Next Article in Journal
Effect of the Modification of Catalysts on the Catalytic Performance
Next Article in Special Issue
Non-Conventional Synthesis and Repetitive Application of Magnetic Visible Light Photocatalyst Powder Consisting of Bi-Layered C-Doped TiO2 and Ni Particles
Previous Article in Journal
Influence of Calcination Temperature on Photocatalyst Performances of Floral Bi2O3/TiO2 Composite
Previous Article in Special Issue
Synthesis of Ce0.1La0.9MnO3 Perovskite for Degradation of Endocrine-Disrupting Chemicals under Visible Photons
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Evaluation of Structural and Functional Properties of La0.6Sr0.4MnO3 Perovskite Prepared by the Fast Solution Combustion Approach

by
Ramón Cobo Rendón
1,*,
Christopher Salvo
2,
Erwin Sepúlveda
1,
Arunachalam Arulraj
3,4,
Felipe Sanhueza
1,
José Jiménez Rodríguez
5 and
Ramalinga Viswanathan Mangalaraja
1,4,*
1
Advanced Ceramics and Nanotechnology Laboratory, Department of Materials Engineering, Faculty of Engineering, University of Concepcion, Concepcion 4030000, Chile
2
Department of Mechanical Engineering, Faculty of Engineering, Universidad del Bío-Bío, Concepción 4030000, Chile
3
Departamento de Física, Facultad de Ciencias Naturales, Universidad de Atacama, Copiapó 1530000, Chile
4
Faculty of Engineering and Sciences, Universidad Adolfo Ibáñez, Diagonal las Torres 2640, Peñalolén, Santiago 1030000, Chile
5
National Center for Metallurgical Research, Superior Council of Scientific Investigations (CENIM-CSIC), X-ray Laboratory, 28040 Madrid, Spain
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(12), 1636; https://doi.org/10.3390/catal12121636
Submission received: 31 October 2022 / Revised: 2 December 2022 / Accepted: 6 December 2022 / Published: 13 December 2022
(This article belongs to the Special Issue Catalytic and Functional Materials for Environment and Energy)

Abstract

:
A series of La0.6Sr0.4MnO3 (LSM) perovskite was made using the rapid solution combustion method, which was calcined by varying the temperatures. In order to determine how the calcination temperature affected the nanopowders produced and calcined at various temperatures, their microstructural, morphological, compositional, optical, and electrical properties were analyzed using corresponding characterization tools. The XRD results showed the coexistence of the rhombohedral polymorphs R-3c and Pm-3m for the perovskite phase under a calcination temperature of 1400 °C, which were eliminated with increased calcination temperature. The average grain size was found to increase with increasing calcination temperature. The EDS analysis showed better agreement of the stoichiometry with the theoretical composition. The apparent porosity decreased with increasing temperature due to the coalescence of sintering pores. The sample obtained after calcination at 1500 °C showed 10.3% porosity. The hardness also improved with increasing calcination temperature and reached a maximum value of 0.4 GPa, which matched the bulk density. A similar trend was observed in the resistivity studies as a function of temperature, and all the samples exhibited a low resistivity of ~1.4 Ω·cm in the temperature range of 500–600 °C. The optical characterization showed broad absorption at 560–660 nm and bandwidth values between 3.70 and 3.95 eV, according to the applied heat treatment.

1. Introduction

Perovskites are a class of functional ceramic materials that can be broadly characterized by the general formula of ABX3, where X denotes a halide (Cl, Br, or I) or an oxygen ion, which can be further subdivided into perovskite halides and/or oxides. Perovskite halides are widely used in optoelectronic devices and in perovskite solar cells due to their great optical absorption, good charge carrier mobilities, and low optical band gap (1.2–1.4 eV) [1]. A perovskite with oxides is generally termed an ABO3 or A2BO4 structure. A larger cation, A, is situated in the center edge of the structure in a perfect cubic crystalline unit cell of perovskite, whereas a smaller cation, B, is situated in the middle of the octahedron. Both cations are bound by the anion O [2]. A cation, which has 12 oxygen atoms for coordination, can be either rare-earth, alkaline, or alkaline-earth. Any transition metal ion from a 3d, 4d, or 5d configuration can be the B cation, which is encircled by six oxygen atoms in octahedral coordination. While the B cation is in charge of the catalytic activity in the perovskite structure, the A cation plays a crucial function in stabilizing the structure. As indicated above, the ability to modify a perovskite structure’s catalytic characteristics through the partial or complete substitution of A and B cations enables a better fit for the required catalytic applications [3,4]. In many fields, including solid oxide fuel and electrolysis cells (SOFCs and SOECs, respectively) [5], solar cells [1,6], ceramic membranes [7,8], heterogeneous catalysis [3,9], photocatalyst [10,11], water splitting [12], magnetic refrigerators [13], biomaterials, and smart drugs [14], this class of material has shown signs of development. Its unique structure and properties include electronic and ionic conductivity, superconductivity, piezoelectricity, magnetic properties, and catalytic activity [15]. Since Libby [16] and Voorhoeve et al. [17] first studied perovskites in the early 1970s, perovskite-type oxides have gained huge attention as viable catalysts for energy conversion and storage applications. Since then, there has been a huge increase in investigations into perovskite as catalysts or catalytic precursors. Furthermore, due to their ability to hold many oxygen vacancies, they are desirable choices for electrodes in SOFC applications.
The microstructure and composition of electrode materials have a big impact on the performance of SOFCs. The materials that can be employed are constrained by high temperature as well as the presence of oxidizing or reducing atmospheres. Perovskite-oxides can function as catalysts in oxidation and reduction processes, which largely rely on the band structure and the density of states [18]. At the moment, the highest-performing SOFC in the commercial sector is primarily composed of oxygen ion conductors made up of perovskite oxides such as La0.6Sr0.4Co0.2Fe0.8O3 (LSCF) or La0.6Sr0.4Co0.2Fe0.8O3 (LSM) and Y2O3-ZrO2 (YSZ), etc., as a composite cathode [19,20,21,22,23]. These cathode materials exhibit polarization, which is the primary cause of energy loss and low operational stability in SOFCs [24]. Moreover, the high cost of SOFCs is mostly a result of the high production cost of LSM [25]. As a result, one of the major areas of research in SOFCs is synthesizing approaches to these perovskite oxide materials in order to offset their many disadvantages [26], which include cost as well as structural, morphological, electrical, thermal, magnetic, and other issues.
In the current generation of SOFCs, a nominal composition of La1−xSrxMnO3 (x = 0.2) is typically utilized. Massive magneto-resistance materials are another use for compounds from the doped lanthanum manganite group [27,28]. The use of materials falling under this category is influenced by their final microstructure, grain size, pore size, and pore size distribution, in addition to their chemical, structural, and thermodynamic properties [29,30,31]. The pore size distribution is crucial for both the permeability of oxygen gas and the conversion of oxygen into oxygen ions, especially in SOFC devices with tubular designs where LSM serves as both a cathode and a support tube. In order to predict the reaction result when the electrolyte interacts with the La–Mn–Zr–O system, the homogeneity of the LSM must also be investigated [31]. In the traditional ceramic synthesis process, which is based on the diffusion of components in their solid state at high temperatures, these features depend on the processing parameters but are hardly adjustable. Hence, the creation of novel, quick, and affordable synthesis methods for usage on a commercial scale, together with the application of tools to assess the effectiveness of processing, is of utmost importance for enhancing the electrode performance in SOFCs and their catalytic activity in energy applications.
Due to its simplicity, cost-effectiveness, and high-quality end product, solution combustion synthesis, which has been used to prepare a variety of compounds for technological applications, particularly in the development of ReSOFC components with the desired composition and structure, is a good option in this situation [32,33,34,35]. Thus, in this investigation, we sought to obtain a perovskite-oxides-structured La0.6Sr0.4MnO3 material using the fast combustion synthesis method by dissolving metal nitrates in an aqueous solution and then directly combusting in a single step for only 10 min, obtaining a powder foam that was ground in an agate mortar and subsequently calcined at temperatures of 1400, 1450, or 1500 °C for 8 h. The prepared LSM (La0.6Sr0.4MnO3) perovskite materials’ phase formation, shape, crystalline structure, stabilization, electrical, and optical characteristics were thoroughly examined.

