Next Article in Journal
Use of Unmanned Aerial Vehicles (UAVs) and Photogrammetry to Obtain the International Roughness Index (IRI) on Roads
Previous Article in Journal
New Ternary Compounds of the Composition Cu2SnTi3 and Their Crystal Structures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Impact of Alkali Ions Codoping on Magnetic Properties of La0.9A0.1Mn0.9Co0.1O3 (A: Li, K, Na) Powders and Ceramics

1
Nanoceramics Inc., Okolna 2, PL-50422 Wroclaw, Poland
2
Institute of Low Temperature and Structural Research PAS, Okolna 2, PL-50422 Wroclaw, Poland
3
Helmholtz-Zentrum Geesthacht, Zentrum für Material- und Küstenforschung GmbH, Max-Planck-Straße 1, DE-21502 Geesthacht, Germany
4
Department of Inorganic Chemistry, Vilnius University, Universiteto g. 3, LT-01513 Vilnius, Lithuania
5
School of Natural Sciences and Mathematics, Ural Federal University, Kuybysheva 48, RU-620026 Ekaterinburg, Russia
6
CICECO-Materials Institute of Aveiro and Physics Dept, University of Aveiro, PT-3810-193 Aveiro, Portugal
7
Scientific and Practical Materials Research Center NAS Belarus, P. Brovki 19, BY-220072 Minsk, Belarus
8
Nanotechnology Research & Education Centre, South Ural State University, Lenin av., 76, RU-454080, Chelyabinsk, Russia
*
Author to whom correspondence should be addressed.
Submission received: 22 October 2020 / Revised: 30 November 2020 / Accepted: 4 December 2020 / Published: 8 December 2020
(This article belongs to the Special Issue Physical Properties of the Nanoscale Materials)

Abstract

:
The aim of the work was to check how the introduction of alkali and cobalt ions into a manganese structure can affect the structural disorder and, in consequence, lead to the changes (improvements) of magnetic properties. The high-pressure sintering technique was applied to check if the external factor can modify the magnetization of manganites. Nanocrystalline La0.9A0.1Mn0.9Co0.1O3 (where A is Li, K, Na) powders were synthesized by the combustion technique. The respective powders were used for nanoceramics preparation by the high-pressure sintering technique. The structure and morphology of the compounds were studied by X-ray powder diffraction, scanning electron microscopy and energy-dispersive X-ray spectroscopy. Magnetization studies for all compounds were performed in order to check the changes induced by either codoping or the sintering pressure. It was found that the type of the dopant ion and sintering pressure produced significant changes to the magnetic properties of the studied compounds. Alkali ions lead to the stabilization of Co ions in the +2 oxidation state and the formation of positive exchange interactions Mn3+–Mn4+ and Co2+–Mn4+ and the subsequent increase in remanent magnetization. High sintering pressure leads to a decrease in grain size and reduction of long-range ferromagnetic order and lower magnetization.

1. Introduction

Multiferroics are compounds which in one phase exhibit more than one primary ferroic order (ferromagnetism, ferroelectricity or ferroelasticity) [1]; however, in recent years this term has been mainly used for magnetoelectric materials. Such materials have both magnetic and ferroelectric properties simultaneously. To obtain magnetic materials where magnetic moments are not cancelled out by each other, it is necessary to exclude all ions where outer orbitals are completely filled. This limits the potential multiferroic materials to systems containing either transition metals with partially filled 3d shells or lanthanides with partially filled 4f shells. In the case of ferroelectricity, the Jahn–Teller effect has to be taken into account [2], where oppositely charged ions form a local dipole with a spatially degenerate electronic ground state. For the purpose of removing this degeneracy in the structure, a geometrical distortion appears to lead to the lowering of the overall energy of the system. The mentioned effect and the character of the chemical bonds favor ferroelectricity in such systems. The research conducted in the last two decades showed that the nature of colossal magnetoresistance effect (CMR) [3,4,5,6,7,8,9] in lanthanum manganates cannot be explained without invoking the mechanism of the Jahn–Teller distortion of MnO6 octahedra. The effect of distortions on magnetic interactions and electron transfer can be induced by structural stresses which are lessened by MnO6 array oxidation and by the doping the structure by ions with different ionic radii [10,11,12].
The combination and control of simultaneous ferroic orderings is still a challenge for scientists all around the world. The studies are conducted mainly towards the development of new materials in which, by changing their chemical composition, morphology or form, it will be possible to intentionally introduce the desired properties. The possibility of designing unique magnetic properties and coexistence of Mn3+ and Mn4+ ions allows for the widening of the range of applications for lanthanum manganates and lanthanum manganates doped with transition metal ions in Li-O2 batteries [6], solid oxide fuel cells [13,14,15], alkaline fuel cells [16], magnetic refrigeration [17], and catalysts in oxidation reaction [8,18,19].
Cobalt doped lanthanum manganite attracts a great deal of attention as catalytic material. The bifunctional catalytic activity of oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) can be tailored by the substitution of manganese ions by cobalt ions and control the Mn3+/Mn4+ ratio [8,9]. The chemical doping of manganese ions by cobalt ions as well as diamagnetic A-site substitution with alkali elements can be used as an effective tool to control the magnetic properties of the LaMnO3-based compounds [7,20]. The Curie temperature, below which material exhibits ferromagnetic properties, is the highest when the concentration of Mn4+ is in the range of 25–30% [20,21]. It was reported that point defects, dependent on the oxygen content, strongly influenced the physical properties of manganites [5,8]. One of the possible ways of inducing defects to the matrix is aliovalent doping which leads also to the oxidation of Mn-array [5,22]. Kim et al. observed that the effect can also tune the electrochemical polarization leading to the dopant segregation in the matrix [23].
This work proposes a new approach to changing the magnetic properties of one of the types of manganites by introducing aliovalent ions and applying pressure. The paper is a continuation of our work on changing the magnetic properties of manganites by doping their structure with alkaline metal ions [24].

