Next Article in Journal
Homogeneous Group IVB Catalysts of New Generations for Synthesis of Ethylene-Propylene-Diene Rubbers: A Mini-Review
Previous Article in Journal
Zero-Valent Copper-Mediated Peroxymonosulfate Activation for Efficient Degradation of Azo Dye Orange G
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Performance A-Site Deficient Perovskite Electrocatalyst for Rechargeable Zn–Air Battery

1
Shen Zhen Polytechnic, Shenzhen 518055, China
2
Aviation Engineering Institute, Civil Aviation Flight University of China, Guanghan 618037, China
3
Faculty of Chemical and Petroleum Engineering, University of Tabriz, Tabriz 5166616471, Iran
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(7), 703; https://doi.org/10.3390/catal12070703
Submission received: 20 April 2022 / Revised: 15 June 2022 / Accepted: 18 June 2022 / Published: 27 June 2022
(This article belongs to the Section Catalytic Materials)

Abstract

:
Zinc–air batteries are one of the most excellent of the next generation energy devices. However, their application is greatly hampered by the slow kinetics of oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) of air electrode. It is of great importance to develop good oxygen electrocatalysts with long durability as well as low cost. Here, A-site deficient (SmSr)0.95Co0.9Pt0.1O3 perovskites have been studied as potential OER electrocatalysts prepared by EDTA–citrate acid complexing method. OER electrocatalytic performance of (SmSr)0.95Co0.9Pt0.1O3 was also evaluated. (SmSr)0.95Co0.9Pt0.1O3 electrocatalysts exhibited good OER activities in 0.1 M KOH with onset potential and Tafel slope of 1.50 V and 87 mV dec−1, similar to that of Ba0.5Sr0.5Co0.8Fe0.2O3 (BSCF-5582). Assembled rechargeable Zn–air batteries exhibited good discharge potential and charge potential with high stability, respectively. Overall, all results illustrated that (SmSr)0.95Co0.9Pt0.1O3 is an excellent OER electrocatalyst for zinc–air batteries. Additionally, this work opens a good way to synthesize highly efficient electrocatalysts from A-site deficient perovskites.

1. Introduction

Rechargeable zinc–air battery has been paid much more attention due to its excellent energy density, safety, and economic costs [1,2,3,4]. OER is restricted due to slow kinetics as well as high overpotential [5]. Until now, IrO2 and RuO2 have been regarded as efficient commercial OER electrocatalysts. However, the high cost and scarcity greatly restricted the wide commercialization of rechargeable zinc–air batteries [6]. Hence, it is urgent to develop non-platinum substances to replace RuO2 and IrO2 electrocatalysts. Among what has been mentioned above, perovskite electrocatalysts have been considered practicable electrocatalysts because of low cost, high electronic conductivity, and electrocatalytic activity [7,8,9,10,11,12].
Perovskite oxides (ABO3) are able to show good OER activities because of increasing ionic, electronic conductivities and flexible compositional diversification properties. Suntivich et al. [13] studied ABO3 electrocatalysts with OER activity and concluded that the increasing OER activity was due to, e.g., orbital of first-row transition metal. Ba0.5Sr0.5Co0.8Fe0.2O3 (BSCF-5582) perovskites [13] have been reported to be highly active catalysts, which could be comparable to IrO2, however, BSCF-5582 perovskite structure could easily become amorphous after a long-time test. Therefore, a lot of perovskite oxides have been studied as potential OER electrocatalysts, such as SrNb0.1Co0.7Fe0.2O3-δ [14], SrCo0.9Ti0.1O3-δ [15], BaCo0.9-xFexSn0.1O3-δ [16], La0.3(Ba0.5Sr0.5)0.7Co0.8Fe0.2O3 [17], and SrCo0.95P0.05O3 (SCP) [18]. Great efforts could be made to develop perovskite OER catalysts with a good activity as well as stability.
There have been many reports regarding tuning A-site deficient perovskites by yielding positive effects on the electrocatalytic activity toward OER [19,20,21,22,23,24,25,26,27,28]. Zhu YL et al. [19] reported that La0.95FeO3-delta (L0.95F) demonstrated the highest OER activity due to surface oxygen vacancies, highlighting the importance of cation deficiency in perovskites by enhancing OER activity. Moreover, Wu XY et al. [20] reported A-site deficient BSCF nanofibers (300 nm diameter) prepared by electrospinning and found OER potential of optimized BSCF, which was stable after long-time tests. It was also reported that it had a great potential in the field of aqueous and flexible zinc–air batteries. However, the universality of deficient effects and mechanisms on ABO3 perovskite were ambiguous and needed to be clarified.
In this paper, A-site deficient (SmSr)0.95Co0.9Pt0.1O3 perovskites with OER performance under alkaline condition in the application of rechargeable zinc–air batteries were studied in detail. (SmSr)0.95Co0.9Pt0.1O3 delivered good OER activity and stability. Assembled initial Zn–air battery by (SmSr)0.95Co0.9Pt0.1O3 exhibited good cycling stability. This work sheds light on a facile method to prepare (SmSr)0.95Co0.9Pt0.1O3 perovskite electrocatalyst and enhance its potential application of rechargeable zinc–air battery.

