Next Article in Journal
Metabolomic Fingerprinting of Potato Cultivars Differing in Susceptibility to Spongospora subterranea f. sp. subterranea Root Infection
Previous Article in Journal
Aptamer Cocktail to Detect Multiple Species of Mycoplasma in Cell Culture
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of the Oxygen Evolution Reaction at Strontium Palladium Perovskite Electrocatalyst in Acidic Medium

by
Areej A. Eskandrani
1,
Shimaa M. Ali
1,2,* and
Hibah M. Al-Otaibi
1
1
Department of Chemistry, Faculty of Science, Taibah University, Madinah 3002, Saudi
2
Department of Chemistry, Faculty of Science, Cairo University, Giza 12613, Egypt
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(11), 3785; https://doi.org/10.3390/ijms21113785
Submission received: 2 May 2020 / Revised: 17 May 2020 / Accepted: 18 May 2020 / Published: 27 May 2020
(This article belongs to the Section Physical Chemistry and Chemical Physics)

Abstract

:
The catalytic activity of Sr2PdO3, prepared through the sol-gel citrate-combustion method for the oxygen evolution reaction (OER) in a 0.1 M HClO4 solution, was investigated. The electrocatalytic activity of Sr2PdO3 toward OER was assessed via the anodic potentiodynamic polarization and electrochemical impedance spectroscopy (EIS). The glassy carbon modified Sr2PdO3 (GC/Sr2PdO3) electrode exhibited a higher electrocatalytic activity, by about 50 times, in comparison to the unmodified electrode. The order of the reaction was close to unity, which indicates that the adsorption of the hydroxyl groups is a fast step. The calculated activation energy was 21.6 kJ.mol−1, which can be considered a low value in evaluation with those of the reported OER electrocatalysts. The Sr2PdO3 perovskite portrayed a high catalyst stability without any probability of catalyst poisoning. These results encourage the use of Sr2PdO3 as a candidate electrocatalyst for water splitting reactions.

1. Introduction

Electrochemical water splitting is the process of the decomposition of water into its components, hydrogen and oxygen, by applying an electric current in the presence of an electrolyte. It provides a green alternative route for producing hydrogen fuel, which can be utilized instead of non-renewable and environmentally harmful fossil fuels [1,2,3,4]. The oxygen evolution reaction (OER) and the anodic reaction of the water splitting process are the major challenges facing researchers. The kinetics of the OER involve complicated multi-electron steps with the prospective existence of several accompanying processes such as the dissolution of the metal oxide catalyst. Thus, the multitude and complexity of the OER mechanism can create difficulty in comprehending it clearly [5,6]. In addition, the OER requires the application of high overpotentials [7,8,9], leading to a decrease in the efficiency of the hydrogen production by the water splitting. The most common catalysts for the OER are IrO2/RuO2, Pt, and HfN which can be efficiently employed but with low stability and high cost [10,11]. Nowadays, researchers seek multi-functional catalysts, which are abundant, cost-effective, stable, and of high catalytic activity. Perovskites, mixed metal oxides possessing the general formula ABO3 where A is a lanthanide and B is a transition metal, have been extensively used as effectual catalysts for OER in alkaline [12,13,14,15,16,17,18,19] and acidic media [20,21,22,23,24]. Lu and co-workers have reported that the LaCoO3 perovskite demonstrated high electrocatalytic behavior due to its unique electronic structure, which facilitates the presence of the oxygen vacancies in a high concentration [25]. Oxygen vacancies can highly increase the intrinsic activity of the perovskite active sites for the electro-oxidation reactions [26,27,28,29,30,31].
The SrPdO3 perovskite, prepared for the first time by the citrate combustion methodology by our group, presented interesting electrocatalytic and sensing applications [32]. Furthermore, SrPdO3 has been used as an efficient catalyst for the hydrogen evolution reaction [32], which was subsequently implemented for the first time by our group as an electrochemical sensor for several neurotransmitters and hydrazine detection [33,34,35]. It was discovered that the nano-dispersed Pd sites within the stable perovskite matrix are more effectual than the electro-deposited Pd. Recently, the JCPDS card of the SrPdO3 perovskite (00-025-0908) has been replaced by the JCPDS card of Sr2PdO3 (00-028-1249). It was reported that the partial substitution at the A-site of Sr2PdO3 and the formation of the nano-composites of the Sr2PdO3/carbon nanotubes and Sr2PdO3/gold nanoparticles can further enhance its electro-sensing ability for glucose and some drugs [36,37,38]. The study of the electronic structure of SrPdO3 disclosed that it can exhibit a spin transition from low to high at a certain temperature, which is similar to the isoelectronic LaCoO3 [39], and is one of the main reasons for the enhanced electrocatalytic activity of LaCoO3 for the OER.
In this work, Sr2PdO3, prepared by the sol-gel citrate-nitrate combustion procedure, was employed as an electrocatalyst for the OER in an acidic medium. The electrocatalytic activity was examined by the potentiodynamic polarization (PP) and electrochemical impedance spectroscopy (EIS). The kinetic and thermodynamic evaluations were performed to identify the reaction order and the activation energy, respectively. The catalyst stability was also investigated to ensure a satisfactory performance.