2. Materials and Methods

To obtain the La0.9A0.1Mn0.9Co0.1O3 (where A is Li, K, Na) compounds, the combustion method was applied [25]. Firstly, lanthanum (REacton®, 99.999%, Karlsruhe, Germany), manganese (Puratronic®, 99.995%, Karlsruhe, Germany) and cobalt (ACS, 98.0–102.0%, Washington, DC, USA) nitrates, and A–precursors (LiNO3-ACS, 99.0% min, NaNO3-ACS, 99.0% min or KNO3 ACS, 99.0% min, Washington, DC, USA) with the molar ratio of La/A and Mn/Co = 0.9/0.1 were dissolved in distilled water. Then 10% of molten stearic acid (Calbiochem, San Diego, CA, USA) was added to the solution in a porcelain reactor. The resulting mixture was continuously stirred and kept at a temperature of 140 °C. After 2 h La-Mn-stearic acid formed a gel. Then, the porcelain crucible reactor was placed in a furnace and heated up to 500 °C. The gel volatilized and auto ignited; with the evolution of a large volume of gases it produced loose grey powder. The obtained powders were ground in the agate mortar and additionally calcined at 700 °C for 1 h to remove all organic residues. After calcination, all powders were grinded and taken for further experiments. Half of the powders were used for ceramics sintering, using the high pressure sintering technique [26]. For a short duration powders were pressed (200 MPa) at room temperature to “green body” (GB). The GB was placed into a CaCO3 container with a graphite heater and the inside was isolated by hexagonal boron nitride. The shape of the container allowed to keep isostatic pressure during axial pressing. After the application of 8 GPa pressure, the pellets were sintered at 500 °C for 1 min into ceramics. After sintering, the as obtained ceramics were polished and taken for further studies.
The structure of the powders and ceramics were characterized and analyzed by powder X-ray diffraction (XRD) using a PANalytical X’Pert diffractometer (Malvern Panalytical, Almelo, The Netherlands) with Ni-filtered CuKλ radiation, λ = 1.5418 Å with the step of 0.02626 2θ for 1 h. The powders and ceramics were measured in the reflection mode. In the case of the powder, it was ground and placed directly in the sample holder, while the ceramics were placed in plasticine so that the surface of the ceramics would be parallel to the surface of the holder. For SEM imaging, the powders were ground in an agate mortar and glued to the carbon tape. The excess powder was removed using compressed air and the prepared material was taken for measurements. In the case of ceramics, the sample was simply glued to the carbon tape on the holder. Scanning electron microscopy (SEM) characterization was conducted using a commercial Tescan Vega3 SB microscope (SEM, Brno, Czech Republic) at 5 kV (acceleration voltage) with 3.5 spot (size of electron beam) and 3.2–3.7 mm working distance in order to analyze grain size and morphology. The chemical homogeneity of the powders and ceramics was measured using a scanning electron microscope FESEM FEI Nova NanoSEM 230 equipped with an energy dispersive spectroscopy (EDS) spectrometer (EDAX Genesis) at 18 kV (acceleration voltage) with 4.0 spot (size of electron beam) and 4.0 mm working distance. The XPS analysis was carried out with a Kratos Axis Supra spectrometer using monochromatic Al K(alpha) source (25 mA, 15 kV). The instrument work function was calibrated to give the binding energy (BE) of 83.96 eV for the Au 4f7/2 line for metallic gold, and the spectrometer dispersion was adjusted to give a BE of 932.62 eV for the Cu 2p3/2 line of metallic copper. The Kratos charge neutralizer system was used on all specimens. Survey scan analyses were carried out with an analysis area of 300 × 700 microns and a pass energy of 160 eV. High resolution analyses were carried out with an analysis area of 300 × 700 microns and a pass energy of 20 eV. The spectra were charge corrected to the main line of the carbon 1s spectrum (adventitious carbon) set to 284.8 eV. The spectra were analyzed using the CasaXPS software (version 2.3.23rev11.1R). The magnetic properties of the powders and ceramics were evaluated, using the Physical Properties Measurement Systems from Cryogenic Ltd. (London, UK) in magnetic fields up to ±14 T at 5 K.

