DOI:
10.1039/C4RA07575G
(Paper)
RSC Adv., 2014,
4, 59740-59746
Characterization of Li-doped WO3 nanowires and their enhanced electrocatalytic oxidation of ascorbic acid†
Received
24th July 2014
, Accepted 14th October 2014
First published on 14th October 2014
Abstract
Li-doped WO3 nanowires have been hydrothermally prepared and characterized mainly via spectroscopic methods. Both the hexagonal structure distortion and morphology evolution induced by Li doping reveal a lattice expansion of about 0.07 Å. Also, the residence of oxygen vacancy and the enlargement of external surface areas positively correlate with the narrowing of energy band. Subsequently, the electrocatalytic oxidation of ascorbic acid using a Li-doped WO3 film-coated electrode performs a 13% and 21% negative shift of the oxidation overpotential compared with a WO3 film-coated electrode and a bare glassy carbon electrode, respectively. A preliminary mechanism has been proposed on the basis of relevant model analyses.
1. Introduction
As an indispensable ingesta that participates in many human physiological biochemical reactions, the detection of ascorbic acid (AA) in human body fluid is of significance in assessing the status of health. An electrochemical sensor adopting a variety of chemically-modified electrodes renders a progressive technology for overcoming the disadvantages of high overpotential, poor reproducibility, low selectivity and sensitivity on conventional electrodes.1–4 Thus, much interest has been focused on the use of mediators and modified electrodes to catalyze the oxidation of AA.5,6
It has been well recognized that oxide-based semiconductors display excellent electrocatalytic activity due to their unique physicochemical properties.7–14 However, the concentration of free electrons in metal oxide semiconductors is low, i.e., ∼1021 electrons per m−3.14 A valid approach to enhancing the electrocatalytic performance is metal doping that achieves a significant raise in free electron concentration exemplified by Nb-doped TiO2/carbon composite and Li-doped tantalum oxide,15–17 wherein the electron conductivity and electrocatalytic activity of both improve.
Tungsten oxide (WO3) is a versatile wide band gap metal oxide that finds extensive application in electrochromic, photochromic and photocatalytic devices.18–25 Besides, WO3 can be utilized as an electrode modifier due to its high chemical stability and considerable semiconductivity. Thus, the electrocatalytic property of WO3 in the morphology of nanorods, nanowires and its aggregates has been highlighted.26 Compared with pristine WO3, metal-doping can further improve the electrochemical properties of WO3. As exemplified, Bathe et al. showed that the cyclic stability, charge storage capacity, and reversibility could be improved by addition of Nb2O5 to WO3 films.27 Also, Ti-doped WO3 film has improved electrical conductivity and reaction kinetics.28 However, the effect of doping on the electroactivity of WO3 is rarely studied. We propose that development of a highly efficient Li-doped WO3 electrocatalyst and its application in the electrochemical detection of ascorbic acid in biological systems will be of interest to a broad range of material scientists, as Li ions doping can affect the oxygen vacancies, energy band gap and surface area of WO3.
In our previous work, we successfully prepared Ta-doped tungsten oxide nanowires and studied their electrocatalytic activity for hydrogen evolution reaction.29 In this paper, Li-doped WO3 film has been fabricated and homogenously deposited on a glassy carbon electrode. A test of its electrocatalytical oxidation of AA has been carried out to investigate the effect of experimental parameters of cyclic voltammetry in order to explore the mechanism.
2. Experimental
2.1. Chemical reagents and synthesis of Li-doped WO3 nanowires
Sodium tungstate dihydrate (NaWO4·2H2O, analytical grade) was purchased from Kelong Company, lithium nitrate (LiNO3, 99.9%, analytical grade) was purchased from Sigma-Aldrich, and ammonium sulfate ((NH4)2SO4) was received from Tianjin Chemical Reagents Corporation (China). All other chemicals were purchased from the Chengdou Chemical Reagents Corporation (China).