2. Results and Discussion

Figure 1 shows the XRD patterns of La0.6Sr0.4MnO3 combusted (LSM500) and calcined at 1400, 1450, and 1500 °C (LSM1400, LSM1450, and LSM1450, respectively). The LSM500 pattern exhibited multiple peaks due to the impurities present in the sample, which were not observed in the calcined samples. Moreover, we noticed an increase in the intensity of the diffraction peaks of the samples calcined at 1450 and 1500 °C and narrowed peaks; this was related to the higher crystallinity [36]. Similarly, the samples LSM1400, LSM1450, and LSM1500 exhibited peaks similar to those in earlier reports [37,38,39] for a single phase, which corresponds to the single-phase perovskite La0.6Sr0.4MnO3 hexagonal R-3c (JCPDS No.: 98-009-9555). Previous experience has shown the existence of polymorphs with identical nominal compositions during the formation of the LSM phase with this synthesis method [40,41]. In order to study the structural properties of the samples in detail, Rietveld refinements were performed using TOPAS software (Bruker AXS) with the double Voigt function [42].
For sample LSM 1400, the XRD data were modeled with the R-3c space group symmetry as the main perovskite phase. The Pm-3m polymorph was included as the second phase in the refinement; this significantly improved the fit (goodness of fit (GoF)). The difference between the experimental data and the fitted simulated pattern is provided as a continuum. Figure S1 of the supplementary material shows the gray line as the best fit suggesting the coexistence of two distinct polymorphs of La0.6Sr0.4MnO3 in the sample LSM1400 presenting the space group, 96.29% of R-3c and 3.71% of Pm-3m phases. This is likely because LaMnO3 perovskites adopt a highly symmetric Pm-3m cubic structure at elevated temperatures [42,43]. Islam et al. explained that these ABO3-type perovskites can deviate from the ideal cubic structure due to distortions in the rigid BO6 octahedra, i.e., by changes in the lengths of the B–O bonds, generating flexibility in the antisymmetric Jahn–Teller-type distortion of the BO6 octahedra [44]. Thus, as the temperature decreases, the MO6 octahedron sharing the corners in this structure can tilt, giving rise to the polymorphic transition to the lower symmetry. These distortions are induced by valence bonding, orbital degenerations, polar distortions, valence fluctuations, etc. Therefore, the heat treatment applied to the samples is a crucial factor in influencing the phase segregation. This was further evidenced in the samples LSM1450 and LSM1500, which exhibited a successful fit with a single uniform R-3c phase. The LSM1500 powder showed an improvement in the GoF from the LSM1450 sample, displaying a variation in lattice parameters with the temperature, in an agreement with the literature [38,41,43,44]. The phase quantification and the microstructural parameters of the Rietveld refinement are shown in Table 1.
Additionally, through this analysis, it was possible to identify the nonexistence of the SrO phase, because it is an undesirable phenomenon highlighted in the literature for perovskites [45,46,47,48], as it can lead to a lower electronic conductivity. Squizzato et al. [48] explained that the transition metal B-site of perovskites plays a fundamental role in the catalytic activity, so it is necessary to preserve the pure LSM structure to optimize its oxygen reduction properties in the ReSOFC. With the absence of the SrO phase in all the materials developed by the fast combustion method, it is evident that this synthesis route, and particularly the LSM1500 material, offers a promising crystalline arrangement to use as a catalytic material in the SOFC.
It is well known that the catalytic properties of a material are closely related to the morphology of its particles [49,50]. The morphological effect plays a fundamental role in the performance of the SOFC constituent materials, and the existence of different morphologies may play an important role in influencing the performance of catalysts in particular perovskites. Because of this, the morphology of the synthesized and calcined powders was observed by scanning electron microscopy (SEM), and the images are shown in Figure 2. In LSM500 (Figure 2A), the lack of ordering of its structure is reflected in the SEM because it does not exhibit grain boundaries, which involves the nonformation of defined grains. However, the effect of calcination brings, as a morphological consequence, evidence of clear grain boundaries, an increase in grain size, and a decrease in porosity, as corroborated in Figure 2B–D. As the calcination temperature increases, the grain size distribution significantly increases to exhibit agglomerates up to 3000 nm in sample LSM1500.
The prepared sample evidenced a more regular grain size distribution than the previously reported material with a different chemical composition. The regularity of the particles would result in improvements in the catalytic activity due to the ease of charge transfer due to the generation of pores with more homogeneous sizes and volumes [48,50]. The average grain sizes agree with the those in the literature for the synthesized perovskite composition [38,39,51]. Additionally, these SEM images revealed that these grains consisted of La, Sr, Mn, and O in stoichiometric percentage by weight (Table 2). The comparison between theoretical and experimental values of the prepared samples analyzed by EDS exhibited a reasonable stoichiometric control in the synthesis process of the materials (Supplementary Material Figure S2). The effect of calcination in the optimization of the desired stoichiometry was corroborated, because the increase in the calcination temperature led to the differences between the theoretical and experimental values, which were considerably reduced [40,41].