2. Results and Discussions

2.1. Phase and Microstructure Characterization

Figure 1 illustrates the XRD patterns of (SmSr)0.95Co0.9Pt0.1O3 sample sintered at 850 °C for 3 h. The sample showed perovskite structure, which was similar to that of published results regarding Sm0.5Sr0.5CoO3. The inverted triangles in the XRD pattern referred to different crystal planes of perovskite structure [29]. ICP-OES test results can indicate that Co was 25.4290 wt%, Pt was 7.2015 wt%, Sm was 22.2669 wt%, Sr was 19.7917 wt%, and O was 25.3109 wt%. Figure 2a shows SEM images of the catalyst for (SmSr)0.95Co0.9Pt0.1O3 sample. Nanoparticles were columnar and the average size was around 150–250 nm. TEM images for (SmSr)0.95Co0.9Pt0.1O3 catalysts are also shown in Figure 2b. The large size was probably due to the sintering and aggregation during heat treatment. As shown in the HADDF images (Figure 3), homogeneity features were found in (SmSr)0.95Co0.9Pt0.1O3, which suggested the distributing of particle sizes in TEM. Sm, Sr, Co, Pt, and O elements were overlapped well and distributed homogeneously within the structure according to the element mapping result. Figure 2c,d demonstrate that the SAED pattern of (SmSr)0.95Co0.9Pt0.1O3 gave Debye–Scherrer rings, confirming the crystalline nature. The d-spacing of 0.297 nm was indexed to the (110) plane of ABO3 perovskites, as shown in Figure 2c. The selected-area electron diffraction (SAED) pattern of nanoparticles is shown in Figure 2d, imparting a crystalline nature.
The electrocatalytic behavior of perovskites was strongly relied on the valence state of transition metal as well as O anion ordering on catalysts’ surface [30,31,32,33]. Figure 4a shows the survey scan of the catalysts for (SmSr)0.95Co0.9Pt0.1O3 electrocatalysts, indicating the presence of Sm, Sr, Co, and O elements. Moreover, some unknown elements? were also detected, which might be due to the contamination of tests. Deconvoluted Co 2p and Pt 4f, O1s spectra of the electrocatalyst are shown in Figure 4b,c. Upon deconvolution, Co2p3/2 (about 780.0 eV) and Co2p1/2 (about 795.2 eV) peaks corresponding to Co3+ were observed. Besides, O 1s spectra (Figure 4c) was deconvoluted into two peaks [31]. The first one was a highly oxidative oxygen species (530.4 eV for O22−/O) and the second one was hydroxyl group or the surface-adsorbed oxygen (531.4 eV for OH or O2) [34]. From what has been mentioned in the paper, good OER activity in an alkaline solution might be attributed to O22−/O on the surface of catalysts [17,35].

