved electrocatalytic activity of Pd3Cu multipods towards FAO can be attributed to the synergistic alloying effect between Pd and Cu, as observed in other Pd-based alloy electrocatalysts.2,43,46,47 More importantly, the highly branched multipods nanostructure with a step surface topology could provide high density of catalytically active sites derived from the atomic steps, ledges and kinks on surface, which may further contribute to the enhanced activity.17–27 In addition, the catalytic activity of Pd3Cu multipods for HCOOH oxidation is comparable to other reported Pd3Cu alloy catalysts.31,36 Fig. 6c shows the chronoamperometric curves measured at a constant potential of 0.2 V in 0.5 M H2SO4 + 0.5 M HCOOH solution. The Pd3Cu multipods exhibits larger current density over the entire time examined than that of Pd black, indicating its enhanced catalytic activity and stability for FAO.
 |
| Fig. 6 (a and b) CV curves measured in 0.5 M H2SO4 + 0.5 M HCOOH at scan rate of 50 mV s−1. (c) Chronoamperometry curves measured in 0.5 M H2SO4 + 0.5 M CH3OH at applied potential of 0.2 V. (d) Nyquist plots recorded in 0.5 M H2SO4 + 0.1 M HCOOH in frequency range of 10−2 to 105 Hz at 0.1 V. | |
The 3D highly branched Pd3Cu multipod nanostructure could promote the accessibility with reactant not only at the surfaces, but also throughout the bulk of the nanocrystals. This will facilitate the mass diffusion and electron transfer during electrocatalysis. The electron-transfer property of Pd3Cu multipods was further investigated by electrochemical impedance spectroscopy (EIS). Fig. 6d shows the Nyquist plots of different materials in 0.5 M H2SO4 + 0.1 M HCOOH solution. The radius of the semicircle impedance loop at high frequency region represents electron-transfer resistance (Rct) of the electrode materials.48 The Rct is approximately 100 and 140 Ω for Pd3Cu multipods and Pd black, respectively. This result confirms that the electron-transfer kinetics for FAO is greatly facilitated on the Pd3Cu multipods due to its advantageous 3D highly branched nanostructure.
Conclusions
In summary, Pd3Cu alloy multipods with a 3D highly branched morphology were synthesized by a facile polyol reduction method, without addition of any other surfactants and additives. This novel nanostructure was constructed many of arthrogenous branches which were jointed together at the center. The arthrogenous branches with a step surface topology are favorable for providing high density of catalytically active sites. Moreover, the 3D highly branched morphology ensures the maximum exposure of surface active sites for electrocatalysis, which greatly promotes the efficient utilization of noble metals. The Pd3Cu multipods catalysts demonstrated remarkably enhanced electrocatalytic activity and would be a promising candidate for wide applications in catalysis.
Acknowledgements
This work was financially supported by the Science & Technology Innovation Talents in Universities of Henan Province (No. 13HASTIT012) and the Nanhu Scholars Program for Young Scholars of XYNU.
Notes and references
- M. Liu, R. Zhang and W. Chen, Chem. Rev., 2014, 114, 5117–5160 CrossRef CAS PubMed.
- C. J. Zhong, J. Luo, P. N. Njoki, D. Mott, B. Wanjala, R. Loukrakpam, S. Lim, L. Wang, B. Fang and Z. Xu, Energy Environ. Sci., 2008, 1, 454–466 CAS.
- Y. Bing, H. Liu, L. Zhang, D. Ghosh and J. Zhang, Chem. Soc. Rev., 2010, 39, 2184–2202 RSC.
- J. Gu, Y. W. Zhang and F. Tao, Chem. Soc. Rev., 2012, 41, 8050–8065 RSC.
- H. You, S. Yang, B. Ding and H. Yang, Chem. Soc. Rev., 2013, 42, 2880–2904 RSC.
- H. L. Liu, F. Nosheen and X. Wang, Chem. Soc. Rev., 2015, 44, 3056–3078 RSC.
- H. Liu, R. R. Adzic and S. S. Wong, ACS Appl. Mater. Interfaces, 2015, 7, 26145–26157 CAS.
- H. Liu, C. Koenigsmann, R. R. Adzic and S. S. Wong, ACS Catal., 2014, 4, 2544–2555 CrossRef CAS.
- B. Jiang, C. Li, V. Malgras, Y. Bando and Y. Yamauchi, Chem. Commun., 2016, 52, 1186–1189 RSC.
- Y. Lu, Y. Jiang and W. Chen, Nanoscale, 2014, 6, 3309–3315 RSC.
- Y. Jia, J. Su, Z. Chen, K. Tan, Q. Chen, Z. Cao, Y. Jiang, Z. Xie and L. Zheng, RSC Adv., 2015, 5, 18153–18158 RSC.
- Q. Lv, J. Chang, W. Xing and C. Liu, RSC Adv., 2014, 4, 32997–33000 RSC.
- Z. Zhang, C. Zhang, J. Sun, T. Kou and C. Zhao, RSC Adv., 2012, 2, 11820–11828 RSC.
- Y. Qi, T. Bian, S. Choi, Y. Jiang, C. Jin, M. Fu, H. Zhang and D. Yang, Chem. Commun., 2014, 50, 560–562 RSC.
- X.-J. Liu, C.-H. Cui, H.-H. Li, Y. Lei, T.-T. Zhuang, M. Sun, M. N. Arshad, H. A. Albar, T. R. Sobahi and S.-H. Yu, Chem. Sci., 2015, 6, 3038–3043 RSC.
- J. Mao, T. Cao, Y. Chen, Y. Wu, C. Chen, Q. Peng, D. Wang and Y. Li, Chem. Commun., 2015, 51, 15406–15409 RSC.