Table 3 shows the apparent density and porosity compared with the grain size and microhardness of the samples. The measurement by the Archimedes method ratified this, which is observed in the SEM and EDS, showing that as the temperature increased, the porosity decreased. The grain size became more prominent, the material was denser, and the microhardness increased up to 0.4 GPa for LSM1500.
The resistivity variation as a function of temperature is shown in Figure 3. The measurement was performed using the four collinear probes technique, with a DC voltage ranging from −3 to 3 mV with a step of 2 × 10−6, measuring the current obtained and obtaining the point-to-point resistivity according to Ohm’s law. All the evaluated samples exhibited a rapid decrease in resistivity with an increase in temperature, which is typical of LSM samples [52,53], associated with the increased mobility of charge carriers due to electron hopping from Mn3+ to Mn2+.
The effect of calcination was also visible, as the LSM1500 sample presented the steepest resistivity decay gradient compared with the samples calcined at a lower temperature (1400 and 1500°), with a lower resistivity of 1.1 Ω·cm at a temperature of 550 °C. This is associated with the intimate relationship between the resistivity and grain size for the LSM samples [41,52,53,54,55]. Additionally, there was no difference observed between the samples synthesized using the fast solution combustion approach.
The optical absorption of the samples and the band gap energy (Eg) were evaluated using Wood–Tauc theory [56,57] represented in Figure 4. The optical response of La0.6Sr0.4MnO3 was clearly shown in the wavelength range from 500 to 800 nm, which closely agrees with other investigations [58], with a second absorbance range from 460 nm to 480 nm.
( α h ν ) 2 = A ( h ν E g )
The sample calcined at 1500 °C (LSM1500) showed a better optical response than those calcined at lower temperatures (1400 and 1450 °C), possibly due to its better crystalline and morphological arrangement [41,56,57,59]. The relationship between the band gap energy and the absorption coefficient according to Wood–Tauc theory is formulated by Equation (1), where α is the absorption coefficient; is the photon energy; and A is a constant that depends on the effective electron mass, hole, and refractive index of the material. The calculated band gap values ranged from 3.71 to 3.95 eV, which are close to the existing reported values for these perovskite materials [57,60,61,62]. However, the change in the band-gap energy is attributed to the difference in the Mn4+ to Mn3+ ratio as a function of the La/Sr ratio and the calcination process. Because Sr3+ substituted La2+, the thermal treatment led to a product out of stoichiometry; hence, Mn3+ must be oxidized to Mn4+ to maintain charge neutrality. It is important to point out that these band gap values deepen the versatility of LSM material synthesized this way. In addition to being a good catalyst for the SOFC or ReSOFC, it could be a suitable candidate for photocatalysis processes or for developing high-frequency optoelectronic devices, because the band gap values of the materials are obtained above 2 eV and at absorption values from 500 to 700 nm [10,11,57,63,64].

3. Materials and Methods

The development of the synthesis of La0.6Sr0.4MnO3 by the fast solution combustion method started with the use of citric acid (C6H8O7, 99.99%, Sigma-Aldrich, San Luis, MO, USA) as a fuel in the desired stoichiometric ratio of La(NO3)3·6H2O (99.9%, Merck, Darmstadt, Germany), Sr(NO3)2 (99.8%, Sigma-Aldrich, San Luis, MO, USA), and Mn(NO3)2·4H2O (97%, Sigma-Aldrich, San Luis, MO, USA) as precursors (Table 4). These materials were mixed in 50 mL of water and stirred at room temperature for 1 h. Then, the solution was combusted at 500 °C for 10 min in a Nabertherm furnace, LT 40/12 (Lilienthal, Germany), generating rapid combustion. Finally, the foamed powder was ground in an agate mortar and calcined at different temperatures of 1400, 1450, or 1500 °C for 8 h.
Additionally, 11.5 mm diameter and 1 mm thick pellets were made from the combusted powder by compacting them at 4500 Kg for 1 min. Finally, the pellets were sintered at the same temperatures of 1400, 1450, or 1500 °C. The coding of the samples is provided in Table 5.
The crystal structure and phase of the elements present in the calcined La0.6Sr0.4MnO3 powders were obtained with a X-ray diffractometer (Bruker AXS, D4 Endeavor, Bremen, DE) with 40 kV, 20 mA, and 0.1542 nm Cu-Kα with an angular range of 2𝜃 from 20° to 70° in 0.02° and 1 s steps. In addition, a Rietveld analysis of XRD patterns was carried out to obtain the microstructural information using TOPAS software (Bruker AXS) version 4.2. The morphology, structure, and composition of the material were observed using a scanning electron microscope (SEM) (Tescan, Vega 3 Easyprobe SBU, Brno, Czech Republic) equipped with an energy-dispersive X-ray detector (EDS). Additionally, the bulk density and porosity were measured using the Archimedes method. The optical property of the materials was measured with a UV–Vis spectrophotometer (Shimadzu UV-2600, Kioto Japan). The resistivity as a function of temperature (50–700 °C) was examined using the conventional four-collinear probe method with a Keithley 4200-SCS (Cleveland, OH, USA). In addition, the room-temperature mechanical microhardness of the materials was evaluated using a Zwick/Roell 8187.5 ZHU (Santiago, Chile) microhardness tester.