3. Results

3.1. Structure and Morphology

The La0.9A0.1Mn0.9Co0.1O3 powders and ceramics were characterized by X-ray diffraction at room temperature (Figure 1). The crystal structure and the unit cell parameters were calculated using the Rietveld refinement with the X’pert HighScore Plus software using the pseudo-Voigt profile shape function, the background profile was fitted by the Lagrangian polynomial of 5th order [27]. This procedure allowed to attest the phase purity of the compounds and also to determine the crystal structure and structural parameters. The analysis of the XRD results indicated the rhombohedral symmetry described by R-3c space group (ICSD 75070) [28], which are similar to the structural results obtained for the La0.9A0.1MnO3 compounds, previously studied by the authors [24], and other LaMnO3-based compounds [29]. It should be noted that the crystal structure of undoped compounds is also characterized by the rhombohedral distortion of the unit cell which implies nominal oxygen excess in LaMnO3+δ (δ > 1%) compounds [30], leading to nonzero remanent magnetization as compared to stoichiometric LaMnO3 having an orthorhombic structure [31].
After the application of pressure, the diffraction peaks became wider, which may indicate a decrease in grain size associated with the decomposition of particle shells from the crystalline to amorphous state, such a phenomenon has already been described in our previous study [32]. In addition, the broadening of reflections is associated with the introduction of additional strains into the structure induced by high pressure during the sintering process [33]. It should be noted that the introduction of the bigger ion and the pressure applied during sintering lead to an increase in the strain in the compounds. It was observed that the introduction of 10 mol.% of alkali ions does not change the structure but clearly shifts the peaks to the higher 2θ degrees and increases the ionic radii [34]. This suggests an expansion of the unit cells and an increase in their volumes (Table 1) as confirmed by the results of the Rietveld refinement analysis. It should be noted that no additional peaks coming from the secondary phase were observed for the initial powder compounds, while the XRD patterns of the ceramic compounds contain distinct extra reflection located at ~26.5 2θ deg. The observed reflection is attributed to h-BN material integrated into tablets during the pressure procedure, which does not affect the crystal structure and the magnetic properties. In the case of ceramics, the applied pressure leads to a decrease in the unit cell volume, which has a significant impact on the magnetic properties of the ceramics as compared to the starting powders [35]. As the XRD measurements were performed in the reflection mode, the ceramics are small and, in some cases, they do not have parallel surfaces, the XRD patterns at high 2Θ degrees become noisy and the baseline rises. However, the main peak observed at lower degrees shows that the structure of the ceramics remains unchanged and is described by the rhombohedral symmetry.
The SEM images show that the combustion synthesis significantly affects grain morphology (Figure 2). Short synthesis time allows one to obtain nanosized crystallites and later calcination leads to their agglomeration. The high temperature generated during self-combustion (above 1500 °C) and relatively short calcination time causes broad grain size distribution and the crystallization of bigger grains [36]. However, even after the calcination of the powder for one hour, it can be seen that bigger aggregates are composed of smaller grains.
For powders and ceramics codoped with different alkali ions, the energy dispersive spectroscopy (EDS) maps were prepared to check the elements distribution and confirm that there is no segregation of particular ions on grain boundaries (Figure 3). The EDS analysis reveals that the concentration of the sodium and potassium is 10 mol% of La, as was assumed at the preparation stage, also cobalt ions substitute about 10% of manganese ions. For the powders and ceramics codoped with lithium, it was impossible to register signal from the dopant alkali ions as these ions are too light to be reliably registered by the analyzer. It can also be concluded that oxygen content in ceramics is about 10% lower than in powders, which was also confirmed in the EDS maps (Figure 3). It should be noted that the assessment of oxygen content by the EDS method has a rather qualitative character; however, at the same time the obtained results can allow one to estimate the evolution in the oxygen content as a function of the dopant ions and to determine a difference in the oxygen content in powder and ceramic samples.

3.2. Determination of Oxidation States of Mn and Co Ions

The high-resolution Mn 2p and Co 2p XPS spectra of La0.9A0.1Mn0.9Co0.1O3 (A: Li, K, Na) ceramics are shown in Figure 4. It can be seen that the Mn 2p spectrum consists of two asymmetric peaks located at 641.8 (Mn 2p3/2) and 653.3 eV (Mn 2p1/2). Both of these peaks can be deconvoluted into two peaks centered at 641.5 and 643.1 eV, and at 653.1 and 654.3 eV, respectively. The peaks located at 641.5 and 653.1 eV correspond to Mn3+ ions, while the peaks located at 643.1 and 654.3 eV are assigned to Mn4+ ions [37,38,39]. The absence of a satellite peak at +5 eV from Mn 2p3/2 indicates that no Mn2+ is present [40]. The Co 2p XPS spectrum exhibited two main peaks, corresponding to the 2p3/2 and 2p1/2 levels, and shake-up satellite peaks centered at around 790.0 and 796.9 eV. Similarly as in the case of Mn, the asymmetric peaks of Co 2p1/2 and Co 2p3/2 can be resolved into components. The results indicate that Co ions exist in the Co2+ and Co3+ oxidation state. The peaks at lower BE were attributed to Co3+ oxidation state, whereas the peaks at higher BE to Co2+ [41]. The atomic ratio of Mn3+/Mn4+ and Co2+/Co3+ is summarized in Table 2.

