logo

Amyloid fibril-supported Pd nanoparticles as electrocatalyst for hydrogen peroxide reduction

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Amyloid fibril-supported Pd nanoparticles as electrocatalyst for hydrogen peroxide reduction

KONG Li-Xia
WANG Lan-Jie
SU Lan-Lan
ZHAO Zhi-Jie
YAN Zhong-Dan
ZHANG Gong-Jun
ZHOU Xing-Fei
Nuclear Science and TechniquesVol.27, No.4Article number 97Published in print 20 Aug 2016Available online 15 Jul 2016
47400

Palladium nanoparticles (Pd NPs) were fabricated by using insulin amyloid fibrils (INSAFs) as biotemplates. Atomic force microscopy measurements showed that ultra-small Pd NPs were well adsorbed and dispersed on surfaces of INSAFs. X-ray photoelectron spectroscopy confirmed the partial reduction of Pd ion into metallic Pd(0) probably due to the presence of Cys groups on surface of the insulin fibrils. The electrochemical performance of Pd/INSAFs to reduction of H2O2 was further evaluated by cyclic voltammetry (CV). The remarkably high electrocatalytic activity, low detection limitation and excellent stability make the Pd/INSAFs a promising bio-nanoelectrocatalyst.

Palladium nanoparticlesAmyloid fibrilsCyclic VoltammetryNanoelectrocatalystHydrogen peroxide

1 Introduction

Platinum nanomaterials are widely used as electrocatalysts in polymer electrolyte membrane fuel cells [1-3]. Nevertheless, one major technical challenge of Pt nanoparticles is that these zero-dimensional materials have frequently large number of lattice boundaries and defect sites, resulting in low catalyst activities [4]. Also, the expensiveness of Pt hinders its wide application in industries [3,5]. Pd-based nanostructured materials can be a promising alternative to Pt, due to its lower cost, its relatively high catalytic activity and its similar intrinsic properties to Pt such as lattice parameters and energy band structures [6,7].

Palladium nanoparticles (Pd NPs) play a significant role in the catalysis of a variety of organic reactions including hydrogenation, C–C coupling reactions and oxygen reduction in fuel cells [8,9]. It has been demonstrated that substantially smaller, uniform and well-dispersed nanoparticles are required for high-yield catalytic applications [10-12]. Meanwhile, due to large surface energy and electrostatic attractive force of the Pd NPs, they are easily to aggregate into bigger clusters, hence decreased catalytic activities [13]. Therefore, carbon-based materials such as carbon nanotubes (CNTs) and graphite are often used as supporting materials for catalytic metal nanoparticles because of their fascinating structure and electrical/mechanical properties [3,14-16]. In order to fix Pd NPs on the substrates, carbon materials shall be functionalized because of their hydrophobic and inert surface. However, chemical treatments on carbon materials often involve harsh experimental conditions, which may destroy the mechanical and electronic properties of CNTs and introduce defects.

Recently, biological supramolecules have attracted dramatic attention as template materials for the controlled synthesis of metallic Pd NPs as a consequence of their well-defined structures [17-19]. For example, in their pioneering work, Nguyen et al. fabricated Pd clusters along DNA molecules by reducing PdO precipitates with gaseous hydrogen [17]. Compared with DNA molecules, amyloid nanofibrils are much stable and rigid, because their internal structures comprise a cross-β sheet core with richer hydrogen bonds formed between the amino and carbonyl groups of the polypeptide chain [20]. Importantly, amyloid fibrils are rich in functional groups that can specifically interact with various nanoparticles. Therefore it provides another green alternative as scaffolds for nanostructures assembly [21,22].

In our previous work, we prepared “amyloid fibrils-Pd chip” by alternatively depositing Pd NPs and amyloid fibrils on a solid substrate for catalyzing Suzuki Coupling reactions [22,23]. In this paper, we describe a facile and mild approach to fabricate Pd-insulin catalyst with controlled Pd NPs size. The Pd-insulin catalyst demonstrated very nice electrocatalytic activities towards the reduction of H2O2.