2. Results and Discussion

2.1. Structural and Surface Characterizations of Sr2PdO3

Figure 1A illustrates the XRD spectrum of the strontium palladium perovskite, prepared by the citrate-combustion method at a pH value of 2 and at a calcination temperature of 750 °C for 3 h. The formation of the orthorhombic Sr2PdO3 perovskite is confirmed by the appearance of its theoretical major peak (110), JCPDS card number: 00-028-1249. However, the major phase of the fashioned sample is SrPd3O4, which is verified via the XRD pattern, displays a major peak at (210), and agrees with the theoretical value, Figure 1A. The formation of SrPd3O4, as a secondary phase, during the synthesis of Sr2PdO3, is well reported, due to the self-regeneration property of the Pd-based perovskites [40]. However, in the prepared sample, SrPd3O4 is the major phase; this is possibly because of the use of Pd2+ rather than the Pd4+ salt as a precursor during the synthesis. The calculated average particle size, according to the Scherrer equation [41], is 24.2 nm. The value of the measured BET surface area is 5.0 m2 g−1.
The morphology of the prepared Sr2PdO3 is observed by the TEM, Figure 1B. It is witnessed that the prepared perovskites consist of agglomerations of orthorhombic particles.

2.2. Examination of the Electrocatalytic Activity of Sr2PdO3 for OER

The electrocatalytic behavior of Sr2PdO3, prepared by the citrate combustion approach for the OER in a 0.1 M HClO4 solution, is appraised by anodic PP. Figure 2 reveals PP curves for the bare GC and GC electrode, modified with the Sr2PdO3 perovskite, in the potential range of 0.9 to 1.5 V vs. Ag/AgCl. It can be detected that the current density, due to the OER, is increased by about 50 times in the presence of the Sr2PdO3 perovskite at a definite potential value of 1.0 V vs. Ag/AgCl. Giordano et al. reported the use of HfN as an efficient catalyst for OER, which gave a current density of about 5 mA cm−2 at 1.5 V [11]. Meanwhile, in our case the value of the current density, at 1.5 V, is 8.2 mA cm−2. This can be attributed to the catalytic activity of the Sr2PdO3 perovskite for the OER, resultant from the highly stabilized nano-Pd sites within the perovskite matrix [32,33,34,35], which utilizes its application as an efficient electrocatalyst for water oxidation.
According to the Tafel equation, Equation (1) [42]:
ɳ = a + b log I I o
where ɳ is the overpotential (volt), I is the current density (A cm−2), b is Tafel slope (volt), and a is the intercept which is related to the exchange current density, Io, by the Equations (2) and (3):
a = b log I o
b = 2.3   R T F
where F is Faraday constant (96,485 C mol−1), R is the universal gas constant (8.314 J.mol−1 K−1), T is the temperature in Kelvin, and α is the transfer coefficient.
The calculated values of the Tafel slope from Figure 2 are 366.3 and 454.1 mV for the bare GC and GC electrode modified with Sr2PdO3, respectively. The value of the Tafel slope usually ranges from 30 to 120 mV and can help to determine whether the slowest step of the reaction is from the first or second electron transfer or the recombination step. In our case, the higher Tafel slope values are observed, indicating additional contributions related to the oxide-surface processes, a potential drop in the anode charge layer, or a blockage of the electrode surface by the bubble accumulations [43,44,45,46]. The high Tafel slope values, 200 to 500 mV, were previously reported in literature for the same mentioned reasons [20]. The calculated values of the logarithm of the exchange current density, log Io, which is directly proportional to the rate of the reaction at equilibrium, are −6.2 and −4.4, (Io = 0.6 and 39.8 µA.cm−2) for the bare GC and GC electrode modified with Sr2PdO3, respectively. In other words, the reaction rate at equilibrium is highly increased upon the perovskite modification.

2.3. Determination of the Reaction Order

Figure 3A indicates PP curves for the GC electrode modified with Sr2PdO3 in different concentrations of aqueous HClO4 (0.05 to 0.4 M) at a constant ionic strength by using Na2SO4. This can demonstrate that the rate of the OER is increased with increasing the concentration of the HClO4 solution. The reaction order can be determined from the slope of the logarithm of the current density at a certain potential value, at which a considerable OER is observed, against the pH value of the solution, Figure 3B. The reaction order is 0.81, which is close to unity [20]. This suggests that no further hydroxide ions are adsorbed from the bulk electrolyte before or during the rate-determining step [46,47,48].

2.4. Temperature Effect and Activation Energy Determination

The temperature effect on the rate of the OER at Sr2PdO3 in a 0.1 M HClO4 aqueous solution is performed by recording the PP curves at different temperatures (T), from 298 to 328 K, Figure 4A. The reaction rate is favored with the temperature rise. The value of the activation energy (Ea) can be determined from the Arrhenius plot, Figure 4B, which equals the slope of log Io vs. 1 T plot, multiplied by −2.3 × R. The calculated value of Ea for the OER at Sr2PdO3 is 21.6 kJ mol−1, which can be considered a decent value in evaluation with those reported for the transition metal based perovskites, LaBO3 (B = Fe, Ni, Mn, Co, or Cr) tested in the same electrolyte, Ea values from 11.3 to 552.9 kJ mol−1 [20,21], or those of other OER perovskite catalysts in an alkaline medium, Ea values from 45.1 to 89.7 kJ mol−1 [49,50,51]. This reflects the high electrocatalytic activity for Sr2PdO3 toward the OER. This can be explained based on the matrix effect of the stable nano-perovskite crystal structure, in which nano-Pd sites are highly distributed and stabilized within the matrix [51,52,53,54].

