Next Article in Journal
Influence of the Valence of Iron on the NO Reduction by CO over Cu-Fe-Mordenite
Previous Article in Journal
Hydrogenation of Carbon Monoxide in the Liquid Phase: Influence of the Synthetic Methods on Characteristics and Activity of Hydrogenation Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Role of the A-Site Cation on the Bifunctional Electrocatalytic Activities of Ln0.5Sr0.5CoO3-δ (Ln = La, Pr and Sm) for Rechargeable Zinc–Air Batteries

1
Institute of New Energy Materials and Devices, School of Materials Science and Engineering, Jingdezhen Ceramic University, Jingdezhen 333001, China
2
School of Mechanical and Electronic Engineering, Jingdezhen Ceramic University, Jingdezhen 333403, China
3
State Key Lab of New Ceramics and Fine Processing, School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China
*
Authors to whom correspondence should be addressed.
Submission received: 27 December 2022 / Revised: 10 February 2023 / Accepted: 20 February 2023 / Published: 27 February 2023
(This article belongs to the Section Electrocatalysis)

Abstract

:
Developing efficient bifunctional electrocatalysts for both the oxygen reduction reaction (ORR) and the oxygen evolution reaction (OER) is crucial for the large-scale application of rechargeable zinc–air batteries. Perovskite oxides are attractive candidates as bifunctional electrocatalysts. Herein, Ln0.5Sr0.5CoO3-δ (Ln = La, Pr and Sm) was prepared by the sol-gel method and evaluated as bifunctional electrocatalysts for both ORR and OER. Compare with La0.5Sr0.5CoO3-δ (LSC) and Sm0.5Sr0.5CoO3-δ (SSC), Pr0.5Sr0.5CoO3-δ (PSC) demonstrated the enhanced catalytic activity towards ORR with the limiting current density of 4.19 mA cm−2 at 0.40 V vs. RHE and OER with the potential of 1.75 V vs. RHE at 10 mA cm−2 as well as the Tafel slope of 105 mV dec−1. The higher Co3+/Co2+ ratio in PSC played an essential part in boosting the bifunctional electrocatalytic activities for ORR and OER. For demonstration, rechargeable zinc–air batteries with PSC as the air electrode displayed the maximum power density of 72 mW cm−2, the low charge-discharge voltage gap (1.01 V) at 10 mA cm−2 over 80 cycles and outstanding cycling stability. This work highlights the importance of A-site cations of the perovskite oxide as a valid strategy to improve ORR and OER activities.

1. Introduction

With the rapid consumption of fossil energy and the aggravation of environmental pollution, the development of efficient, green and sustainable energy storage and conversion technologies has attracted widespread attention due to the increasing demands for renewable and environmental-friendly energy. Particularly, rechargeable zinc–air batteries as a promising candidate for future energy application have attracted intensive attention due to the high safety, cost effective, high theoretical volumetric and gravimetric energy densities as well as environment-friendly [1,2]. Nevertheless, the performance improvement and commercialization of rechargeable zinc–air batteries were restricted by the sluggish kinetics of the oxygen reduction reaction (ORR) during the discharge process as well as the oxygen evolution reaction (OER) during the charge process at the air electrode [3,4].
Hitherto, noble-metal Pt catalyst has been regarded as the benchmark for ORR, and the oxides of Ru or Ir are regarded as the benchmark towards OER, which were greatly impeded by the exorbitant price, scarcity and unsatisfactory stability [5,6]. Furthermore, noble-metal-based catalysts usually exhibit selective activity for either ORR or OER, which poses a significant challenge to the ideal air electrode design [7]. Consequently, it is highly desirable to explore novel bifunctional catalysts for both ORR and OER with earth-abundant, environmentally friendly and high durability.
Growing attention has been given to perovskites with the general formula of “ABO3” (A = alkaline-earth or rare-earth metals, B = transition metals) due to the high compositional flexibility, abundance on earth as well as the relatively high ionic and/or electronic conductivities [8,9,10]. The binding strength of reaction intermediates of ORR/OER, closely related to the intrinsic electronic structure of the catalyst, is the key factor affecting electrocatalytic activity. The superior catalytic activity could be achieved when the adsorbed species lightly bind to the surface neither too strongly nor too weakly according to the Sabatier’s principle [11]. Therefore, electronic structure descriptors were applied to correlate the ORR and/or OER, such as the O 2p band center, the charge-transfer energy and the eg-orbital filling [12,13,14,15]. The eg-orbital filling could be the most suitable descriptor to describe ORR and OER activities of perovskite oxides according to the molecular orbital theory considerations and experimental observations [16]. A volcano-type relationship between the eg-orbital filling and catalytic activity of transition metals was reported by Suntivich et al. [14]. In addition, the eg-orbital filling of the catalyst with highest activities was usually close to 1.2. The binding energy of *OH− is too strong when the eg-orbital filling is higher than 1, and too weak when the eg-orbital filling is lower than 1. Perovskites allows the isovalent and aliovalent doping to adjust the eg-orbital filling due to the high Goldschmidt tolerance factor, thereby regulating ORR/OER catalytic activities [15,17]. Many efforts have been devoted to improving the bifunctional catalytic activity and cycling stability of perovskite oxides by means of doping in A-site and/or B-site cations [18,19,20]. LaMnO3, LaCoO3, La0.6Sr0.4Co0.4Fe0.6O3 and La0.5Sr0.5CoO3 perovskite oxides have been evaluated as bifunctional catalysts, and the results demonstrated that the change of B-site cations would lead to the change of the oxygen adsorption energy on the surface of bifunctional catalysts, thus affecting catalytic activities [21,22]. Deficiency and/or doping of A-site cations could regulate the oxygen vacancy concentration, the mixed ionic and electronic conductivity and the valence of B-site cations, which is closely related to catalytic activities of ORR and OER [23,24]. Perovskites with moderate A-site-deficiency and/or doping, such as La1-xNiO3, La1-xFeO3, La1-xCoO3, and La1-xSrxCoO3, displayed enhanced oxygen redox catalytic activities due to the mixed-valence states of B-site cations, the mixed ionic and electronic conductivity, the enhanced σ* electron-transferring ability as well as the hydroxide/oxygen surface-adsorbing ability [25,26,27,28]. However, excessive A-site defects in perovskite will lead to the formation of the second phase, which is unfavorable to bifunctional catalytic activities.
Several previous reports have shown that the A-site cations have a great influence on properties of perovskite oxides such as LnBaCo2O5+δ (Ln = La, Pr, Nd and Sm), Ln0.6Sr0.4CoO3-δ (Ln = La, Pr, Nd, Sm and Gd) and Ln0.5Sr0.5Fe0.9Mo0.1O3-δ (Ln = La, Pr, Nd, Sm and Gd) [29,30,31]. The electronic and ionic conductivities, oxygen vacancy concentration and the electrocatalytic activity of perovskite oxides were affected by A-site cations [27,31]. Therefore, it is necessary to systematically explore the influence of various A-site cations on ORR and OER catalytic activities of perovskite oxides.
Given the foregoing, the effect of A-site rare earth cations on the crystal structure, the valence state of Co and bifunctional catalytic activities of Ln0.5Sr0.5CoO3-δ (Ln = La, Pr and Sm) prepared by the facial sol-gel method were systematically investigated in this study. The phase structure and compositions transformation of Ln0.5Sr0.5CoO3-δ were analyzed in detail. The electrocatalytic properties of the as-prepared Ln0.5Sr0.5CoO3-δ for both ORR and OER were systematically tested in the alkaline medium. Furthermore, electrocatalytic properties and the stability of as-prepared catalysts were also evaluated in home-made rechargeable zinc–air batteries.