2.2. Electrocatalytic Performance

Figure 5a shows OER activity of (SmSr)0.95Co0.9Pt0.1O3 electrodes under O2–0.1 M KOH at 5 mV/s at 1600 rpm. (SmSr)0.95Co0.9Pt0.1O3 showed good OER activity with onset potential of ~1.58 V, which was similar to that of BSCF (1.61 V). The maximum current density measured at 2.0 V was 27 mA cm−2 for (SmSr)0.95Co0.9Pt0.1O3, which was similar to that of BSCF (26.7 mA cm−2). Table 1 lists the OER activity of (SmSr)0.95Co0.9Pt0.1O3 with BSCF perovskite-based catalysts in 0.1 M KOH. The results showed its great potential as a highly efficient OER electrocatalyst. Nitrogen adsorption/desorption isotherm patterns of (SmSr)0.95Co0.9Pt0.1O3 and BSCF-5582 are also shown in Figure 6.
Figure 5b shows Tafel slope plots for (SmSr)0.95Co0.9Pt0.1O3 and BSCF, measured in 0.1 M KOH solution at 1 mV/s. Tafel slope of (SmSr)0.95Co0.9Pt0.1O3 was 82 mV dec−1, which was similar to that of 80 mV dec−1 obtained on BSCF. Zhu et al. reported Tafel slope of 76 and 94 mV dec−1 at low overpotentials for SrNb0.1Co0.7Fe0.2O3 and BSCF [14]. Figure 5c further compares the mass activity of (SmSr)0.95Co0.9Pt0.1O3 and BSCF catalysts. Mass activity for (SmSr)0.95Co0.9Pt0.1O3 at (η = 0.77 V) was 66 A g−1, while for BSCF was 65 A g−1. According to the results, the electrochemical activity of (SmSr)0.95Co0.9Pt0.1O3 catalysts was also comparable with that of BSCF. The BET surface area of (SmSr)0.95Co0.9Pt0.1O3 sample is shown in Figure 6. Relatively lower Tafel slope might indicate higher increase of current density, lower overpotential, faster reaction coefficient, which can indicate the relatively higher electrochemical performance.
The electrochemically active surface area (ECSA) of (SmSr)0.95Co0.9Pt0.1O3 and BSCF was calculated from CDL of catalytic surface (a general specific capacitance Cs = 0.040 mF cm−2 was used to estimate ESA for carbon proposed by McCrory et al. [36]). Figure 7 shows that based on the slopes, ECSA was 1.79, and 1.56 m2 g−1 for BSCF-5582 and (SmSr)0.95Co0.9Pt0.1O3, respectively. The ECSA of catalysts was smaller than BET surface area.
Figure 8a shows chronopotentiometry results at different current densities for BSCF and (SmSr)0.95Co0.9Pt0.1O3 catalysts measured in 0.1 M KOH. Potential to achieve 2, 5, and 10 mA/cm2 for (SmSr)0.95Co0.9Pt0.1O3, which was fired at 850 °C, was 1.534, 1.568, and 1.590 V, respectively. The potential of BSCF was 1.630, 1.646, and 1.694 V, which was higher than (SmSr)0.95Co0.9Pt0.1O3. The durability of (SmSr)0.95Co0.9Pt0.1O3 fired at 850 °C was studied (Figure 8b), and the potential under 10 mA/cm2 was 1.605 V, and the potential was around 1.615 V after 10 h. The schematic diagram of (SmSr)0.95Co0.9Pt0.1O3 catalysts for OER in alkaline solution can be seen in Figure 9. Therefore, (SmSr)0.95Co0.9Pt0.1O3 is a good OER catalyst in alkaline condition.

2.3. Cell Performance

The practical applicability of (SmSr)0.95Co0.9Pt0.1O3 electrocatalyst was tested in aqueous rechargeable Zn–air battery. Figure 9 shows the stability of Zn–air battery, which was evaluated by charging for 10 min and discharging for 10 min over repeated cycles at 5 mA cm−2 for nearly 100 h. Initially, it showed the discharge potential and charge potential of 1.10 V and 2.02V. Therefore, it could be calculated that the voltage gap (Δη) was 0.92 V. After 100 h, voltage gap increased to 1.0 V, respectively. Such a phenomenon was probably due to irreversible Zn plating–stripping process. It can be seen that (SmSr)0.95Co0.9Pt0.1O3-Zn–air battery had potential recharge ability. Moreover, (SmSr)0.95Co0.9Pt0.1O3 is cheaper than commercial IrO2, which shows its high economic potential.

3. Materials and Methods

3.1. Powder Synthesis

(SmSr)0.95Co0.9Pt0.1O3 powders were synthesized by EDTA–citrate acid complexing method. Stoichiometric mole amounts of Sm(NO3)3·6H2O, Sr(NO3)2, Co(NO3)3·9H2O, and H2PtClO4 (CAS: 26023-84-7) were mixed in distilled water with calculated molar ratios of total metal ions/citric acid/EDTA(1:1.5:1). Sol was prepared after mixing and stirring at 80 °C for several hours, and then gel was put into an oven for 12 h at 200 °C to form precursor. Then, the precursor was calcined in air for 3 h at 850 °C. BSCF-5582 powders were prepared as mentioned in our previous paper [37].