4. Conclusions

The synthesis of rapid solution combustion was proven as an alternative way to obtain the perovskite compound La0.6Sr0.4MnO3 through thermal treatment at 1400, 1450, or 1500 °C for 8 h. The results obtained by the XRD indicated that all the samples presented the hexagonal primary crystalline structure (R-3c). The single-phase sample with the best crystalline arrangement was found in LSM1500, with crystallite size >150 nm. The sample LSM1400, analyzed through Rietveld’s analysis of the XRD data, showed the coexistence of R-3c and Pm-3m phases, corroborating the description provided for other compositions with the same synthesis method. The SEM images showed the influence of the heat treatment applied to the samples because there was a close relationship between calcination temperature, grain size, and agglomerate formation. The average grain size for sample LSM1500 was 1010 nm, and agglomerate formation was up to 3 µm. As expected, the porosity decreased with increased calcination temperature, and the hardness reached a maximum value of 0.39 GPa. The EDS analysis showed a close relationship between the heat treatment applied and the stoichiometry optimization. The dependence of resistivity on temperature for the LSM samples showed a decrease in resistivity with an increase in temperature, reaching a resistivity of approximately 1.1 Ω·cm for the sample calcined at 1500 °C, which may be a suitable electrode material for ReSOFCs. Furthermore, the optical response of the fabricated samples occurs in the wavelength range from 500 to 800 nm, and the calculated band gap values ranged from 3.71 to 3.95 eV, which proved that the synthesized perovskite compound may be a suitable candidate for the development of high-frequency optoelectronic devices.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12121636/s1, Figure S1: Rietveld refinement of XRD patterns of La0.6Sr0.4MnO3 by considering (a) only the rhombohedral R-3c and (b) also the cubic Pm-3m polymorphs of the La0.6Sr0.4MnO3 perovskite. The difference between experimental data and the fitted simulated pattern is plotted as a continuous gray line at the bottom. (Blue open circles: Experimental data; Green solid line: Cubic phase; Orange solid line: Rhombohedral phase; Red solid line: Fitted data); Figure S2: EDS Mapping of the combusted and calcined LSM powders (A) LSM500, (B) LSM1400, (C) LSM1450, (D) LSM1500.

Author Contributions

Conceptualization, R.V.M. and R.C.R.; methodology, R.C.R.; software, R.C.R., J.J.R., E.S. and C.S.; formal analysis, R.C.R., J.J.R. and C.S.; investigation, R.C.R.; resources, R.V.M., F.S. and J.J.R.; writing—original draft preparation, R.C.R., R.V.M., C.S. and A.A.; writing—review and editing, R.C.R., R.V.M., C.S. and A.A.; visualization, R.C.R. and J.J.R.; supervision, R.V.M.; project administration, R.V.M.; funding acquisition, R.V.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by FONDECYT-ANID (project No. 1181703) for the financial support, University of Concepcion, Concepcion, Chile.

Data Availability Statement

Data is available on request from the corresponding authors. The data are not publicly available due to the possible extension of the carried work.