3.3. Magnetic Properties

The results of magnetization measurements clarified the modifications which occurred in the magnetic structure of lanthanum manganates driven by chemical codoping by alkali elements and cobalt ions (Figure 5). The results of magnetization measurements obtained for the undoped compound indicate the inhomogeneous magnetic state with a strong ferromagnetic component which is observed for the compounds regardless the preparation procedure. The complex magnetic structure of the initial LaMnO3 compound is associated with a lack in the cation content and thus the nominal excess of oxygen content leading to the formation of competing positive (Mn3+–O–Mn4+) and negative (Mn3+–O–Mn3+; Mn4+–O–Mn4+) exchange interactions associated with ferromagnetic and antiferromagnetic orders, respectively. The competing antiferromagnetic and ferromagnetic phases lead to the formation of a mixed magnetic state, which is characteristic for LnMnO3-based systems [24,42,43,44,45]. Chemical substitution with alkali elements and cobalt ions leads to an increase in the magnetization of the compounds which is caused by the higher number of manganese ions having +4 oxidation state and thus leading to the strengthening of positive exchange interactions Mn3+–Mn4+. Certain deviations from the above mentioned tendency were ascribed to the modification in the oxidation during the synthesis process [24] and thus to variations in the exchange interactions formed between Co and Mn ions.
The powder samples codoped with sodium and cobalt ions possess the maximal value of magnetization among the compounds of both series. The difference in the magnetization at the magnetic field of 14 T of the doped compound as compared to the initial LaMnO3 is about 30% (Figure 5). The observed increase in magnetization is most probably associated with the increase in the amount of Mn4+ ions and the stabilization of Co ions in +2 oxidation state. The rise in the amount of Mn4+ strengthens ferromagnetism in the compounds, due to positive exchange interactions Mn3+–Mn4+. Simultaneously, the presence of Co2+ ions causes the formation of new positive interactions Co2+–Mn4+. The stabilization of the cobalt ions in +2 oxidation state is in accordance with the notable increase in magnetic anisotropy observed for the doped compound (HC ~ 0.2 T for the compound La0.9Li0.1Mn0.9Co0.1O3 as compared to HC ~ 0.02 T of the initial compound). It should be noted that the ionic radius of the Co2+ ions (0.53Å for ion in high spin state and octahedral coordination) nearly coincides with the radius of Mn4+ ions with the same coordination, which facilitates the occupation of B- position of the perovskite lattice. The stabilization of +2 oxidation state of the cobalt ions is also in accordance with the structural data showing a gradual increase in the unit cell volume with chemical doping. One can conclude about the optimal concentration of the amount of Mn3+ and Mn4+ ions in the compounds codoped with sodium (potassium) and cobalt ions which show maximal magnetization due to the dominance of strongly positive exchange interactions Mn3+–Mn4+.
It should be noted that the magnitude of magnetization observed for the compound codoped with Li and Co ions is notably lower than the magnetization observed for other doped compounds in the powder form. The coercivity of the compound codoped with Li and Co ions is about 40% larger than the values observed for other doped compounds, which indicates a larger amount of the cobalt ions in +2 oxidation state as confirmed by XPS measurements. Examining isothermal magnetization dependences, one can assume that the magnetic state of the Li- and Co- codoped compounds is determined by the coexistence of positive exchange interactions formed between Co2+–Mn4+ ions and negative interactions Mn3+–Mn3+ and Co2+–Mn3+.
The ceramic compounds are characterized by a lower value of magnetization as compared to the related powder samples, which is caused by a reduced amount of oxygen content occurring during the high pressure procedure, as indirectly confirmed by the EDS and XPS data. The smaller amount of oxygen content results in the increase in the number of structural defects, which frustrates exchange interactions thus leading to the mixed magnetic state. As in the case of the compounds in the powder form, chemical doping leads to the stabilization of +2 oxidation state of the cobalt ions, which leads to an increase in the value of coercivity. The stabilization of cobalt ions in +2 oxidation state leads to an increase in the magnetization of the compounds doped with sodium (lithium) and cobalt ions, which can be caused by the formation of positive exchange interactions Co2+–Mn4+ ions. Magnetization estimated for the potassium and cobalt codoped compound is notably lower in comparison to the other codoped compounds, especially in the low field region, which can be explained by a more distinct reduction in oxygen content that occurred in these compounds during the high pressure procedure. The deficit in the oxygen content frustrates the dominant long-range ferromagnetic order, thus reducing spontaneous magnetization; moreover, an increase in the amount of oxygen vacancies provides negative Mn3+–Mn3+ and Co2+–Mn3+ interactions and thus fosters the stabilization of the mixed magnetic state characteristic of the initial LaMnO3 compound (Figure 5). The nonsystematic evolution of magnetization observed for the ceramic compounds can be explained by the absorption of oxygen by a sample holder during high pressure treatment. The formation of nanoscale crystallites with an average size of a few tens of nanometers, due to the application of high pressure, causes a reduction in magnetization for all ceramic compounds as compared to powders because of the disturbance of the long range magnetic order associated with ferromagnetism.

4. Discussion and Conclusions

The combustion method may be applied for fast and effective manganese powders production, also with a different type of codopants. The analysis of the structure and morphology of the manganites doped with cobalt ions and codoped with alkali ions shows the expansion of the unit cell with an increase in the alkali metal ion radius. It does not have an impact on the crystallization in the proper structure although it shows a great influence on the magnetic properties of the powders. It may also be concluded that applying high pressure during the sintering leads to a decrease in grains size, amorphization of grain shells and changes in the magnetic properties of the ceramics. The results of magnetization measurements point at the mixed magnetic state of the undoped compounds LaMnO3, regardless of the preparation procedure which is caused by the nominal excess of oxygen content and thus the formation of competing positive Mn3+–O–Mn4+ exchange interactions and negative Mn3+–O–Mn3+; Mn4+–O–Mn4+ interactions. Chemical doping with alkali and cobalt ions leads to the stabilization of Co ions in +2 oxidation state thus leading to the formation of positive Mn3+–Mn4+ and Co2+–Mn4+ exchange interactions and an increase in remanent magnetization. The difference in the magnetization values observed for the compounds in the powder form is caused by a certain deviation in the oxygen content formed during the preparation process. Ceramic compounds are characterized by lower magnetization as compared to the compounds in the powder form having the same chemical compositions, which is caused by a more intense reduction in oxygen content that occurred in the surface layer of the samples near the sample-hBN interface in the high pressure matrix and the nanometer size of the crystallites causing a reduction in the long range ferromagnetic order.