The synthetic method of Li-doped WO3 is as follows: 2 g of Na2WO4·2H2O and LiNO3 (at varying Li/W mole ratios of 0, 0.01, 0.03, 0.05, 0.07, and 0.1) was dissolved in 45 mL of deionized water under stirring at room temperature, then 5 mL of 3 mol L−1 HCl solution was added to the above solution under continuous stirring until tungstenic acid was heavily precipitated, and finally 30 mL of 0.5 mol L−1 (NH4)2SO4 solution was added to this solution, which was then transferred to a Teflon-lined autoclave with a capacity of 100 mL. Hydrothermal treatments were carried out at 200 °C for 24 h. After that, the autoclave was allowed to cool naturally. The final products were collected, washed with deionized water and ethanol several times, and dried in air at 80 °C. The Li-doped WO3 nanowires were finally obtained.
2.2. Characterization of Li-doped WO3 nanowires
Powder X-ray diffraction (PXRD) analysis was conducted to characterize the crystalline phase of the Li-doped WO3 samples on a X'Pert PRO PANalytical diffractometer (Almelo, Netherlands) with Cu Kα radiation (λ 0.15406 nm), operating at 40 kV and 45 mA. The size and shape of the samples were observed on a field-emission scanning electron microscope (FESEM, Philips XL30 FEG, Eindhoven, Netherlands) and transmission electron microscope (TEM, JEM200CX, 120 kV). X-ray photoelectron spectroscopy (XPS) was performed on a RBD upgraded PHI-5000C ESCA system (PerkinElmer) using an Mg-monochromatic X-ray at a power of 25 W and an X-ray-beam diameter of 10 mm, and a pass energy of 29.35 eV. The binding energy was calibrated using the C1s hydrocarbon peak at 284.8 eV. UV-vis diffuse reflectance spectroscopy was carried out on an UV-vis spectrophotometer (Lambda 850, PerkinElmer, USA), equipped with an integrating sphere and a BaSO4 reference. Nitrogen adsorption and desorption isotherms were measured at 77 K with a Beckman Coulter SA 3100 surface area analyzer. To determine the surface area, the Brunauer–Emmett–Teller (BET) method was used. The doped amount of lithium in WO3 was measured with inductively coupled plasma (VISTA-MPX CCD Simultaneous ICP-OES, Varian, USA).
2.3. Electrochemical measurements
The Li-doped WO3 nanowires were further characterized using cyclic voltammetry (CV) in 0.1 mol L−1 KCl containing 0.4 mmol L−1 AA as the electrolyte solution. Prior to the electrochemical measurements, high-purity Ar2 was bubbled in the above solution for 20 min. During the experiments, high-purity Ar2 was continually bubbled at the surface of the electrolyte solution. The fabrication of the Li-doped WO3 film- and pure WO3 film-coated electrodes is as follows: 20 mg of Li-doped WO3 or WO3 powder was dispersed ultrasonically in a mixture of 1.0 mL of ethanol and 8 μL of nafion solution to obtain a suspension. The suspension was spread on the glassy carbon electrode (GCE), and dried for 10 min at 25 °C. This film-coated GCE was used as the working electrode. A Pt coil and an Ag/AgCl electrode (3 mol L−1 KCl) were used as the counter and reference electrodes, respectively. The CV measurements were carried out on a CHI 660C electrochemical workstation (Shanghai Chenhua Instrument Factory, China) at room temperature (22 ± 2 °C).