Acknowledgments

The author Ramón Cobo-Rendón is grateful to ANID-Chile Grant Nº: 21210463, University of Concepcion, Concepcion, Chile.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pitaro, M.; Tekelenburg, E.K.; Shao, S.; Loi, M.A. Tin halide perovskites: From fundamental properties to solar cells. Adv. Mater. 2022, 34, 2105844. [Google Scholar] [CrossRef] [PubMed]
  2. Zhu, H.; Zhang, P.; Dai, S. Recent Advances of Lanthanum-Based Perovskite Oxides for Catalysis. ACS Catal. 2015, 5, 6370–6385. [Google Scholar] [CrossRef]
  3. Zhu, J.; Li, H.; Zhong, L.; Xiao, P.; Xu, X.; Yang, X.; Zhao, Z.; Li, J. Perovskite Oxides: Preparation, Characterizations, and Applications in Heterogeneous Catalysis. ACS Catal. 2014, 4, 2917–2940. [Google Scholar] [CrossRef]
  4. Royer, S.; Duprez, D.; Can, F.; Courtois, X.; Batiot-Dupeyrat, C.; Laassiri, S.; Alamdari, H. Perovskites as Substitutes of Noble Metals for Heterogeneous Catalysis: Dream or Reality. Chem. Rev. 2014, 114, 10292–10368. [Google Scholar] [CrossRef]
  5. Hussain, S.; Yangping, L. Review of solid oxide fuel cell materials: Cathode, anode, and electrolyte. Energy Transit. 2020, 4, 113–126. [Google Scholar] [CrossRef]
  6. Su, C.; Wang, W.; Shao, Z. Cation-Deficient Perovskites for Clean Energy Conversion. Accounts Mater. Res. 2021, 2, 477–488. [Google Scholar] [CrossRef]
  7. Athayde, D.D.; Souza, D.F.; Silva, A.M.; Vasconcelos, D.; Nunes, E.H.; da Costa, J.C.; Vasconcelos, W.L. Review of perovskite ceramic synthesis and membrane preparation methods. Ceram. Int. 2016, 42, 6555–6571. [Google Scholar] [CrossRef] [Green Version]
  8. Wang, Z.; Li, Z.; Cui, Y.; Chen, T.; Hu, J.; Kawi, S. Highly efficient NO decomposition via dual-functional catalytic perovskite hollow fiber membrane reactor coupled with partial oxidation of methane at medium-low temperature. Environ. Sci. Technol. 2019, 53, 9937–9946. [Google Scholar] [CrossRef]
  9. Hwang, J.; Rao, R.R.; Giordano, L.; Katayama, Y.; Yu, Y.; Shao-Horn, Y. Perovskites in catalysis and electrocatalysis. Science 2017, 358, 751–756. [Google Scholar] [CrossRef] [Green Version]
  10. Xu, Y.F.; Yang, M.Z.; Chen, B.X.; Wang, X.D.; Chen, H.Y.; Kuang, D.B.; Su, C.Y. A CsPbBr3 Perovskite Quantum Dot/Graphene Oxide Composite for Photocatalytic CO2 Reduction. J. Am. Chem. Soc. 2017, 139, 5660–5663. [Google Scholar] [CrossRef]
  11. Zhu, Y.; Liu, Y.; Miller, K.A.; Zhu, H.; Egap, E. Lead Halide Perovskite Nanocrystals as Photocatalysts for PET-RAFT Polymerization under Visible and Near-Infrared Irradiation. ACS Macro Lett. 2020, 9, 725–730. [Google Scholar] [CrossRef] [PubMed]
  12. Sun, H.; Dai, J.; Zhou, W.; Shao, Z. Emerging Strategies for Developing High-Performance Perovskite-Based Materials for Electrochemical Water Splitting. Energy Fuels 2020, 34, 10547–10567. [Google Scholar] [CrossRef]
  13. Yu, B.F.; Gao, Q.; Zhang, B.; Meng, X.Z.; Chen, Z. Review on research of room temperature magnetic refrigeration. Int. J. Refrig. 2003, 26, 622–636. [Google Scholar] [CrossRef]
  14. Biswas, S.; Keshri, S.; Goswami, S.; Isaac, J.; Ganguly, S.; Perov, N. Antibiotic loading and release studies of LSMO nanoparticles embedded in an acrylic polymer. Phase Transit. 2016, 89, 1203–1212. [Google Scholar] [CrossRef]
  15. Žužić, A.; Ressler, A.; Macan, J. Perovskite oxides as active materials in novel alternatives to well-known technologies: A review. Ceram. Int. 2022, 48, 27240–27261. [Google Scholar] [CrossRef]
  16. Libby, W.F. Promising Catalyst for Auto Exhaust. Science 1971, 171, 499–500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Voorhoeve, R.J.H.; Remeika, J.P.; Trimble, L.E. Defect Chemistry and Catalysis in Oxidation and Reduction over Perovskite-Type Oxides. Ann. N. Y. Acad. Sci. 2006, 272, 3–21. [Google Scholar] [CrossRef]
  18. Hanif, M.B.; Rauf, S.; Motol, M.; Babar, Z.; Li, C.J.; Li, C.X. Recent progress of perovskite-based electrolyte materials for solid oxide fuel cells and performance optimizing strategies for energy storage applications. Mat. Res. Bull. 2022, 146, 111612. [Google Scholar] [CrossRef]
  19. Yang, G.; Su, C.; Shi, H.; Zhu, Y.; Song, Y.; Zhou, W.; Shao, Z. Towards reducing operation temperature of solid oxide fuel cells: Our past fifteen years of efforts in cathode development. Energ. Fuel. 2020, 34, 15169–15194. [Google Scholar] [CrossRef]
  20. Park, B.-K.; Barnett, S.A. Boosting solid oxide fuel cell performance via electrolyte thickness reduction and cathode infiltration. J. Mater. Chem. A. 2020, 8, 11626–11631. [Google Scholar] [CrossRef]
  21. Wang, W.; Mogensen, M. High-performance lanthanum-ferrite based cathode for SOFC. Solid State Ion. 2005, 176, 457–462. [Google Scholar] [CrossRef]
