Next Article in Journal
Modified Ammonium Polyphosphate and Its Application in Polypropylene Resins
Previous Article in Journal
Phase Transition and Optical Properties of VO2 and Al: ZnO/VO2 Thin Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancement in the Electrochemical Performance of Strontium (Sr)-Doped LaMnO3 as Supercapacitor Materials

1
Institute of Material Science and Engineering, Jiangsu University of Science and Technology, Zhenjiang 212003, China
2
School of Science, Jiangsu University of Science and Technology, Zhenjiang 212003, China
*
Authors to whom correspondence should be addressed.
Coatings 2022, 12(11), 1739; https://doi.org/10.3390/coatings12111739
Submission received: 8 October 2022 / Revised: 8 November 2022 / Accepted: 9 November 2022 / Published: 13 November 2022
(This article belongs to the Topic Properties of the Corroding Interface)

Abstract

:
In this study, Strontium (Sr)-doped perovskite lanthanum manganite (La1−xSrxMnO3) nanoparticles were prepared by the sol–gel method and used as electrode materials of supercapacitors. Microstructures, morphologies, and electrochemical properties of the samples were characterized by X-ray diffraction (XRD), scanning electron microscopy (SEM), a transmission electron microscope (TEM), Brunauer–Emmett–Teller (BET) surface area measurements, cyclic voltammetry (CV), and galvanostatic charge/discharge (GCD) cycling. Investigations demonstrated that the La0.85Sr0.15MnO3 nanoparticles had a maximum specific capacitance of 185.5 F/g at a current density of 0.5 A/g and a low charge transfer resistance (0.38 Ω) in 3 M KOH aqueous electrolyte solutions. La0.85Sr0.15MnO3 electrode yields the highest capacitance behavior because of the larger specific surface area, lower charge transfer resistance, and higher concentration of oxygen vacancy. This result demonstrates that Sr doping significantly improved the electrochemical properties of the LaMnO3 system. The anion-intercalation mechanism was examined by a charge–discharge process. This provides a promising electrode material for supercapacitors.

1. Introduction

Due to the escalating global energy crisis and environmental pollution, the development of innovative and high-efficiency energy storage technologies has drawn substantial interest. To this end, electrochemical energy storage systems, such as lithium batteries and supercapacitors, offer enormous opportunities for lowering the use of fossil fuels and pollution [1,2,3] because of their quick discharging and charging times, significant power density, and extended cycle life. Specifically, supercapacitors have been identified as the best energy storage technology method, attracting substantial interest. However, compared to batteries, the practical applications of supercapacitors are limited because of their lower energy density. Therefore, considerable effort has been made toward the design of electrode materials for high-energy-density supercapacitors to address this limitation [4,5,6].
On the other hand, studies have shown that supercapacitor electrode materials considerably influence their electrochemical performance, prompting the extensive exploration of carbon compounds, conductive polymers, and metal oxides which have been extensively explored as electrochemical devices [7,8]. Among these compounds, metal oxides have attracted great interest because of their high energy concentration and more consistent electrochemical characteristics [9]. Metal oxides such as Ru2O, MnO2, NiCo2O4, Co3O4, V2O5, and Fe3O4 have also been considered as important primary materials because of their tunable surface morphology, different flexible oxidation states, and high structural stability [10,11,12,13,14,15,16]. Perovskite materials have become an important component in solid oxide fuel cells as well due to their excellent electrical and ionic properties [17]. Due to these materials having many oxygen vacancies that allow charge storage, their electronic and ionic conductivity satisfy the requirements of supercapacitors; thus, perovskite materials are even more preferable as supercapacitor electrode materials.
In a previous study, Mefford et al. [18] reported the anion insertion technique for LaMnO3 perovskite as a supercapacitor electrode material for the first time. Based on their findings, researchers have been drawn to lanthanide-based perovskite materials due to their excellent thermal stability, simple manufacturing, oxygen preservation, low price, and high electrical conductivity [19,20,21]. The oxygen vacancies in perovskites are considered as charge storage sites for pseudocapacitance [20]. Therefore, the energy densities of perovskite supercapacitors could be maximized by increasing the oxygen vacancies. The oxygen-vacancy concentration in perovskite oxides can be tailored by partial substitution on A-site and/or B-site, resulting in effective control over the crystal structure, electrical conductivity, and electrochemical performance [22,23,24,25,26]. Doping is an effective way to alter the oxygen vacancies in a LaMnO3 matrix [22,23,24,25,27]. For instance, Ma et al. [24] showed that the specific capacitance of the La0.85Sr0.15MnO3 electrode (129.0 mF/cm2) was greater than that of LaMnO3 (32.6 mF/cm2). Then, Alexander et al. [28] used a low-valence A-site doping concept to synthesize La1−xSrxMnO3−δ perovskite materials with high charge storage, which proved that La0.2Sr0.8MnO2.7 had the highest specific capacitance. Lu et al. [29] reported the results of a preliminary study of the pseudocapacitance features of strontium-doped lanthanum manganite. They reported no capacitance degradation after 1000 cycles and a specific capacitance of 56 F g−1 at a scan rate of 5 mV s−1. The above-mentioned results show that Strontium (Sr) doping varied when LaMnO3 had maximum specific capacitance. Therefore, exploring the impact of Sr doping on the electrochemical performance of LaMnO3 perovskite materials is interesting and necessary.
Due to the sol–gel method with the advantages of rapid preparation of nanomaterials, it was easy to evenly and quantitatively incorporate some trace elements. To this end, the sol–gel method was used to create several La1−xSrxMnO3 (x = 0, 0.05, 0.1, 0.15, and 0.2) formulations, after which the impact of the properties and structural morphology of La1−xSrxMnO3 (x = 0, 0.05, 0.1, 0.15, and 0.2) on its electrochemical performance was determined. Investigations revealed that 185.5 F/g was the highest specific capacitance of the investigated La0.85Sr0.15MnO3 nanoparticles.