2. Results and Discussion

The XRD measurement was applied to reveal the crystal structure of as-synthesized samples. Room temperature XRD patterns of La0.5Sr0.5CoO3-δ (LSC) and Pr0.5Sr0.5CoO3-δ (PSC) were displayed in Figure S1a,b, which agreed well with the corresponding standard cards (PDF#48-0122, PDF#53-0110), without any impurity peaks, indicating that the pure powders were synthesized successfully. The XRD pattern of Sm0.5Sr0.5CoO3-δ (SSC) (Figure S1c) displayed that the main diffraction lines coincide with the corresponding standard card (PDF#53-0112). However, the non-indexed peak labeled with an asterisk at 2θ of 33.87° was related to a small amount of SmCoO3 (PDF#25-1071) [32]. To reveal the characterization clearly, XRD patterns and the enlarged view at the main peak with 2θ = 32–35° of LSC, PSC and SSC (Figure 1). Importantly, some peaks shift positive in the order of LSC, PSC and SSC. It is clearly to be seen that the diffraction peaks gradually shift to higher angles in the order of LSC, PSC and SSC. The crystal structure distortions increase as the lanthanide ionic radii decreases from La to Sm. LSC displayed only a single peak in the (110) direction. Similarly, PSC exhibited only a single peak in the (112) direction. However, the peak of SSC split into two peaks, namely the (200), (121) and (002) directions, which belongs to phase transition from tetragonal to orthorhombic, indicating the distortion of crystal structure toward lower symmetry as previous similar results [29]. Perovskite-type structure with cubic symmetry is regarded as ideal perovskite structure with the highest phase stability. The structural symmetry of ABO3 perovskite can be evaluated by the Goldschmidt tolerance factor t defined as follows Equation (1) [33]:
t = r A + r O 2 ( r B + r O )
where r A , r O , and r B are the mean ionic radius of Ln3+ and Sr2+, oxygen ionic radius and the ionic radius of cobalt, respectively. The ionic radii of La3+, Pr3+ and Sm3+ are 1.36, 1.30 and 1.24 Å, respectively [27,34]. With a smaller ion as the A-site cation, i.e., for Pr3+ and Sm3+, the structure of perovskite oxides deviated from cubic and was closer to orthorhombic [27]. The gradual decrease of t value caused by the decrease of Ln3+ ionic radius from La3+ to Sm3+, suggesting the lower crystal symmetry and the decrease of the cell volume [31].
The morphologies of the as-prepared LSC, PSC and SSC powders were presented in Figure S2 and Figure 2. It is clearly seen that LSC, PSC and SSC powders with the size of about several hundred nanometers were agglomerated into larger particles of several micrometers with porous structure due to the high sintering temperature. It should be noted that PSC exhibits a porous structure, which is conductive to the adequate exposure of active sites during ORR and OER.
The cation oxidation state and the surface chemical composition of LSC, PSC and SSC were analyzed by XPS [35]. The carbon peak located at 284.71 eV could be the residual carbon after calcination of as-prepared samples [15]. The survey spectrum of LSC displayed peaks of La 3d, Sr 3d, Co 2p, O 1s and C 1s peaks according to XPS results in Figure S3a. Similarly, the survey spectrum of PSC and SSC displayed peaks of Pr 3d and Sm 3d, respectively. The corresponding binding energies of various peaks were consistent with previously reported values [36,37].
The oxygen-related species as well as the transition metals on the surface of the perovskite oxide were usually regarded as active sites to ORR and OER [38]. Two main peaks of O 1s spectra for LSC, PSC and SSC indicated that more than one type of oxygen species presented on surface of various catalysts as shown in Figure 3a–c.
Although in the O 1s spectra the positions of peaks of various catalysts remain unchanged, the relative intensity of peaks has obviously changed. The O 1s spectra of LSC, PSC and SSC were deconvoluted into four peaks. The peak located at the lower binding energy around 528.25 eV was ascribed to the lattice oxygen species O2− in LSC, PSC and SSC (Figure 3a–c). The relative intensity of the lattice oxygen species O2− for PSC was higher than that of LSC and SSC. Previous studies have suggested that the stronger covalency of lattice oxygen species was helpful to improve the ORR kinetics [39]. Taking PSC as an example, the higher binding energy peak for PSC was deconvoluted into three peaks, which ascribed to characteristics of O22−/O (529.99 eV), hydroxyl groups OH/O2 (531.18 eV) and the molecular water (532.35 eV), respectively [40,41]. Generally, the O 2 2 / O was regarded as reactive species for ORR and OER [42].
The best OER activity was predicted to be at the eg-orbital filling close to 1, which could promote the binding of oxygen-related intermediate species on B-sites and thus improve OER activities [14]. The eg-orbital filling of Co3+ ( t 2 g 5 e g 1 ) and Co2+ ( t 2 g 5 e g 2 ) was 1 and 2, respectively. The surface electronic structure is affected by the oxidation state of the surface transition metal cations [36]. Hence, a high Co3+/Co2+ ratio can achieve a near-unity of eg-orbital filling [43,44]. Higher oxidation state of Co in the perovskite oxide is regarded as potential sites for oxygen redox reactions [45,46]. The Co 2p spectrum of LSC, PSC and SSC can be deconvoluted into major characteristic peaks and two shake-up satellites (Figure 3d–f). For PSC, the peaks located at 779.50 eV and 794.53 eV are assigned t°Co3+, and the peaks at 780.69 eV and 795.76 eV indicate the existence of Co2+ as shown in Figure 3e. The area under the XPS peaks was fitted, which corresponds to their relative amounts. Among the as-prepared catalysts, displayed the Co3+/Co2+ ratio of PSC was 1.36 as shown in Table S1, which was much higher than that of LSC (1.02) and SSC (0.98). The higher Co3+/Co2+ ratio was conductive to the promotion of the electron transfer process during ORR and OER, which led to the enhancement of ORR and OER catalytic activities [35,36,47].
LSV curves were tested by rotating disk electrode in O2-saturated 0.1 M KOH to evaluate catalytic activities of LSC, PSC and SSC for ORR. Cyclic voltammetry curves of LSC, PSC and SSC in 0.1 M KOH saturated with O2 were displayed in Figure 4a. It can be seen that PSC displayed more positive oxygen reduction peak potential (0.58 V vs. RHE) than that of the LSC (0.57 V vs. RHE) and SSC (0.56 V vs. RHE). ORR activities of various samples were evaluated by LSV curves at a rotating rate of 1600 rpm as shown in Figure 4b. PSC displayed the half-wave potentials (E1/2) about 0.55 V vs. RHE. Furthermore, the limiting current density of various samples follows the trend of SSC (3.34 mA cm−2) < LSC (4.04 mA cm−2) < PSC (4.19 mA cm−2) at 0.40 V vs. RHE. Therefore, PSC displayed significantly enhanced ORR catalytic activity than that of LSC and SSC. LSV curves were measured at various rotating rates as shown in Figure 4c to further explain the mechanism of PSC for ORR. The electron transferred numbers (n) calculated by K-L plots, as shown in Figure 4d, was about 3.94, which indicated the good selectivity for four electron ORR pathways.
In addition to the outstanding ORR activity discussed above, the OER activity of various samples was also evaluated in O2-saturated 0.1 M KOH using a 10 mV s−1 scan rate and 1600 rpm rotation rate. The corresponding polarization plots were iR-corrected to compensate for the solution resistance and depicted in Figure 5a. It can be seen that PSC displayed satisfactory OER activities with the potential of 1.75 V vs. RHE at the current density of 10 mA cm−2, which was lower than that of LSC (1.77 V vs. RHE) and SSC (1.79 V vs. RHE). Therefore, PSC exhibited better OER performance than LSC and SSC. Moreover, the OER electrocatalytic kinetics of various bifunctional electrocatalytic were further investigated by the Tafel slope as shown in Figure 5b. PSC displayed the Tafel slope of 105 mV dec−1, which was smaller than that of LSC (125 mV dec−1) and SSC (120 mV dec−1), indicating its superior OER activities. The potential gap (ΔE = EOER@10E1/2) between the OER potential at 10 mA cm−2 (EOER@10) and the ORR half-wave potential was used to evaluate the electrochemical reversibility of the oxygen-related electrode [48,49]. The ΔE value of PSC was 1.20 V (Figure 5c), which implied its reversibility and better bifunctional activity [50].
Home-made zinc–air batteries were fabricated to assess the bifunctional activity and stability of as-prepared catalysts under actual operating conditions (Figure S4). All home-made zinc–air batteries were conducted in ambient conditions (room temperature) using atmospheric air. The primary zinc–air battery with PSC demonstrated a stable open circuit voltage of 1.47 V as illustrated in Figure 6a, indicating PSC as promising candidate in practical implementation [51]. As shown in Figure 6b, the primary zinc–air battery with PSC displayed the maximum power density of 72 mW cm−2 at the current density of 120 mA cm−2, which was higher than that of battery with SSC (54 mW cm−2 at the current density of 92 mA cm−2) and LSC (57 mW cm−2 at the current density of 93 mA cm−2).
In addition to the galvanodynamic behavior, the kinetics of the as-prepared air electrodes were evaluated by the electrochemical impedance spectroscopy as shown in Figure 6c. Nyquist plots of various samples for primary zinc–air batteries at 1.2 V vs. Zn consists of two semicircles, which is similar to the previous reports [52,53]. Nyquist plots were fitted by an equivalent circuit (Figure 6c, inset) with five elements (Rohm, Rint, Rct, CPEint, CPEct) [53,54,55]. Rohm represents the electrolyte and contact resistance of primary zinc–air batteries. Rint represents the solid–liquid electrolyte interface resistance, which was related to the electrical conductivity of catalyst as well as its wettability in the liquid electrolyte [56]. In addition, the charge-transfer resistance (Rct) was related to the catalytic activity of the catalyst during electrochemical reactions [56]. Constant phase elements (CPEint and CPEct) were related to the capacitance at the interface of the air electrode and the liquid electrolyte [56]. On the basis of the equivalent circuit, the resistance values (Rohm, Rint and Rct) were listed in Table S2. The primary zinc–air battery with PSC displayed lower Rohm (1.46 Ω) than that of the primary zinc–air battery with LSC (1.70 Ω) or SSC (1.64 Ω), which was mainly attribute to the higher electronic conductivity of PSC than LSC and SSC as previously reported [24,57]. PSC displayed the lowest charge-transfer resistance (0.929 Ω) than that of LSC and SSC, which indicated the improvement of ORR kinetics.
Similarly, primary zinc–air batteries with PSC as the air electrode displayed the discharge time about 23 h at the constant current density of 10 mA cm−2 (Figure 6d). Being as a practical example, the LED lamp was successfully light up with two zinc–air connected in series assembled with PSC air electrode (inset of Figure 6d). The recharge ability and stability of various catalysts as the air electrodes were also evaluated by rechargeable zinc–air batteries in an atmospheric air environment. The overpotential for ORR and OER of zinc–air battery with PSC as bifunctional catalysts is 1.48 V at 50 mA cm−2 as discerned in Figure 7a, which is significantly lower than that of SSC and LSC. The charge-discharge cycling stability measurements of rechargeable zinc–air batteries with LSC, PSC and SSC as air electrodes were performed at 10 mA cm−2 for 30 min discharge and 30 min charge for 80 cycles (80 h). After 80 cycles, rechargeable zinc–air batteries with PSC as the air electrode exhibited the lowest charge-discharge voltage gap (1.01 V, Figure 7b), further indicating the promising bifunctional catalytic activities and excellent cycling stability.