3.2. Catalysts Characterization

XRD (Bruker, D8 Advances) was used to test the prepared samples’ phase in the range of 5–80° (2θ). N2 adsorption/desorption isotherms were tested by Micromeritics TriStar II instrument at P/P0 from 0.05 to 0.35. The samples were analyzed by XPS (Thermo Fisher company, ESCALAB 250Xi instrument). SEM from ZEISS was used to characterize the prepared powders’ morphology. TEM and EDS were conducted by Titan G2 60–300 microscope.

3.3. Electrochemical Tests

(SmSr)0.95Co0.9Pt0.1O3 (4 mg) was mixed with 1 mL of an ethanol–Nafion mixture (with ethanol:Nafion = 9:1) to form suspension. Pt wire was used as a counter electrode, and Hg/HgO electrode was used as a reference electrode. Suspension (10 μL) was dipped into glassy carbon rotating disk electrode (GC-RDE, 0.196 cm2, Pine Research Instrumentation, USA). The catalyst loading was 0.2 mg/cm2. CV measurements [36] were tested as mentioned in the procedure, which was the same as our previous published paper [37]. The flow rate of oxygen supply was 10 mL/min. Before starting the electrochemical test, oxygen was flowed for 30 min to fill the reaction tank with saturated oxygen. CV was conducted at a 5 mV/s between −1 and 1 V for 5 cycles to obtain stable data. LSV was conducted at a 5 mV/s between 0 and 1 V. Tafel plots were tested at 1 mV/s with rotating speed of 1600 rpm. Chronopotentiometry was conducted at different current densities.

3.4. Battery Assembly and Test

Catalyst inks were prepared as mentioned above and then it was distributed uniformly on the carbon cloth with Ni-foam (current collector). Then, carbon cloth (cathode) and polished zinc plate (anode) were assembled in rechargeable Zn air battery by 6M KOH including 0.2 M ZnCl2. Aqueous Zn–air battery tests were carried out by a LAND CT2001A testing device. Charge and discharge data were obtained at 5 mA cm−2.

4. Conclusions

In conclusion, A-site efficient (SmSr)0.95Co0.9Pt0.1O3 oxygen electrocatalysts were successfully prepared by EDTA–citrate complexing sol-gel approach and adopted as a potential OER electrode in Zn–air battery. Optimized (SmSr)0.95Co0.9Pt0.1O3 electrocatalysts showed a higher OER intrinsic activity and durability, which was comparable with that of BSCF. Moreover, it had a high durability, good rechargeable Zn–air battery discharge performance and high cycle stability, which could be due to the efficient mass transferring and higher active sites. The results clarified a facile way to prepare non-platinum electrocatalysts for energy storage devices.

Author Contributions

C.W.: writing—original draft, B.H.: formal analysis, X.W.: data curation, Z.Y.: methodology, D.L.: validation, M.G.: writing—review and editing, X.F.: funding acquisition, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

The authors also acknowledge 2022 Research Funding of Shenzhen Polytechnic (6022310007K,6021310004K0), Shenzhen 2022 collaborative innovation technology program international science and technology, Cooperation Project the Science and Technology Innovation Project of Guangdong Provincial Department of Education (2021KTSCX278) and Sichuan science and technology program (2020YJ0501,2022YFH0044).