2.5. Stability

The stability of the proposed Sr2PdO3 catalyst for the OER in a 0.1 M HClO4 solution is examined by performing the reaction at a constant potential of 0.5 V and monitoring the current with the operation time. Figure 5 displays the current-time response of the GC electrode casted with Sr2PdO3 in a 0.1 M HClO4 aqueous solution by being subjected to a constant potential of 0.5 V for 150 min. A sharp current decrease is discerned during the first minutes, which arises from the decrease in the capacitive property. Subsequently, the current due to the OER is almost constant with the % decrease in the current at 2.5% after two and a half hours of the catalyst operation. This result reflects the high catalyst stability and excludes any probability of catalyst poisoning. Furthermore, the Sr2PdO3 perovskite presents an enhanced catalytic performance under potentiostatic conditions, as indicated by the increased current value 47.6 µA cm−2 in assessment with the potentiodynamic experiment, 31.6 µA cm−2. This can be explicated on the basis that the catalyst undergoes “self-activation” under potentiostatic conditions [53].

2.6. Electrochemical Impedance Spectroscopy (EIS)

The mechanism by which the OER occurs at the Sr2PdO3 surface can be analyzed by performing the EIS measurements, according to the classical Equation (4); the real part of the impedance, Z’, can be expressed as a function of the frequency ω:
Z = R s + R P + R C T + σ W   ω
where Rs, Rp, and RCT are solution, polarization, and charge-transfer resistances, respectively. σW is the Warburg impedance.
Figure 6 portrays the EIS spectra, in terms of the (A) Bode and (B) Nyquist plots, of the GC electrode casted with Sr2PdO3 in a 0.1 M HClO4 aqueous solution. The selected overpotential is 0.50 V in the case of the Bode plot and different values for the Nyquist plots: 0.45, 0.50, and 0.55 V. The EIS data reveals the presence of two time constants, as indicated by the two semicircles appearing in the Nyquist plots. The first semicircle, at high frequency regions, is related to the polarization and the double layer capacitance at the perovskite/electrolyte interfaces, Rp and CPE1, while the second semicircle, appearing at intermediate and low frequencies, is related to the charge-transfer and adsorption–desorption processes at the perovskite, RCT and CPE2 [54,55,56,57]. The electrical equivalent circuit, illustrated in Figure 6C, is utilized for data fitting and exhibits a good agreement between experimental (circles) and fitted (lines) data. Similar circuits are employed successively for the study of the OER at metals and metal oxides surfaces [58,59]. The electrocatalytic activity of Sr2PdO3 for the OER was explored through the impact of the applied overpotential on the Nyquist plot, Figure 6B. The fitted parameters are presented in Table 1. It can be noted that with increasing the value of the applied overpotential, the value of RCT is decreased, reflecting the high catalytic performance of Sr2PdO3 for the OER. It is worth mention that by increasing the applied overpotential, the diffusion within the perovskite matrix decreases quickly.

3. Materials and Method

3.1. Chemicals

Strontium nitrate (99%), palladium (II) chloride (99%), ammonium hydroxide (28–30% NH3 basis), perchloric acid (ACROS, 70%), nitric acid (70%), citric acid Anhydrous (99.5%), N,N-dimethyl formamide (DMF) (99%), and deionized water. All chemicals were purchased from the Sigma Aldrich Company.

3.2. Synthesis of Sr2PdO3 by the Citrate-Nitrate Combustion Method

The preparation through the sol-gel procedure has been reported in detail in a previous work [32,60,61,62,63], via mixing Sr(II) and Pd(II) ions in the same molar ratio. Subsequently, citric acid was added to the homogeneous metal ions solution, pH = 2. The solution was kept stirred while heating until complete vaporization, and then it was ignited. The resultant powder was calcined at 750 °C for 3 h to obtain a crystalline perovskite phase.

3.3. Electrochemical Cell and Measurements

The electrochemical measurements were performed in a one-compartment, three-electrode cell in which the working electrode was a glassy carbon (GC) electrode (area = 0.071 cm2), the auxiliary electrode was a platinum coil, and a saturated Ag/AgCl (3.5 M) was used as a reference electrode. The modified electrode, GC/Sr2PdO3, was prepared by casting 25 μL of the perovskite suspension in DMF, which was prepared by homogenously mixing 10 mg of perovskite in 1 mL of DMF.
The anodic PP measurements were done in a 0.1 M HClO4 aqueous solution by first operating an open circuit potential experiment for 10 min until a steady state was reached. This was followed by the conditioning of the working electrode under two-step potentiostatic conditions at 0.4 V for 300 s and at 0.5 V for 600 s, respectively. Finally, the appearance of the PP measure occurred from 0.5 to 1.5 V at a scan rate of 5 mV s−1. All the electrochemical parameters were calculated based on the ohmic drop correction.
The EIS was recorded at different applied overpotentials at a frequency range from 100,000 to 0.1 Hz with an amplitude of 5 mV. All the electrochemical measurements were performed employing a Gamry 1000 potentiostat.