3. Experimental

3.1. Material Synthesis

All reagents in this work were the analytical grade, which were used as received without further purification. Ln0.5Sr0.5CoO3 (LnSC, Ln = La, Pr and Sm) were synthesized via the sol-gel method according to previously reported synthesis procedures [54]. Specifically, La(NO3)·6H2O (99.9%), Pr(NO3)3·6H2O (99.9%), Sm(NO3)3·6H2O (99.9%), Sr(NO3)2 (99.9%) and Co(NO3)2·H2O (99.9%) were purchased from Innochem (Beijing, China). Citric acid (C6H8O7·H2O) and zinc acetate ((Zn(CH3COO)2·2H2O, 99%)) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Typically, the nitrates of the various metals (La, Pr, Sm, Sr and Co) in the required stoichiometric concentration were dissolved in distilled water. Then, citric acid was added in the molar ratio of 1.5:1 to the metal ion. The solution was stirred continuously at 80 °C under the water bath for 4 h to obtain the viscous wet gel. After that, the solid precursor was obtained by drying the wet gel at 200 °C for 12 h. The resulting black powder was received by calcination of the solid precursor at 1100 °C for 4 h under the air atmosphere.

3.2. Physical Characterization

X-ray diffraction (XRD, Rigaku D/max2200, Tokyo, Japan) with Cu-Kα radiation (λ = 0.15418 nm) was applied to analyze the phase structure of LSC, PSC and SSC. The contents and valence state information were analyzed by the X-ray photoelectron spectroscopy (XPS, Thermo Scientific Escalab 250XI, Waltham, MA, USA), Al Kα radiation, = 1486.6 eV). The C 1s binding energy (284.8 eV) was applied to compensate the binding energy of all XPS peaks in this work. The morphology of various catalysts was evaluated by the transmission electron microscopy (TEM, Tecnai G2 20ST, Amsterdam, The Netherlands) and a field emission scanning electron microscope (SEM, Hitachi-SU800, Tokyo, Japan).

3.3. Electrochemical Measurements

All electrochemical measurements in this work were performed with the CHI 660E electrochemical workstation (Chenhua, Shanghai, China). The RRDE-3A Apparatus (ALS, Tokyo, Japan) was applied to test the electrochemical properties in the three-electrode half-cell consisting of the Hg/HgO reference electrode, a working electrode based on the rotating disk glassy carbon electrode (diameter: 5 mm, geometric area: 0.196 cm2) and the Pt wire counter electrode. 5 mg of catalyst as well as 5 mg of conductive carbon (Super P Li, TIMCAL) were ultrasonically dispersed into the mixture of 50 μL of Nafion solution (5 wt %) and 500 μL of ethanol and sonicated for at least 1 h to form a uniformly dispersed ink as previously reported [54,58]. Subsequently, 5 μL of as-prepared catalysts catalyst ink was dropped onto the surface of glassy carbon disk electrode. The nominal catalyst loading of the as-prepared catalyst was 232 μg cm−2. All electrochemical properties toward ORR and OER of as-prepared catalysts were evaluated by the linear sweep voltammetry (LSV) at various rotating rates in O2-saturated 0.1 M KOH with the scan rate of 10 mV s−1.
The results of potentials were given versus reversible hydrogen electrode (RHE) according to the following Equation (2) [59]:
E(vs. RHE) = E(vs. Hg / HgO) + 0.098 + 0.059 × pH
The electron-transfer number(n) during ORR can be calculated based on the Koutechy–Levich (K-L) Equations (3)–(5) given below [60,61]:
1 j = 1 j L + 1 j K = 1 B ω 1 / 2 + 1 j K
B = 0.62 n F C 0 ( D 0 ) 2 / 3 ν 1 / 6
j K = n F k C 0
Here, jL, jK and j are the kinetic-limited, mass-transfer-limited and measured current densities, respectively. F is the Faraday constant (F = 96485 C mol−1). k is the Boltzmann constant. n is the number of transferred electrons. C0 is the saturated concentration of oxygen in electrolyte (C0 = 1.26 × 10-6 mol cm−3). ν is the kinetic viscosity of solution (ν = 0.01009 cm2 s−1), D0 is the diffusion coefficient of oxygen. ω is the rotating rate (rad s−1). The proportionality coefficient (B) is determined from the slope of K-L Equation. jK is assumed to be the constant at the certain potential [62]. jL is proportional to ω−1/2. Hence, the electro-transfer number (n) can be calculated according to the slope of the linear relationship of j−1 versus ω−1/2.
The Tafel slope (b) was calculated depending on the Tafel curve derived from the LSVs of OER measurements based on the Tafel Equation (6) [63]:
η = a + b × log j
where a and j was the Tafel constant and measured current density, respectively. The overpotential (η) was obtained based on η = Evs. RHE − 1.23.

3.4. Characterization of Rechargeable Zinc–Air Batteries

Home-made zinc–air batteries were assembled to characterize the bifunctional catalytic activities under ambient atmosphere. The home-made air electrode was composed of a catalyst layer, a current collector and a hydrophobic gas diffusion layer. The catalyst layer slurry was prepared by mixing the active carbon, the as-prepared catalyst powder, acetylene black and polytetrafluoroethylene emulsion (PTFE, 60 wt %) with the weight ratio of 3:3:1:3 into ethanol as per the previous report [64]. The gas diffusion layer slurry consisted of acetylene black and PTFE with the mass ratio of 2:5. The catalyst layer slurry and the gas diffusion slurry were continuously magnetically stirred in the water bath at 80 °C until the formation of a dough-like mixture, respectively. Afterwards, the gas diffusion layer slurry and catalyst layer slurry were evenly coated on each side of the foam nickel. The air electrode was dried at 80 °C for 12 h. The introduction of acetylene black as the composite partner is conducive to the formation of heterogeneous interface and the electronic transfer at the interface, which helps to improve the efficiency and stability of the air electrode [65,66]. The polished zinc plate with the thickness of 1 mm was applied as the anode of zinc–air batteries. The separator was not used in this work. The electrolyte of primary zinc–air batteries was 6 M KOH aqueous solution. However, the electrolyte of rechargeable zinc–air batteries was the solution containing 6 M KOH and 0.2 M zinc acetate. The polarization and power density curves as well as the electrochemical impedance spectroscopy of zinc–air batteries were recorded by the electrochemical workstation (DH7000, Jiangsu Donghua Analytical Instruments Co., Ltd., Taizhou, China). The discharge as well as charge-discharge cycling tests were recorded by the Neware Battery test instrument (CT-4008-5V6A-S1, Shenzhen Neware Instrument Company, Shenzhen, China).

4. Conclusions

In summary, LSC, PSC and SSC were fabricated by the so-gel method and systematically evaluated as bifunctional catalysts for ORR and OER. A-site cations of perovskite oxides not only causes the distortion of crystal structure, but also changes the valence state of Co. Compared with LSC and SSC, PSC demonstrated the enhanced catalytic activity towards ORR (the limiting current density of 4.19 mA cm−2 at 0.40 V vs. RHE) and OER (the potential of 1.75 V vs. RHE at 10 mA cm−2 and the Tafel slope of 105 mV dec−1). The higher Co3+/Co2+ ratio of PSC was conductive to the promotion of the electron transfer process during ORR and OER, thereby improving ORR and OER catalytic activities. Rechargeable zinc–air batteries with PSC as the air electrode displayed the maximum power density of 72 mW cm−2 and the low charge-discharge voltage gap (1.01 V) over 80 cycles, which mainly due to the improvement of charge-transfer process. This work not only unveils A-site cations effects on the bifunctional catalytic activities of perovskite oxides but also provides a valid strategy to explore novel bifunctional catalysts for rechargeable zinc air batteries.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal13030483/s1, Figure S1. X-ray diffraction patterns of (a) LSC, (b) PSC and (c) SSC powders; Figure S2. TEM images of (a) LSC, (b) PSC and (c) SSC powders; Figure S3. XPS survey spectrum of LSC, PSC and SSC; Figure S4. (a) Digital photos of home-made air electrodes and (b) the zinc-air battery assembled with the home-made air electrode; Table S1. The Co3+/Co2+ ratio of LSC, PSC and SSC powders in XPS test; Table S2. The values of the equivalent circuit elements resulting from fitting the electrochemical impedance spectroscopy data of LSC, PSC, and SSC electrodes in the primary zinc-air batteries.

Author Contributions

Conceptualization, P.L.; Methodology, P.L.; Formal analysis, P.L., Q.H., C.T., Y.L., W.Z., Z.C., C.W. and Z.X.; Investigation, P.L., Q.H. and X.L.; Resources, Y.L. and X.L.; Data curation, Q.H.; Writing—review & editing, P.L.; Supervision, W.Y.; Funding acquisition, P.L. and C.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science and Technology Project of Jiangxi Provincial Education Department (GJJ211311, GJJ201349, GJJ211312), the Jiangxi Provincial Natural Science Foundation of China (20224BAB214004, 20224BAB204001, 20202BABL204071), the National Natural Science Foundation of China (51962015, 52162028).

Data Availability Statement

The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, J.; Wang, Y.; Yin, Z.; He, R.; Wang, Y.; Qiao, J. In-situ growth of CoNi bimetal anchored on carbon nanoparticle/nanotube hybrid for boosting rechargeable Zn-air battery. J. Energy Chem. 2022, 66, 348–355. [Google Scholar] [CrossRef]
  2. Li, G.; Tang, Y.; Fu, T.; Xiang, Y.; Xiong, Z.; Si, Y.; Guo, C.; Jiang, Z. S, N co-doped carbon nanotubes coupled with CoFe nanoparticles as an efficient bifunctional ORR/OER electrocatalyst for rechargeable Zn-air batteries. Chem. Eng. J. 2022, 429, 132174. [Google Scholar] [CrossRef]
  3. Wang, X.; Tang, Y.; Lee, J.-M.; Fu, G. Recent advances in rare-earth-based materials for electrocatalysis. Chem. Catal. 2022, 2, 967–1008. [Google Scholar] [CrossRef]
  4. Kundu, A.; Mallick, S.; Ghora, S.; Raj, C.R. Advanced Oxygen Electrocatalyst for Air-Breathing Electrode in Zn-Air Batteries. ACS Appl. Mater. Interfaces 2021, 13, 40172–40199. [Google Scholar] [CrossRef] [PubMed]
  5. Bhatt, M.D.; Lee, J.Y. Advancement of Platinum (Pt)-Free (Non-Pt Precious Metals) and/or Metal-Free (Non-Precious-Metals) Electrocatalysts in Energy Applications: A Review and Perspectives. Energy Fuels 2020, 34, 6634–6695. [Google Scholar] [CrossRef]
  6. Liang, J.; Li, F.; Cheng, H.-M. On energy from fuel cells to metal-air batteries: Catalysts with two functions. Energy Storage Mater. 2018, 10, A1–A3. [Google Scholar] [CrossRef]
  7. Liu, Y.; Huang, H.; Xue, L.; Sun, J.; Wang, X.; Xiong, P.; Zhu, J. Recent advances in the heteroatom doping of perovskite oxides for efficient electrocatalytic reactions. Nanoscale 2021, 13, 19840–19856. [Google Scholar] [CrossRef]
  8. Beall, C.E.; Fabbri, E.; Schmidt, T.J. Perovskite Oxide Based Electrodes for the Oxygen Reduction and Evolution Reactions: The Underlying Mechanism. ACS Catal. 2021, 11, 3094–3114. [Google Scholar] [CrossRef]
  9. Matienzo, D.J.D.; Kutlusoy, T.; Divanis, S.; Bari, C.; Instuli, E. Benchmarking Perovskite Electrocatalysts’ OER Activity as Candidate Materials for Industrial Alkaline Water Electrolysis. Catalysts 2020, 10, 1387. [Google Scholar] [CrossRef]
  10. Li, F.; Mushtaq, N.; Su, T.; Cui, Y.; Huang, J.; Sun, M.; Singh, M.; Zhao, X.; Maliutina, K.; Zhang, Y.; et al. NCNT grafted perovskite oxide as an active bifunctional electrocatalyst for rechargeable zinc-air battery. Mater. Today Nano 2023, 21, 100287. [Google Scholar] [CrossRef]
  11. Mondal, R.; Ratnawat, H.; Mukherjee, S.; Gupta, A.; Singh, P. Investigation of the Role of Sr and Development of Superior SrDoped Hexagonal BaCoO3-δ Perovskite Bifunctional OER/ORR Catalysts in Alkaline Media. Energy Fuels 2022, 36, 3219–3228. [Google Scholar] [CrossRef]
  12. Hong, W.T.; Risch, M.; Stoerzinger, K.A.; Grimaud, A.; Suntivich, J.; Shao-Horn, Y. Toward the rational design of non-precious transition metal oxides for oxygen electrocatalysis. Energy Environ. Sci. 2015, 8, 1404–1427. [Google Scholar] [CrossRef] [Green Version]
  13. 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] [PubMed]
  14. 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]
  15. Sheelam, A.; Balu, S.; Muneeb, A.; Bayikadi, K.S.; Namasivayam, D.; Siddharthan, E.E.; Inamdar, A.I.; Thapa, R.; Chiang, M.-H.; Huang, S.-J.I.; et al. Improved Oxygen Redox Activity by High-Valent Fe and Co3+ Sites in the Perovskite LaNi1−xFe0.5xCo0.5xO3. ACS Appl. Energy Mater. 2022, 5, 343–354. [Google Scholar] [CrossRef]
  16. Suntivich, J.; Hong, W.T.; Lee, Y.-L.; Rondinelli, J.M.; Yang, W.; Goodenough, J.B.; Dabrowski, B.; Freeland, J.W.; Shao-Horn, Y. Estimating Hybridization of Transition Metal and Oxygen States in Perovskites from O K-edge X-ray Absorption Spectroscopy. J. Phys. Chem. C 2014, 118, 1856–1863. [Google Scholar] [CrossRef]
  17. Peña, M.A.; Fierro, J.L.G. Chemical Structures and Performance of Perovskite Oxides. Chem. Rev. 2001, 101, 1981–2017. [Google Scholar] [CrossRef]
  18. Flores-Lasluisa, J.X.; Huerta, F.; Cazorla-Amoros, D.; Morallon, E. Structural and morphological alterations induced by cobalt substitution in LaMnO3 perovskites. J. Colloid Interface Sci. 2019, 556, 658–666. [Google Scholar] [CrossRef] [Green Version]
  19. Zhang, C.; Cui, Y.; Ke, W.; Liang, Y.; Chao, Y.; Han, N.; Liang, P.; He, X. Enhanced cycling stability of La0.2Sr0.8CoO3−δ for oxygen evolution reaction via trace doping of Nb. Ceram. Int. 2022, 48, 36992–36999. [Google Scholar] [CrossRef]
  20. Bao, B.; Liu, Y.; Sun, M.; Huang, B.; Hu, Y.; Da, P.; Ji, D.; Xi, P.; Yan, C.-H. Boosting the Electrocatalytic Oxygen Evolution of Perovskite LaCo1- x Fex O3 by the Construction of Yolk-Shell Nanostructures and Electronic Modulation. Small 2022, 18, e2201131. [Google Scholar] [CrossRef]
  21. Seo, M.H.; Park, H.W.; Lee, D.U.; Park, M.G.; Chen, Z. Design of Highly Active Perovskite Oxides for Oxygen Evolution Reaction by Combining Experimental and ab Initio Studies. ACS Catal. 2015, 5, 4337–4344. [Google Scholar] [CrossRef]
  22. Xiong, D.; Rasaki, S.A.; Li, Y.; Fan, L.; Liu, C.; Chen, Z. Enhanced cathodic activity by tantalum inclusion at B-site of La0.6Sr0.4CO0.4Fe0.6O3 based on structural property tailored via camphor-assisted solid-state reaction. J. Adv. Ceram. 2022, 11, 1330–1342. [Google Scholar] [CrossRef]
  23. Kolla, P.; Nasymov, G.; Rajappagowda, R.; Smirnova, A. Bi-functionality of samarium- and praseodymium-based perovskite catalysts for oxygen reduction and oxygen evolution reactions in alkaline medium. J. Power Sources 2020, 446, 227234. [Google Scholar] [CrossRef]
  24. Yoshii, K.; Nakamura, A.; Abe, H.; Mizumaki, M.; Muro, T. Magnetism and transport of Ln0.5Sr0.5CoO3 (Ln = Pr, Nd, Sm, Eu and Gd). J. Magn. Magn. Mater. 2002, 239, 85–87. [Google Scholar] [CrossRef]
  25. Qiu, Y.; Gao, R.; Yang, W.; Huang, L.; Mao, Q.; Yang, J.; Sun, L.; Hu, Z.; Liu, X. Understanding the Enhancement Mechanism of A-Site-Deficient LaxNiO3 as an Oxygen Redox Catalyst. Chem. Mater. 2020, 32, 1864–1875. [Google Scholar] [CrossRef]
  26. 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]
  27. Ren, Y.; Küngas, R.; Gorte, R.J.; Deng, C. The effect of A-site cation (Ln = La, Pr, Sm) on the crystal structure, conductivity and oxygen reduction properties of Sr-doped ferrite perovskites. Solid State Ion. 2012, 212, 47–54. [Google Scholar] [CrossRef]
  28. Bousnina, M.A.; Dujardin, R.; Perriere, L.; Giovannelli, F.; Guegan, G.; Delorme, F. Synthesis, sintering, and thermoelectric properties of the solid solution La1−xSr xCoO3 ± δ (0 ≤ x ≤ 1). J. Adv. Ceram. 2018, 7, 160–168. [Google Scholar] [CrossRef] [Green Version]
  29. Wang, S.; Zan, J.; Qiu, W.; Zheng, D.; Li, F.; Chen, W.; Pei, Q.; Jiang, L. Evaluation of perovskite oxides LnBaCo2O5 + δ (Ln = La, Pr, Nd and Sm) as cathode materials for IT-SOFC. J. Electroanal. Chem. 2021, 886, 115144. [Google Scholar] [CrossRef]
  30. Wu, M.; Cai, H.; Jin, F.; Sun, N.; Xu, J.; Zhang, L.; Han, X.; Wang, S.; Su, X.; Long, W.; et al. Assessment of cobalt–free ferrite–based perovskite Ln0.5Sr0.5Fe0.9Mo0.1O3−δ (Ln = lanthanide) as cathodes for IT-SOFCs. J. Eur. Ceram. Soc. 2021, 41, 2682–2690. [Google Scholar] [CrossRef]
  31. Lee, K.T.; Manthiram, A. Comparison of Ln0.6Sr0.4CoO3−δ (Ln = La, Pr, Nd, Sm, and Gd) as Cathode Materials for Intermediate Temperature Solid Oxide Fuel Cells. J. Electrochem. Soc. 2006, 153, A794–A798. [Google Scholar] [CrossRef]
  32. Yi, L.; Le, C.; Chunhua, L.; Yaru, N.; Zhongzi, X. Effects of oxygen defects on structure and properties of Sm0.5Sr0.5CoO3−δ annealed in different atmospheres. J. Rere Earths 2013, 31, 1183–1190. [Google Scholar]
  33. Goldschmidt, V.M. Die Gesetze der Krystallochemie. Naturwissenschaften 1926, 14, 477–485. [Google Scholar] [CrossRef]
  34. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst 1976, A32, 751. [Google Scholar] [CrossRef]
  35. Zhang, Z.; Sun, H.; Li, J.; Shi, Z.; Fan, M.; Bian, H.; Wang, T.; Gao, D. S-doped CoMn2O4 with more high valence metallic cations and oxygen defects for zinc-air batteries. J. Power Sources 2021, 491, 229584. [Google Scholar] [CrossRef]
  36. Wang, Q.; Xue, Y.; Sun, S.; Li, S.; Miao, H.; Liu, Z. La0.8Sr0.2Co1−xMnxO3 perovskites as efficient bi-functional cathode catalysts for rechargeable zinc-air batteries. Electrochim. Acta 2017, 254, 14–24. [Google Scholar] [CrossRef]
  37. Kotha, V.; Karajagi, I.; Ghosh, P.C.; Panchakarla, L.S. Potassium-Substituted LaMnO3 as a Highly Active and Exceptionally Stable Electrocatalyst toward Bifunctional Oxygen Reduction and Oxygen Evolution Reactions. ACS Appl. Energy Mater. 2022, 5, 7297–7307. [Google Scholar] [CrossRef]
  38. Cheng, X.; Fabbri, E.; Nachtegaal, M.; Castelli, I.E.; Kazzi, M.E.; Haumont, R.; Marzari, N.; Schmidt, T.J. Oxygen Evolution Reaction on La1−xSrxCoO3 Perovskites: A Combined Experimental and Theoretical Study of Their Structural, Electronic, and Electrochemical Properties. Chem. Mater. 2015, 27, 7662–7672. [Google Scholar] [CrossRef]
  39. Wang, X.; Liu, H.; Li, M.; Li, J.; Lu, Y.; Wang, L.; Wang, Z.; Zhang, X.; Ding, X. Modulation of electronic structure and oxygen vacancies of perovskites SrCoO3−δ by sulfur doping enables highly active and stable oxygen evolution reaction. Electrochim. Acta 2021, 390, 138872. [Google Scholar] [CrossRef]
  40. Kim, S.; Jung, J.-W.; Song, D.; Cho, S.-H.; Kim, J.; Kim, J.K.; Oh, D.; Sun, H.; Cho, E.; Kim, I.-D.; et al. Exceptionally durable CoFe-exsolved Sr0.95Nb0.1Co0.7Fe0.2O3−δ catalyst for rechargeable Zn–air batteries. Appl. Catal. B Environ. 2022, 315, 121553. [Google Scholar] [CrossRef]
  41. Gui, L.; Wang, Z.; Zhang, K.; He, B.; Liu, Y.; Zhou, W.; Xu, J.; Wang, Q.; Zhao, L. Oxygen vacancies-rich Ce0.9Gd0.1O2−δ decorated Pr0.5Ba0.5CoO3−δ bifunctional catalyst for efficient and long-lasting rechargeable Zn-air batteries. Appl. Catal. B Environ. 2020, 266, 118656. [Google Scholar] [CrossRef]
  42. Da, Y.; Zeng, L.; Wang, C.; Gong, C.; Cui, L. A simple approach to tailor OER activity of SrxCo0.8Fe0.2O3 perovskite catalysts. Electrochim. Acta 2019, 300, 85–92. [Google Scholar] [CrossRef]
  43. Hua, B.; Li, M.; Sun, Y.-F.; Zhang, Y.-Q.; Yan, N.; Chen, J.; Thundat, T.; Li, J.; Luo, J.-L. A coupling for success: Controlled growth of Co/CoOx nanoshoots on perovskite mesoporous nanofibres as high-performance trifunctional electrocatalysts in alkaline condition. Nano Energy 2017, 32, 247–254. [Google Scholar] [CrossRef]
  44. Maitra, U.; Naidu, B.S.; Govindaraj, A.; Rao, C.N. Importance of trivalency and the e (g) configuration in the photocatalytic oxidation of water by Mn and Co oxides. Proc. Natl. Acad. Sci. USA 2013, 110, 11704–11707. [Google Scholar] [CrossRef] [Green Version]
  45. Huang, S.J.; Muneeb, A.; Sabhapathy, P.; Sheelam, A.; Bayikadi, K.S.; Sankar, R. Tailoring the Co4+/Co3+ active sites in a single perovskite as a bifunctional catalyst for the oxygen electrode reactions. Dalton Trans. 2021, 50, 7212–7222. [Google Scholar] [CrossRef]
  46. Yamada, I.; Kinoshita, M.; Oda, S.; Tsukasaki, H.; Kawaguchi, S.; Oka, K.; Mori, S.; Ikeno, H.; Yagi, S. Enhanced Catalytic Activity and Stability of the Oxygen Evolution Reaction on Tetravalent Mixed Metal Oxide. Chem. Mater. 2020, 32, 3893–3903. [Google Scholar] [CrossRef]
  47. Nie, R.; Deng, Y.; Yang, H.; Tan, Y.; Yuan, H.; Sagar, R.U.R.; Liang, T. Efficient oxygen evolution reaction in SrCo0.8Fe0.2O3−δ perovskite and surface reconstruction for practical zinc-air batteries. Appl. Surf. Sci. 2021, 552, 149509. [Google Scholar] [CrossRef]
  48. Wei, H.-l.; Tan, A.-d.; Hu, S.-z.; Piao, J.-h.; Fu, Z.-y. Efficient spinel iron-cobalt oxide/nitrogen-doped ordered mesoporous carbon catalyst for rechargeable zinc-air batteries. Chin. J. Catal. 2021, 42, 1451–1458. [Google Scholar] [CrossRef]
  49. Han, X.; Li, N.; Xiong, P.; Jung, M.G.; Kang, Y.; Dou, Q.; Liu, Q.; Lee, J.Y.; Park, H.S. Electronically coupled layered double hydroxide/MXene quantum dot metallic hybrids for high-performance flexible zinc–air batteries. InfoMat 2021, 3, 1134–1144. [Google Scholar] [CrossRef]
  50. Zhang, Z.; Liang, X.; Li, J.; Qian, J.; Liu, Y.; Yang, S.; Wang, Y.; Gao, D.; Xue, D. Interfacial Engineering of NiO/NiCo2O4 Porous Nanofibers as Efficient Bifunctional Catalysts for Rechargeable Zinc-Air Batteries. ACS Appl. Mater. Interfaces 2020, 12, 21661–21669. [Google Scholar] [CrossRef] [PubMed]
  51. Ye, X.; Han, Z.; Yan, Q.; Feng, J.-J.; Mei, L.-P.; Zhang, L.; Wang, A.-J. CoNi/MoC nanoparticles entrapped into N, P-codoped carbon nanotubes-on-nanosheets: A synergy of 1D@2D heterostructures with multiple active sites for rechargeable Zn-air battery. J. Power Sources 2021, 506, 230225. [Google Scholar] [CrossRef]
  52. Xiong, M.; Ivey, D.G. Sequentially Electrodeposited MnOX/Co-Fe as Bifunctional Electrocatalysts for Rechargeable Zinc-Air Batteries. J. Electrochem. Soc. 2017, 164, A1012–A1021. [Google Scholar] [CrossRef]
  53. Chen, Z.; Yu, A.; Higgins, D.; Li, H.; Wang, H.; Chen, Z. Highly active and durable core-corona structured bifunctional catalyst for rechargeable metal-air battery application. Nano Lett. 2012, 12, 1946–1952. [Google Scholar] [CrossRef] [PubMed]
  54. Li, P.; Wei, B.; Lü, Z.; Wu, Y.; Zhang, Y.; Huang, X. La1.7Sr0.3Co0.5Ni0.5O4+δ layered perovskite as an efficient bifunctional electrocatalyst for rechargeable zinc-air batteries. Appl. Surf. Sci. 2019, 464, 494–501. [Google Scholar] [CrossRef]
  55. Lee, D.U.; Choi, J.-Y.; Feng, K.; Park, H.W.; Chen, Z. Advanced Extremely Durable 3D Bifunctional Air Electrodes for Rechargeable Zinc-Air Batteries. Adv. Energy Mater. 2014, 4, 1301389. [Google Scholar] [CrossRef]
  56. Lee, D.U.; Park, H.W.; Park, M.G.; Ismayilov, V.; Chen, Z. Synergistic Bifunctional Catalyst Design based on Perovskite Oxide Nanoparticles and Intertwined Carbon Nanotubes for Rechargeable Zinc-Air Battery Applications. ACS Appl. Mater. Interfaces 2014, 7, 902–910. [Google Scholar] [CrossRef] [PubMed]
  57. Dwivedi, G.D.; Shukla, K.K.; Shahi, P.; Ghosh, A.K.; Nigam, A.K.; Chatterjee, S. Effect of Y-doping on the transport and magnetic properties of La0.5Sr0.5CoO3 and La0.7Sr0.3CoO3. J. Mater. Sci. 2013, 48, 1997–2001. [Google Scholar] [CrossRef]
  58. Li, P.; Tian, C.; Yang, W.; Zhao, W.; Lü, Z. LaNiO3 modified with Ag nanoparticles as an efficient bifunctional electrocatalyst for rechargeable zinc–air batteries. Front. Mater. Sci. 2019, 13, 277–287. [Google Scholar] [CrossRef]
  59. Ye, C.; Liu, J.; Zhang, Q.; Jin, X.; Zhao, Y.; Pan, Z.; Chen, G.; Qiu, Y.; Ye, D.; Gu, L.; et al. Activating Metal Oxides Nanocatalysts for Electrocatalytic Water Oxidation by Quenching-Induced Near-Surface Metal Atom Functionality. J. Am. Chem. Soc. 2021, 143, 14169–14177. [Google Scholar] [CrossRef]
  60. Yang, G.; Zhu, J.; Yuan, P.; Hu, Y.; Qu, G.; Lu, B.-A.; Xue, X.; Yin, H.; Cheng, W.; Cheng, J.; et al. Regulating Fe-spin state by atomically dispersed Mn-N in Fe-N-C catalysts with high oxygen reduction activity. Nat. Commun. 2021, 12, 1734. [Google Scholar] [CrossRef]
  61. Li, X.; Liu, Y.; Chen, H.; Yang, M.; Yang, D.; Li, H.; Lin, Z. Rechargeable Zn-Air Batteries with Outstanding Cycling Stability Enabled by Ultrafine FeNi Nanoparticles-Encapsulated N-Doped Carbon Nanosheets as a Bifunctional Electrocatalyst. Nano Lett. 2021, 21, 3098–3105. [Google Scholar] [CrossRef] [PubMed]
  62. Zhou, R.; Zheng, Y.; Jaroniec, M.; Qiao, S.Z. Determination of the Electron Transfer Number for the Oxygen Reduction Reaction: From Theory to Experiment. ACS Catal. 2016, 6, 4720–4728. [Google Scholar] [CrossRef]
  63. Zhou, S.; Zang, J.; Gao, H.; Tian, X.; Tian, P.; Song, S.; Wang, Y. Deflagration method synthesizing N, S co-doped oxygen-functionalized carbons as a bifunctional catalyst for oxygen reduction and oxygen evolution reaction. Carbon 2021, 181, 234–245. [Google Scholar] [CrossRef]
  64. Wang, M.; Lei, X.; Hu, L.; Zhang, P.; Hu, H.; Fang, J. High-performance Waste Biomass-derived Microporous Carbon Electrocatalyst with a Towel-like Surface for Alkaline Metal/air batteries. Electrochim. Acta 2017, 250, 384–392. [Google Scholar] [CrossRef]
  65. Zhou, Q.; Zhang, L.; Wang, X. Bimetallic ZIFs derived 3D acetylene black loading La2O3/Co bifunctional ORR/OER catalysts. Appl. Surf. Sci. 2023, 610, 155551. [Google Scholar] [CrossRef]
  66. Joshi, A.; Sood, P.; Gaur, A.; Rani, D.; Madaan, V.; Singh, M. Improved OER performance of an Anderson-supported cobalt coordination polymer by assembling with acetylene black. J. Mater. Chem. A 2022, 10, 12805–12810. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the as-prepared LSC, PSC and SSC powders.
Figure 1. XRD patterns of the as-prepared LSC, PSC and SSC powders.
Catalysts 13 00483 g001
Figure 2. SEM patterns of (a) LSC, (b) PSC and (c) SSC powders.
Figure 2. SEM patterns of (a) LSC, (b) PSC and (c) SSC powders.
Catalysts 13 00483 g002
Figure 3. XPS high-resolution spectra of the O 1s (ac) and Co 2p (df) for LSC, PSC and SSC. Red lines and circles—spectra before deconvolution.
Figure 3. XPS high-resolution spectra of the O 1s (ac) and Co 2p (df) for LSC, PSC and SSC. Red lines and circles—spectra before deconvolution.
Catalysts 13 00483 g003
Figure 4. (a) Cyclic voltammetry curves of various samples in O2-saturated 0.1 M KOH at a scan rate of 10 mV s−1; (b) LSV curves for ORR of LSC, PSC and SSC; (c) LSV curves of the PSC at various rotating rates from 400 to 2500 rpm; (d) K-L plots of PSC at various potentials and the calculated electron transfer number (n).
Figure 4. (a) Cyclic voltammetry curves of various samples in O2-saturated 0.1 M KOH at a scan rate of 10 mV s−1; (b) LSV curves for ORR of LSC, PSC and SSC; (c) LSV curves of the PSC at various rotating rates from 400 to 2500 rpm; (d) K-L plots of PSC at various potentials and the calculated electron transfer number (n).
Catalysts 13 00483 g004
Figure 5. (a) OER polarization curves on various catalysts in O2-saturated 0.1 M KOH at 1600 rpm with a scan rate of 10 mV s−1 and (b) corresponding Tafel plots derived from LSV polarization curves; (c) Overall polarization curves in the full OER/ORR region of PSC in O2-saturated 0.1 M KOH.
Figure 5. (a) OER polarization curves on various catalysts in O2-saturated 0.1 M KOH at 1600 rpm with a scan rate of 10 mV s−1 and (b) corresponding Tafel plots derived from LSV polarization curves; (c) Overall polarization curves in the full OER/ORR region of PSC in O2-saturated 0.1 M KOH.
Catalysts 13 00483 g005
Figure 6. (a) The open circuit voltage curves of primary zinc–air batteries with PSC; (b) Discharge polarization curves and corresponding power density plots; (c) Nyquist plots obtained by the electrochemical impedance spectroscopy for primary zinc–air batteries at 1.2 V vs. Zn; (d) The long-term discharge performance (inset: a photograph of the working-LED lighted by two assembled primary zinc–air batteries in series connection).
Figure 6. (a) The open circuit voltage curves of primary zinc–air batteries with PSC; (b) Discharge polarization curves and corresponding power density plots; (c) Nyquist plots obtained by the electrochemical impedance spectroscopy for primary zinc–air batteries at 1.2 V vs. Zn; (d) The long-term discharge performance (inset: a photograph of the working-LED lighted by two assembled primary zinc–air batteries in series connection).
Catalysts 13 00483 g006
Figure 7. (a) Charge and discharge polarization curves; (b) Galvanostatic discharge-charge cycling curves at 10 mA cm−2 of rechargeable zinc–air batteries.
Figure 7. (a) Charge and discharge polarization curves; (b) Galvanostatic discharge-charge cycling curves at 10 mA cm−2 of rechargeable zinc–air batteries.
Catalysts 13 00483 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, P.; Huang, Q.; Yang, W.; Tian, C.; Liu, Y.; Zhao, W.; Lu, X.; Cao, Z.; Wang, C.; Xie, Z. The Role of the A-Site Cation on the Bifunctional Electrocatalytic Activities of Ln0.5Sr0.5CoO3-δ (Ln = La, Pr and Sm) for Rechargeable Zinc–Air Batteries. Catalysts 2023, 13, 483. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13030483

AMA Style

Li P, Huang Q, Yang W, Tian C, Liu Y, Zhao W, Lu X, Cao Z, Wang C, Xie Z. The Role of the A-Site Cation on the Bifunctional Electrocatalytic Activities of Ln0.5Sr0.5CoO3-δ (Ln = La, Pr and Sm) for Rechargeable Zinc–Air Batteries. Catalysts. 2023; 13(3):483. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13030483

Chicago/Turabian Style

Li, Pengzhang, Qing Huang, Wei Yang, Chuanjin Tian, Yumin Liu, Wenyan Zhao, Xiaojie Lu, Zhenbao Cao, Changan Wang, and Zhipeng Xie. 2023. "The Role of the A-Site Cation on the Bifunctional Electrocatalytic Activities of Ln0.5Sr0.5CoO3-δ (Ln = La, Pr and Sm) for Rechargeable Zinc–Air Batteries" Catalysts 13, no. 3: 483. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13030483

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