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge the facilities, scientific and technical assistance from Hoffman Research Institute in Shenzhen Polytechnic.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cao, R.; Lee, J.S.; Liu, M.L.; Cho, J. Recent Progress in Non-Precious Catalysts for Metal-Air Batteries. Adv. Energy Mater. 2012, 2, 816–829. [Google Scholar] [CrossRef]
  2. Li, H.; Ma, L.; Han, C.; Wang, Z.; Liu, Z.; Tang, Z.; Zhi, C. Advanced rechargeable zinc-based batteries: Recent progress and future perspectives. Nano Energy 2019, 62, 550–587. [Google Scholar] [CrossRef]
  3. Pan, J.; Xu, Y.Y.; Yang, H.; Dong, Z.; Liu, H.; Xia, B.Y. Advanced Architectures and Relatives of Air Electrodes in Zn-Air Batteries. Adv. Sci. 2018, 5, 1700691–1700721. [Google Scholar] [CrossRef] [PubMed]
  4. Park, J.; Park, M.; Nam, G.; Lee, J.S.; Cho, J. All-Solid-State Cable-Type Flexible Zinc-Air Battery. Adv. Mater. 2015, 27, 1396. [Google Scholar] [CrossRef]
  5. Shao, M.; Chang, Q.; Dodelet, J.-P.; Chenitz, R. Recent Advances in Electrocatalysts for Oxygen Reduction Reaction. Chem. Rev. 2016, 116, 3594–3657. [Google Scholar] [CrossRef] [Green Version]
  6. Debe, M.K. Electrocatalyst approaches and challenges for automotive fuel cells. Nature 2012, 486, 43–51. [Google Scholar] [CrossRef]
  7. Lefèvre, M.; Proietti, E.; Jaouen, F.; Dodelet, J.-P. Iron-Based Catalysts with Improved Oxygen Reduction Activity in Polymer Electrolyte Fuel Cells. Science 2009, 324, 71–74. [Google Scholar] [CrossRef]
  8. Wu, G.; More, K.L.; Johnston, C.M.; Zelenay, P. High-performance electrocatalysts for oxygen reduction derived from poly-aniline, iron, and cobalt. Science 2011, 332, 443–447. [Google Scholar] [CrossRef] [Green Version]
  9. Pan, Y.; Liu, S.; Sun, K.; Chen, X.; Wang, B.; Wu, K.; Cao, X.; Cheong, W.C.; Shen, R.; Han, A.; et al. A bimetallic Zn/Fe polyphthalocyanine-derived single-atom FeN4 catalytic site: A superior trifunctional catalyst for overall water splitting and Zn-air batteries. Angew. Chem. Int. Ed. 2018, 130, 8750–8754. [Google Scholar] [CrossRef]
  10. Chen, Y.; Li, Z.; Zhu, Y.; Sun, D.; Liu, X.; Xu, L.; Tang, Y. Atomic Fe dispersed on Ndoped carbon hollow nanospheres for high-effificiency electrocatalytic oxygen reduction. Adv. Mater. 2019, 31, 1806312. [Google Scholar] [CrossRef]
  11. Jasinski, R. A New Fuel Cell Cathode Catalyst. Nature 1964, 201, 1212–1213. [Google Scholar] [CrossRef]
  12. He, W.; Wang, Y.; Jiang, C.; Lu, L. Structural effects of a carbon matrix in non-precious metal O2-reduction electrocatalysts. Chem. Soc. Rev. 2016, 45, 2396–2409. [Google Scholar] [CrossRef]
  13. 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]
  14. Zhu, Y.; Zhou, W.; Chen, Z.G.; Chen, Y.; Su, C.; Tade, M.O.; Shao, Z. SrNb0.1Co0.7Fe0.2O3-delta perovskite as a next-generation electrocatalyst for oxygen evolution in alkaline solution. Angew. Chem. Int. Ed. Engl. 2015, 54, 3897–3901. [Google Scholar] [CrossRef]
  15. Su, C.; Wang, W.; Chen, Y.; Yang, G.; Xu, X.; Tade, M.O.; Shao, Z. SrCo0.9Ti0.1O3-delta as a New Electrocatalyst for the Oxygen Evolution Reaction in Alkaline Electrolyte with Stable Performance. ACS Appl. Mater. Interfaces 2015, 7, 17663–17670. [Google Scholar] [CrossRef]
  16. Xu, X.; Su, C.; Zhou, W.; Zhu, Y.; Chen, Y.; Shao, Z. Co-doping Strategy for Developing Perovskite Oxides as Highly Efficient Electrocatalysts for Oxygen Evolution Reaction. Adv. Sci. 2016, 3, 1500187–1500193. [Google Scholar] [CrossRef]