  22. Liu, J.; Zhou, M.; Zhang, Y.; Liu, P.; Liu, Z.; Xie, Y.; Cai, W.; Yu, F.; Zhou, Q.; Wang, X.; et al. Electrochemical Oxidation of Carbon at High Temperature: Principles and Applications. Energy Fuels 2017, 32, 4107–4117. [Google Scholar] [CrossRef]
  23. Wang, G.; Zhang, Y.; Han, M. Densification of Ce0.9Gd0.1O2-δ interlayer to improve the stability of La0.6Sr0.4Co0.2Fe0.8O3-δ/Ce0.9Gd0.1O2-δ interface and SOFC. J. Electroanal. Chem. 2020, 857, 113591–113598. [Google Scholar] [CrossRef]
  24. Sun, C.; Hui, R.; Roller, J. Cathode materials for solid oxide fuel cells: A review. J. Solid State Electrochem. 2010, 14, 1125–1144. [Google Scholar] [CrossRef]
  25. Saha, S.; Ghanawat, S.J.; Purohit, R.D. Solution combustion synthesis of nano particle La0.9Sr0.1MnO3 powder by a unique oxidant-fuel combination and its characterization. J. Mater. Sci. 2006, 41, 1939–1943. [Google Scholar] [CrossRef]
  26. Tsvetkov, N.; Lu, Q.; Sun, L.; Crumlin, E.J.; Yildiz, B. Improved chemical and electrochemical stability of perovskite oxides with less reducible cations at the surface. Nat. Mater. 2016, 15, 1010–1016. [Google Scholar] [CrossRef] [Green Version]
  27. Jin, S.; Tiefel, T.H.; McCormack, M.; Fastneach, R.A.; Ramesh, R.; Clien, L.H. Thousandfold change in resistivity in magnetoresistive La-Ca-Mn-O films. Science 1994, 264, 413. [Google Scholar] [CrossRef] [Green Version]
  28. Mahesh, R.; Mahendiran, R.; Raychaudhuri, A.K.; Rao, C.N.R. Effect of particle size on the giant magnetoresistance of La0.7Ca0.3MnO3. Appl. Phys. Lett. 1996, 68, 2291. [Google Scholar] [CrossRef]
  29. Huang, Y.H.; Xu, Z.G.; Yan, C.H.; Wang, Z.M.; Zhu, T.; Liao, C.S.; Gao, S.; Xu, G.X. Soft chemical synthesis and transport properties of La0.7Sr0.3MnO3 granular perovskites. Solid State Commun. 2000, 114, 43. [Google Scholar] [CrossRef]
  30. Zhang, N.; Ding, W.; Zhong, W.; Xing, D.; Du, Y. Tunnel-type giant magnetoresistance in the granular perovskite La0.85Sr0.15MnO3. Phys. Rev. B 1997, B56, 8138. [Google Scholar] [CrossRef]
  31. Yokokawa, H.; Sakai, N.; Kawada, T.; Dokiwa, M. Thermodynamic analysis on interface between perovskite electrode and YSZ electrolyte. Solid State Ion. 1990, 40, 398. [Google Scholar] [CrossRef]
  32. Mangalaraja, R.; Mouzon, J.; Hedström, P.; Camurri, C.P.; Ananthakumar, S.; Odén, M. Microwave assisted combustion synthesis of nanocrystalline yttria and its powder characteristics. Powder Technol. 2009, 191, 309–314. [Google Scholar] [CrossRef]
  33. Marinšek, M.; Zupan, K.; Maèek, J. Ni–YSZ cermet anodes prepared by citrate/nitrate combustion synthesis. J. Power Sources 2002, 106, 178–188. [Google Scholar] [CrossRef]
  34. Wang, W.; Huang, Y.; Jung, S.; Vohs, J.M.; Gorte, R.J. A Comparison of LSM, LSF, and LSCo for Solid Oxide Electrolyzer Anodes. J. Electrochem. Soc. 2006, 153, A2066–A2070. [Google Scholar] [CrossRef] [Green Version]
  35. Gan, L.M.; Chan, H.S.; Zhang, L.H.; Chew, C.H.; Loo, B.H. Preparation of fine LaNiO3 powder from oxalate precursors via reactions in inverse micro emulsions. Mater. Chem. Phys. 1994, 37, 263–268. [Google Scholar] [CrossRef]
  36. Di Florio, G.; Macchi, E.G.; Mongibello, L.; Baratto, M.C.; Basosi, R.; Busi, E.; Caliano, M.; Cigolotti, V.; Testi, M.; Trini, M. Comparative life cycle assessment of two different SOFC-based cogeneration systems with thermal energy storage integrated into a single family house nanogrid. Appl Energ. 2021, 285–304, 116378–116397. [Google Scholar] [CrossRef]
  37. Neagu, D.; Irvine, J.T. Structure and properties of La0.4Sr0.4TiO3 ceramics for use as anode materials in solid oxide fuel cells. Chem. Mater. 2010, 22, 5042–5053. [Google Scholar] [CrossRef]
  38. Raoufi, T.; Ehsani, M.; Khoshnoud, D.S. Critical behavior near the paramagnetic to ferromagnetic phase transition temperature in La0.6Sr0.4MnO3 ceramic: A comparison between sol-gel and solid state process. Ceram. Int. 2017, 43, 5204–5215. [Google Scholar] [CrossRef]
  39. Sanna, C.; Squizzato, E.; Costamagna, P.; Holtappels, P.; Glisenti, A. Electrochemical study of symmetrical intermediate temperature-solid oxide fuel cells based on La0.6Sr0.4MnO3/Ce0.9Gd0.1O1.95 for operation in direct methane/air. Electrochim. Acta 2022, 409, 139939. [Google Scholar] [CrossRef]
  40. Durango-Petro, J.; Salvo, C.; Usuba, J.; Abarzua, G.; Sanhueza, F.; Mangalaraja, R.V. Fast Solution Synthesis of NiO-Gd0.1Ce0.9O1.95 Nanocomposite via Different Approach: Influence of Processing Parameters and Characterizations. Materials 2021, 14, 3437. [Google Scholar] [CrossRef]
  41. Rendón, R.C.; Udayabhaskar, R.; Salvo, C.; Sepúlveda, E.; Rodríguez, J.J.; Camurri, C.P.; Viswanathan, M.R. Evaluation of La0.8Sr0.2MnO3 perovskite prepared by fast solution combustion. Ceram. Int. 2022, 48, 35100–35107. [Google Scholar] [CrossRef]