3. Results and discussion
3.1. Structure and morphology of the pristine and Li-doped WO3
The XRD patterns of pristine and Li-doped WO3 presented in Fig. 1 demonstrate a hexagonal phase of good crystallinity with lattice constants a = 7.299 Å and c = 3.899 Å (JCPDS Card no. 33-1387). No lithium-bearing impurities were found and the diffraction peak of (001) facet shifts to lower degree as the lithium doping amount increases. The observations indicate that the lithium ion has been incorporated into the WO3 lattice and therefore induced local distortions as a result of a slight lattice expansion.
 |
| Fig. 1 XRD patterns of the Li-doped WO3 nanowires at varying Li/W molar ratios of (a)–(f) 0, 0.01, 0.03, 0.05, 0.07, and 0.10, respectively. | |
Both pristine and Li-doped WO3 illustrated in Fig. 2 unfold coral-like tangled nanowires forming accessible hierarchical pores. The diameter of those nanowires is in the range of 20 to 30 nm while the length of Li-doped nanowires is roughly longer than that of the pristine nanowires. The high-resolution transmission electron microscopy presented in Fig. 2c and f indicated a marked fringe with the spacing of 0.375, 0.382 nm for the pristine and doped, respectively. Since the labeled fringe corresponds to the (001) facet, the lattice spacing expands 0.07 Å according to the XRD analysis results. Both crystal structure and morphology representations imply that a lattice defect induced by the substitution of W5+/W6+ with Li+ proceeds to cause local lattice distortion and possibly incurs more exposed oxygen vacancies.
 |
| Fig. 2 (a) SEM, (b) TEM and (c) HRTEM images of the pristine WO3 nanowires. (d) SEM, (e) TEM and (f) HRTEM images of Li-doped WO3 nanowires at a Li/W molar ratio of 0.05. | |
3.2. Oxygen vacancies, energy band gap and surface area of the pristine and Li-doped WO3
A survey of the oxide-based electrocatalyst revealed that the oxygen vacancies, energy band level, and surface textures of the electrode materials regulate their electrocatalytic performance.30 In this part, the overall XPS spectra of Li-doped WO3 presented in Fig. 3a demonstrates the residence of W, O, and Li elements, the contents of which identify with the composition of the hydrothermal mixture. The O1s XPS spectra of pristine and Li-doped WO3 illustrated by Fig. 3b and c feature an asymmetrical peak. It has been deconvoluted into two separate peaks using Gaussian distributions,31 which are lattice O2− at 530.30 eV and surface OH− at 531.10 eV. The intensity of the hydroxyl group is significantly higher for the Li-doped WO3 compared with the pristine WO3. Although the OH− resides as a minor oxygen vacancy, it is beneficial for inhibiting the electron-hole recombination process, thereby improving the electron transmission efficiency,32,33 further enhancing the electrocatalytic activities of WO3. Moreover, the real ratios of Li/W in Li-doped WO3 samples were measured by ICP-OES (see ESI, Table 1†).
 |
| Fig. 3 XPS spectra of (a) Li-doped WO3 (Li/W = 0.05). XPS O1s spectra of (b) undoped and (c) doped WO3. | |
The energy band level has been assessed via conventional UV-vis spectroscopy. As illustrated in Fig. 4, the maximum reflectance shifts to higher wavelength after Li doping. The band gap energy (Eg) of the semiconductor can be determined accordingly:
where
α is the absorption coefficient,
hν is the incident photon energy,
A is a constant, and
n is either 2 or 1/2 for direct and indirect transitions, respectively.
34,35 It is known that WO
3 crystal is an indirect-gap semiconductor, thus a
n value of 2 is assumed here. The band gaps determined by the linear extrapolation presented in the inset are approximately 3.05 and 2.88 eV for the pristine and Li-doped WO
3 (Li/W mole ratio of 0.05), respectively. The narrowing of Li-doped WO
3 energy band gap could be interpreted in the two mentioned aspects: (1) the local lattice distortion possibly changes the electronic structure and carrier density of WO
3, which results in the improvement of its electrocatalytic activity;
14,17 (2) the oxygen vacancies incur a defect band with a low energy.