2 Experimental

2.1. Materials

Insulin peptide with a purity of 98% was purchased from ProSpec (ProSpec-Tany TechnoGene Ltd). The stock solution was prepared by dissolving insulin powder in Milli-Q water (18.2 MΩ.cm) to achieve a final concentration of 1.0 mg/mL and adjusting to pH 1.6 with HCl. The solution was stored at 20°C before use. Sodium tetrachloropalladate (II) (Na2PdCl4) were purchased from Sigma Co., Ltd (Shanghai, China). PdC was purchased from Aladdin Co. All these chemicals were used in our experiments without further purification.

2.2. Preparation of Pd-insulin catalyst

To prepare INSAFs in solution, 200 µL of insulin peptide (1.0 mg/mL, pH 1.6) was incubated at 80°C for two days. After that, a drop of incubated solution was deposited on a fresh cleaved mica surface for AFM imaging to check the formation of INSAFs. The PdO NPs was achieved by aging the Na2PdCl4 solution (5 mM) at 25°C for two days. Then the incubated insulin solution was mixed with PdO NPs for about 30 minutes.

2.3. Preparation of the electrodes

To immobilize the INSAFs and Pd NPs on the electrode, the glassy carbon electrode (GCE) was polished with 0.05 μm alumina slurry and sonicated in distilled water. Then cyclic voltammetry (CV) was run in 1.0 mM HAuCl4 solution in the range of −0.6~0.8V at a scan rate of 100 mV·s−1 to coat Au on the surface of GCE. The Au-coated GCE electrode was dipped into 1.0 mg/mL INSAFs solution for 40 min followed by air-blow drying to remove excess solution. Finally the INSAFs adsorbed electrode was put into a 5 mM aged Na2PdCl4 solution for 120 min. After drying in air, the Pd/INSAFs modified GCE (Pd/INSAFs-GCE) could be used as a working electrode.

2.4 Physical characterization of catalysts

Surface morphologies of all samples were collected by A Multimode 8/Nanoscope V system (Bruker AXS, Germany) equipped with RTESP silicon cantilevers at room temperature. AFM data were analyzed with the NanoScope Analysis V1.20 software. X-ray photoelectron spectroscopy (XPS) analysis was conducted using an AXIS Ultra DLD spectrometer (Kratos Analytical Inc., Manchester, UK) with a monochromated Al Ka X-ray source (1486.7 eV). The XPS was calibrated at the Au 4f7/2 with a binding energy of 84.0 eV.

2.5 Electrocatalytic measurement

The electrocatalytic performance of Pd/INSAFs was measured on a CHI660C electrochemical workstation (Shanghai Chenhua Instrument Co., Shanghai, China) with three-electrode cell at room temperature. The three-electrode setup consisted of Pd/INSAFs-GCE working electrode, a saturated calomel reference electrode (SCE) and a platinum wire electrode. CV was carried out in phosphate buffer solution (PBS) (0.1 M, pH 7.0) from −0.4~0.8 V at a scan rate of 100 mV·s−1 at room temperature.

3 Results and Discussion

The INSAFs was obtained by incubating insulin peptides solution (1.0 mg/mL, pH 1.6) at 80°C for two days. The PdO NPs were obtained by aging the Na2PdCl4 solution (5 mM) at 25°C for two days [23]. After mixing with the as-prepared INSAFs, the PdO NPs were adsorbed onto the INSAFs (Fig.1a and 1b) and gradually reduce to metallic palladium probably due to the existence of Cys groups on fibrils surface as confirmed by XPS measurement (Fig.1c). Interestingly, two shoulder peaks can be seen clearly at 335.6 and 340.8 eV, which are very close to the reported values of Pd(0) species [17,24]. The light shift of Pd may be caused by the interaction of Pd NPs with fibrils. It is worthy to stress that, unlike previous reports [25,26], no external reduce agency such as NaBH4 was introduced in our system. Here, we attribute the reduction of palladium precursor into Pd (0) to the presence of Cys residue along INSAFs surface as suggested by Kurouski et al. [27]. Several groups reported that palladium precursors could self-mineralize on Cys mutant TMV [28,29].