  42. Balzar, D.; Ledbetter, H. Voigt-function modeling in Fourier analysis of size- and strain-broadened X-ray diffraction peaks. J. Appl. Crystallogr. 1993, 26, 97–103. [Google Scholar] [CrossRef]
  43. Coey, J.M.D.; Viret, M.; Von Molnár, S. Mixed-valence manganites. Adv. Phys. 1999, 48, 167–293. [Google Scholar] [CrossRef]
  44. Islam, M.A.; Rondinelli, J.M.; Spanier, J.E. Normal mode determination of perovskite crystal structures with octahedral rotations: Theory and applications. J. Phys. Condens. Matter 2013, 25, 175902. [Google Scholar] [CrossRef] [Green Version]
  45. Wu, Q.-H.; Liu, M.; Jaegermann, W. X-ray photoelectron spectroscopy of La0.5Sr0.5MnO3. Mater. Lett. 2005, 59, 1980–1983. [Google Scholar] [CrossRef]
  46. Caillol, N.; Pijolat, M.; Siebert, E. Investigation of chemisorbed oxygen, surface segregation and effect of post-treatments on La0.8Sr0.2MnO3 powder and screen-printed layers for solid oxide fuel cell cathodes. Appl. Surf. Sci. 2007, 253, 4641–4648. [Google Scholar] [CrossRef]
  47. Jiang, S. The electrochemical performance of LSM/zirconia–yttria interface as a function of a-site non-stoichiometry and cathodic current treatment. Solid State Ionics 1999, 121, 1–10. [Google Scholar] [CrossRef]
  48. Squizzato, E.; Sanna, C.; Glisenti, A.; Costamagna, P. Structural and catalytic characterization of La0.6Sr0.4MnO3 nanofibers for application in direct methane intermediate temperature solid oxide fuel cell anodes. Energies 2021, 14, 3602. [Google Scholar] [CrossRef]
  49. Xie, X.; Li, Y.; Liu, Z.-Q.; Haruta, M.; Shen, W. Low-temperature oxidation of CO catalysed by Co3O4 nanorods. Nature 2009, 458, 746–749. [Google Scholar] [CrossRef]
  50. Arandiyan, H.; Dai, H.; Deng, J.; Liu, Y.; Bai, B.; Wang, Y.; Li, X.; Xie, S.; Li, J. Three-dimensionally ordered macroporous La0.6Sr0.4MnO3 with high surface areas: Active catalysts for the combustion of methane. J. Catal. 2013, 307, 327–339. [Google Scholar] [CrossRef]
  51. Huang, F.; Sun, X.; Zheng, Y.; Xiao, Y.; Zheng, Y. Facile co precipitation synthesis of La0.6Sr0.4MnO3 perovskites with high surface area. Mater. Lett. 2018, 210, 287–290. [Google Scholar] [CrossRef]
  52. Saleem, M.; Varshney, D. Structural, thermal, and transport properties of La0.67Sr0.33MnO3 nanoparticles synthesized via the sol–gel auto-combustion technique. RSC Adv. 2018, 8, 1600–1609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Shinde, K.P.; Thorat, N.D.; Pawar, S.S.; Pawar, S.H. Combustion synthesis and characterization of perovskite La0.9Sr0.1MnO3. Mater. Chem. Phys. 2012, 134, 881–885. [Google Scholar] [CrossRef]
  54. Varshney, D.; Dodiya, N. Electrical resistivity of the hole doped La0.8Sr0.2MnO3 manganites: Role of electron–electron/phonon/magnon interactions. Mater. Chem. Phys. 2011, 129, 896–904. [Google Scholar] [CrossRef]
  55. Zhu, N.; Liu, Y.J. Prediction of the Magneto-Resistance of La0.65Ca0.35MnO3 and La0.8Sr0.2MnO3 via Temperature and Magnetic Field. In Advanced Materials Research; Trans Tech Publications Ltd.: Bäch, Switzerland, 2013. [Google Scholar]
  56. Rashad, M.M.; Turky, A.O.; Kandil, A.T. Optical and electrical properties of Ba1−xSrxTiO3 nanopowders at different Sr2+ ion content. J. Mater. Sci. Mater. Electron. 2013, 24, 3284–3291. [Google Scholar] [CrossRef]
  57. Turky, A.O.; Rashad, M.M.; Zaki, Z.I.; Ibrahim, I.A.; Bechelany, M. Tuning the optical and dielectric properties of calcium copper titanate CaxCu3−xTi4O12 nanopowders. RSC Adv. 2015, 5, 18767–18772. [Google Scholar] [CrossRef]
  58. Cui, K.; Cheng, Y.; Dai, J.; Liu, J. Synthesis, characterization and microwave absorption properties of La0.6Sr0.4MnO3/polyaniline composite. Mater. Chem. Phys. 2013, 138, 810–816. [Google Scholar] [CrossRef]
  59. Afje, F.R.; Ehsani, M. Size-dependent photocatalytic activity of La0.8Sr0.2MnO3 nanoparticles prepared by hydrothermal synthesis. Mater. Res. Express 2018, 5, 045012. [Google Scholar] [CrossRef]
  60. Cesaria, M.; Caricato, A.P.; Leggieri, G.; Martino, M.; Maruccio, G. Optical response of oxygen deficient La0.7Sr0.3MnO3 thin films deposited by pulsed laser deposition. Thin Solid Film. 2013, 545, 592–600. [Google Scholar] [CrossRef]
  61. Busse, P.; Yin, Z.; Mierwaldt, D.; Scholz, J.; Kressdorf, B.; Glaser, L.; Miedema, P.S.; Rothkirch, A.; Viefhaus, J.; Jooss, C.; et al. Probing the surface of La0.6Sr0.4MnO3 in water vapor by in situ photon-in/photon-out spectroscopy. J. Phys. Chem. C 2020, 124, 7893–7902. [Google Scholar] [CrossRef]
  62. Liu, X.L.; Machida, A.M.; Moritomo, Y.M.; Ichida, M.I.; Nakamura, A.N. Room-temperature photo switching in La0.6Sr0.4MnO3 film. Jpn. J. Appl. Phys. 2000, 39, L670. [Google Scholar] [CrossRef]