- K. Eid, V. Malgras, P. He, K. Wang, A. Aldalbahi, S. M. Alshehri, Y. Yamauchi and L. Wang, RSC Adv., 2015, 5, 31147–31152 RSC.
- B. Lim and Y. Xia, Angew. Chem., Int. Ed., 2011, 50, 76–85 CrossRef CAS PubMed.
- J. Watt, S. Cheong, M. F. Toney, B. Ingham, J. Cookson, P. T. Bishop and R. D. Tilley, ACS Nano, 2010, 4, 396–402 CrossRef CAS PubMed.
- Z. Niu, D. Wang, R. Yu, Q. Peng and Y. Li, Chem. Sci., 2012, 3, 1925–1929 RSC.
- S. Chen, H. Su, Y. Wang, W. Wu and J. Zeng, Angew. Chem., Int. Ed., 2015, 54, 108–113 CrossRef CAS PubMed.
- X. Qiu, R. Zhao, Y. Li, Y. Tang, D. Sun, S. Wei and T. Lu, RSC Adv., 2014, 4, 57144–57147 RSC.
- B. D. Adams, G. Wu, S. Nigro and A. Chen, J. Am. Chem. Soc., 2009, 131, 6930–6931 CrossRef CAS PubMed.
- Y. Kuang, Z. Cai, Y. Zhang, D. He, X. Yan, Y. Bi, Y. Li, Z. Li and X. Sun, ACS Appl. Mater. Interfaces, 2014, 6, 17748–17752 CAS.
- J. J. Lv, L. P. Mei, X. Weng, A. J. Wang, L. L. Chen, X. F. Liu and J. J. Feng, Nanoscale, 2015, 7, 5699–5705 RSC.
- K. Zhang, D. Bin, B. Yang, C. Wang, F. Ren and Y. Du, Nanoscale, 2015, 7, 12445–12451 RSC.
- Y. Jiang, T. Bian, F. Lin, H. Zhang, C. Jin, Z. Y. Li, D. Yang and Z. Zhang, J. Mater. Chem. A, 2015, 3, 21284–21289 CAS.
- B. Y. Xia, W. T. Ng, H. B. Wu, X. Wang and X. W. Lou, Angew. Chem., Int. Ed., 2012, 51, 7213–7216 CrossRef CAS PubMed.
- J. Cai, Y. Zeng and Y. Guo, J. Power Sources, 2014, 270, 257–261 CrossRef CAS.
- J. J. Lv, S. S. Li, A. J. Wang, L. P. Mei, J. J. Feng, J. R. Chen and Z. Chen, J. Power Sources, 2014, 269, 104–110 CrossRef CAS.
- C. Xu, Y. Liu, J. Wang, H. Geng and H. Qiu, J. Power Sources, 2012, 199, 124–131 CrossRef CAS.
- Z. Yin, W. Zhou, Y. Gao, D. Ma, C. J. Kiely and X. Bao, Chem.–Eur. J., 2012, 18, 4887–4893 CrossRef CAS PubMed.
- J. Mao, Y. Liu, Z. Chen, D. Wang and Y. Li, Chem. Commun., 2014, 50, 4588–4591 RSC.
- L. Wang, J. J. Zhai, K. Jiang, J. Q. Wang and W. B. Cai, Int. J. Hydrogen Energy, 2015, 40, 1726–1734 CrossRef CAS.
- Y. Fan, P. F. Liu, Z. W. Zhang, Y. Cui and Y. Zhang, J. Power Sources, 2015, 296, 282–289 CrossRef CAS.
- C. Xu, A. Liu, H. Qiu and Y. Liu, Electrochem. Commun., 2011, 13, 766–769 CrossRef CAS.
- S. Koh and P. Strasser, J. Am. Chem. Soc., 2007, 129, 12624–12625 CrossRef CAS PubMed.
- L. Dai and S. Zou, J. Power Sources, 2012, 199, 124–131 CrossRef.
- S. Tominaka, T. Hayashi, Y. Nakamura and T. Osaka, J. Mater. Chem., 2010, 20, 7175–7182 RSC.
- P. Strasser, S. Koh, T. Anniyev, J. Greeley, K. More, C. Yu, Z. Liu, S. Kaya, D. Nordlund, H. Ogasawara, M. F. Toney and A. Nilsson, Nat. Chem., 2010, 2, 454–460 CrossRef CAS PubMed.
- L. Gan, C. Cui, M. Heggen, F. Dionigi, S. Rudi and P. Strasser, Science, 2014, 346, 1502–1506 CrossRef CAS PubMed.
- S. Hu, L. Scudiero and S. Ha, Electrochim. Acta, 2012, 83, 354–358 CrossRef CAS.
- Y. Lu and W. Chen, ACS Catal., 2012, 2, 84–90 CrossRef CAS.
- S. Sun, F. Jaouen and J. P. Dodelet, Adv. Mater., 2008, 20, 3900–3904 CrossRef CAS.
- S. Sun, G. Zhang, D. Geng, Y. Chen, M. N. Banis, R. Li, M. Cai and X. Sun, Chem.–Eur. J., 2010, 16, 829–835 CrossRef CAS PubMed.
- Y. Lu and W. Chen, J. Phys. Chem. C, 2010, 114, 21190–21200 CAS.
- M. Liu, Y. Lu and W. Chen, Adv. Funct. Mater., 2013, 23, 1289–1296 CrossRef CAS.
- H. T. Lu, Z. J. Yang, X. Yang and Y. Fan, Electrocatalysis, 2015, 6, 255–262 CrossRef CAS.
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.