Fig. 1.
AFM images of pure INSAFs (a) and Pd NPs-coated INSAFs (b); and XPS spectra of Pd/INSAFs (c), the peaks at 335.6 and 340.8 eV suggest the presence of metallic Pd.
pic

Quantitative detection of H2O2 is of great importance in recent years as it is increasingly used in food, chemical and biochemical industries. Direct electrochemical detection of H2O2 has been achieved via its reduction at a variety of electrode materials. Some metal or metal oxide-modified electrodes have exhibited high electrocatalytic activities towards the reduction of H2O2. Here, the electrocatalytic activity of Pd/INSAFs was investigated by CV. Details of experiments were described in the experimental section. Briefly, the obtained Pd/INSAFs were adsorbed on the Au-coated GCE. The modified electrodes were immersed in 0.1 M PBS, and the potential was scanned from −0.2 to 0.8 V at a scan rate of 100 mV/s to obtain CV curves. As shown in Fig. 2(a), no apparent reduction current could be observed in the presence of 0.4 mM of H2O2 for the INSAFs-GCE, while commercial Pd/C demonstrated a remarkable increased activity in comparison with INSAFs-GCE (Fig. 2b). For the Pd/INSAFs-GCE, we observed a strong and sharp reduction peak in the absence of H2O2 between −0.1~ +0.1 mV (Fig. 2c), which can be attributed to the reduction of palladium oxide [30]. Interestingly, with 0.4 mM H2O2, an obvious catalytic reduction peak was seen in even stronger intensity. The peak position of the CV curve shows characteristics similar to those of commercial Pd/C, indicating that Pd/INSAFs were effectively loaded on the GCE surface. The remarkable increase of peak intensity suggests a faster electron transfer rate and higher electrocatalytic activity toward the reduction of H2O2 at the Pd/INSAFs [30]. We further estimated the electrochemical active surface area (ECSA), which provides important information regarding the number of available active sites. As reported in Ref. [31, 32], the electrochemical active sites increase with the ECSA. The ECSA was calculated by ECSA =Qo/0.424, where Qo is the coulombic charge for the reduction of PdO, i.e. the surface charge obtained from the area under the CV trace, and 0.424 mC·cm−2 charge is required for desorption of monolayer of oxygen on the Pd surface [32]. The Pd/INSAFs displays an outstandingly high ECSA of 13.82 m2·g−1, which is much higher than that of the commercial Pd/C catalyst (3.34 m2·g−1). This substantial difference indicates that Pd/INSAFs possess larger number of electrochemical active sites, which probably stems from the overall smaller size of Pd nanoparticle around INSAFs [33].

Fig. 2.
CV curves of insulin-GCE (a), commercial Pd/C-GCE (b) and Pd/INSAFs-GCE (c) in PBS buffer without (red) or with (blue) 0.4 mM H2O2.
pic

Figure 3 shows typical CV curves recorded for the Pd/INSAFs in PBS solution containing 4 mM H2O2 at scan rates of 20−260 mV/s. The reduction current increases with the scan rate. For all the scan rates, the Pd/INSAFs exhibited excellent electrochemical performance. A plot of the cathodic peak currents versus the square root of the scan rate (ν1/2) for Pd/INSAFs is also given in the inset of Fig. 3. The peak current is linearly proportional to ν1/2, indicating that the diffusion limited electron transfer kinetics for the reduction of H2O2 at Pd/INSAFs [31]. Moreover, we notice that the reduction peaks shift slightly to more negative potentials at higher scan rates, suggesting that the electrochemical reduction process for H2O2 is irreversible [25].