Author Contributions

P.G. conceived and planned the experiments and wrote the manuscript with support from other coauthors, R.N. synthesized powders, R.T. prepared ceramics, W.S. supervised the project, T.S. performed SEM/EDS measurements of powders, M.S. analyzed SEM/EDS measurements, T.M. performed XPS measurements, A.Z. analyzed XPS spectra, A.P. synthesized powders, R.S. supervised the synthesis and analyzed structural measurements, A.K. (Aivaras Kareiva) supervised the synthesis and analyzed the results, A.K. (Andrei Kholkin) analyzed and discussed the results, M.B. measured the magnetic properties of powders and ceramics, S.L. XRD data refinement, D.K. analyzed the magnetic properties of powders and ceramics. All authors provided critical feedback and helped shape the research, analysis and manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This project has received funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No 778070–TransFerr–H2020-MSCA-RISE-2017. Part of this work was developed within the scope of the project CICECO-Aveiro Institute of Materials, UIDB/50011/2020 and UIDP/50011/2020, financed by national funds through the Portuguese Foundation for Science and Technology/MCTES. The equipment of the Ural Center for Shared Use “Modern nanotechnology” UrFU was used. The work has been supported in part by the Ministry of Science and Higher Education of the Russian Federation under Project #FEUZ-2020-0054.
Applsci 10 08786 i001