  17. Jung, J.-I.; Jeong, H.Y.; Lee, J.-S.; Kim, M.G.; Cho, J. A Bifunctional Perovskite Catalyst for Oxygen Reduction and Evolution. Angew. Chem. Int. Ed. 2014, 53, 4582–4586. [Google Scholar] [CrossRef]
  18. Zhu, Y.; Zhou, W.; Sunarso, J.; Zhong, Y.; Shao, Z. Phosphorus-Doped Perovskite Oxide as Highly Efficient Water Oxidation Elec-trocatalyst in Alkaline Solution. Adv. Funct. Mater. 2016, 26, 5862–5872. [Google Scholar] [CrossRef]
  19. Bian, J.J.; Li, Z.P.; Li, N.W.; Sun, C.W. Oxygen Deficient LaMn0.75Co0.25O3-Delta Nanofibers as an Efficient Electrocatalyst for Oxygen Evolution Reaction and Zinc-Air Batteries. Inorg. Chem. 2019, 58, 8208–8214. [Google Scholar] [CrossRef]
  20. Choi, M.J.; Kim, T.L.; Kim, J.K.; Lee, T.H.; Lee, S.A.; Kim, C.; Hong, K.; Bark, C.W.; Ko, K.T.; Jang, H.W. Enhanced Oxygen Evolution Electrocatalysis in Strained a-Site Cation Deficient LaNiO3 Perovskite Thin Films. Nano Lett. 2020, 20, 8040–8045. [Google Scholar] [CrossRef]
  21. Jiang, Y.L.; Geng, Z.B.; Yuan, L.; Sun, Y.; Cong, Y.G.; Huang, K.K.; Wang, L.; Zhang, W. Nanoscale Architecture of RuO2/La0.9Fe0.92Ru0.08-xO3-Delta Composite Via Manipulating the Exsolution of Low Ru-Substituted a-Site Deficient Perovskite. Acs Sustain. Chem. Eng. 2018, 6, 11999–12005. [Google Scholar] [CrossRef]
  22. Ma, J.J.; Geng, Z.B.; Jiang, Y.L.; Hou, X.Y.; Ge, X.; Wang, Z.Z.; Huang, K.K.; Zhang, W.; Feng, S.H. Exsolution Manipulated Local Surface Cobalt/Iron Alloying and Dealloying Conversion in La0.95Fe0.8Co0.2O3 Perovskite for Oxygen Evolution Reaction. J. Alloy. Compd. 2021, 854, 157154–157161. [Google Scholar] [CrossRef]
  23. Majee, R.; Islam, Q.A.; Mondal, S.; Bhattacharyya, S. An electrochemically reversible lattice with redox active A-sites of double perovskite oxide nanosheets to reinforce oxygen electrocatalysis. Chem. Sci. 2020, 11, 10180–10189. [Google Scholar] [CrossRef]
  24. Miao, H.; Wu, X.Y.; Chen, B.; Wang, Q.; Wang, F.; Wang, J.T.; Zhang, C.F.; Zhang, H.C.; Yuan, J.L.; Zhang, Q.J. A-Site De-ficient/Excessive Effects of LaMnO3 Perovskite as Bifunctional Oxygen Catalyst for Zinc-Air Batteries. Electrochim. Acta 2020, 333, 135566–135577. [Google Scholar] [CrossRef]
  25. Wu, X.; Miao, H.; Hu, R.; Chen, B.; Yin, M.; Zhang, H.; Xia, L.; Zhang, C.; Yuan, J. A-site deficient perovskite nanofibers boost oxygen evolution reaction for zinc-air batteries. Appl. Surf. Sci. 2021, 536, 147806–147813. [Google Scholar] [CrossRef]
  26. Xu, W.; Yan, L.; Teich, L.; Liaw, S.; Zhou, M.; Luo, H. Polymer-assisted chemical solution synthesis of La0.8Sr0.2MnO3-based perovskite with A-site deficiency and cobalt-doping for bifunctional oxygen catalyst in alkaline media. Electrochim. Acta 2018, 273, 80–87. [Google Scholar] [CrossRef]
  27. Yao, X.L.; Liu, J.Y.; Wang, W.H.; Lu, F.; Wang, W.C. Origin of Oer Catalytic Activity Difference of Oxygen-Deficient Perovskites A2Mn2O5 (a = Ca, Sr): A Theoretical Study. J. Chem. Phys. 2017, 146, 1404–1427. [Google Scholar] [CrossRef]
  28. Zhu, Y.; Zhou, W.; Yu, J.; Chen, Y.; Liu, M.; Shao, Z. Enhancing Electrocatalytic Activity of Perovskite Oxides by Tuning Cation Deficiency for Oxygen Reduction and Evolution Reactions. Chem. Mater. 2016, 28, 1691–1697. [Google Scholar] [CrossRef]
  29. Xia, C.; Rauch, W.; Chen, F.; Liu, M. Sm0.5Sr0.5CoO3 cathodes for low-temperature SOFCs. Solid State Ion. 2002, 148, A795. [Google Scholar] [CrossRef] [Green Version]