Fig. 3.
CV curves of Pd/INSAFs for the catalystic reduction of 0.4mM of H2O2 in PBS (0.1M, pH: 7.0) at different scan rates between 20 to 260 mV·s−1. The inset shows plot of cathodic peak currents vs the square root of the scan rate.
pic

Typical CVs for Pd/INSAFs modified GCE in PBS buffer with H2O2 of various concentrations is shown Fig. 4. The Pd/INSAFs-GCE shows a wider linear range from ~0.2 μM to 0.8 mM for H2O2 [34,35]. When the concentration of H2O2 was less than 0.2 μM, the current peak gradually deviated from linear region. Thus the detection limit is estimated at 0.2 μM, which is comparable to some similar types of sensors [36]. This wide linear range and low detection limit are likely due to the 3D network structure of the Pd/INSAFs and the high electrocatalytic activity of Pd NPs dispersed on INSAFs [21].

Fig. 4.
The current responses of the Pd/INSAFs on different H2O2 concentrations (a) in PBS; and the cathodic peak current as function of the H2O2 concentration, the inset shows the cathodic peak current at low H2O2 concentration (b).
pic

Stability of catalysts is an important factor affecting the durability of Pd-based electrocatalyst. Fig. 5(a) shows the current responses of the Pd/INSAFs after different days at room temperature. The initial current response was maintained in 15 days, suggesting a remarkable resistance of the Pd/INSAFs-GCE to environmental fouling. These advantages make Pd/INSAFs a promising material for electrochemical applications.

Fig. 5.
CV curves of Pd/INSAFs-GCE in PBS (0.1M, pH 7.0) on different days (a) and at different coverage of Pd/INSAFs (b).
pic

We further investigated the effect of surface coverage of Pd/INSAFs on its electrochemical performance. Surface coverage was controlled by tuning the dipping duration of Au coated GCE in Na2PdCl4 solution. As expected, the reduction peak current enhanced with an increase of Pd/INSAFs coverage on GCE (Fig.5b). At low particle coverage (dipping duration: 40 minutes), the peak for the reduction of H2O2 at modified electrode was very broad. With an increase of dipping time, the current peak became much shaper. When the dipping duration was increased to 120 minutes, the peak intensity achieved at 63.77 μA, which was about 4 times as much as that at low coverage. This is likely due to the change of ensemble behavior of Pd/INSAFs on the electrode [37]. Meanwhile, the peak position was shift from 0.058 to −0.19 V, probably because the particle coverage increases the rate constant kinetically as reported by Kumar where they observed the positivity peak position shift when Pd/C was used to catalyze the H2O2 [38].

4 Conclusions

In conclusions, we have presented a simple and green approach for the synthesis of ultra-small Pd NPs supported by INSAFs. The key for the successful synthesis of Pd NPs is the use of INSAFs which poss special amino acid that can specific interaction with Pd NPs. The Pd-INSAFs are demonstrated to have very high electrocatalytic activity and stability toward the reduction of H2O2. We believe that this approach can be used for the development of metallic nanoparticles based ensemble electrode for a wide range electrocatalytic and electroanalytical applications.

References
[1] C. Koenigsmann, W.P. Zhou, R.R. Adzic, et al.

Size-dependent enhancement of electrocatalytic performance in relatively defect-free, processed ultrathin platinum nanowires

. Nano. Lett. 10, 2806-2811 (2010). doi: 10.1021/nl100718k
Baidu ScholarGoogle Scholar
[2] C. Li, S. Takaaki, Y. Yusuke,

Electrochemical synthesis of one-dimensional mesoporous Pt nanorods using the assembly of surfactant micelles in confined space

. Angew. Chem. Int. Ed. 52, 8050-8053 (2013). doi: 10.1002/anie.201303035
Baidu ScholarGoogle Scholar
[3] Q.L. Wu, B. Song,

Interaction of Al-induced peptide backbone ring structure with the sidechains of His, Phe, Trp and Tyr

. Nucl. Sci. Tech. 26, 040504 (2015). doi: 10.13538/j.1001-8042/nst.26.040504
Baidu ScholarGoogle Scholar
[4] J.D. Maeseneer, D. Egilman, W.P. Burdick, et al.