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schmid, H. Multi-ferroic magnetoelectrics. Ferroelectrics 1994, 162, 317–338. [Google Scholar] [CrossRef]
  2. Jahn, H.A.; Teller, E. Stability of polyatomic molecules in degenerate electronic states—I—Orbital degeneracy. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1937, 161, 220–235. [Google Scholar] [CrossRef]
  3. Rivero, P.; Meunier, V.; Shelton, W. Electronic, structural, and magnetic properties of LaMnO3 phase transition at high temperature. Phys. Rev. B 2016, 93, 024111. [Google Scholar] [CrossRef] [Green Version]
  4. Norby, P.; Andersen, I.; Andersen, E.; Andersen, N.H. The crystal structure of lanthanum manganate(iii), LaMnO3, at room temperature and at 1273 K under N2. J. Solid State Chem. 1995, 119, 191–196. [Google Scholar] [CrossRef]
  5. Malavasi, L.; Ritter, C.; Mozzati, M.C.; Tealdi, C.; Islam, M.S.; Azzoni, C.B.; Flor, G. Effects of cation vacancy distribution in doped LaMnO3+δ perovskites. J. Solid State Chem. 2005, 178, 2042–2049. [Google Scholar] [CrossRef] [Green Version]
  6. Lv, Y.; Li, Z.; Yu, Y.; Yin, J.; Song, K.; Yang, B.; Yuan, L.; Hu, X. Copper/cobalt-doped LaMnO3 perovskite oxide as a bifunctional catalyst for rechargeable Li-O2 batteries. J. Alloy. Compd. 2019, 801, 19–26. [Google Scholar] [CrossRef]
  7. Liu, X.; Gong, H.; Wang, T.; Guo, H.; Song, L.; Xia, W.; Gao, B.; Jiang, Z.; Feng, L.; He, J. Cobalt-Doped Perovskite-Type Oxide LaMnO3 as Bifunctional Oxygen Catalysts for Hybrid Lithium-Oxygen Batteries. Chem. Asian J. 2018, 13, 528–535. [Google Scholar] [CrossRef]
  8. Zhu, W.; Chen, X.; Liu, Z.; Liang, C. Insight into the Effect of Cobalt Substitution on the Catalytic Performance of LaMnO3 Perovskites for Total Oxidation of Propane. J. Phys. Chem. C 2020, 124, 14646–14657. [Google Scholar] [CrossRef]
  9. Flores-Lasluisa, J.X.; Huerta, F.; Cazorla-Amorós, 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]
  10. Krishnan, R.V.; Banerjee, A. Evidence of dynamic Jahn-Teller effects in ferromagnetism of rhombohedral Al-substituted lanthanum manganite. J. Phys. Condens. Matter 2000, 12, 3835–3847. [Google Scholar] [CrossRef]
  11. Golenishchev-Kutuzov, A.V.; Golenishchev-Kutuzov, V.A.; Kalimullin, R.I.; Semennikov, A.V. Ordered states of Jahn-Teller distorted MnO6 octahedra in weakly doped lanthanum-strontium manganites. Phys. Solid State 2015, 57, 1633–1638. [Google Scholar] [CrossRef]
  12. Bogdanova, K.G.; Bulatov, A.R.; Golenishchev-Kutuzov, V.A.; Potapov, A.A. Influence of the Jahn-Teller effect on the structural, magnetic, and electrical properties of lightly doped manganites. Bull. Russ. Acad. Sci. Phys. 2013, 77, 275–277. [Google Scholar] [CrossRef]
  13. shihara, T. Perovskite Oxide for Solid Oxide Fuel Cells; Springer: New York, NY, USA, 2009; ISBN 78-0-387-77707-8. [Google Scholar]
  14. Jin, S.; Tiefel, T.H.; McCormack, M.; Fastnacht, R.A.; Ramesh, R.; Chen, L.H. Thousandfold Change in Resistivity in Magnetoresistive La-Ca-Mn-O Films. Science 1994, 264, 413–415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Tao, S.; Irvine, J.T.S.; Kilner, J.A. An Efficient Solid Oxide Fuel Cell Based upon Single-Phase Perovskites. Adv. Mater. 2005, 17, 1734–1737. [Google Scholar] [CrossRef]
  16. Ahmad, E.A.; Mallia, G.; Kramer, D.; Kucernak, A.R.; Harrison, N.M. The stability of LaMnO3 surfaces: A hybrid exchange density functional theory study of an alkaline fuel cell catalyst. J. Mater. Chem. A 2013, 1, 11152–11162. [Google Scholar] [CrossRef]
  17. Morelli, D.T.; Mance, A.M.; Mantese, J.V.; Micheli, A.L. Magnetocaloric properties of doped lanthanum manganite films. J. Appl. Phys. 1996, 79, 373–375. [Google Scholar] [CrossRef]
  18. 18. Jia, M.; Li, X.; Zhaorigetu; Shen, Y.; Li, Y. Activity and deactivation behavior of Au/LaMnO3 catalysts for CO oxidation. J. Rare Earths 2011, 29, 213–216. [Google Scholar] [CrossRef]
  19. Zhang, C.; Guo, Y.; Guo, Y.; Lu, G.; Boreave, A.; Retailleau, L.; Baylet, A.; Giroir-Fendler, A. LaMnO3 perovskite oxides prepared by different methods for catalytic oxidation of toluene. Appl. Catal. B Environ. 2014, 148–149, 490–498. [Google Scholar] [CrossRef]
  20. Schiffer, P.; Ramirez, A.P.; Bao, W.; Cheong, S.-W. Low Temperature Magnetoresistance and the Magnetic Phase Diagram of La1−xCaxMnO3. Phys. Rev. Lett. 1995, 75, 3336–3339. [Google Scholar] [CrossRef]
  21. Mahendiran, R.; Tiwary, S.; Raychaudhuri, A.; Ramakrishnan, T.; Mahesh, R.; Rangavittal, N.; Rao, C. Structure, electron-transport properties, and giant magnetoresistance of hole-doped systems. Phys. Rev. B Condens. Matter Mater. Phys. 1996, 53, 3348–3358. [Google Scholar] [CrossRef]
  22. Bara, M.; Dzik, J.; Feliksik, K.; Kozielski, L.; Wodecka-Duś, B.; Goryczka, T.; Zarycka, A.; Adamczyk-Habrajska, M. Influence of calcium ions on the structure and properties of LaMnO3. Arch. Metall. Mater. 2020, 65, 1189–1195. [Google Scholar]
  23. Kim, D.; Bliem, R.; Hess, F.; Gallet, J.-J.; Yildiz, B. Electrochemical Polarization Dependence of the Elastic and Electrostatic Driving Forces to Aliovalent Dopant Segregation on LaMnO3. J. Am. Chem. Soc. 2020, 142, 3548–3563. [Google Scholar] [CrossRef] [PubMed]