29
 |
| Fig. 4 UV-vis reflectance spectra of undoped and Li-doped WO3 (Li/W = 0.05). The inset is a plot of (αhν)1/2 as a function of hν for the undoped and doped WO3 samples. | |
The active surface area of the nanoparticles that is also crucial to determining the electrocatalytic property was evaluated. The nitrogen adsorption–desorption isotherms of the pristine and Li-doped WO3 nanowires are shown in Fig. 5. The BET surface areas of the pristine and Li-doped WO3 nanowires were determined as 19.42 and 74.92 m2 g−1, respectively. As observed, WO3 featured an enhanced surface area following doping with Li ions, indicative of the presence of a larger number of active sites that would be beneficial towards improving the electrocatalytic activity of the doped sample.
 |
| Fig. 5 Nitrogen sorption isotherm curves of undoped WO3 and Li-doped WO3 (Li/W = 0.05). | |
3.3. Electrocatalytic activity of the Li-doped WO3 film-coated electrode
In this section, the electrocatalytic experimental parameters including the type of working electrode, the concentration of AA and the scan rate have been adopted to evaluate the electrocatalytic activity of the Li-doped WO3 film-coated electrode. The cyclic voltammogram illustrated by Fig. 6 left displays an anodic peak (Ep) at 0.57 V (Curve a), 0.52 V (Curve c), 0.45 V (Curve b) for the bare GCE, WO3 film-coated electrode and Li-doped WO3 film-coated electrode, respectively. Thus, there is a negative shift of 0.12 V of Ep on the Li-doped WO3 film-coated electrode. The negative shift of the anodic peak position indicates that the Li-doped WO3 film-coated electrode has excellent electrocatalytic activity toward AA oxidation. Meanwhile, the Li doping amount has been optimized as a Li/W molar ratio of 0.05 since its anodic current is the maximum among those different doping ratios (see Fig. 6 right).
 |
| Fig. 6 (Left) Cyclic voltammograms of 4 mmol L−1 AA adopting (a) bare GCE, (b) Li-doped WO3, and (c) WO3 electrode in 0.1 mol L−1 KCl solution. (Right) Variation of anodic current as a function of the Li/W molar ratios. | |
The Li-doped WO3 film-coated electrode with the optimum doping amount is selected and the effect of AA concentration and scan rate on the anodic current have been assessed in Fig. 7. Fig. 7a shows the cyclic voltammograms of AA at varying concentrations in 0.1 mol L−1 KCl solution using the Li-doped WO3 film-coated electrode. As observed, the oxidation peak current increased with increasing AA concentrations. The inset of Fig. 7a shows a linear relationship between the anodic peak current and AA concentration in the range of 2–8 mmol L−1, with a correlation coefficient of 0.998. At a fixed potential of 0.5 V the catalytic current observed was also linear dependent on the AA concentration in the range 0.01–4 mmol L−1 and the detection limit was 0.04 mmol L−1. Furthermore, the as-prepared Li-doped WO3 modified electrode is compared with other modified electrodes on aspects such as linear range and detection limit, which are shown in Table 1; this information demonstrates that the sensor behaves well with a wide linear range and low detection limit towards the oxidation of AA.