Enhanced electrocatalytic performance of processed, ultrathin, supported Pd-Pt core-shell nanowire catalysts for the oxygen reduction reaction

. J. Am. Chem. Soc. 133, 9783-9795 (2011). doi: 10.1021/ja111130t
Baidu ScholarGoogle Scholar
[5] H. Huang, D. Sun, X. Wang,

PtCo alloy nanoparticles supported on graphene nanosheets with high performance for methanol oxidation

. Chin. Sci. Bull. 57, 3071-3079 (2012). doi: 10.1007/s11434-012-5327-4
Baidu ScholarGoogle Scholar
[6] M. Homma, S.J. Kim,

Supercritical fluid synthesis and characterization of catalytic metal nanoparticles on carbon nanotubes

. J. Mater. Chem. 14, 908-913 (2004). doi: 10.1039/b308124a
Baidu ScholarGoogle Scholar
[7] E. Antolini,

Palladium in fuel cell catalysis

. Energy Environ. Sci. 2, 915-931 (2009). doi: 10.1039/b820837a
Baidu ScholarGoogle Scholar
[8] N. Miyaura, A. Suzuki,

Palladium-catalyzed cross-coupling reactions of organoboron compounds

. Chem. Rev. 95, 2457-2483 (1995). doi: 0009-2665/95/0795
Baidu ScholarGoogle Scholar
[9] Y. Lin, X. Cui, X. Ye,

Electrocatalytic reactivity for oxygen reduction of palladium-modified carbon nanotubes synthesized in supercritical fluid

. Electrochem. Commun. 7, 267-274 (2005). doi: 10.1016/j.elecom.2005.01.007
Baidu ScholarGoogle Scholar
[10] R. Kannan, A.R. Kim, D.J. Yoo,

Synthesis, characterization and electrocatalytic studies of palladium-manganese oxyhydroxide nanocomposite towards direct ethylene glycol fuel cell

. Chin. Sci. Bull. 59, 3413-3419 (2014). doi: 10.1007/s11434-014-0459-3
Baidu ScholarGoogle Scholar
[11] H. Erikson, A. Sarapuu, N. Alexeyeva, et al.

Electrochemical reduction of oxygen on palladium nanocubes in acid and alkaline solutions

. Electrochim. Acta. 59, 329-335 (2012). doi: 10.1016/j.electacta.2011.10.074
Baidu ScholarGoogle Scholar
[12] L. Zhang, K. Lee, J. Zhang,

Effect of synthetic reducing agents on morphology and ORR activity of carbon-supported nano-Pd-Co alloy electrocatalysts

. Electrochim. Acta. 52, 7964-7971 (2007). doi: 10.1016/j.electacta.2007.06.056
Baidu ScholarGoogle Scholar
[13] V. Mazumder, S. Sun,

Oleylamine-mediated synthesis of Pd nanoparticles for catalytic formic acid oxidation

. J. Am. Chem. Soc. 131, 4588-4589 (2009). doi: 10.1021/ja9004915
Baidu ScholarGoogle Scholar
[14] J. Yang, Y. Xi, R. Wang, et al.

Synergistic effect of tungsten carbide and palladium on graphene for promoted ethanol electrooxidation

. ACS Appl. Mater. Interfaces. 5, 6571-6579 (2013). doi: 10.1021/am401216s
Baidu ScholarGoogle Scholar
[15] B. Yoon, C.M. Wai,

Microemulsion-templated synthesis of carbon nanotube-supported Pd and Rh nanoparticles for catalytic applications

. J. Am. Chem. Soc. 127, 17174-17175 (2005). doi: 10.1021/ja055530f
Baidu ScholarGoogle Scholar
[16] X.L. Li, Y.Q. Liu, L. Fu, et al.