  24. Głuchowski, P.; Nikonkov, R.; Tomala, R.; Stręk, W.; Shulha, T.; Serdechnova, M.; Zheludkevich, M.; Pakalaniškis, A.; Skaudžius, R.; Kareiva, A.; et al. Magnetic Properties of La0.9A0.1MnO3 (A: Li, Na, K) Nanopowders and Nanoceramics. Materials 2020, 13, 1788. [Google Scholar] [CrossRef]
  25. Ekambaram, S.; Patil, K.C.; Maaza, M. Synthesis of lamp phosphors: Facile combustion approach. J. Alloy. Compd. 2005, 393, 81–92. [Google Scholar] [CrossRef] [Green Version]
  26. Fedyk, R.; Hreniak, D.; Łojkowski, W.; Stręk, W.; Matysiak, H.; Grzanka, E.; Gierlotka, S.; Mazur, P. Method of preparation and structural properties of transparent YAG nanoceramics. Opt. Mater. 2007, 29, 1252–1257. [Google Scholar] [CrossRef]
  27. Degen, T.; Sadki, M.; Bron, E.; König, U.; Nénert, G. The high score suite. In Powder Diffraction; Cambridge University Press: Cambridge, UK, 2014; Volume 29, pp. S13–S18. [Google Scholar]
  28. Van Roosmalen, J.A.M.; Cordfunke, E.H.P.; Helmholdt, R.B.; Zandbergen, H.W. The defect chemistry of lamno3±δ. 2. structural aspects of lamno3+δ. J. Solid State Chem. 1994, 110, 100–105. [Google Scholar] [CrossRef]
  29. Ma, F.; Jiao, Y.; Jiang, Z.; Du, A. Rhombohedral Lanthanum Manganite: A New Class of Dirac Half-Metal with Promising Potential in Spintronics. ACS Appl. Mater. Interfaces 2018, 10, 36088–36093. [Google Scholar] [CrossRef]
  30. Markovich, V.; Fita, I.; Mogilyansky, D.; Wisniewski, A.; Puzniak, R.; Titelman, L.; Vradman, L.; Herskowitz, M.; Gorodetsky, G. Effect of particle size on magnetic properties of LaMnO3 + δ nanoparticles. Superlattices Microstruct. 2008, 44, 476–482. [Google Scholar] [CrossRef]
  31. Handbook on the Physics and Chemistry of Rare Earths, Volume 33—1st Edition. Available online: https://0-www-elsevier-com.brum.beds.ac.uk/books/handbook-on-the-physics-and-chemistry-of-rare-earths/gschneidner/978-0-444-51323-6 (accessed on 20 November 2020).
  32. Gluchowski, P.; Stręk, W. Luminescence and excitation spectra of Cr3+:MgAl2O4 nanoceramics. Mater. Chem. Phys. 2013, 140, 222–227. [Google Scholar] [CrossRef]
  33. Zhao, Y.; Zhang, J. Microstrain and grain-size analysis from diffraction peak width and graphical derivation of high-pressure thermomechanics. J. Appl. Crystallogr. 2008, 41, 1095–1108. [Google Scholar] [CrossRef] [Green Version]
  34. Gálvez, M.E.; Jacot, R.; Scheffe, J.; Cooper, T.; Patzke, G.; Steinfeld, A. Physico-chemical changes in Ca, Sr and Al-doped La–Mn–O perovskites upon thermochemical splitting of CO2via redox cycling. Phys. Chem. Chem. Phys. 2015, 17, 6629. [Google Scholar] [CrossRef] [PubMed]
  35. Mikhaylov, V.I.; Zubov, E.E.; Pashchenko, A.V.; Varyukhin, V.N.; Shtaba, V.A.; Dyakonov, V.P.; Szewczyk, A.; Abal’oshev, A.; Piotrowski, K.; Lewandowski, S.J.; et al. Comparison of pressure, magnetic-field, and excess manganese effects on transport properties of film and bulk ceramic La–Ca manganites. Low Temp. Phys. 2006, 32, 139–147. [Google Scholar] [CrossRef] [Green Version]
  36. Guarieiro, L.L.N.; Guarieiro, A.L.N. Impact of the Biofuels Burning on Particle Emissions from the Vehicular Exhaust. In Biofuels—Status and Perspective; InTech: London, UK, 2015. [Google Scholar]
  37. Zhang, C.; Wang, C.; Zhan, W.; Guo, Y.; Guo, Y.; Lu, G.; Baylet, A.; Giroir-Fendler, A. Catalytic oxidation of vinyl chloride emission over LaMnO3 and LaB0.2Mn0.8O3 (B = Co, Ni, Fe) catalysts. Appl. Catal. B Environ. 2013, 129, 509–516. [Google Scholar] [CrossRef]
  38. Qi, G.; Yang, R.T. Characterization and FTIR studies of MnOx-CeO2 catalyst for low-temperature selective catalytic reduction of NO with NH3. J. Phys. Chem. B 2004, 108, 15738–15747. [Google Scholar] [CrossRef]
  39. Wang, F.; Dai, H.; Deng, J.; Bai, G.; Ji, K.; Liu, Y. Manganese Oxides with Rod-, Wire-, Tube-, and Flower-Like Morphologies: Highly Effective Catalysts for the Removal of Toluene. Environ. Sci. Technol. 2012, 46, 4034–4041. [Google Scholar] [CrossRef] [PubMed]
  40. Kucharczyk, B.; Tylus, W. Partial substitution of lanthanum with silver in the LaMnO3 perovskite: Effect of the modification on the activity of monolithic catalysts in the reactions of methane and carbon oxide oxidation. Appl. Catal. A: Gen. 2008, 335, 28–36. [Google Scholar] [CrossRef]
  41. Zhou, T.; Fan, W.; Liu, Y.; Wang, X. Comparative assessment of the chronic effects of five nano-perovskites on Daphnia magna: A structure-based toxicity mechanism. Environ. Sci. Nano 2018, 5, 708–719. [Google Scholar] [CrossRef]
  42. Figueiras, F.G.; Karpinsky, D.; Tavares, P.B.; Gonçalves, J.N.; Yañez-Vilar, S.; Moreira Dos Santos, A.F.; Franz, A.; Tovar, M.; Agostinho Moreira, J.; Amaral, V.S. Novel multiferroic state and ME enhancement by breaking the AFM frustration in LuMn1-xO3. Phys. Chem. Chem. Phys. 2017, 19, 1335–1341. [Google Scholar] [CrossRef]
  43. Troyanchuk, I.O.; Bushinsky, M.V.; Karpinsky, D.V.; Sikolenko, V.V.; Gavrilov, S.A.; Silibin, M.V.; Franz, A.; Ritter, C. Magnetic and magnetotransport properties of La1−xSrxMn0.5Co0.5O3 perovskites. Ceram. Int. 2018, 44, 1432–1437. [Google Scholar] [CrossRef]