  30. 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]
  31. Sun, H.; Chen, G.; Zhu, Y.; Liu, B.; Zhou, W.; Shao, Z. B-Site Cation Ordered Double Perovskites as Efficient and Stable Electrocatalysts for Oxygen Evolution Reaction. Chem. Eur. J. 2017, 23, 5722–5728. [Google Scholar] [CrossRef] [PubMed]
  32. Gui, L.; Huang, Z.; Li, G.; Wang, Q.; He, B.; Zhao, L. Insights into Ni-Fe couple in perovskite electrocatalysts for highly efficient electrochemical oxygen evolution. Electrochim. Acta 2019, 293, 240–246. [Google Scholar] [CrossRef]
  33. Zhu, Y.; Zhang, L.; Zhao, B.; Huijun, C.; Liu, X.; Zhao, R.; Wang, X.; Liu, J.; Chen, Y.; Liu, M. Oxygen Defect Engineering: Improving the Activity for Oxygen Evolution Reaction by Tailoring Oxygen Defects in Double Perovskite Oxides. Adv. Funct. Mater. 2019, 29, 1970236. [Google Scholar] [CrossRef] [Green Version]
  34. Zhu, Y.; Zhou, W.; Chen, Y.; Yu, J.; Liu, M.; Shao, Z. ChemInform Abstract: A High-Performance Electrocatalyst for Oxygen Evolution Reaction: LiCo0.8Fe0.2O2. ChemInform 2015, 47, 7150–7155. [Google Scholar] [CrossRef]
  35. Liu, R.; Liang, F.; Zhou, W.; Yang, Y.; Zhu, Z. Calcium-doped lanthanum nickelate layered perovskite and nickel oxide nanohybrid for highly efficient water oxidation. Nano Energy 2015, 12, 115–122. [Google Scholar] [CrossRef]
  36. McCrory, C.C.L.; Jung, S.; Peters, J.C.; Jaramillo, T.F. Benchmarking Heterogeneous Electrocatalysts for the Oxygen Evolution Reaction. J. Am. Chem. Soc. 2013, 135, 16977–16987. [Google Scholar] [CrossRef] [PubMed]
  37. Wang, C.C.; Cheng, Y.; Ianni, E.; Lin, B. A highly active and stable La0.5Sr0.5Ni0.4Fe0.6O3-δ perovskite electrocatalyst for oxygen evolution reaction in alkaline media. Electrochim. Acta 2017, 246, 997–1003. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of as-prepared (SmSr)0.95Co0.9Pt0.1O3.
Figure 1. XRD patterns of as-prepared (SmSr)0.95Co0.9Pt0.1O3.
Catalysts 12 00703 g001
Figure 2. (a) SEM image of (SmSr)0.95Co0.9Pt0.1O3. (b,c) HRTEM images for (SmSr)0.95Co0.9Pt0.1O3. (d) Corresponding higher resolution SAED patterns.
Figure 2. (a) SEM image of (SmSr)0.95Co0.9Pt0.1O3. (b,c) HRTEM images for (SmSr)0.95Co0.9Pt0.1O3. (d) Corresponding higher resolution SAED patterns.
Catalysts 12 00703 g002
Figure 3. HADDF image and EDS element mappings of Sm, Sr, Co, Pt, and O.
Figure 3. HADDF image and EDS element mappings of Sm, Sr, Co, Pt, and O.
Catalysts 12 00703 g003
Figure 4. (a) Survey scan of (SmSr)0.95Co0.9Pt0.1O3, XPS spectra of (b) the deconvolution result of Co 2p. (c) The deconvolution result of O 1 s. (d) The deconvolution result of Pt 4f.
Figure 4. (a) Survey scan of (SmSr)0.95Co0.9Pt0.1O3, XPS spectra of (b) the deconvolution result of Co 2p. (c) The deconvolution result of O 1 s. (d) The deconvolution result of Pt 4f.
Catalysts 12 00703 g004
Figure 5. (a) LSV curves of (SmSr)0.95Co0.9Pt0.1O3, Ba0.5Sr0.5Co0.8Fe0.2O3 catalysts for OER at 5 mV s−1 scan rate. (b) Tafel plots and corresponding fitted slopes at 1 mV s−1. (c) MA comparison.
Figure 5. (a) LSV curves of (SmSr)0.95Co0.9Pt0.1O3, Ba0.5Sr0.5Co0.8Fe0.2O3 catalysts for OER at 5 mV s−1 scan rate. (b) Tafel plots and corresponding fitted slopes at 1 mV s−1. (c) MA comparison.
Catalysts 12 00703 g005
Figure 6. Nitrogen adsorption/desorption isotherm patterns of (SmSr)0.95Co0.9Pt0.1O3 (a) and Ba0.5Sr0.5Co0.8Fe0.2O3 (BSCF-5582) (b).
Figure 6. Nitrogen adsorption/desorption isotherm patterns of (SmSr)0.95Co0.9Pt0.1O3 (a) and Ba0.5Sr0.5Co0.8Fe0.2O3 (BSCF-5582) (b).
Catalysts 12 00703 g006
Figure 7. Anodic charging currents measured at 0.01 V plotted as a function of scan rate for (SmSr)0.95Co0.9Pt0.1O3 and Ba0.5Sr0.5Co0.8Fe0.2O3 powders in 0.1M KOH solution.
Figure 7. Anodic charging currents measured at 0.01 V plotted as a function of scan rate for (SmSr)0.95Co0.9Pt0.1O3 and Ba0.5Sr0.5Co0.8Fe0.2O3 powders in 0.1M KOH solution.
Catalysts 12 00703 g007
Figure 8. Chronopotentiometry of (SmSr)0.95Co0.9Pt0.1O3 and Ba0.5Sr0.5Co0.8Fe0.2O3 (a). Stability plots for (SmSr)0.95Co0.9Pt0.1O3 and Ba0.5Sr0.5Co0.8Fe0.2O3, measured at different current densities. (b) Chronopotentiometry responses at 10 mA cm−2 of (SmSr)0.95Co0.9Pt0.1O3. The measurements were carried out in 0.1 M KOH solution and the rotating rate of 1600 rpm. Note: testing time from 0 to 1200 s, current density was 2 mA cm−2; testing time from 1200 to 2400 s, current density was 5 mA cm−2; testing time from 2400 to 3600 s, current density was 10 mA cm−2.
Figure 8. Chronopotentiometry of (SmSr)0.95Co0.9Pt0.1O3 and Ba0.5Sr0.5Co0.8Fe0.2O3 (a). Stability plots for (SmSr)0.95Co0.9Pt0.1O3 and Ba0.5Sr0.5Co0.8Fe0.2O3, measured at different current densities. (b) Chronopotentiometry responses at 10 mA cm−2 of (SmSr)0.95Co0.9Pt0.1O3. The measurements were carried out in 0.1 M KOH solution and the rotating rate of 1600 rpm. Note: testing time from 0 to 1200 s, current density was 2 mA cm−2; testing time from 1200 to 2400 s, current density was 5 mA cm−2; testing time from 2400 to 3600 s, current density was 10 mA cm−2.
Catalysts 12 00703 g008
Figure 9. The galvanostatic charge–discharge profiles of rechargeable Zn–air batteries incorporated with catalyst (SmSr)0.95Co0.9Pt0.1O3 mixture at 5 mA cm−2.
Figure 9. The galvanostatic charge–discharge profiles of rechargeable Zn–air batteries incorporated with catalyst (SmSr)0.95Co0.9Pt0.1O3 mixture at 5 mA cm−2.
Catalysts 12 00703 g009
Table 1. Physical and electrocatalytic properties of the perovskite-based catalyst materials studied in 0.1 M KOH including Ba0.5Sr0.5Co0.8Fe0.2O3 and (SmSr)0.95Co0.9Pt0.1O3.
Table 1. Physical and electrocatalytic properties of the perovskite-based catalyst materials studied in 0.1 M KOH including Ba0.5Sr0.5Co0.8Fe0.2O3 and (SmSr)0.95Co0.9Pt0.1O3.
SampleBa0.5Sr0.5Co0.8Fe0.2O3(SmSr)0.95Co0.9Pt0.1O3
BET surface area (m2 g−1)10.84 ± 0.0214.77 ± 0.03
ECSA(m2 g−1)1.79 ± 0.011.56 ± 0.02
Onset potential (V)1.61V~1.58 V
Jmax (mA cm−2) (η = 0.77 V)26.727
Tafel slope (mV dec−1)8082
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, C.; Hou, B.; Wang, X.; Yu, Z.; Luo, D.; Gholizadeh, M.; Fan, X. High-Performance A-Site Deficient Perovskite Electrocatalyst for Rechargeable Zn–Air Battery. Catalysts 2022, 12, 703. https://doi.org/10.3390/catal12070703

AMA Style

Wang C, Hou B, Wang X, Yu Z, Luo D, Gholizadeh M, Fan X. High-Performance A-Site Deficient Perovskite Electrocatalyst for Rechargeable Zn–Air Battery. Catalysts. 2022; 12(7):703. https://doi.org/10.3390/catal12070703

Chicago/Turabian Style

Wang, Chengcheng, Bingxue Hou, Xintao Wang, Zhan Yu, Dawei Luo, Mortaza Gholizadeh, and Xincan Fan. 2022. "High-Performance A-Site Deficient Perovskite Electrocatalyst for Rechargeable Zn–Air Battery" Catalysts 12, no. 7: 703. https://doi.org/10.3390/catal12070703

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