Efficient synthesis of carbon nanotube-nanoparticle hybrids

. Adv. Funct. Mater. 16, 2431-2437 (2006). doi: 10.1002/adfm.200600339
Baidu ScholarGoogle Scholar
[17] K. Nguyen, M. Monteverde, A. Filoramo, et al.

Synthesis of thin and highly conductive DNA-based palladium nanowires

. Adv. Mater. 20, 1099-1104 (2008). doi: 10.1002/adma.200701803
Baidu ScholarGoogle Scholar
[18] Y.S. Nam, A.P. Magyar, D. Lee, et al.

Biologically templated photocatalytic nanostructures for sustained light-driven water oxidation

. Nature Nanotech. 5, 340-344 (2010). doi: 10.1038/NNANO.2010.57
Baidu ScholarGoogle Scholar
[19] K.N. Avery, J.E. Schaak, R.E. Schaak,

M13 bacteriophage as a biological scaffold for magnetically-recoverable metal nanowire catalysts: Combining specific and nonspecific interactions to design multifunctional nanocomposites

. Chem. Mater. 21, 2176-2178 (2009). doi: 10.1021/cm900869u
Baidu ScholarGoogle Scholar
[20] J.H. Hou, Q.G. Du, R.B. Zhong, et al.

Temperature manipulating peptide self-assembly in water nanofilm

. Nucl. Sci. Tech. 25, 060502 (2014). doi: 10.13538/j.1001-8042/nst.25.060502
Baidu ScholarGoogle Scholar
[21] L. Zhang, N. Li, F. Gao, et al.

Insulin amyloid fibrils: an excellent platform for controlled synthesis of ultrathin superlong platinum nanowires with high electrocatalytic activity

. J. Am. Chem. Soc. 134, 11326-11329 (2012). doi: 10.1021/ja302959e
Baidu ScholarGoogle Scholar
[22] X. Zhou, L. Zheng, R. Li, et al.

Biotemplated fabrication of size controlled palladium nanoparticle chains

. J. Mater. Chem. 22, 8862-8867 (2012). doi: 10.1039/c2jm16411f
Baidu ScholarGoogle Scholar
[23] X. Zhou, Y. Wei, X. Zhao, et al.

Palladium nanoparticles supported by amyloid fibrils: From size controllable synthesis to extremely high catalytic performance

. Colloids Surf. A. 482, 416-421 (2015). doi: 10.1016/j.colsurfa.2015.07.007
Baidu ScholarGoogle Scholar
[24] A.F. Lee, J.N. Naughton, Z. Liu, et al.

High-pressure XPS of crotyl alcohol selective oxidation over metallic and oxidized Pd (111)

. ACS Catal. 2, 2235-2241 (2012). doi: 10.1021/cs300450y
Baidu ScholarGoogle Scholar
[25] X. Bian, K. Guo, L. Liao, et al.

Nanocomposites of palladium nanoparticle-loaded mesoporous carbon nanospheres for the electrochemical determination of hydrogen peroxide

. Talanta. 99, 256-261 (2012). doi: 10.1016/j.talanta.2012.05.048
Baidu ScholarGoogle Scholar
[26] K.C. Poon, D.C. Dan, T.D. Vo, et al.

Newly developed stepwise electroless deposition enables a remarkably facile synthesis of highly active and stable amorphous Pd nanoparticle electrocatalysts for oxygen reduction reaction

. J. Am. Chem. Soc. 136, 5217-5220 (2014). doi: 10.1021/ja500275r
Baidu ScholarGoogle Scholar
[27] D. Kurouski, T. Deckert-Gaudig, V. Deckert, et al.