3.4. Sample Characterizations

The X-ray diffractogram was acquired by XRD, Shimadzu, XRD-7000, Japan, at 40 kV and 30 mA, utilizing the CuKα incident beam. The transmission electron microscope (TEM) image is taken by the TEM, JEOL JEM 1400, Japan.

4. Conclusions

Sr2PdO3, synthesized through the citrate-combustion methodology, was effectually employed as an electrocatalyst for the OER. The current density was amplified by about 50 times by casting the Sr2PdO3 perovskite on the GC substrate and values of RCT, calculated from the Nyquist plots fitting, and was decreased with the upsurge in the applied overpotentials, reflecting a powerful catalytic performance of Sr2PdO3 for the OER in 0.1 M HClO4 solution. This funding is substantiated by the low values of the calculated activation energy 21.6 kJ mol−1. Sr2PdO3 exhibited a high operation stability with the self-activation property, which promotes its use in water splitting fuel cells.

Author Contributions

S.M.A.: Supervision of catalyst preparation, characterizations and electrochemical experiments. Performing data analysis, and writing manuscript. Revision and approval the submitted version of the manuscript. H.M.A.-O.: Performing catalyst preparation, characterizations and electrochemical experiments. A.A.E.: Revision and approval the submitted version of the manuscript. All authors have read and agree to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hurst, J.K. In Pursuit of Water Oxidation Catalysts for Solar Fuel Production. Science 2010, 328, 315–316. [Google Scholar] [CrossRef] [PubMed]
  2. Yue, X.; Huang, S.; Cai, J.; Jin, Y.; Shen, P.K. Heteroatoms dual doped porous graphene nanosheets as efficient bifunctional metal-free electrocatalysts for overall water-splitting. J. Mater. Chem. A 2017, 5, 7784–7790. [Google Scholar] [CrossRef]
  3. Shu, C.; Kang, S.; Jin, Y.; Yue, X.; Shen, P.K. Bifunctional porous non-precious metal WO2 hexahedral networks as an electrocatalyst for full water splitting. J. Mater. Chem. A 2017, 5, 9655–9660. [Google Scholar] [CrossRef]
  4. Xiao, Z.; Wang, Y.; Huang, Y.C.; Wei, Z.; Dong, C.L.; Ma, J.; Shen, S.; Li, Y.; Wang, S. Filling the oxygen vacancies in Co3O4 with phosphorus: An ultra-efficient electrocatalyst for overall water splitting. Energy Environ. Sci. 2017, 10, 2563–2569. [Google Scholar] [CrossRef]
  5. Feng, J.R.; Lv, F.; Zhang, W.; Li, P.; Wang, K.; Yang, C.; Wang, B.; Yang, Y.; Zhou, J.; Lin, F.; et al. Iridium-Based Multimetallic Porous Hollow Nanocrystals for Efficient Overall-Water-Splitting Catalysis. Adv. Mater. 2017, 29, 1703798. [Google Scholar] [CrossRef]
  6. Pi, Y.; Shao, Q.; Wang, P.; Guo, J.; Huang, X. General Formation of Monodisperse IrM (M = Ni, Co, Fe) Bimetallic Nanoclusters as Bifunctional Electrocatalysts for Acidic Overall Water Splitting. Adv. Funct. Mater. 2017, 27, 1700886. [Google Scholar] [CrossRef]
  7. Fabbri, E.; Habereder, A.; Waltar, K.; Kötz, R.; Schmidt, T.J. Developments and perspectives of oxide-based catalysts for the oxygen evolution reaction. Catal. Sci. Technol. 2014, 4, 3800–3821. [Google Scholar] [CrossRef] [Green Version]
  8. Herranz, J.; Durst, J.; Fabbri, E.; Pătru, A.; Cheng, X.; Permyakova, A.A.; Schmidt, T.J. Interfacial effects on the catalysis of the hydrogen evolution, oxygen evolution and CO2-reduction reactions for (co-)electrolyzer development. Nano Energy 2016, 29, 4–28. [Google Scholar] [CrossRef] [Green Version]
  9. Zeng, K.; Zhang, D. Recent progress in alkaline water electrolysis for hydrogen production and applications. Prog. Energy Combust. Sci. 2010, 36, 307–326. [Google Scholar] [CrossRef]
  10. McCrory, C.C.L.; Jung, S.; Peters, J.C.; Jaramillo, T. Benchmarking Heterogeneous Electrocatalysts for the Oxygen Evolution Reaction. J. Am. Chem. Soc. 2013, 135, 16977–16987. [Google Scholar] [CrossRef]
  11. Defilippi, C.; Shinde, D.V.; Dang, Z.; Manna, L.; Hardacre, C.; Greer, A.J.; D’Agostino, C.; Giordano, C. HfN nanoparticles: An Unexplored Catalyst for the Electrocatalytic Oxygen Evolution Reaction. Angew. Chem. Int. Ed. 2019, 58, 15464–15470. [Google Scholar] [CrossRef] [PubMed]