 |
| Fig. 7 The concentrations of AA studied are (a)–(d) 2, 4, 6, and 8 mmol L−1, respectively. The different potential scan rates of (a)–(f) 10, 30, 50, 70, 100 and 150 mV s−1, respectively. | |
Table 1 Comparison of electrochemical parameters of the Li-doped WO3 modified GCE with those of other modified electrodes reported in the literature
Modified electrode |
Linear range (μmol L−1) |
Detection limit (μmol L−1) |
Reference |
Li-doped Ta2O5/GCE |
5000–12 000 |
— |
17 |
Li doped Bi2O3/CNT/GCE |
20–5000 |
50 |
36 |
Fe3O4@Au-s-Fc/GS-chitosan/GCE |
6–350 |
5 |
37 |
MgO nanobelts/GCE |
25–150 |
0.2 |
38 |
Nano-Cu/PSA III/GCE |
0.3–730 |
0.15 |
39 |
Nano-CuO/GCE |
0.1–3100 |
0.095 |
40 |
Nano-NiO/GCE |
50–3500 |
860 |
41 |
Li-doped WO3/GCE |
2000–8000 |
40 |
This work |
The cyclic voltammograms of the Li-doped film-coated electrode at various scan rates in the presence of 4 mmol L−1 AA in 0.1 mol L−1 KCl (Fig. 7b) reveal that the catalytic effect of the Li-doped film-coated electrode appears at the higher scan rates because of the considerably higher catalytic reaction rate. Moreover, the catalytic oxidation peak potential shifts to more positive values with increasing scan rate, indicative of a kinetics limitation in the reaction between the redox sites of Li-doped WO3 film and AA. This result also proves that the electrochemical oxide reaction of AA is an irreversible process. The inset of Fig. 7b shows the linear correlation between the anodic peak currents of AA and the square-root of scan rates, indicative of a diffusion controlled electrode process.
In order to explore the diffusion mechanism, two models have been adopted to determine the rate-limiting step of the process, which are the scan rate-normalized current (I/v1/2) as a function of scan rate (v) and Ep as a function of logarithm of the scan rate (log
v). The fitting results are illustrated in Fig. 8. The shape of the curve in Fig. 8a is typical of an electrochemical–chemical (EC) process.42 To obtain information on the rate determining step, the following function involving Ep and logarithm of the scan rates (log
v) for an irreversible diffusion-controlled process was assessed:43
|
 | (2) |
where
K is a constant,
α is the transfer coefficient,
nα is the number of electrons transferred,
v is the scan rate, and
b is the Tafel slope.
Fig. 8b shows the variation of
Ep with log
v. The slope of the
Ep–log
v plot was determined as 0.067 and a Tafel slope of 0.134 was obtained. The value of 2.303
RT/
αnαF is equivalent to the Tafel slope. The value of the transfer coefficient was then determined to be 0.429. Another method that is used for calculating the transfer coefficient (
α) was also employed, using the following equation for an irreversible system:
17,44 |
 | (3) |
where
na is the number of electrons transferred during the electrochemical process. The transfer coefficients obtained from the above fitting are in good agreement and indicate that the electrocatalytic oxidation of AA is an electrochemical–chemical (EC) process, simultaneously controlled by diffusion of solution AA and cross-exchange through the Li-doped WO
3 film.
 |
| Fig. 8 (a) Plot of the anodic peak current (I/ν1/2) as a function of scan rate (ν). (b) Dependence of the peak potential Ep on log ν for the oxidation of AA on Li-doped WO3 film-coated electrode. | |
4. Conclusions
Creating an electron deficit by substituting some of the cations of the host oxide with cations having a lower valence is a versatile approach to tailor catalysts. Herein, the insertion of Li+ into the WO3 causes a lattice distortion, surface area enlargement and oxygen vacancy formation. The integrated effect of those structure determinants facilitates the oxidation of AA on the electrocatalyst surface, enhancing its electrocatalytic performance. The anodic peak potential of AA shifted from 0.57 V (versus Ag|AgCl) on a bare GCE to 0.45 V on the lithium–WO3 film coated electrode. Despite the preliminary results demonstrating that the Li-doped WO3 film-coated electrode is a promising electrochemical sensor for the detection of AA, the hydrothermal synthesis factors dominating the crystal structure and morphology, the advanced characterizations combined with quantum computation, and electrocatalytic experiments carried out in complex simulant bearing AA are in need of in-depth investigation for the development of a highly selective, sensitive and effective electrochemical detector.
Acknowledgements
This work was financially supported by the Institute of Nuclear Chemistry and Physics, China Academy of Engineering Physics.
References
- L. Fernandez and H. Carrero, Electrochim. Acta, 2005, 50, 1233 CrossRef CAS PubMed.