Surface characterization of insulin protofilaments and fibril polymorphs using tip-enhanced Raman spectroscopy (TERS)

. Biophys. J. 106, 263-271 (2014). doi: 10.1016/j.bpj.2013.10.040
Baidu ScholarGoogle Scholar
[28] S.Y. Lee, J.S. Lim, M.T. Harris,

Synthesis and application of virus-based hybrid nanomaterials

. Biotechnol. Bioeng. 109, 16-30 (2012). doi: 10.1002/bit.23328
Baidu ScholarGoogle Scholar
[29] C. Yang, C.H. Choi, C.S. Lee, et al.

A facile synthesis-fabrication strategy for integration of catalytically active viral-palladium nanostructures into polymeric hydrogel microparticles via replica molding

. ACS Nano. 7, 5032-5044 (2013). doi: 10.1021/nn4005582
Baidu ScholarGoogle Scholar
[30] J. Huang, D. Wang, H. Hou, et al.

Electrospun palladium nanoparticle-loaded carbon nanofibers and their electrocatalytic activities towards hydrogen peroxide and NADH

. Adv. Funct. Mater. 18, 441-448 (2008). doi: 10.1002/adfm.200700729
Baidu ScholarGoogle Scholar
[31] W. Hong, J. Wang, E. Wang,

Facile synthesis of highly active PdAu nanowire networks as self-supported electrocatalyst for ethanol electrooxidation

. ACS Appl. Mater. Interfaces. 6, 9481-9487 (2014). doi: 10.1021/am501859k
Baidu ScholarGoogle Scholar
[32] M.S. Ahmed, S. Jeon,

Synthesis and electrocatalytic activity evaluation of nanoflower shaped Ni-Pd on alcohol oxidation reaction

. J. Electrochem. Soc. 161, F1300-F1306 (2014). doi: 10.1149/2.1041412jes
Baidu ScholarGoogle Scholar
[33] F. Yang, S.C. Kung, M. Cheng M, et al.

Smaller is faster and more sensitive: The effect of wire size on the detection of hydrogen by single palladium nanowires

. ACS Nano. 2010, 4, 5233-5244 (2010). doi: 10.1021/nn101475c
Baidu ScholarGoogle Scholar
[34] Y. Liu, G. Sun, C. Jiang, et al.

Highly sensitive detection of hydrogen peroxide at a carbon nanotube fiber microelectrode coated with palladium nanoparticles

. Microchim. Acta. 181, 63-70 (2013). doi: 10.1007/s00604-013-1066-8
Baidu ScholarGoogle Scholar
[35] F. Jiang, R. Yue, Y. Du, et al.

A one-pot 'green' synthesis of Pd-decorated PEDOT nanospheres for nonenzymatic hydrogen peroxide sensing

. Biosens. Bioelectron. 44, 127-131 (2013). doi: 10.1016/j.bios.2013.01.003
Baidu ScholarGoogle Scholar
[36] S. Momeni, I. Nabipour,

A simple green synthesis of palladium nanoparticles with Sargassum alga and their electrocatalytic activities towards hydrogen peroxide

. Appl. Biochem. Biotechnol. 176, 1937-1949 (2015). doi: 10.1007/s12010-015-1690-3
Baidu ScholarGoogle Scholar
[37] S. Chakraborty, C.R. Retna,

Electrocatalytic performance of carbon nanotube-supported palladium particles in the oxidation of formic acid and the reduction of oxygen

. Carbon. 48, 3242-3249 (2010). doi: 10.1016/j.carbon.2010.05.014
Baidu ScholarGoogle Scholar
[38] S. Kumar, S. Zou,

Electrooxidation of carbon monoxide on gold nanoparticle ensemble electrodes: Effects of particle coverage

. J. Phys. Chem. B. 109, 15707-15713 (2005). doi: 10.1021/jp051981a
Baidu ScholarGoogle Scholar