  12. Suntivich, J.; May, K.J.; Gasteiger, H.A.; Goodenough, J.B.; Shao-Horn, Y. A Perovskite Oxide Optimized for Oxygen Evolution Catalysis from Molecular Orbital Principles. Science 2011, 334, 1383–1385. [Google Scholar] [CrossRef] [PubMed]
  13. Yagi, S.; Yamada, I.; Tsukasaki, H.; Seno, A.; Murakami, M.; Fujii, H.; Chen, H.; Umezawa, N.; Abe, H.; Nishiyama, N.; et al. Covalency-reinforced oxygen evolution reaction catalyst. Nat. Commun. 2015, 6, 8249. [Google Scholar] [CrossRef] [PubMed]
  14. Grimaud, A.; May, K.J.; Carlton, C.E.; Lee, Y.-L.; Risch, M.; Hong, W.T.; Zhou, J.; Shao-Horn, Y. Double perovskites as a family of highly active catalysts for oxygen evolution in alkaline solution. Nat. Commun. 2013, 4, 2439. [Google Scholar] [CrossRef] [PubMed]
  15. Suntivich, J.; Gasteiger, H.A.; Yabuuchi, N.; Nakanishi, H.; Goodenough, J.B.; Shao-Horn, Y. Design principles for oxygen-reduction activity on perovskite oxide catalysts for fuel cells and metal–air batteries. Nat. Chem. 2011, 3, 546–550. [Google Scholar] [CrossRef]
  16. Zhu, Y.; Zhou, W.; Chen, Z.-G.; Chen, Y.; Su, C.; Tadé, M.O.; Shao, Z. SrNb0.1Co0.7Fe0.2O3−δ Perovskite as a Next-Generation Electrocatalyst for Oxygen Evolution in Alkaline Solution. Angew. Chem. Int. Ed. 2015, 54, 3897–3901. [Google Scholar] [CrossRef]
  17. Takeguchi, T.; Yamanaka, T.; Takahashi, H.; Watanabe, H.; Kuroki, T.; Nakanishi, H.; Orikasa, Y.; Uchimoto, Y.; Takano, H.; Ohguri, N.; et al. Layered Perovskite Oxide: A Reversible Air Electrode for Oxygen Evolution/Reduction in Rechargeable Metal-Air Batteries. J. Am. Chem. Soc. 2013, 135, 11125–11130. [Google Scholar] [CrossRef]
  18. Zhu, Y.; Zhou, W.; Chen, Y.; Yu, J.; Xu, X.; Su, C.; Tadé, M.O.; Shao, Z. Boosting Oxygen Reduction Reaction Activity of Palladium by Stabilizing Its Unusual Oxidation States in Perovskite. Chem. Mater. 2015, 27, 3048–3054. [Google Scholar] [CrossRef]
  19. Zhou, W.; Zhao, M.; Liang, F.; Smith, S.C.; Zhu, Z. High activity and durability of novel perovskite electrocatalysts for water oxidation. Mater. Horiz. 2015, 2, 495–501. [Google Scholar] [CrossRef]
  20. Ali, S.; Al-Rahman, Y.M.A.; Galal, A. Catalytic Activity toward Oxygen Evolution of LaFeO3Prepared by the Microwave Assisted Citrate Method. J. Electrochem. Soc. 2012, 159, F600–F605. [Google Scholar] [CrossRef]
  21. Ali, S.; Al-Rahman, Y.M.A. Catalytic Activity of LaBO3 for OER in HClO4 Medium: An Approach to the Molecular Orbital Theory. J. Electrochem. Soc. 2015, 163, H81–H88. [Google Scholar] [CrossRef]
  22. Diaz-Morales, O.; Raaijman, S.; Kortlever, R.; Kooyman, P.J.; Wezendonk, T.; Gascon, J.; Fu, W.T.; Koper, M.T.M. Iridium-based double perovskites for efficient water oxidation in acid media. Nat. Commun. 2016, 7, 12363. [Google Scholar] [CrossRef]
  23. Seitz, L.; Dickens, C.F.; Nishio, K.; Hikita, Y.; Montoya, J.; Doyle, A.; Kirk, C.; Vojvodic, A.; Hwang, H.Y.; Norskov, J.K.; et al. A highly active and stable IrOx/SrIrO3 catalyst for the oxygen evolution reaction. Science 2016, 353, 1011–1014. [Google Scholar] [CrossRef] [PubMed]
  24. Kumari, S.; Ajayi, B.P.; Kumar, B.; Jasinski, J.B.; Sunkara, M.K.; Spurgeon, J.M. A Low-Noble-Metal W1−xIrxO3−δ Water Oxidation Electrocatalyst for Acidic Media via Rapid Plasma Synthesis. Energy Environ. Sci. 2017, 10, 2432–2440. [Google Scholar] [CrossRef]
  25. Lu, Y.; Ma, A.; Yu, Y.; Tan, R.; Liu, C.; Zhang, P.; Liu, D.; Gui, J.Z. Engineering Oxygen Vacancies into LaCoO3 Perovskite for Efficient Electrocatalytic Oxygen Evolution. ACS Sustain. Chem. Eng. 2018, 7, 2906–2910. [Google Scholar] [CrossRef]
  26. Goldschmidt, V.M. Die Gesetze der Krystallochemie. Naturwissenschaften 1926, 14, 477–485. [Google Scholar] [CrossRef]
  27. Abbes, L.; Noura, H. Perovskite oxides MRuO3 (M = Sr, Ca and Ba): Structural distortion, electronic and magnetic properties with GGA and GGA-modified Becke–Johnson approaches. Results Phys. 2015, 5, 38–52. [Google Scholar] [CrossRef] [Green Version]