2. Experimental Section

2.1. Material Preparation

The specific experimental procedure for synthesizing the La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) samples using the sol–gel method is as follows: Firstly, the mass of the citric acid monohydrate (≥99.5%) is 0.4224 g, and the mass of La(NO3)3·6H2O (≥99.9%), MnCl2·4H2O (≥99%), and Sr(NO3)2 (≥99.5%) according to stoichiometric ratios of La1−xSrxMnO3 were simultaneously solubilized in 30 mL of deionized water in turn and stirred for 24 h to obtain a clear solution. Next, the transparent solution was then dehydrated in an oven at 80°C for 72 h. Finally, the obtained solid powders were thoroughly ground, moved to a tube furnace, calcined in air at a heating rate of 5°C/min from room temperature to 700°C, and then held at 700°C for 2 h to yield La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2).

2.2. Characterization

The thermal decomposition temperature of the samples was determined by employing a thermogravimetry analysis (Diamond TG/DTA, Perkin Elmer S. A., Waltham, MA, USA). First, with continuous ventilation, a part of every item weighing 4–6 mg was treated at 800 °C and a scan temperature of 15 °C min–1. Then, the stages of generated nanoparticles were assessed through X-ray powder diffraction ( D8 Advance, Bruker, Karlsruhe, Germany) with a checking width of 20°–60° and a reading speed of 6°/min. Subsequently, microstructures of materials were studied utilizing a field-emission scanning electron microscopy (FE-SEM, Merlin Compact, Carl Zeiss, Oberkochen, Germany) and a transmission electron microscope (TEM, 2100F, JEOL, Shoshima City, Japan). The Brunauer–Emmett–Teller (BET) surface area and pore size distribution were measured using a N2 adsorption–desorption physisorption analyzer (ASAP 2020, Micromeritics, Norcross, Ga, USA). The compositions’ items, as well as molecular valency of the specimens, were evaluated through X-ray photoelectron spectroscopy (XPS, Thermo Fisher Scientific K-Alpha, Waltham, MA, USA). XPS with an Al K alpha X-ray source was used to examine the valance distributions of the samples.

2.3. Electrochemical Measurements

The active material (80 wt.%), acetylene black (10 wt%), and polyvinylidene fluoride (PVDF, 10 wt.%) were blended in N-methylpyrrolidone (NMP) for 24 h. The mass of the active materials used is 100 mg, and the volume of NMP is about 0.5 mL. A 1 mg active material was pasted onto carbon paper (1 cm × 1 cm) current collector and baked in a vacuum stove at 80 °C for 24 h to produce electrodes that need evaluation. Afterward, a three-electrode system with a Hg/HgO reference electrode, comprising a 3 M KOH mixture as the electrolyte, was utilized to investigate the electrochemical behavior of electrodes, followed by recordings of the electrochemical impedance spectra (EIS), cyclic voltammetric (CV) curves, and constant current charge–discharge (GCD) curves of the manufactured electrodes from an analyzer (CHI660E). The specific capacity of an electrode was calculated from the GCD curves using Equation (1):
C s = I Δ t m Δ V
where Cs is the specific capacitance (F g−1), I is the specific current (A), Δt is the time for discharge (s), m is the mass of the active material (g), and ΔV is the potential range (V).