  63. de Jong, M.P.; Dediu, V.A.; Taliani, C.; Salaneck, W.R. Electronic structure of La0.7Sr0.3MnO3 thin films for hybrid organic/inorganic spintronics applications. J. Appl. Phys. 2003, 94, 7292–7296. [Google Scholar] [CrossRef]
  64. Takenaka, K.; Sawaki, Y.; Shiozaki, R.; Sugai, S. Electronic structure of the double-exchange ferromagnet La0.825Sr0.175MnO3 studied by optical reflectivity. Phys. Rev. B 2000, 62, 13864. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of La0.6Sr0.4MnO3 combusted and calcined at 1400, 1450, and 1500 °C for 8 h.
Figure 1. XRD patterns of La0.6Sr0.4MnO3 combusted and calcined at 1400, 1450, and 1500 °C for 8 h.
Catalysts 12 01636 g001
Figure 2. SEM images and corresponding grain size distribution of the combusted and calcined LSM powders: (A) LSM500, (B) LSM1400, (C) LSM1450, and (D) LSM1500.
Figure 2. SEM images and corresponding grain size distribution of the combusted and calcined LSM powders: (A) LSM500, (B) LSM1400, (C) LSM1450, and (D) LSM1500.
Catalysts 12 01636 g002
Figure 3. Variation in resistivity with the temperature of the calcined La0.6Sr0.4MnO3 samples.
Figure 3. Variation in resistivity with the temperature of the calcined La0.6Sr0.4MnO3 samples.
Catalysts 12 01636 g003
Figure 4. Optical absorbance and band gap spectra of the La0.6Sr0.4MnO3 samples calcined at different temperatures (1400, 1450, or 1500 °C).
Figure 4. Optical absorbance and band gap spectra of the La0.6Sr0.4MnO3 samples calcined at different temperatures (1400, 1450, or 1500 °C).
Catalysts 12 01636 g004
Table 1. Quantitative analysis of phases and microstructural parameters obtained using Rietveld refinement of LSM samples.
Table 1. Quantitative analysis of phases and microstructural parameters obtained using Rietveld refinement of LSM samples.
SamplesPhasesWeight %Lattice
Parameters (Ǻ)
O (18e)-
Position
(x)
Crystallite
Size (nm)
R-Factors
LSM1400La0.6Sr0.4MnO3
S.G: R-3c
96.29%a = b = 5.487 (2)
c = 13.352 (3)
0.4516 (1)>150Rexp = 5.51
Rwp =9.13
GoF = 1.66
S.G: Pm-3m3.71%a = 3.873 (1) >150
LSM1450La0.6Sr0.4MnO3
S.G: R-3c
100%a = b = 5.486 (2)
c = 13.354 (3)
0.4665 (1)>150Rexp = 2.61
Rwp =4.41
GoF = 1.69
LSM1500La0.6Sr0.4MnO3
S.G: R-3c
100%a = b = 5.484 (2)
c = 13.346 (3)
0.4686 (1)>150Rexp = 2.90
Rwp = 4.39
GoF = 1.51
Table 2. Stoichiometric percentage by weight of La0.6Sr0.4MnO3 powder samples calcined at different temperatures.
Table 2. Stoichiometric percentage by weight of La0.6Sr0.4MnO3 powder samples calcined at different temperatures.
SamplesAtomLaSrMnO
Stoichiometric0.60.413
LSM1400Wt. % (Expe.)
Wt. % (Theo.)
Difference (%)
38.59
37.66
2.47
15.04
15.84
5.05
24.86
24.82
0.16
21.51
21.68
0.78
LSM1450Wt. % (Expe.)
Wt. % (Theo.)
Difference (%)
35.45
37.66
5.87
15.05
15.84
4.99
24.17
24.82
2.62
25.33
21.68
16.84
LSM1500Wt. % (Expe.)
Wt. % (Theo.)
Difference (%)
38.33
37.66
1.78
15.44
15.84
2.53
24.34
24.82
1.93
21.89
21.68
0.97
Table 3. Mechanical and physical properties of La0.6Sr0.4MnO3 samples obtained at different calcination temperatures.
Table 3. Mechanical and physical properties of La0.6Sr0.4MnO3 samples obtained at different calcination temperatures.
SampleLSM1400LSM1450LSM1500
Hardness (GPa)0.2 ± 0.10.3 ± 0.10.4 ± 0.1
Bulk density (g/cm3)6.204.584.75
Apparent porosity (%)3.921.910.3
Average grain size (nm)5286961010
Table 4. Preparation parameters of La0.6Sr0.4MnO3.
Table 4. Preparation parameters of La0.6Sr0.4MnO3.
PrecursorsGrams (g)
La(NO3)3.6H2O2.992
Sr(NO3)20.366
Mn(NO3)2.4H2O2.168
C6H8O72.212
Table 5. La0.6Sr0.4MnO3 sample nomenclature and parameters.
Table 5. La0.6Sr0.4MnO3 sample nomenclature and parameters.
Temperature (°C)Time (min/h)Code Name
5005 minLSM500
14008 hLSM1400
14508 hLSM1450
15008 hLSM1500
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rendón, R.C.; Salvo, C.; Sepúlveda, E.; Arulraj, A.; Sanhueza, F.; Rodríguez, J.J.; Mangalaraja, R.V. Evaluation of Structural and Functional Properties of La0.6Sr0.4MnO3 Perovskite Prepared by the Fast Solution Combustion Approach. Catalysts 2022, 12, 1636. https://doi.org/10.3390/catal12121636

AMA Style

Rendón RC, Salvo C, Sepúlveda E, Arulraj A, Sanhueza F, Rodríguez JJ, Mangalaraja RV. Evaluation of Structural and Functional Properties of La0.6Sr0.4MnO3 Perovskite Prepared by the Fast Solution Combustion Approach. Catalysts. 2022; 12(12):1636. https://doi.org/10.3390/catal12121636

Chicago/Turabian Style

Rendón, Ramón Cobo, Christopher Salvo, Erwin Sepúlveda, Arunachalam Arulraj, Felipe Sanhueza, José Jiménez Rodríguez, and Ramalinga Viswanathan Mangalaraja. 2022. "Evaluation of Structural and Functional Properties of La0.6Sr0.4MnO3 Perovskite Prepared by the Fast Solution Combustion Approach" Catalysts 12, no. 12: 1636. https://doi.org/10.3390/catal12121636

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