  28. Zhou, S.; Miao, X.; Zhao, X.; Ma, C.; Qiu, Y.; Hu, Z.; Zhao, J.; Shi, L.; Zeng, J. Engineering electrocatalytic activity in nanosized perovskite cobaltite through surface spin-state transition. Nat. Commun. 2016, 7, 11510. [Google Scholar] [CrossRef] [PubMed]
  29. Khan, S.; Oldman, R.J.; Corà, F.; Catlow, C.R.A.; French, S.A.; Axon, S.A. A computational modelling study of oxygen vacancies at LaCoO3 perovskite surfaces. Phys. Chem. Chem. Phys. 2006, 8, 5207–5222. [Google Scholar] [CrossRef]
  30. Kim, G.J.; Lee, S.M.; Hong, S.C.; Kim, S.S. Active oxygen species adsorbed on the catalyst surface and its effect on formaldehyde oxidation over Pt/TiO2 catalysts at room temperature; role of the Pt valence state on this reaction? RSC Adv. 2018, 8, 3626–3636. [Google Scholar] [CrossRef] [Green Version]
  31. Zheng, Z.; Jia, J.; Zhong, Z. Revisiting the CO oxidation reaction on various Au/TiO2 catalysts: Roles of the surface OH groups and the reaction mechanism. J. Nanosci. Nanotechnol. 2014, 14, 6885–6893. [Google Scholar] [CrossRef] [PubMed]
  32. Galal, A.; Atta, N.F.; Darwish, S.A.; Fatah, A.A.; Ali, S. Electrocatalytic evolution of hydrogen on a novel SrPdO3 perovskite electrode. J. Power Sources 2010, 195, 3806–3809. [Google Scholar] [CrossRef]
  33. Atta, N.F.; Ali, S.; El-Ads, E.H.; Galal, A. The Electrochemistry and Determination of Some Neurotransmitters at SrPdO3 Modified Graphite Electrode. J. Electrochem. Soc. 2013, 160, G3144–G3151. [Google Scholar] [CrossRef]
  34. Atta, N.F.; Ali, S.; El-Ads, E.H.; Galal, A. Nano-perovskite carbon paste composite electrode for the simultaneous determination of dopamine, ascorbic acid and uric acid. Electrochim. Acta 2014, 128, 16–24. [Google Scholar] [CrossRef]
  35. Ali, S.M.; Al Lehaibi, H.A. Smart Perovskite Sensors: The Electrocatalytic Activity of SrPdO3 for Hydrazine Oxidation. J. Electrochem. Soc. 2018, 165, B345–B350. [Google Scholar] [CrossRef]
  36. El-Ads, E.H.; Atta, N.F.; Galal, A. The effect of A-site doping in a strontium palladium perovskite and its applications for non-enzymatic glucose sensing. RSC Adv. 2016, 6, 16183–16196. [Google Scholar] [CrossRef]
  37. El-Ads, E.H.; Galal, A.; Galal, A.; El-Gohary, A.R. Nano-perovskite decorated carbon nanotubes composite for ultrasensitive determination of a cardio-stimulator drug. J. Electroanal. Chem. 2018, 816, 149–159. [Google Scholar] [CrossRef]
  38. El-Ads, E.H.; Galal, A.; Atta, N.F. Electrochemistry of glucose at gold nanoparticles modified graphite/SrPdO3 electrode—Towards a novel non-enzymatic glucose sensor. J. Electroanal. Chem. 2015, 749, 42–52. [Google Scholar] [CrossRef]
  39. He, J.; Franchini, C. Structural determination and electronic properties of the 4d perovskite SrPdO3. Phys. Rev. B 2014, 89, 45104–45108. [Google Scholar] [CrossRef] [Green Version]
  40. Nishihata, Y.; Mizuki, J.; Akao, T.; Tanaka, H.; Uenishi, M.; Kimura, M.; Okamoto, T.; Hamada, N. Self-regeneration of a Pd-perovskite catalyst for automotive emissions control. Nature 2002, 418, 164–167. [Google Scholar] [CrossRef]
  41. Monshi, A.; Foroughi, M.R. Modified Scherrer Equation to Estimate More Accurately Nano-Crystallite Size Using XRD. World J. Nano Sci. Eng. 2012, 2, 154–160. [Google Scholar] [CrossRef] [Green Version]
  42. Fernandes, K.C.; Da Silva, L.M.; Boodts, J.F.; De Faria, L.A. Surface, kinetics and electrocatalytic properties of the Ti/(Ti + Ru + Ce)O2-system for the oxygen evolution reaction in alkaline medium. Electrochim. Acta 2006, 51, 2809–2818. [Google Scholar] [CrossRef]
  43. Balogun, M.S.; Qiu, W.; Yang, H.; Fan, W.; Huang, Y.; Fang, P.; Li, G.R.; Ji, H.; Tong, Y. A monolithic metal-free electrocatalyst for oxygen evolution reaction and overall water splitting. Energy Environ. Sci. 2016, 9, 3411–3416. [Google Scholar] [CrossRef]
  44. Reier, T.; Oezaslan, M.; Strasser, P. Electrocatalytic Oxygen Evolution Reaction (OER) on Ru, Ir, and Pt Catalysts: A Comparative Study of Nanoparticles and Bulk Materials. ACS Catal. 2012, 2, 1765–1772. [Google Scholar] [CrossRef]