  44. Coey, J.M.D.; Viret, M.; Von Molnár, S. Mixed-valence manganites. Adv. Phys. 1999, 48, 167–293. [Google Scholar] [CrossRef]
  45. García-Muñoz, J.; Fontcuberta, J.; Martinez, B.; Seffar, A.; Pinol, S.; Obradors, X. Magnetic frustration in mixed valence manganites. Phys. Rev. B Condens. Matter Mater. Phys. 1997, 55, R668–R671. [Google Scholar] [CrossRef]
Figure 1. XRD patterns recorded for La0.9A0.1Mn0.9Co0.1O3 powders and ceramics (a). The right panel shows enlarged region of the pattern, with shift of the peak and with change of the dopant ionic radius observed in powders (b) and broadening of the peak observed for ceramics (c).
Figure 1. XRD patterns recorded for La0.9A0.1Mn0.9Co0.1O3 powders and ceramics (a). The right panel shows enlarged region of the pattern, with shift of the peak and with change of the dopant ionic radius observed in powders (b) and broadening of the peak observed for ceramics (c).
Applsci 10 08786 g001
Figure 2. SEM images and particle size distribution of La0.9A0.1Mn0.9Co0.1O3 (A: Li, Na, K) powders.
Figure 2. SEM images and particle size distribution of La0.9A0.1Mn0.9Co0.1O3 (A: Li, Na, K) powders.
Applsci 10 08786 g002
Figure 3. Energy dispersive spectroscopy (EDS) maps of La0.9A0.1Mn0.9Co0.1O3 powders (left) and ceramics (right).
Figure 3. Energy dispersive spectroscopy (EDS) maps of La0.9A0.1Mn0.9Co0.1O3 powders (left) and ceramics (right).
Applsci 10 08786 g003
Figure 4. Mn 2p and Co 2p XPS spectra of La0.9Li0.1Mn0.9Co0.1O3 (a,b), La0.9Na0.1Mn0.9Co0.1O3 (c,d) La0.9K0.1Mn0.9Co0.1O3 (e,f) ceramics.
Figure 4. Mn 2p and Co 2p XPS spectra of La0.9Li0.1Mn0.9Co0.1O3 (a,b), La0.9Na0.1Mn0.9Co0.1O3 (c,d) La0.9K0.1Mn0.9Co0.1O3 (e,f) ceramics.
Applsci 10 08786 g004
Figure 5. Isothermal magnetization curves measured for the initial LaMnO3 and codoped La0.9A0.1Mn0.9Co0.1O3 (A: Li, Na, K) powders (left) and ceramics (right) at T = 5 K (the graphs presented using the same scales).
Figure 5. Isothermal magnetization curves measured for the initial LaMnO3 and codoped La0.9A0.1Mn0.9Co0.1O3 (A: Li, Na, K) powders (left) and ceramics (right) at T = 5 K (the graphs presented using the same scales).
Applsci 10 08786 g005
Table 1. Average crystallite size and the cell parameters calculated for La0.9A0.1Mn0.9M0.1O3 powders and ceramics based on the XRD data.
Table 1. Average crystallite size and the cell parameters calculated for La0.9A0.1Mn0.9M0.1O3 powders and ceramics based on the XRD data.
Crystallite Sizea, bcVStrain
nmÅÅÅ3%
Powders
La0.9Li0.1Mn0.9Co0.1O3
Rwp = 2.897; Rexp = 1.991; GOF = 2.119
225.5073(3)13.3145(3)349.73(4)
(58.29)
0.008
La0.9Na0.1Mn0.9Co0.1O3
Rwp = 3.216; Rexp = 1.993; GOF = 3.113
265.5041(1)13.3375(4)349.92(9)
(58.32)
0.045
La0.9K0.1Mn0.9Co0.1O3
Rwp = 4.285; Rexp = 2.012; GOF = 4.899
335.5127(1)13.3567(3)351.52(8)
(58.59)
0.054
Ceramics
La0.9Li0.1Mn0.9Co0.1O3
Rwp = 11.004; Rexp = 8.322; GOF = 9.849
95.5060(8)13.3123(7)348.64(3)
(58.11)
0.063
La0.9Na0.1Mn0.9Co0.1O3
Rwp = 13.415; Rexp = 9.071; GOF = 10.185
115.5080(3)13.3198(7)349.47(6)
(58.25)
0.065
La0.9K0.1Mn0.9Co0.1O3
Rwp = 17.080; Rexp = 12.096; GOF = 14.003
125.5058(7)13.3755(8)351.15(2)
(58.55)
0.077
Rwp—Weighted Profile Rietveld R-factor, Rexp—Expected Rietveld R-factor, GOF—goodness of fit.
Table 2. Oxidation state of Mn and Co ions in synthesized powders and ceramics.
Table 2. Oxidation state of Mn and Co ions in synthesized powders and ceramics.
Mn Oxidation State (%)Co Oxidation State (%)
3+4+2+3+
Powders
La0.9Li0.1Mn0.9Co0.1O373.226.861.138.9
La0.9Na0.1Mn0.9Co0.1O375.224.842.957.1
La0.9K0.1Mn0.9Co0.1O367.232.839.960.1
Ceramics
La0.9Li0.1Mn0.9Co0.1O375.824.260.239.8
La0.9Na0.1Mn0.9Co0.1O365.134.941.758.3
La0.9K0.1Mn0.9Co0.1O366.333.737.063.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Głuchowski, P.; Nikonkov, R.; Tomala, R.; Stręk, W.; Shulha, T.; Serdechnova, M.; Zarkov, A.; Murauskas, T.; Pakalaniškis, A.; Skaudžius, R.; et al. Impact of Alkali Ions Codoping on Magnetic Properties of La0.9A0.1Mn0.9Co0.1O3 (A: Li, K, Na) Powders and Ceramics. Appl. Sci. 2020, 10, 8786. https://0-doi-org.brum.beds.ac.uk/10.3390/app10248786

AMA Style

Głuchowski P, Nikonkov R, Tomala R, Stręk W, Shulha T, Serdechnova M, Zarkov A, Murauskas T, Pakalaniškis A, Skaudžius R, et al. Impact of Alkali Ions Codoping on Magnetic Properties of La0.9A0.1Mn0.9Co0.1O3 (A: Li, K, Na) Powders and Ceramics. Applied Sciences. 2020; 10(24):8786. https://0-doi-org.brum.beds.ac.uk/10.3390/app10248786

Chicago/Turabian Style

Głuchowski, Paweł, Ruslan Nikonkov, Robert Tomala, Wiesław Stręk, Tatsiana Shulha, Maria Serdechnova, Aleksej Zarkov, Tomas Murauskas, Andrius Pakalaniškis, Ramūnas Skaudžius, and et al. 2020. "Impact of Alkali Ions Codoping on Magnetic Properties of La0.9A0.1Mn0.9Co0.1O3 (A: Li, K, Na) Powders and Ceramics" Applied Sciences 10, no. 24: 8786. https://0-doi-org.brum.beds.ac.uk/10.3390/app10248786

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