- N. F. Atta and M. F. El-Kady, Anal. Biochem., 2010, 400, 78 CrossRef CAS PubMed.
- M. G. Hosseini, M. Faraji and M. M. Momeni, Thin Solid Films, 2011, 519, 3457 CrossRef CAS PubMed.
- L. Zhang, Electrochim. Acta, 2007, 52, 6969 CrossRef CAS PubMed.
- L. Zhang and S. J. Dong, J. Electroanal. Chem., 2004, 568, 189 CrossRef CAS PubMed.
- M. H. Pournaghi-Azar and H. Razmi-Nerbin, J. Electroanal. Chem., 2000, 488, 17 CrossRef CAS.
- E. Tsuji, A. Imanishi, K. Fukui and Y. Nakato, Electrochim. Acta, 2011, 56, 2009 CrossRef CAS PubMed.
- J. Masud, M. T. Alam, Z. Awaludin, M. S. El-Deab, T. Okajima and T. Ohsaka, J. Power Sources, 2012, 220, 399 CrossRef CAS PubMed.
- A. Moulahi, F. Sediri and N. Gharbi, Mater. Res. Bull., 2012, 47, 667 CrossRef CAS PubMed.
- B. Klapste, J. Vondrak and J. Velicka, Electrochim. Acta, 2002, 47, 2365 CrossRef CAS.
- D. Das, P. K. Sen and K. Das, Electrochim. Acta, 2008, 54, 289 CrossRef CAS PubMed.
- Y. J Feng and Y. H. Cui, Water Res., 2003, 37, 2399 CrossRef.
- J. Zhoo, N. S. Xu, S. Z. Deng, J. Chen and J. C. She, Chem. Phys. Lett., 2003, 382, 433 Search PubMed.
- Q. Q. Du, X. F. Cao, B. Chi, J. Zhang, C. M. Zhang, T. Liu, X. J. Wang and Y. G. Su, J. Electroanal. Chem., 2012, 681, 139 CrossRef PubMed.
- Y. H. Zhao, W. Y. Wang, Q. Y. Jia, Y. Gao, X. J. Wang and Y. G. Su, J. Inn. Mong. Univ., 2010, 41, 307 CAS.
- K. Senevirathne, V. Neburchilov, V. Alzate, R. Baker, R. Neagu, J. Zhang, S. Camp-bell and S. Ye, J. Power Sources, 2012, 220, 1 CrossRef CAS PubMed.
- Y. H. Zhao, C. Y. Li, W. Zhao, Q. Q. Du, B. Chi, J. J. Sun, Z. L. Chai and X. J. Wang, Electrochim. Acta, 2013, 107, 52 CrossRef CAS PubMed.
- X. J. Comini, C. Baratto, G. Faglia, M. Ferroni, A. Vomiero and G. Sberveglieri, Prog. Mater. Sci., 2009, 54, 1 CrossRef PubMed.
- Y. Xie, F. C. Cheong, B. Varghese, Y. W. Zhu, R. Mahendiran and C. H. Sow, Sens. Actuators, B, 2011, 151, 320 CrossRef CAS PubMed.
- J. Yu, L. Qi, B. Cheng and X. Zhao, J. Hazard. Mater., 2008, 160, 621 CrossRef CAS PubMed.
- L. E. Bevers, L. E. Hagedoorn and W. R. Hagen, Coord. Chem. Rev., 2009, 253, 269 CrossRef CAS PubMed.
- S. Salmaoui, F. Sediri, N. Gharbi, C. Perruchot and M. Jouini, Electrochim. Acta, 2013, 108, 634 CrossRef CAS PubMed.
- C. Santato, M. Odziemkowski, M. Ulmann and J. Augustynski, J. Am. Chem. Soc., 2001, 123, 10639 CrossRef CAS PubMed.
- R. Solarska, A. Krolikowska and J. Augustynski, Angew. Chem., Int. Ed., 2010, 49, 7980 CrossRef CAS PubMed.
- J. Z. Su, X. J. Feng, J. D. Sloppy, L. J. Guo and C. A. Grimes, Nano Lett., 2011, 11, 203 CrossRef CAS PubMed.
- J. Rajeswari, P. S. Kishore, B. Viswanathan and T. K. Varadarajan, Nanorods Nanoscale Res. Lett., 2007, 2, 496 CrossRef CAS.
- S. R. Bathe and P. S. Patil, J. Phys. D: Appl. Phys., 2007, 40, 74231 CrossRef.
- G. F. Cai, X. L. Wang, D. Zhou, J. H. Zhang, Q. Q. Xiong, C. D. Guo and J. P. Tu, RSC Adv., 2013, 3, 6896 RSC.
- X. Xie, W. J. Mu, X. L. Li, H. Y. Wei, Y. Jian, Q. H. Yu, R. Zhang, K. Lv, H. Tang and S. Z. Luo, Electrochim. Acta, 2014, 134, 201 CrossRef CAS PubMed.
- R. S. Mulliken, J. Chem. Phys., 1955, 23, 1841 CrossRef CAS PubMed.
- E. A. Davis and N. F. Mott, Philos. Mag., 1970, 22, 903 CrossRef CAS.
- J. Tauc, A. Menth and D. L. Wood, Phys. Rev. Lett., 1970, 25, 749 CrossRef CAS.
- S. Polarz, J. Strunk, V. Ischenko, M. W. E. Van den Berg, O. Hinrichsen, M. Muhler and M. Driess, Angew. Chem., Int. Ed., 2006, 45, 2965 CrossRef CAS PubMed.
- H. T. Sun, C. Cantalini, L. Lozzi, M. Passacantando, S. Santucci and M. Pelino, Thin Solid Films, 1996, 287, 258 CrossRef CAS.
- F. S. Manciu, J. L. Enriquez, W. G. Durrer, Y. Yun, C. V. Ramana and S. K. Gullapalli, J. Mater. Res., 2010, 25, 2401 CrossRef CAS.
- Z. Mohammed, T. W. Tan, Z. Zainal, A. H. Abdullah and J. K. Goh, Int. J. Electrochem. Sci., 2010, 5, 501 Search PubMed.
- M. L. Liu, Q. Chen, C. L. Lai, Y. Y. Zhang, J. H. Deng, H. T. Li and S. Z. Yao, /Biosensors and Bioelectronics, 2013, 48, 75 CrossRef CAS PubMed.
- M. J. Li, W. L. Guo, H. J. Li, W. Dai and B. Yang, Sens. Actuators, B, 2014, 204, 629 CrossRef CAS PubMed.
- L. Zhang, W. J. Yuan and B. Q. Hou, J. Electroanal. Chem., 2013, 689, 135 CrossRef CAS PubMed.
- C. X. Wang, J. Liu, X. Huang, H. H. Wang, Y. D. Zheng, L. Li, S. Y. Wang, S. Chen and Y. Jin, Appl. Surf. Sci., 2014, 292, 291 CrossRef CAS PubMed.
- C. Xia, X. Yanjun and W. Ning, Sens. Actuators, B, 2011, 153, 434 CrossRef PubMed.
- A. P. Shpak, A. M. Korduban, M. M. Medvedskij and V. O. Kandyba, J. Electron Spectrosc. Relat. Phenom., 2007, 156, 172 CrossRef PubMed.
- G. Erdogdu and A. E. Karagozler, Talanta, 1997, 44, 2011 CrossRef CAS.
- A. J. Bard and L. R. Faulkner, Electrochemical Methods: Fundamentals and Applications, New York, 2nd edn, 2011 Search PubMed.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra07575g |
|
This journal is © The Royal Society of Chemistry 2014 |
Click here to see how this site uses Cookies. View our privacy policy here.