  45. Beck, F.; Krohn, H.; Kaiser, W.; Fryda, M.; Klages, C.; Schäfer, L. Boron doped diamond/titanium composite electrodes for electrochemical gas generation from aqueous electrolytes. Electrochim. Acta 1998, 44, 525–532. [Google Scholar] [CrossRef]
  46. Eigeldinger, J.; Vogt, H. The bubble coverage of gas-evolving electrodes in a flowing electrolyte. Electrochim. Acta 2000, 45, 4449–4456. [Google Scholar] [CrossRef]
  47. Katsuki, N.; Takahashi, E.; Toyoda, M.; Kurosu, T.; Iida, M.; Wakita, S.; Nishiki, Y.; Shimamune, T. Water Electrolysis Using Diamond Thin-Film Electrodes. J. Electrochem. Soc. 1998, 145, 2358. [Google Scholar] [CrossRef]
  48. Vogt, H.; Balzer, R. The bubble coverage of gas-evolving electrodes in stagnant electrolytes. Electrochim. Acta 2005, 50, 2073–2079. [Google Scholar] [CrossRef]
  49. Singh, R.; Singh, N.; Singh, J. Electrocatalytic properties of new active ternary ferrite film anodes for O2 evolution in alkaline medium. Electrochim. Acta 2002, 47, 3873–3879. [Google Scholar] [CrossRef]
  50. Nikolov, I.; Darkaoui, R.; Zhecheva, E.; Stoyanova, R.; Dimitrov, N.; Vitanov, T. Electrocatalytic Activity of Spinel Related Cobaltites MxCo3−xO4 (M = Li, Ni, Cu) in the Oxygen Evolution Reaction. J. Electroanal. Chem. 1997, 429, 157–168. [Google Scholar] [CrossRef]
  51. Fakhroueian, Z.; Farzaneh, F.; Afrookhteh, N. Oxidative coupling of methane catalyzed by Li, Na and Mg doped BaSrTiO3. Fuel 2008, 87, 2512–2516. [Google Scholar] [CrossRef]
  52. Qi, A.; Wang, S.; Fu, G.; Ni, C.; Wu, D. La–Ce–Ni–O monolithic perovskite catalysts potential for gasoline autothermal reforming system. Appl. Catal. A Gen. 2005, 281, 233–246. [Google Scholar] [CrossRef]
  53. Tanaka, K.I. Catalysts working by self-activation. Appl. Catal. A Gen. 1999, 188, 37–52. [Google Scholar] [CrossRef]
  54. Harrington, D.A.; Conway, B.E. ac Impedance of Faradaic Reactions Involving Electrosorbed Intermediates-I. Kinetic Theory. Electrochim. Acta 1987, 32, 1703–1712. [Google Scholar] [CrossRef] [Green Version]
  55. Galal, A.; Darwish, S.A.; Atta, N.F.; Ali, S.; El Fatah, A.A.A.; Elzatahry, A. Synthesis, structure and catalytic activity of nano-structured Sr–Ru–O type perovskite for hydrogen production. Appl. Catal. A Gen. 2010, 378, 151–159. [Google Scholar] [CrossRef]
  56. Atta, N.F.; Galal, A.; Ali, S.; El-Said, D.M. Improved host–guest electrochemical sensing of dopamine in the presence of ascorbic and uric acids in a β-cyclodextrin/Nafion®/polymer nanocomposite. Anal. Methods 2014, 6, 5962–5971. [Google Scholar] [CrossRef]
  57. Atta, N.F.; Galal, A.; Ali, S.; Hassan, S.H. Electrochemistry and detection of dopamine at a poly(3,4-ethylenedioxythiophene) electrode modified with ferrocene and cobaltocene. Ionics 2015, 21, 2371–2382. [Google Scholar] [CrossRef]
  58. Conway, B.E.; Liu, T.C. Characterization of electrocatalysis in the oxygen evolution reaction at platinum by evaluation of behavior of surface intermediate states at the oxide film. Langmuir 1990, 6, 268–276. [Google Scholar] [CrossRef]
  59. Wu, G.; Li, N.; Zhou, D.R.; Mitsuo, K.; Xu, B.Q. Anodically electrodeposited Co+Ni mixed oxide electrode: Preparation and electrocatalytic activity for oxygen evolution in alkaline media. J. Solid State Chem. 2004, 177, 3682–3692. [Google Scholar] [CrossRef]
  60. Galal, A.; Atta, N.F.; Ali, S. Investigation of the catalytic activity of LaBO3 (B = Ni, Co, Fe or Mn) prepared by the microwave-assisted method for hydrogen evolution in acidic medium. Electrochim. Acta 2011, 56, 5722–5730. [Google Scholar] [CrossRef]
  61. Galal, A.; Atta, N.F.; Ali, S. Optimization of the synthesis conditions for LaNiO3 catalyst by microwave assisted citrate method for hydrogen production. Appl. Catal. A Gen. 2011, 409, 202–208. [Google Scholar] [CrossRef]
  62. Atta, N.F.; Galal, A.; Ali, S.M. The Catalytic Activity of Ruthenates ARuO3 (A = Ca, Sr or Ba) for the Hydrogen Evolution Reaction in Acidic Medium. Int. J. Electrochem. Sci. 2012, 7, 725–746. [Google Scholar]