3. Results and Discussion

3.1. Microstructure and Morphology

While this study assessed the thermal decomposition temperature of La1−xSrxMnO3 precursors (x = 0.0, 0.05, 0.1, 0.15, and 0.2) using thermogravimetric assessment, that of the LaMnO3 precursors were examined at an environment temperature of 50–800 °C with the temperature being increased at 15 °C/min. Figure 1a demonstrates the TG curve of the LaMnO3 precursor. Investigations revealed a 15 wt.% drop in the weight of the LaMnO3 precursors at 250 °C, mostly attributable to the evaporation of the natural solvents, gas, and humidity from the surface. At 380 °C, however, the precursor’s weight reduced rapidly due to the breakdown of the citric acid monohydrate, La(NO3)3·6H2O, MnCl2·4H2O, and Sr(NO3)2; then, at temperatures of more than 650 °C, the weight of the precursor remained unchanged. These findings indicate that thermal decomposition temperatures should exceed 650 °C. Therefore, a calcination temperature of 650 °C was adopted for the La1−xSrxMnO3 precursors. However, XRD studies indicated that LaMnO3 was not synthesized, causing the calcination temperature to be raised to 700 °C, with subsequent XRD patterns proving that LaMnO3 had been produced. In this investigation, La1−xSrxMnO3 precursors were thus synthesized at a calcination temperature of 700 °C.
On the other hand, Figure 1b illustrates the XRD system of La1−xSrxMnO3 (x = 0, 0.05, 0.1, 0.15, and 0.2) samples held at 700°C for 2 h. In contrast to the standard card (PDF 50–0297), all peaks at 23.0°, 32.6°, 40.1°, 46.7°, 52.7°, and 58.2°, which belonged to the (110), (200), (022), (220), (130), and (024) crystal planes, respectively, appeared to correlate with those of LaMnO3; no additional peaks were visible. This finding suggests that samples of La1−xSrxMnO3 were achieved at the specified calcination temperature. Investigations also revealed that the (200) deviation peaks of La1−xSrxMnO3 examples moved to the left as the Sr doping concentration increased. In addition, per the Bragg formula, 2dsinθ = nλ (where d is the crystal plane area, θ is the angle of incidence, λ is the wavelength, and n is the count of reflection levels), the lattice constant of La1−xSrxMnO3 samples was also raised with increasing Sr doping because the ionic radius of Sr (1.18 Å) was greater than that of La (1.06 Å) [24].
In contrast, the morphologies of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) specimens are illustrated in Figure 2a–e. The graphs demonstrate that the La1−xSrxMnO3 samples had a homogeneous particle distribution and tiny grain size; the grain sizes of the La1−xSrxMnO3 nanoparticles reduced as Sr doping increased because the lattice flaws in the La1−xSrxMnO3 nanoparticles rose as Sr doping increased, negatively affecting grain expansion and lowering grain size [30]. Subsequently, TEM was used to assess the effect of Sr doping on the La1−xSrxMnO3 nanoparticle’s structure by characterizing the morphology of LaMnO3 and La0.85Sr0.15MnO3 samples. The results are shown in Figure 2f–i. Figure 2f,h shows that LaMnO3 and La0.85Sr0.15MnO3 samples comprised fine nanoparticle stacks with small grain sizes and uniform distribution; the grain size of the La0.85Sr0.15MnO3 sample (~25 nm) was considerably smaller than that of the LaMnO3 sample (~50 nm). Furthermore, the electron diffraction patterns of both the LaMnO3 {Figure 2g} and La0.85Sr0.15MnO3 {Figure 2i} samples showed concentric ring patterns, indicating the polycrystalline nature of the LaMnO3 and La0.85Sr0.15MnO3 samples. Then, the XRD data proposed that each ring can be assigned to various planes, as seen in the relevant images.
The surface area and pore size distribution of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) nanoparticles were measured using a N2 adsorption–desorption physisorption analyzer. Figure 3a,b depicts the isotherm description and orifice size distribution curve. Investigations revealed that the isothermal curves of the La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) samples were a type IV hysteresis coil, indicating the mesoporous region of the synthesized samples [31]. Figure 3 shows the specific surface area and average pore diameters of La1−xSrxMnO3. The specific surface area of the La0.95Sr0.05MnO3, La0.9Sr0.1MnO3, La0.85Sr0.15MnO3, and La0.8Sr0.2MnO3 sample was 24.82, 29.61, 31.56, and 44.33 m2 g−1, respectively, which is larger than that of the LaMnO3 sample (22.25 m2 g−1). The average pore diameters of the LaMnO3, La0.95Sr0.05MnO3, La0.9Sr0.1MnO3, La0.85Sr0.15MnO3, and La0.8Sr0.2MnO3 samples were 2.81, 2.03, 2.04, 3.05, and 3.14 nm, respectively. Note that the La0.85Sr0.15MnO3 sample had a larger specific surface area.
Figure 4 shows the oxidation states and composition of the La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) specimens. In particular, Figure 4a reveals the XPS survey scan, which indicated that the prepared samples included La, Sr, Mn, and O. However, Figure 4b displays the La 3d XPS spectra, with the peaks at 834.4 and 838 eV being the spin–orbit peaks of La 3d5/2 and including extra peaks at 850.8 and 855.4 eV representing that of La 3d3/2 [32]. These results indicate that La in La1−xSrxMnO3 was present in the +3 valence state, which is identical to the results reported by Zhang et al. [33]. On the other hand, the Mn 2p spectra of La1−xSrxMnO3 illustrated a wide outflow line width and a unique top range, confirming the existence of Mn in various oxidation conditions [34]. Accordingly, studies have widely utilized a simple tip-differential examination of Mn 2p to determine the manganese oxidation status of manganese with combined valences [22,24,35]. In contrast, Figure 4c illustrates the peak-differentiating evaluation of the Mn 2p spectra for the La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) samples. The two main peaks at 639.98 and 651.78 eV correlated with Mn 2p3/2 and Mn 2p1/2 [24], and the various numbers were kept at ~11.8 eV. Every peak was adjusted to three Mn oxidation states—Mn2+, Mn3+, and Mn4+. Investigations also revealed that while the features of the peaks at 641.2 eV and 652.3 eV related to Mn2+, those at 642.3 and 653.9 eV related to Mn3+, and those at 644.0 and 655.5 eV related to Mn4+ [22]. Conversely, Figure 4d shows oxygen’s high-resolution XPS spectra. The results illustrate that the three spikes adjusted to the greater fields in the O 1s part at 529.29, 530.9, and 533.1 eV correlated with O1, O2, and O3 [32], where O1 correlated with lattice oxygen (O2−), O2 correlated with the exterior extracted oxygen (O, O2, and O22−), and O3 correlated with the oxygen-holding groups (OH). The proportional amounts of the three oxygen models are shown in Table 1. The oxygen molecules of the electrode items were inextricably connected with the conductivity, as well as the increased content of the O2 types; the chemical absorption possibility of OH− became smoother, proposing that this connection optimized the activity of outward redox processes and their electrochemical qualities [36]. Furthermore, while the chemical oxygen vacancy of O22− or O corresponded with oxygen deficiency, the density of oxygen vacancy was represented by the ratio of the area. The outside O2/O1 molar proportion reduced in the following arrangement: La0.85Sr0.15MnO3 (1.34) > La0.8Sr0.2MnO3 (1.3) > La0.9Sr0.1MnO3 (1.16) > La0.95Sr0.05MnO3 (1.08) > LaMnO3 (1.06), illustrating that La0.85Sr0.15MnO3 had the highest electrochemical efficiency among the supercapacitors [37]. Finally, Figure 4e demonstrates the high-resolution XPS spectrum of Sr 3d, which revealed that the spike features at 131.08 and 132.8 eV correlated with the 3d5/2 and 3d3/2 spin–orbit peaks, confirming that the sample chemical valence state of Sr was +2 [38].

3.2. Electrochemical Properties

Figure 5 illustrates the CV curves of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) electrodes at various testing voltage situations using 3 M KOH as the electrolyte. As the scan rate rose, the CV curves remained the same, suggesting great reversibility. However, the CV curves revealed a typical pseudocapacitor behavior, indicating a relationship between the maximum redox at 0.15 V and −0.15 V and a shift in the valence state of manganese ions from trivalent to divalent [22,24], respectively. Therefore, as per the anion insertion pathway, the redox reactions of LaMnO3 during charging and discharging trials were illustrated as follows [24]:
La [ Mn 2 δ 2 + ;   Mn 1 2 δ 3 + ] O 3 δ + 2 δ OH LaMn 3 + O 3 + 2 δ e + δ H 2 O
LaMn 3 + O 3 + 2 δ OH La [ Mn 2 δ 4 + ;   Mn 1 2 δ 3 + ] O 3 + δ + 2 δ e + δ H 2 O
The oxygen intercalation process in LaMnO3 samples: First, La [ Mn 2 δ 2 + ; Mn 1 2 δ 3 + ] O 3 δ extracts OH from the electrolyte to create O2− and H2O. Then, O2− is pushed into the perovskite lattice, thereby diffusing from t the borders of the O2 octahedron and finally filling oxygen openings in the composition. As a result, the required Mn2+ changes to Mn3+ and finally creates a neutral LaMn3+O3. Second, manganese arises from the center of the oxygen octahedron, after which excessive oxygen is incorporated into the exterior, thus causing the generation of La [ Mn 2 δ 4 + ; Mn 1 2 δ 3 + ] O 3 + δ [18]. Consequently, the divalent component (Sr2+) replaces a part of La3+ in the A site due to the presence of extra oxygen vacancies. The reaction then changes to the following:
La 0.85 Sr 0.15 [ Mn 2 δ 2 + ;   Mn 1 2 δ 3 + ] O 2.925 δ + 2 δ OH La 0.85 Sr 0.15 Mn 3 + O 2.925 + 2 δ e + δ H 2 O
La 0.85 Sr 0.15 Mn 3 + O 2.925 + 2 δ OH La [ Mn 2 δ 4 + ; Mn 1 2 δ 3 + ] O 2.925 + δ + 2 δ e + δ H 2 O
For the La0.85Sr0.15MnO3 samples, the oxidation process from Mn2+ to Mn3+ was similar to that of Mefford et al. [18], where the oxygen vacancy was filled by O2− intercalation. Although all Mn2+ were oxidized to Mn3+, La0.85Sr0.15Mn3+O2.925 remained hypoxic. Therefore, two stages were identified in the process of Mn3+ oxidation to Mn4+. In the first stage, when δ ≤ 0.075, Mn3+ was oxidized to Mn4+ by constantly extracting O2− to occupy the residual oxygen vacancies. As a result, many oxygen vacancies moved to the material surface to form La 0.85 Sr 0.15 [ Mn 0.15 4 + ; Mn 0.85 3 + ] O 3 ( δ = 0.075 ) . The second stage is similar to that in LaMnO3, in which the oxidation of Mn3+ to Mn4+ caused the formation of La 0.85 Sr 0.15 [ Mn 2 δ 4 + ; Mn 1 2 δ 3 + ] O 2.925 + δ .
Figure 5 shows that the La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) electrodes had a large voltage window (−0.5–0.5 V). As the Sr content rose, the closed area of the CV curves developed and eventually declined at the same scanning speed. La0.85Sr0.15MnO3 had the largest closed region. However, when the sample was doped with Sr, the oxygen vacancies increased, in addition to the pseudocapacitance of the oxide, thereby increasing the specific capacitance value. The specific capacitance rose and dropped with increasing Sr doping levels due to the ion exchange process; the results are shown in Figure 5f. The process of oxygen admission may illustrate this capacity trend by raising Sr doping. As Sr doping increased, the concentrations of oxygen vacancies in the samples of Table 1 increased and then decreased. This changing trend was consistent with the specific capacitance. Furthermore, while the different radii of Sr2+ and La3+ resulted in lattice mismatch, which decreased the tolerance factor, t, it also affected the perovskite structure, indicating that excessive Sr doping could result in the lattice distortion of La1−xSrxMnO3 because oxygen ions were not transported through the oxygen vacancies, which reduces the charge storage efficiency and specific capacitance [24]. Hence, increasing the specific surface area is useful for improving electrochemical performance.
Figure 6 shows the constant current charge–discharge (GCD) curves for the La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) electrodes at diverse current concentrations. Results revealed that the samples discharged faster as the charge–discharge current density rose. This finding was due to the fact that the specific capacitance decreased as the internal resistance rose with increasing current density. Consequently, the discharge time will be lowered. Alterations in the curve slope were also proposed to result from the alterations in the valence state from Mn2+ to Mn3+. The specific capacitances of synthesized La1−xSrxMnO3 at a current density of 0.5 A/g are 67, 97, 115, 185.5, and 137.5 F/g, respectively. Figure 6f displays the specific capacitance of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) as a function of current density. Due to the ion-exchange process, particular capacitance values fell as the current density rose, indicating that the trend of specific capacitance corresponded to the density of oxygen vacancies in the mechanism. These results also demonstrate that the capacitance of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) specimens could be elevated by a La site exchange.
Figure 7 shows the Nyquist plots of the La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) electrodes measured by EIS within the frequency variety of 100 kHz to 0.01 Hz. Generally, while impedance diagrams were divided into linear and semicircular parts, the semicircular curve region was in the AC impedance’s high-frequency region, indicating that a redox reaction occurred on the electrode surface. Investigations also revealed that the straight-line region was in the low-frequency region of AC impedance, which shows the capacitive resistance generated at the electrode surface. Subsequently, the resistive range was adjusted to a circuit (displayed in the inset), usually containing an equivalent series resistance (Rs), charge transfer resistance (Rct), Warburg element (Wo), and constant phase element (CPE). Rs is a combination of the electrolyte ionic resistance, the intrinsic resistance of active material, and the contact resistance between the active material and the current collector. According to the Nyquist plot, the Rct values were 0.49, 0.47, 0.41, 0.38, and 4.4 Ω (as shown in Figure 7b), respectively, with the La0.85Sr0.15MnO3 sample having a minimal charge transfer impedance, demonstrating excellent electrochemical features. Moreover, La1−xSrxMnO3 (x = 0.05, 0.1, 0.15, and 0.2), which had extra oxygen vacancies, was better for the insertion of oxygen than LaMnO3, thus resulting in a smaller impedance to ion diffusion. Hence, these electrochemical records confirmed that the La0.85Sr0.15MnO3 nanoparticles were favorable negative electrode materials for supercapacitors.

4. Conclusions

La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) nanoparticles were successfully synthesized using the sol–gel approach. As the Sr doping content increased, the grain size decreased, the specific surface area increased, and the density of oxygen vacancy in the LaMnO3 regulation initially rose and then reduced, indicating that the particular capacitance of the La1−xSrxMnO3 nanoparticles followed a similar trend. Investigations also revealed that the La0.85Sr0.15MnO3 nanoparticles with additional intrinsic oxygen vacancies demonstrated a maximum specific capacitance (185.5 F/g). This finding demonstrated that a particular capacitance of the LaMnO3 system can be enhanced by La site substitution. As a result, this study successfully developed a potential electrode material for supercapacitor applications.

Author Contributions

Conceptualization and data curation, X.Y.; methodology and formal analysis, X.J.; investigation, J.W.; investigation, L.W.; writing—original draft preparation, S.D.; writing—review and editing, Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 51702132) and the Postgraduate Research & Practice Innovation Program of Jiangsu Province (Grant No. KYCX20_3132).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge financial support from the National Natural Science Foundation of China (51702132). Song-Tao Dong acknowledges the open project of the National Laboratory of Solid State Microstructures at Nanjing University. Jun-Lin Wei acknowledges financial support from the Postgraduate Research & Practice Innovation Program of Jiangsu Province (Grant No. KYCX20_3132).

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Simon, P.; Gogotsi, Y.; Dunn, B. Where do batteries end and supercapacitors begin. Science 2014, 343, 1210–1211. [Google Scholar] [CrossRef] [Green Version]
  2. Chu, S.; Majumdar, A. Opportunities and challenges for a sustainable energy future. Nature 2012, 488, 294–303. [Google Scholar] [CrossRef] [PubMed]
  3. Zhong, C.; Deng, Y.; Hu, W.; Qiao, J.; Zhang, L.; Zhang, J. A review of electrolyte materials and compositions for electrochemical supercapacitors. Chem. Soc. Rev. 2015, 44, 7484–7539. [Google Scholar] [CrossRef] [PubMed]
  4. Bose, S.; Kuila, T.; Mishra, A.K.; Rajasekar, R.; Kim, N.H.; Lee, J.H. Carbon-based nanostructured materials and their composites as supercapacitor electrodes. J. Mater. Chem. 2011, 22, 767–784. [Google Scholar] [CrossRef]
  5. Snook, G.A.; Kao, P.; Best, A.S. Conducting-polymer-based supercapacitor devicesand electrodes. J. Power Sources 2011, 196, 1–12. [Google Scholar] [CrossRef]
  6. Yuan, C.; Yang, L.; Hou, L.; Shen, L.; Zhang, X.; Lou, X.W.D. Growth of ultrathin mesoporous Co3O4 nanosheet arrays on Ni foam for high-performance electrochemical capacitors. Energy Environ. Sci. 2012, 5, 7883–7887. [Google Scholar] [CrossRef]
  7. Wang, G.P.; Zhang, L.; Zhang, J.J. A review of electrode materials for electrochemical supercapacitors. Chem. Soc. Rev. 2012, 41, 797–828. [Google Scholar] [CrossRef] [Green Version]
  8. Xue, Q.; Sun, J.; Huang, Y.; Zhu, M.; Pei, Z.; Li, H.; Wang, Y.; Li, N.; Zhang, H.; Zhi, C. Recent progress on flexible and wearable supercapacitors. Small 2017, 13, 1701827. [Google Scholar] [CrossRef] [PubMed]
  9. Chen, H.; Jiang, G.; Yu, W.; Liu, D.; Liu, Y.; Li, L.; Huang, Q.; Tong, Z. Electrospun carbon nanofibers coated with urchinlike ZnCo2O4 nanosheets as a flexible electrode material. J. Mater. Chem. 2016, 4, 5958–5964. [Google Scholar] [CrossRef]
  10. Zhu, Y.; Ji, X.; Pan, C.; Sun, Q.; Song, W.; Fang, L.; Chen, Q.; Banks, C.E. A carbon quantum dot decorated RuO2 network: Outstanding supercapacitances under ultrafast charge and discharge. Energy Environ. Sci. 2013, 6, 3665–3675. [Google Scholar] [CrossRef]
  11. Chang, J.K.; Huang, C.H.; Tsai, W.T.; Deng, M.J.; Sun, I.-W. Ideal pseudocapacitive performance of the Mn oxide anodized from the nanostructured and amorphous Mn thin film electrodeposited in BMP-NTf2 ionic liquid. J. Power Sources 2008, 179, 435–440. [Google Scholar] [CrossRef]
  12. Zhang, G.; Lou, X.W. Controlled Growth of NiCo2O4 Nanorods and Ultrathin Nanosheets on Carbon Nanofibers for High-Performance Supercapacitors. Sci. Rep. 2013, 3, 1470–1475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wang, B.; Zhu, T.; Wu, H.B.; Xu, R.; Chen, J.S.; Lou, X.W. Porous Co3O4 nanowires derived from long Co(CO3)0.5(OH)0.11H2O, nanowires with improved supercapacitive properties. Nanoscale 2012, 4, 2145–2149. [Google Scholar] [CrossRef] [PubMed]
  14. Pan, A.Q.; Wu, H.B.; Zhang, L.; Lou, X.W. Uniform V2O5 nanosheet-assembled hollow microflowers with excellent lithium storage properties. Energy Environ. Sci. 2013, 6, 1476–1479. [Google Scholar] [CrossRef]
  15. Wu, Z.S.; Yang, S.; Sun, Y.; Parvez, K.; Feng, X.; Müllen, K. 3D nitrogen-doped graphene aerogel-supported Fe3O4 nanoparticles as efficient electrocatalysts for the oxygen reduction reaction. J. Am. Chem. Soc. 2012, 134, 9082–9085. [Google Scholar] [CrossRef]
  16. Hussain, S.; Javed, M.S.; Ullah, N.; Shaheen, A.; Aslam, N.; Ashraf, I.; Abbas, Y.; Wang, M.; Liu, G.; Qiao, G. Unique hierarchical mesoporous LaCrO3 perovskite oxides for highly efficient electrochemical energy storage applications. Ceram. Int. 2019, 45, 15164–15170. [Google Scholar] [CrossRef]
  17. Jin, T.; Wang, X.; Jiao, L. Recent progress in electrospinning method for secondary ion batteries and electrocatalysis. Sci. Sin. Chim. 2019, 49, 692–703. [Google Scholar] [CrossRef] [Green Version]
  18. Mefford, J.T.; Hardin, W.G.; Dai, S.; Johnston, K.P.; Stevenson, K.J. Anion charge storage through oxygen intercalation in LaMnO3 perovskite pseudocapacitor electrodes. Nat. Mater. 2014, 13, 726–732. [Google Scholar] [CrossRef]
  19. Ho, K.H.; Wang, J. Hydrazine reduction of LaNiO3 for active materials in supercapacitors. J. Am. Ceram. Soc. 2017, 100, 4629–4637. [Google Scholar] [CrossRef]
  20. Elsiddig, Z.A.; Xu, H.; Wang, D.; Zhang, W.; Guo, X.; Zhang, Y.; Sun, Z.; Chen, J. Modulating Mn4+ ions and oxygen vacancies in nonstoichiometric LaMnO3 perovskite by a facile sol–gel method as high-performance supercapacitor electrodes. Electrochim. Acta 2017, 253, 422–429. [Google Scholar] [CrossRef]
  21. Ma, P.P.; Zhu, B.; Lei, N.; Liu, Y.K.; Yu, B.; Lu, Q.L.; Dai, J.M.; Li, S.H.; Jiang, G.H. Effect of Sr substitution on structure and electrochemical properties of perovskite-type LaMn0.9Ni0.1O3 nanofibers. Mater. Lett. 2019, 252, 23–26. [Google Scholar] [CrossRef]
  22. Mo, H.Y.; Nan, H.S.; Lang, X.Q.; Liu, S.J.; Qiao, L.; Hu, X.Y.; Tian, H.W. Influence of calcium doping on performance of LaMnO3 supercapacitors. Ceram. Int. 2018, 44, 9733–9741. [Google Scholar] [CrossRef]
  23. Zhang, Y.H.; Zhu, X.B.; Wei, B.; Lv, Z.; Huang, Q. Strontium doped lanthanum manganite/manganese dioxide composite electrode for supercapacitor with enhanced rate capability. Electrochim. Acta 2016, 222, 1585–1591. [Google Scholar]
  24. Ma, P.P.; Lu, Q.L.; Lei, N.; Liu, Y.K.; Yu, B.; Dai, J.M.; Li, S.H.; Jiang, G.H. Effect of A-site substitution by Ca or Sr on the structure and electrochemical performance of LaMnO3 perovskite. Electrochim. Acta 2020, 332, 135489. [Google Scholar] [CrossRef]
  25. Lv, J.; Zhang, Y.; Zhe, L.; Huang, X.; Wang, Z.; Bo, W. Strontium doped lanthanum manganite (LSM) effects on electrochemical performance of LSM/MnO2 composites for supercapacitor. J. Mater. Sci. Mater. Electron. 2017, 28, 17020–17025. [Google Scholar] [CrossRef]
  26. Rezanezhad, A.; Rezaie, E.; Ghadimi, L.S.; Hajalilou, A.; Abouzari-Lotf, E.; Arsalani, N. Outstanding supercapacitor performance of Nd-Mn co-doped perovskite LaFeO3@nitrogen-doped graphene oxide nanocomposites. Electrochim. Acta 2020, 335, 135699. [Google Scholar] [CrossRef]
  27. Younis, A.; Loucif, A. Defects mediated enhanced catalytic and humidity sensing performance in ceria nanorods. Ceram. Int. 2021, 47, 15500–15507. [Google Scholar] [CrossRef]
  28. Alexander, C.T.; Mefford, J.T.; Saunders, J.; Forslund, R.P.; Johnston, K.P.; Stevenson, K.J. Anion-based pseudocapacitance of the perovskite library La1−xSrxBO3−δ (B = Fe, Mn, Co). ACS Appl. Mater. Interfaces 2019, 11, 5084–5094. [Google Scholar] [CrossRef]
  29. Lu, J.; Zhang, Y.; Lu, Z.; Huang, X.; Wang, Z.; Zhu, X.; Wei, B. A preliminary study of the pseudo-capacitance features of strontium doped lanthanum manganite. RSC Adv. 2015, 5, 5858. [Google Scholar] [CrossRef]
  30. Cao, Y.; Lin, B.P.; Sun, Y.; Yang, H.; Zhang, X.Q. Synthesis, structure and electrochemical properties of lanthanum manganese nanofibers doped with Sr and Cu. J. Alloys Compd. 2015, 638, 204–213. [Google Scholar] [CrossRef]
  31. Jiang, R.R.; Wang, Y.K.; Cheng, W.; Yu, S.; Zhang, J.M. Hollow ZnSnO3/reduced graphene oxide ternary composite as anode of lithium ion batteries with enhanced electrochemical performance. Ceram. Int. 2017, 43, 11556–11562. [Google Scholar] [CrossRef]
  32. Nagamuthu, S.; Vijayakumar, S.; Ryu, K. Cerium oxide mixed LaMnO3 nanoparticles as the negative electrode for aqueous asymmetric supercapacitor devices. Mater. Chem. Phys. 2017, 199, 543–551. [Google Scholar] [CrossRef]
  33. Sunding, M.F.; Hadidi, K.; Diplas, S.; Lovvik, O.M.; Norby, T.E.; Gunnæs, A.E. XPS characterisation of in situ treated lanthanum oxide and hydroxide using tailored charge referencing and peak fifitting procedures. J. Electron. Spectrosc. Relat. Phenom. 2011, 184, 399–409. [Google Scholar] [CrossRef]
  34. Najjar, H.; Batis, H. Development of Mn-based perovskite materials: Chemical structure and applications. Catal. Rev. 2016, 58, 371–438. [Google Scholar] [CrossRef]
  35. Zhu, Q.; Zhao, D.; Cheng, M.; Zhou, J.; Owusu, K.A.; Mai, L.; Yu, Y. A new view of supercapacitors: Integrated supercapacitors. Adv. Energy Mater. 2019, 33, 1901081. [Google Scholar] [CrossRef]
  36. Yan, D.; Wang, W.; Luo, X.; Chen, C.; Zeng, Y.; Zhu, Z. NiCo2O4 with oxygen vacancies as better performance electrode material for supercapacitor. Chem. Eng. J. 2018, 334, 864–872. [Google Scholar] [CrossRef]
  37. Tian, H.W.; Lang, X.Q.; Nan, H.S.; An, P.; Zhang, W.; Hu, X.Y.; Zhang, J.S. Nanosheet-assembled LaMnO3@NiCo2O4 nanoarchitecture growth on Ni foam for high power density supercapacitors. Electrochim. Acta 2019, 318, 651–659. [Google Scholar] [CrossRef]
  38. Komai, S.; Hirano, M.; Ohtsu, N. Spectral analysis of Sr 3d XPS spectrum in Sr-containing hydroxyapatite. Surf. Interface Anal. 2020, 52, 823–828. [Google Scholar] [CrossRef]
Figure 1. (a) TG curve of the LaMnO3 precursor and (b) XRD pattern of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2).
Figure 1. (a) TG curve of the LaMnO3 precursor and (b) XRD pattern of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2).
Coatings 12 01739 g001
Figure 2. SEM images of (a) LaMnO3, (b) La0.95Sr0.05MnO3, (c) La0.9Sr0.1MnO3, (d) La0.85Sr0.15MnO3, and (e) La0.8Sr0.2MnO3; (f,g) TEM image and selected area electron diffraction images of LaMnO3; (h,i) TEM image and selected area electron diffraction patterns of La0.85Sr0.15MnO3.
Figure 2. SEM images of (a) LaMnO3, (b) La0.95Sr0.05MnO3, (c) La0.9Sr0.1MnO3, (d) La0.85Sr0.15MnO3, and (e) La0.8Sr0.2MnO3; (f,g) TEM image and selected area electron diffraction images of LaMnO3; (h,i) TEM image and selected area electron diffraction patterns of La0.85Sr0.15MnO3.
Coatings 12 01739 g002
Figure 3. (a) N2 adsorption–desorption isotherms of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) nanoparticles, (b) BJH pore size distributions of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) nanoparticles, (c) The specific surface area and average pore diameters of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2).
Figure 3. (a) N2 adsorption–desorption isotherms of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) nanoparticles, (b) BJH pore size distributions of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) nanoparticles, (c) The specific surface area and average pore diameters of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2).
Coatings 12 01739 g003
Figure 4. (a) XPS survey spectrum of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2); high-resolution spectrum of (b) La3d, (c) Mn2p, (d) O1s, and (e) Sr3d.
Figure 4. (a) XPS survey spectrum of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2); high-resolution spectrum of (b) La3d, (c) Mn2p, (d) O1s, and (e) Sr3d.
Coatings 12 01739 g004
Figure 5. CV curves of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2): (a) LaMnO3; (b) La0.95Sr0.05MnO3; (c) La0.9Sr0.1MnO3; (d) La0.85Sr0.15MnO3; (e) La0.8Sr0.2MnO3; (f) CVs recorded at a scan rate of 100 mV/s for all samples in 3 M KOH electrolyte.
Figure 5. CV curves of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2): (a) LaMnO3; (b) La0.95Sr0.05MnO3; (c) La0.9Sr0.1MnO3; (d) La0.85Sr0.15MnO3; (e) La0.8Sr0.2MnO3; (f) CVs recorded at a scan rate of 100 mV/s for all samples in 3 M KOH electrolyte.
Coatings 12 01739 g005
Figure 6. Charge–discharge curves of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2): (a) LaMnO3; (b) La0.95Sr0.05MnO3; (c) La0.9Sr0.1MnO3; (d) La0.85Sr0.15MnO3; (e) La0.8Sr0.2MnO3; and (f) specific current against the specific capacitance of La1−xSrxMnO3.
Figure 6. Charge–discharge curves of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2): (a) LaMnO3; (b) La0.95Sr0.05MnO3; (c) La0.9Sr0.1MnO3; (d) La0.85Sr0.15MnO3; (e) La0.8Sr0.2MnO3; and (f) specific current against the specific capacitance of La1−xSrxMnO3.
Coatings 12 01739 g006
Figure 7. (a) The impedance comparison diagram of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) electrodes; (b) The Rct of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) electrodes.
Figure 7. (a) The impedance comparison diagram of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) electrodes; (b) The Rct of La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) electrodes.
Coatings 12 01739 g007
Table 1. The proportional density of the three oxygen molecules of O 1s in the La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) samples.
Table 1. The proportional density of the three oxygen molecules of O 1s in the La1−xSrxMnO3 (x = 0.0, 0.05, 0.1, 0.15, and 0.2) samples.
LaMnO3La0.95Sr0.05MnO3La0.9Sr0.1MnO3La0.85Sr0.15MnO3La0.8Sr0.2MnO3
O 1(%)44.844.643.139.940.6
O 2(%)47.648.350.153.652.8
O 3(%)7.67.16.86.56.6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ye, X.; Dong, S.; Jin, X.; Wei, J.; Wang, L.; Zhang, Y. Enhancement in the Electrochemical Performance of Strontium (Sr)-Doped LaMnO3 as Supercapacitor Materials. Coatings 2022, 12, 1739. https://doi.org/10.3390/coatings12111739

AMA Style

Ye X, Dong S, Jin X, Wei J, Wang L, Zhang Y. Enhancement in the Electrochemical Performance of Strontium (Sr)-Doped LaMnO3 as Supercapacitor Materials. Coatings. 2022; 12(11):1739. https://doi.org/10.3390/coatings12111739

Chicago/Turabian Style

Ye, Xin, Songtao Dong, Xiaoyun Jin, Junlin Wei, Lei Wang, and Yamei Zhang. 2022. "Enhancement in the Electrochemical Performance of Strontium (Sr)-Doped LaMnO3 as Supercapacitor Materials" Coatings 12, no. 11: 1739. https://doi.org/10.3390/coatings12111739

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