  63. Atta, N.F.; Galal, A.; Ali, S.M. The Effect of the Lanthanide Ion-Type in LnFeO3 on the Catalytic Activity for the Hydrogen Evolution in Acidic Medium. Int. J. Electrochem. Sci. 2014, 9, 2132–2148. [Google Scholar]
Figure 1. The (A) XRD pattern and (B) TEM image of Sr2PdO3 prepared by the citrate-method, miller indices (h, l, k) are showed.
Figure 1. The (A) XRD pattern and (B) TEM image of Sr2PdO3 prepared by the citrate-method, miller indices (h, l, k) are showed.
Ijms 21 03785 g001aIjms 21 03785 g001b
Figure 2. Potentiodynamic polarization of the unmodified and modified GC electrodes with Sr2PdO3 in a 0.1 M HClO4 solution.
Figure 2. Potentiodynamic polarization of the unmodified and modified GC electrodes with Sr2PdO3 in a 0.1 M HClO4 solution.
Ijms 21 03785 g002
Figure 3. The (A) potentiodynamic polarization of GC/Sr2PdO3 in different concentrations of the HClO4 solutions and (B) the dependence of the logarithm of the current density at a definite potential on the pH of the solution.
Figure 3. The (A) potentiodynamic polarization of GC/Sr2PdO3 in different concentrations of the HClO4 solutions and (B) the dependence of the logarithm of the current density at a definite potential on the pH of the solution.
Ijms 21 03785 g003aIjms 21 03785 g003b
Figure 4. (A) The potentiodynamic polarization of GC/Sr2PdO3 at different temperatures in 0.1 M HClO4 solution, and (B) the corresponding Arrhenius plot.
Figure 4. (A) The potentiodynamic polarization of GC/Sr2PdO3 at different temperatures in 0.1 M HClO4 solution, and (B) the corresponding Arrhenius plot.
Ijms 21 03785 g004aIjms 21 03785 g004b
Figure 5. The current-time response of the potentiostatic experiment of GC/Sr2PdO3 in 0.1 M HClO4 subjected to a potential of 0.5 V for 150 min.
Figure 5. The current-time response of the potentiostatic experiment of GC/Sr2PdO3 in 0.1 M HClO4 subjected to a potential of 0.5 V for 150 min.
Ijms 21 03785 g005
Figure 6. The (A) Bode and (B) Nyquist plots GC/Sr2PdO3 in 0.1 M HClO4 at 0.50 V and various overpotentials, respectively. Symbols are experimental and solid lines are modeled data. (C) is the electrical equivalent circuit used for fitting.
Figure 6. The (A) Bode and (B) Nyquist plots GC/Sr2PdO3 in 0.1 M HClO4 at 0.50 V and various overpotentials, respectively. Symbols are experimental and solid lines are modeled data. (C) is the electrical equivalent circuit used for fitting.
Ijms 21 03785 g006
Table 1. Fitting parameters obtained by using the electrical equivalent circuit, shown in Figure 6C, for Nyquist plots of GC/Sr2PdO3 in 0.1 M HClO4 solution, at different applied overpotentials.
Table 1. Fitting parameters obtained by using the electrical equivalent circuit, shown in Figure 6C, for Nyquist plots of GC/Sr2PdO3 in 0.1 M HClO4 solution, at different applied overpotentials.
Applied Overpotential (V)Rs (Ω cm2)CPE1 (µF cm−2)nRP
(Ω.cm2)
CPE2 (µF cm−2)mRCT
(kΩ.cm2)
W (mΩ s−0.5)
0.452.73456.30.6813.45172.30.8017.5323.97
0.502.63524.30.6614.69167.00.827.149.33
0.552.61572.50.6224.75148.60.851.90-

Share and Cite

MDPI and ACS Style

Eskandrani, A.A.; Ali, S.M.; Al-Otaibi, H.M. Study of the Oxygen Evolution Reaction at Strontium Palladium Perovskite Electrocatalyst in Acidic Medium. Int. J. Mol. Sci. 2020, 21, 3785. https://doi.org/10.3390/ijms21113785

AMA Style

Eskandrani AA, Ali SM, Al-Otaibi HM. Study of the Oxygen Evolution Reaction at Strontium Palladium Perovskite Electrocatalyst in Acidic Medium. International Journal of Molecular Sciences. 2020; 21(11):3785. https://doi.org/10.3390/ijms21113785

Chicago/Turabian Style

Eskandrani, Areej A., Shimaa M. Ali, and Hibah M. Al-Otaibi. 2020. "Study of the Oxygen Evolution Reaction at Strontium Palladium Perovskite Electrocatalyst in Acidic Medium" International Journal of Molecular Sciences 21, no. 11: 3785. https://doi.org/10.3390/ijms21113785

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop