Next Article in Journal
Low-Temperature SCR Catalyst Development and Industrial Applications in China
Previous Article in Journal
Continuous Flow Biocatalysis: Synthesis of Coumarin Carboxamide Derivatives by Lipase TL IM from Thermomyces lanuginosus
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Impact of Sintering Temperature on the Electrical Properties of La0.9Sr0.1MnO3 Manganite

1
Unité de Recherche Matériaux Avancés et Nanotechnologies (URMAN), Institut Supérieur des Sciences Appliquées et de Technologie de Kasserine, BP 471, Kasserine 1200, Tunisia
2
School of Engineering, Renewable Energies, Dundalk Institute of Technology, A91 K584 Dundalk, Ireland
3
Department of Chemistry, College of Sciences, Imam Mohammad Ibn Saud Islamic University (IMSIU), P.O. Box 5701, Riyadh 11432, Saudi Arabia
4
Laboratoire de Physique Appliquée, Faculté des Sciences, Université de Sfax, B.P. 1171, Sfax 3000, Tunisia
5
Laboratoire de Physique des Matériaux et des Nanomatériaux Appliquée à l’Environnement, Faculté des Sciences de Gabès Cité Erriadh, Université de Gabès, Gabès 6079, Tunisia
*
Author to whom correspondence should be addressed.
Submission received: 12 February 2022 / Revised: 28 February 2022 / Accepted: 8 March 2022 / Published: 17 March 2022
(This article belongs to the Topic Nanomaterials for Sustainable Energy Applications)

Abstract

:
La0.9Sr0.1MnO3 nanoparticles were prepared using the citrate–gel route and sintered at different temperatures (TS = 600 °C, 800 °C, and 1000 °C). The x-day diffraction patterns reveal that the samples exhibit a single phase with a rhombohedral ( R 3 ¯ C ) structure. The transmission electron microscopy technique shows an increase in the grain size when the sintering temperature (TS) rises. The obtained values are approximately similar to that of crystallite size calculated from x-ray diffraction patterns. The impact of sintering temperature (TS) on the electrical properties of La0.9Sr0.1MnO3manganite is examined using the impedance spectroscopy technique. A metal-semi-conductor transition at a specific temperature (TM-SC) is observed for all samples. Indeed, the sintering temperature increase induces the shift of this transition temperature toward higher temperatures. Such a behavior is explained by the increase in the grain size. An agreement between the metal-semi-conductor transition values coming from the DC resistivity and the grain boundaries analyses is observed. This agreement proves the contribution of the grain boundaries in the electrical properties of the studied samples. In addition, the presence of the relaxation phenomenon is confirmed. The fitted Nyquist plots show the correlation between the microstructure of the material and the electrical properties using an electrical equivalent circuit model. The DC resistivity and the impedance analyses reveal the thermal activation of the transport properties in the investigated system.

1. Introduction

Manganites constitute one of the most popular families of oxide-perovskites. This popularity comes from its distinguished physical properties. The colossal magnetoresistance effect is pronounced by the close interplay between the magnetic and the electrical transport properties of manganite compounds [1]. In addition, manganites are characterized by phase separation and charge ordering, which make them competitive compounds in magnetic applications, photonic and magnetoelectronic devices, and spintronic technology [2]. The strong correlation between their structural, magnetic, and electrical properties makes these kinds of materials compatible for numerous technological applications [3,4,5,6,7,8,9]. Among the manganite systems appropriate for these useful applications, lanthanum manganites (La1−xAxMnO3) have been extensively studied due totheir rich physical properties [10,11,12,13,14,15,16,17,18,19,20]. A focus on their structural properties attests to the fact that they have a general formula—La1−xAxMnO3—where La is a rare earth and A is a divalent or monovalent ion. The parent compound LaMnO3 is known as an anti-ferromagnetic insulator material. Its conductivity is enhanced by the incorporation of the divalent ion Sr2+ in LaMnO3 at the La site. This incorporation produces the proportional converts of Mn3+ to Mn4+ ion. The formation of the Mn4+ ions is generated due to the difference in charge between La3+ and Sr2+ ions. Indeed, the occurrence of Mn ions is responsible for the double exchange interaction manifested by the whole hops from Mn4+ to Mn3+ cations [21,22]. These hops are more likely to take place in the ferromagnetic phase where the localized spins of these ions are parallel [21]. This later synchronization between the ferromagnetic behavior and the enhancement of the double exchange mechanism gives an explanation to the metallic ferromagnetic character below the metal-semi-conductor transition temperature [23]. Furthermore, the bond angle Mn–O–Mn and bond length Mn–O have a strong dependence on the grain size in manganites [24,25,26,27,28]. So, the electrical information associated with the effects of grains and grain boundaries makes it very important to understand the overall properties of polycrystalline materials. Indeed, the charge carriers transport in these materials can occur by the contribution of grains, the grain boundary, and the sample–electrode interface. By means of the impedance spectroscopy technique, we have the ability to clarify these effects independently [29,30,31,32,33,34].
Lanthanum–strontium manganites (LSM) currently remain as commonly exploited compounds. Thus, they are used in solid oxide fuel cells as cathode materials due to their mixed conduction properties [35,36,37]. Vuk Uskokovic et al. [38] have explored their implication in biomedicine by benefiting from their use in hyperthermia treatment. In fact, the use of this compound allowed us the ability to distinguish malignant cells from healthy ones, which are less vulnerable to the increase in temperature. As a consequence, our studied compounds may be exploited in cancer cure [38,39]. Furthermore, given their good catalytic activity and high stability, LSM are used as electro-catalysts. They exhibit electro-conductivity, which is combined to the ionic conductivity provided by the ionic conductor backbones to form high functional cathodes [40]. LSM systems attract the attention of numerous research groups as they undergo a series of structural, electronic, and magnetic phase transitions when they are exposed to variable temperature [1]. Taking into consideration these tremendous usefulness and special properties, LSM were widely studied theoretically [41,42] and experimentally [30,31,32,43,44,45,46,47,48,49,50,51,52] from all sides (i.e., morphologically, structurally, magnetically, and electrically) and in different forms (i.e., bulk, as nanoparticles and thin films [53,54]). The effect of several factors was investigated such as vacancies at the La site [55] and the Mn site [56] as well as oxygen stoichiometry [38,57,58]. In addition, the effect of different amounts of Sr has been explored [59,60]. Old and recent studies have been discussed the effect of introducing an amount of 10% of Sr in an LaMnO3 system [37,45,46,47,48,49,59]. This interest about the La0.9Sr0.1MnO3 compound is due to its distinctive physical properties. In fact, such a compound is frequently used in solid oxide fuel cells (SOFC) as the cathode. For that reason, the interaction between the La0.9Sr0.1MnO3 cathode and metallic interconnect for a solid oxide fuel cell at reduced temperature has been investigated [37]. In order to more understand the physical properties of this system, Shinde et al. [49] have reported the influence of the annealing temperature on its morphological and magnetic properties. Furthermore, the preparation route determines the electrical and magnetic behavior of LSM. Among the preparation route conditions, the circumstances of the heat treatment (the sintering temperature [1,49,61,62,63,64,65,66], the sintering time [67,68], the steps followed in sintering, microwave sintering [69,70], the fully or partially annealing [71]) affect the response of lanthanum manganite systems. In fact, the increase in the sintering temperature (TS) induces a decrease in the electrical resistivity, which leads to the improving of the electrical transport in these materials [62]. Such a result seems to be explained by the increase of the mobility of charge carriers with the increase in TS. Moreover, the rise in this temperature (TS) conferred a modification on the metal-semi-conductor transition. Several researchers of manganites have reported the possibility of the shift of this transition temperature (TM-SC) toward the lower [66] and the higher [25,65,66] temperatures with the rise of Ts. The effects of sintering temperature on the physical properties of manganites have been recently reported [72,73]. Accordingly, it is found that TS affects the grain size, leading to a variation in the metal–insulator transition temperature [72]. Dey and Nath [72] confirm that the grain size variation does not affects the ferromagnetic–paramagnetic transition temperature. In some cases, this thermal excitation causes the disappearance of the transition temperature [26]. Whereas Baaziz et al. [61] found that the metal-semi-conductor transition temperature was not affected and estimated it at 160 K for all the sintered samples. In the literature, a complete study on the La0.67Sr0.33MnO3 compound has been investigated [61,74,75,76]. They started to explore the effect of sintering temperature on the structural and magnetic properties of La0.9Sr0.1MnO3 brought by the variation of particle size [77,78,79]. In terms of variation as a function of particle size, the evolution of the magnetic properties has been explained by the core–shell model [77]. In addition, the decrease in the Curie temperature (TC) with the increase in the grain size has been justified by the strain effect of grains induced by the distortion at grain boundaries. In addition, the RCP (Relative Cooling Power) values at a low field are comparable with those of the commercial magnetic refrigerant materials. Such comparison indicates that our compound can act as a candidate for magnetic refrigeration [79]. In the aim of completing the enlightenment of the remaining sides of the whole study, we opt to advance a study on the effect of sintering temperature on the electrical properties of the La0.9Sr0.1MnO3 system.

2. Experimental Details

La0.9Sr0.1MnO3 material was synthesized by the Citrate–Gel route. The nitrate precursors are La(NO3)3 6H2O, Mn(NO3)2 4H2O, and Sr(NO3)2. The stoichiometric amounts of the precursors are dissolved beforehand in water. Then, they are mixed with ethylene glycol (C2H6O2) and citric acid (C6H8O7), forming a stable solution. The metal/citric acid molar ratio was 1:1. Then, the solution was heated on a thermal plate under constant sintering at 80 °C to remove excess water and obtain a viscous gel. The gel was dried at 130 °C and then calcinated at 600 °C for 12 h. Eventually, the obtained powder is divided into three portions. The first one is the sample sintered at 600 °C. The second and the third portions are sintered at 800 and 1000 °C respectively for 12 h. The phase purity and the structure of the prepared samples were checked by x-ray diffraction (XRD) using the Cu-Kα1 radiation source (D5000 diffractometer, BRUKER). The recording of the x-ray powder diagrams was carried out at room temperature in an angular range varying from 10 to 100° with a recording time of 10 s by a step of 0.02°. Transmission electron microscopy (TEM) is used to determine the samples’ morphology. Then, the samples are pressed into pellets using a hydraulic press by applying a pressure of 4000 Psi. By means of impedance spectroscopy, the sintered samples were electrically characterized by the following steps. First, thin silver films of some nanometers were deposited on the two opposite faces of the pellets to serve as electrodes. Then, the samples were branched in a Janis VPF 800-cryostat to pick out the electrical measurements using an impedance analyzer 4294A whose frequency range varies from 40 Hz to 10 MHz. We performed the measurements in parallel mode for the equivalent circuit and with an amplitude of 20 mV for the signal. Furthermore, the electrical parameters were recorded in darkness and along the range of temperature extended from 80 to 400 K, using the liquid nitrogen to reach such low temperatures.

3. Results and Discussion

3.1. Structural and Morphological Study

Figure 1a–c show the XRD diagrams of the studied material La0.9Sr0.1MnO3 sintered at 600 (a), 800 (b), and 1000 °C (c). The diagrams present a typical Rietveld refinement including the observed and the calculated data as well as the difference profile. The formation of the perovskite phase La0.9Sr0.1MnO3 is confirmed. The prepared samples were found to be single phase without any detectable impurity, which are crystallized in the rhombohedral symmetry with the R 3 _ c space group. The refined lattice parameters are summarized in Table 1.
The unit cell volumes are 350.14(7), 351.11(4), and 351.92(8) Å3 for TS = 600, 800, and 1000 °C, respectively. For each sample, the average crystallite size was determined based on the full width at half maximum (FWHM) of the Bragg peaks (the intense peaks) and the peak position using the Debye–Scherrer expression as follows:
D SC = 0.9 λ β cos θ
where β is the full width at half maximum of the Bragg peak, λ is the wavelength of CuKα1 (λ = 1.5406Å), and θ is the diffraction angle of the most intense peak. Table 1 shows the sensibility of the crystallite size (d(XRD)) on the sintering process. Indeed, d(XRD) increases with the sintering temperature rise (d(XRD) = 45, 75, and 85 nm, respectively for TS = 600, 800, and 1000 °C). Such a result is in good agreement with comparable systems [61]. The increase in the unit cell volume may be due to a strong correlation between the lattice parameters and the grain size. Figure 1d–f shows the representative TEM micrographs for the particles of our samples. The micrographs reveal a variable grain size as a function of the sintering temperature. It exhibits an increase in the grain size (d(TEM)) from 43 to 72 to 84 nm with the increase in TS (Table 1). The calculated crystallite size is practically similar to the grain size. This indicates that each grain is composed by a single crystallite. Figure 1d–f shows also that these nanoparticles combine in agglomerates. As the sintering temperature rises, the agglomeration rate increases. In addition, it is observed that the prepared nanoparticles, with a form that is not completely spherical, have an inhomogeneous size.

3.2. DC-Resistivity Analysis

Figure 2a–c show the temperature dependence of the electrical DCresistivity (ρDC) for the three studied samples sintered at 600 (a), 800 (b), and 1000 °C (c). There is a similarity in each curve. Then, a metal-semi-conductor transition at TM-SC was observed. For T > TM-SC, the decrease in ρDC can be related to the fact that as the temperature rises, more charge carriers are released from the trapped centers and participate in the conduction. In perovskite systems (for T > TM-SC), the electrical conductivity is usually generated by hopping conduction processes [80,81]. The activation of such processes induces the observed resistivity decrease. The inferred values of TM-SC are depicted in Figure 2a–c. It can be seen that TM-SC increases with the increase in TS, as shown in Figure 2d. Such behavior has been also detected by the increase in the annealing time reported by Banerjee et al. [82] for the La0.5Pb0.5MnO3 compound. For the La0.67Sr0.33MnO3 system, TM-SC was found by Baaziz et al. [61] to be unaffected by the sintering temperature excitation. For the same system reported by Venkataiah et al. [83], TM-SC increases from 200 to 270 K with the rise in TS. Such a difference in results for the same investigated compound (La0.67Sr0.33MnO3) is probably due to the sintering and the synthesis conditions.
The metal-semi-conductor transition temperature can be explained by the DE mechanism, which can be affected by the grain boundary region. In fact, TM-SC may be correlated to a variation made on the stoichiometric ratio of Mn3+/Mn4+, the bond angle Mn–O–Mn (θMn–O–Mn), or the bond length Mn–O (dMn–O). In the present thermal excitation, the shift of TM-SC toward higher temperatures can be attributed to the reduction of the double exchange interactions as TS rises. This reduction may be a result of an increase in θMn–O–Mn or dMn–O caused by the detected increase in the unit cell volume (Table 1) as a function of particle size [24,25,26,27] or the fact that parts of the Mn3+–O2−–Mn4+ network are broken. The representative TEM (transmission electron microscopy) micrographs (Figure 1d–f) exhibit an increase in the grain size from 43 to 72 to 84 nm with the increase in TS (Table 1). Such an increase indicates that the particle size might have a considerable influence on these previously mentioned parameters, leading to its role in the shift of TM-SC. In a microstructure view of the samples, manganites are composed of two regions: a ferromagnetic-metallic one linked with the paramagnetic-insulating region. In addition, it is known that the electrical resistivity is strongly influenced by the region of the grain boundaries. This region can behave as an improved scattering center for the conductive electron. Consequently, the increase in grain size causes a reduction in the number of grain boundaries, leading to the decrease in this insulating region as well as the enhancement of the grain connectivity. The ascending of TM-SC values is expected [25,27,82,83]. An agreement with the literature [25,27,63,64,65,82,83,84,85] for comparable cases in manganites was confirmed. For the La0.85K0.15MnO3 compound [65], this heat treatment results an increase in the crystallite size (from 31.5 to 78.8 nm) as well as the grain size (from 85 to 490 nm). In addition, as TS rises, the unit cell volume increase leads to the detected increase indMn–O with a constant value of θMn–O–Mn. These factors cause the mentioned shift of TM-SC toward room temperature (from 281 to 290K) [65].

3.3. Impedance Analysis

The spectra of the normalized imaginary part of the impedance Z″ (Figure 3a–c) are characterized by the appearance of a peak at a specific frequency (fr) for each temperature curve.
This specific frequency is known as “the relaxation frequency”. The displacement of fr, the center of each peak, with temperature allows visualizing clearly the imprints of the metal-semi-conductor transition. Such behavior has been also observed for the La0.6Sr0.2Na0.2MnO3 compound [14]. Below TM-SC for each sample, the peak shifts toward low frequencies, as depicted in the inset of Figure 3a–c. Such an observation constitutes a good concretization of the metallic behavior. Above TM-SC, an opposite displacement takes place, reflecting a semi-conductor behavior proved by a thermally activated process (Figure 3a–c). Accordingly, the shift of this peak as a function of the temperature indicates the presence of the relaxation phenomenon in the investigated compound. Several groups of researchers have observed the contribution of this phenomenon in the conduction for different perovskite-type manganites [29,30,31,32,86]. However, the presence of this relaxation phenomenon is correlated to the presence of electrons/immobile species at low temperatures and defects/vacancies at the higher temperature side [86]. At lower frequencies, a second relaxation peak with a lower peak magnitude than the previous one is observed. Such an observation reveals the presence of two relaxation processes, which is clearly shown for the sample sintered at 800 °C. Numerous studies have found this experimental result in manganites [87,88,89]. Indeed, for the BaMnO3 manganite compound [88], the appearance of this second peak at low frequencies is confirmed. Accordingly, the relaxation frequency fr and its corresponding relaxation time τ values are deduced from each peak using the following relation:
fr · τ= 1
As shown in Figure 3d, the relaxation time ln(τ) varies linearly against the inverse of temperature (1/kB·T). Such a variation confirms that it obeys the Arrhenius law [90]:
τ ( T ) = τ 0 exp ( Ea / ( k B · T ) )
where τ0 is constant. The activation energy (Ea) values deducted from the slopes of these linear variations are shown in Figure 3d.
Figure 4a–c present the Nyquist diagrams of the studied samples. The diagrams are characterized by the presence of two overlapped semi-circle arcs with a conserved shape when the temperature increases. The overlapped nature of these arcs corresponds to the existence of more than one relaxation phenomenon, in which they are associated to different electro-active regions [91]. The overlap degree of the two arcs suggests that the relaxation frequencies of each contribution are very close. The evolution of the diameter of these arcs with the temperature excitation is another manifestation of the metal-semi-conductor transition. Indeed, for the observed semi-conductor behavior, the diameter of the semi-circle decreases as the temperature rises (Figure 4a–c). The insets of Figure 4 (a–c) show an increase in this diameter with temperature. Such behavior confirms the metallic character. So, the TM-SC values are detected at 170 K, 180 K, and 210 K for TS = 600, 800, and 1000 °C, respectively. Such values confirm the strong agreement between the impedance results and the resistivity evolution.
In order to examine the participation of the grains and grain boundaries regions as well as the electrodes in the electrical conduction, we have modeled the investigated samples to an electrical equivalent circuit. [Rg + (Rgb//CPEgb) + (Re//CPEe)] is the most appropriate equivalent circuit, which is plotted in the insets of Figure 4a–c (where Rg, Rgb, and Re are the grain, grain boundary, and electrodes resistances, respectively, and CPEgb and CPEe are the constant phase elements of the grain boundary and electrodes, respectively). Such a model allowed us to correlate between the electrical properties and the microstructure of this material. As shown in Figure 4a–c, the obtained experimental data were well fitted using Z-view software. According to this model, the first semi-circle presents the contribution of grains and grain boundaries [88,89], and the second one at low frequencies can be associated to the electrodes effect [88,89]. Moreover, the impedance response of grains dominates at high frequencies. Indeed, the left intercept of the first cited semi-circle with the real impedance axis is attributed to the grain resistance. The diameter of this semi-circle provides the grain boundary resistance. Accordingly, the right intercept at low frequencies is ascribed to the total resistance of the samples RT = Rg + Rgb + Re. The fitted values of these resistances are shown in Table 2.
Table 2 shows the reduction of electrode response with the rise of temperature for different TS. Such evolution may be related to the fact that the charge carriers’ mobility increases at the sample–electrode interface as the temperature rises. For the sample sintered at 600 °C, the electrodes region is much more resistive than the grain boundary (Re >> Rgb), as the sample starts to behave as a semi-conductor (T > TM-SC). Therefore, the electrodes response is responsible for the reduction of the grain boundary effect. So, in this temperature range, the contribution of the electrodes region in the conduction process is more important than the grain boundaries. Then, as the temperature rises, the grain boundary returns to be more resistive than the electrodes region for 340 K–400 K. As TS rises from 600, 800, or 1000 °C (and grain size increases), this observation is suppressed, and the grain boundaries recover their higher resistance values (Table 2). As it can be seen for the sample sintered at 800 °C, the electrodes resistance is lower than the grains and the grain boundary ones (Re < Rg < Rgb). As the sintering temperature increases from 800 to 1000 °C, accordingly, the grain size increases, and the electrodes resistance becomes higher than the grain one (Rg < Re < Rgb). Furthermore, Rg values are very low compared to Rgb (Rg << Rgb) [13,29,30,31,32]. In manganites, the grain boundaries are more resistive than the grains [13,29,30,31,32,91]. Such a result can be related to the existence of different conventional electro-active phases [91].
The grain boundary plays a decisive role in the determination of transport properties in manganite oxides [13,14,15,29,30,31,32]. Its resistance variation with temperature is represented in Figure 5a. Indeed, Rgb increases, sculpting the traces of a metallic behavior. Then, it decreases following a thermally activated process in a semi-conductor behavior. This decrease can be related to the fact that the grain boundaries effect has assisted in the reduction of the barrier against the charge carriers hopping. Thereby, it has paved the way to enhance the electrical transport with the temperature increase [30,32]. Hence, the previous TM-SC evolution is confirmed (Figure 5b). The evolution already discussed Rgb values with temperature for the sample sintered at 600 °C, which is represented in Figure 5a. This evolution is also observed in Figure 5c, which depicts the plot of the resistivity against temperature for different frequencies.
In the temperature range of 80–170 K (region 1), the resistivity is practically frequency independent in which the curves are combined. Then, as the temperature rises in region 2, the resistivity is strongly dependent on frequency. Indeed, a significant decrease in the electrical resistivity is observed. Such a result may be due to the pumping force of the applied frequency that excites the trapped centers of charge carriers for the gradual evacuation when the frequency increases [92]. In turn, this causes the participation of more charge carriers in the conduction process and therefore the improving of transport properties. Such a variation confirms the evolution mentioned above of the Rgb values, which are also obtained at high frequencies (Figure 5a). This similarity in their behaviors proves the contribution of the grain boundary region in conduction. In region 3, the resistivity is frequency independent, which can be employed in suitable applications. Figure 5d depicted the plots of ln (Rgb/T) against (1/kB·T). The curves are well fitted by a straight line for T > TM-SC. From this linear variation, we deduce the activation energy value for each sinteredsample. In fact, in manganites, the electrical transport properties are governed by the grain boundary region effects [13,15].

4. Conclusions

In this study, the effects of sintering temperature on the electrical properties of the La0.9Sr0.1MnO3 compound are explored. The citrate–gel method is used to prepare the studied samples. The XRD study shows the samples crystallization in the rhombohedral symmetry with the space group. The XRD and TEM analyses prove the increase in the crystallite and grain size under the sintering temperature effect. A metal-semi-conductor transition is observed for all samples, whereas the metallic character extends over a longer temperature range as TS rises. Such a shift of TM-SC toward high temperatures seems to be assigned to structural and morphological changes, and it is comparable to that reported in the literature. The impedance analysis confirms the existence of the electrical relaxation phenomenon. [Rg + (Rgb//CPEgb) + (Re//CPEe)] is the most suitable equivalent circuit model to describe our material. The resemblance in the resistivity and the grain boundaries variations with temperature confirms the contribution of the grain boundaries in the electrical transport properties.

Author Contributions

W.H., paper writing; H.R., data manipulation; M.G., data manipulation; A.G., data manipulation; N.B.H., paper editing and English correction; L.K., English correction E.D., paper editing and English correction; K.K., data manipulation. N.E.G. contributed in conceptualization, experiments, manuscript review. All authors have read and agreed to the published version of the manuscript.

Funding

Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University.

Data Availability Statement

Not applicable.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University for funding this work through Re-search Group no. RG-21-09-67.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dutta, A.; Gayathri, N.; Ranganathan, R. Effect of particle size on the magnetic and transport properties of La0.875Sr0.125MnO3. Phys. Rev. B 2003, 68, 054432. [Google Scholar] [CrossRef]
  2. Jin, S.; Tifel, T.H.; Cormack, M.M.; Fastenacht, 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] [Green Version]
  3. Kolat, V.S.; Gencer, H.; Gunes, M.; Atalay, S. Effect of B-doping on the structural, magnetotransport and magnetocaloric properties of La0.67Ca0.33MnO3 compounds. Mater. Sci. Eng. B 2007, 140, 212–217. [Google Scholar] [CrossRef]
  4. Zhang, Q.; Yin, L.; Mi, W.; Wang, X. Large Spatial Spin Polarization at Benzene/La2/3Sr1/3MnO3 Spinterface: Toward Organic Spintronic Devices. J. Phys. Chem. C 2016, 120, 6156–6164. [Google Scholar] [CrossRef]
  5. Yang, Q.; Yao, J.; Zhang, K.; Wang, W.; Zuo, X.; Tang, H.; Wu, M.; Li, G. Perovskite-type La1−xCaxMnO3 manganese oxides as effective counter electrodes for dye-sensitized solar cells. J. Electroanal. Chem. 2019, 833, 1–8. [Google Scholar] [CrossRef]
  6. Pankhurst, Q.A.; Connolly, J.; Jones, S.K.; Dobson, J. Applications of magnetic nanoparticles in biomedicine. J. Phys. D Appl. Phys. 2003, 36, R167. [Google Scholar] [CrossRef] [Green Version]
  7. Zhang, S.; Dong, G.; Liu, Y.; Li, H.; Chu, K.; Pu, X.; Yu, X.; Liu, X. Effect of Na-doping on structural, electrical, and magnetoresistive properties of La0.7(Ag0.3–xNax)0.3MnO3 polycrystalline ceramics. Ceram. Int. 2020, 46, 584–591. [Google Scholar] [CrossRef]
  8. Assoudi, N.; Walha, I.; Dhahri, E. Physical properties of double-doping Lanthanum manganite for bolometer applications. Chem. Phys. Lett. 2019, 731, 136609. [Google Scholar] [CrossRef]
  9. Patra, A.S.; Kumar, N.V.; Barpuzary, D.; De, M.; Qureshi, M. Strontium doped lanthanum manganites for efficient and robust photocatalytic water oxidation coupled with graphene oxide. Mater. Lett. 2014, 131, 125–127. [Google Scholar] [CrossRef]
  10. Rahmouni, H.; Cherif, B.; Baazaoui, M.; Khirouni, K. Effects of iron concentrations on the electrical properties of La0.67Ba0.33Mn1−xFexO3. J. Alloys Compd. 2013, 575, 5–9. [Google Scholar] [CrossRef]
  11. Rahmouni, H.; Cherif, B.; Jemai, R.; Dhahri, A.; Khirouni, K. Europium substitution for lanthanium in LaBaMnO–The structural and electrical properties of La0.7–xEuxBa0.3MnO3 perovskite. J. Alloys Compd. 2017, 690, 890–895. [Google Scholar] [CrossRef]
  12. Rahmouni, H.; Dhahri, A.; Khirouni, K. The effect of tin addition on the electrical conductivity of Sn-doped LaBaMnO3. J. Alloys Compd. 2014, 591, 259–262. [Google Scholar] [CrossRef]
  13. Rahmouni, H.; Smari, M.; Cherif, B.; Dhahri, E.; Khirouni, K. Conduction mechanism, impedance spectroscopic investigation and dielectric behavior of La0.5Ca0.5−xAgxMnO3 manganites with compositions below the concentration limit of silver solubility in perovskites (0 ≤ x ≤ 0.2). Dalton Trans. 2015, 44, 10457. [Google Scholar] [CrossRef]
  14. Charguia, R.; Hcini, S.; Boudard, M.; Dhahri, A. Microstructural properties, conduction mechanism, dielectric behavior, impedance and electrical modulus of La0.6Sr0.2Na0.2MnO3 manganite. J. Mater. Sci. Mater. Electron. 2019, 30, 2975–2984. [Google Scholar] [CrossRef]
  15. Elghoul, N.; Wali, M.; Kraiem, S.; Rahmouni, H.; Dhahri, E.; Khirouni, K. Sodium deficiency effect on the transport properties of La0.8Na0.2-xxMnO3 manganites. Phys. B Phys. Condens. Matter 2015, 478, 108–112. [Google Scholar] [CrossRef]
  16. Choura-Maatar, S.; Nofal, M.; M’nassri, R.; Cheikhrouhou-Koubaa, W.; Chniba-Boudjada, N.; Cheikhrouhou, A. Enhancement of the magnetic and magnetocaloric properties by Na substitution for Ca of La0.8Ca0.2MnO3 manganite prepared via the Pechini-type sol–gel process. J. Mater. Sci. Mater. Electron. 2020, 31, 1634–1645. [Google Scholar] [CrossRef]
  17. Mahjoub, S.; M’nassri, R.; Baazaoui, M.; Hlil, E.K.; Oumezzine, M. Tuning magnetic and magnetocaloric properties around room temperature via chromium substitution in La0.65Nd0.05Ba0.3MnO3 system. J. Magn. Magn. Mater. 2019, 481, 29–38. [Google Scholar] [CrossRef]
  18. M’nassri, R.; Cheikhrouhou-Koubaa, W.; Boudjada, N.C.; Cheikhrouhou, A. Effect of barium-deficiency on the structural, magnetic, and magnetocaloric properties of La0.6Sr0.2Ba0.2–xxMnO3 (0 ≤ x ≤ 0.15). J. Appl. Phys. 2013, 113, 073905. [Google Scholar] [CrossRef]
  19. Baazaoui, M.; Oumezzine, M.; Cheikhrouhou-Koubaa, W. Critical Behavior in Ga-Doped Manganites La0.65Bi0.05Sr0.3Mn1−xGaxO3 (x = 0 and 0.06). Phys. Solid State 2020, 62, 278–284. [Google Scholar] [CrossRef]
  20. Liu, Y.; Sun, T.; Dong, G.; Zhang, S.; Chu, K.; Pu, X.; Li, H.; Liu, X. Dependence on sintering temperature of structure, optical and magnetic properties of La0.625Ca0.315Sr0.06MnO3 perovskite nanoparticles. Ceram. Int. 2019, 45, 17467–17475. [Google Scholar] [CrossRef]
  21. Zener, C. Interaction between the d-Shells in the Transition Metals. II. Ferromagnetic Compounds of Manganese with Perovskite Structure. Phys. Rev. 1951, 82, 403–405. [Google Scholar] [CrossRef]
  22. de Gennes, P.G. Effects of Double Exchange in Magnetic Crystals. Phys. Rev. 1960, 118, 141–154. [Google Scholar] [CrossRef]
  23. Lakshmi, Y.K.; Venkataiaha, G.; Vithal, M.; Reddy, P.V. Magnetic and electrical behavior of La1−xAxMnO3 (A= Li, Na, K and Rb) manganites. Phys. B Condens. Matter 2008, 403, 3059–3066. [Google Scholar] [CrossRef]
  24. Zhang, N.; Yang, W.; Ding, W.; Xing, D.; Du, Y. Grain size-dependent magnetism in fine particle perovskite, La1−xSrxMnOz. Solid State Commun. 1999, 109, 537–542. [Google Scholar] [CrossRef]
  25. Oumezzine, M.; Hcini, S.; Baazaoui, M.; Sales, H.B.; Santos, I.M.G.d.; Oumezzine, M. Crystallite size effect on the structural, microstructure, magnetic and electrical transport properties of Pr0.7Sr0.3Mn0.9Cr0.1O3 nanocrystalline via a modified Pechini method. J. Alloys Compd. 2013, 571, 79–84. [Google Scholar] [CrossRef]
  26. Oumezzine, M.; Peña, O.; Guizouarn, T.; Lebullenger, R.; Oumezzine, M. Impact of the sintering temperature on the structural, magnetic and electrical transport properties of doped La0.67Ba0.33Mn0.9Cr0.1O3 manganite. J. Magn. Magn. Mater. 2012, 324, 2821–2828. [Google Scholar] [CrossRef]
  27. Venkataiah, G.; Reddy, P.V. Electrical behavior of sol–gel prepared Nd0.67Sr0.33MnO3 manganite system. J. Magn. Magn. Mater. 2005, 285, 343–352. [Google Scholar] [CrossRef]
  28. M’nassri, R.; Boudjada, C.; Cheikhrouhou, A. Impact of sintering temperature on the magnetic and magnetocaloric properties in Pr0.5Eu0.1Sr0.4MnO3 manganites. J. Alloys Compd. 2015, 626, 20–28. [Google Scholar] [CrossRef]
  29. Moualhi, Y.; M’nassri, R.; Nofal, M.M.; Rahmouni, H.; Selmi, A.; Gassoumi, M.; Chniba-Boudjada, N.; Khirouni, K.; Cheikrouhou, A. Influence of Fe doping on physical properties of charge ordered praseodymium–calcium–manganite material. Eur. Phys. J. Plus 2020, 135, 809. [Google Scholar] [CrossRef]
  30. Rahmouni, H.; Nouiri, M.; Jemai, R.; Kallel, N.; Rzigua, F.; Selmi, A.; Khirouni, K.; Alaya, S. Electrical conductivity and complex impedance analysis of 20% Ti-doped La0.7Sr0.3MnO3 perovskite. J. Magn. Magn. Mater. 2007, 316, 23–28. [Google Scholar] [CrossRef]
  31. Rahmouni, H.; Jemai, R.; Nouiri, M.; Kallel, N.; Rzigua, F.; Selmi, A.; Khirouni, K.; Alaya, S. Admittance spectroscopy and complex impedance analysis of Ti-modified La0.7Sr0.3MnO3. J. Cryst. Growth 2008, 310, 556–561. [Google Scholar] [CrossRef]
  32. Rahmouni, H.; Selmi, A.; Khirouni, K.; Kallel, N. Chromium effects on the transport properties in La0.7Sr0.3Mn1−xCrxO3. J. Alloys Compd. 2012, 533, 93–96. [Google Scholar] [CrossRef]
  33. Kumar, V.P.; Dayal, V.; Hadimani, R.L.; Bhowmik, R.N.; Jiles, D.C. Magnetic and electrical properties of Ti-substituted lanthanum bismuth manganites. J. Mater. Sci. 2015, 50, 3562–3575. [Google Scholar] [CrossRef]
  34. Khan, A.A.; Fayaz, M.U.; Khan, M.N.; Iqbal, M.; Majeed, A.; Bilkees, R.; Mukhtar, S.; Javed, M. Investigation of charge transport mechanism in semiconducting La0.5Ca0.5Mn0.5Fe0.5O3 manganite prepared by sol–gel method. J. Mater. Sci. Mater. Electron. 2018, 29, 13577–13587. [Google Scholar] [CrossRef]
  35. Sacanell, J.; Sánchez, J.H.; López, A.E.R.; Martinelli, H.; Siepe, J.; Leyva, A.G.; Ferrari, V.; Juan, D.; Pruneda, M.; Gómez, A.M.; et al. Oxygen Reduction Mechanisms in Nanostructured La0.8Sr0.2MnO3 Cathodes for Solid Oxide Fuel Cells. J. Phys. Chem. 2017, 121, 6533–6539. [Google Scholar] [CrossRef] [Green Version]
  36. Carthy, B.P.M.; Pederson, L.R.; Chou, Y.S.; Zhou, X.; Surdoval, W.A.; Wilson, L.C. Low-temperature sintering of lanthanum strontium manganite-based contact pastes for SOFCs. J. Power Sources 2008, 180, 294–300. [Google Scholar]
  37. Wu, J.; Yan, D.; Pu, J.; Chi, B.; Jian, L. The investigation of interaction between La0.9Sr0.1MnO3 cathode and metallic interconnect for solid oxide fuel cell at reduced temperature. J. Power Sources 2012, 202, 166–174. [Google Scholar] [CrossRef]
  38. Uskoković, V.; Košak, A.; Drofenik, M. Silica-coated lanthanum–strontium manganites for hyperthermia treatments. Mater. Lett. 2006, 60, 2620–2622. [Google Scholar] [CrossRef] [Green Version]
  39. Manh, D.H.; Phong, P.T.; Namb, P.H.; Tung, D.K.; Phuc, N.X.; Lee, I.J. Structural and magnetic study of La0.7Sr0.3MnO3 nanoparticles and AC magnetic heating characteristics for hyperthermia applications. Phys. B 2014, 444, 94–102. [Google Scholar] [CrossRef]
  40. Ju, J.; Lin, J.; Wang, Y.; Zhang, Y.; Xia, C. Electrical performance of nanostructured strontium-doped lanthanum manganite impregnated onto yttria-stabilized zirconia backbone. J. Power Sources 2016, 302, 298–307. [Google Scholar] [CrossRef]
  41. Qiu, J.; Jin, K.; Han, P.; Lu, H.; Hu, C.; Wang, B.; Yang, G. A theoretical study on the transport property of the La0.7Sr0.3MnO3/Si p-n heterojunction. Europhys. Lett. 2007, 79, 57004. [Google Scholar] [CrossRef] [Green Version]
  42. Hu, C.-L.; Jin, K.-J.; Han, P.; Lu, H.-B.; Liao, L.; Yang, G.-Z. Theoretical study on the positive magnetoresistance in perovskite oxide p–n junctions. Solid State Commun. 2009, 149, 334–336. [Google Scholar] [CrossRef]
  43. Rahmouni, H.; Jemai, R.; Kallel, N.; Selmi, A.; Khirouni, K. Titanium effects on the transport properties in La0.7Sr0.3Mn1−xTixO3. J. Alloys Compd. 2010, 497, 1–5. [Google Scholar] [CrossRef]
  44. Venkataiah, G.; Prasad, V.; Reddy, P.V. Structure and electrical transport of some Cd-doped La0.67Sr0.33MnO3 manganites. Phys. Status Solidi 2006, 203, 2478–2487. [Google Scholar] [CrossRef]
  45. Choi, J.H.; Jang, J.H.; Ryu, J.H.; Oh, S.M. Microstructure and cathodic performance of La0.9Sr0.1MnO3 electrodes according to particle size of starting powder. J. Power Sources 2000, 87, 92–100. [Google Scholar] [CrossRef]
  46. 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]
  47. Korolyov, A.V.; Arkhipov, V.Y.; Gaviko, V.S.; Mukovskii, Y.; Arsenov, A.A.; Lapina, T.P.; Bader, S.D.; Jiang, J.S.; Nizhankovskii, V.I. Magnetic properties and magnetic states in La0.9Sr0.1MnO3. J. Magn. Magn. Mater. 2000, 213, 63–74. [Google Scholar] [CrossRef]
  48. Arkhipov, V.E.; Gaviko, V.S.; Korolyov, A.V.; Naish, V.E.; Marchenkov, V.V.; Mukovskii, Y.M.; Karabashev, S.G.; Shulyatev, D.A.; Arsenov, A.A. Structural and magnetic phase transition in La0.9Sr0.lMnO3. J. Magn. Magn. Mater. 1999, 196, 539–540. [Google Scholar] [CrossRef]
  49. Shinde, K.P.; Pawar, S.S.; Pawar, S.H. Influence of annealing temperature on morphological and magnetic properties of La0.9Sr0.1MnO3. Appl. Surf. Sci. 2011, 257, 9996–9999. [Google Scholar] [CrossRef]
  50. Das, S.; Chowdhury, P.; Rao, T.K.G.; Das, D.; Bahadur, D. Influence of grain size and grain boundaries on the properties of La0.7Sr0.3CoxMn1−xO3. Solid State Commun. 2002, 121, 691–695. [Google Scholar] [CrossRef]
  51. Hizi, W.; Rahmouni, H.; Gassoumi, M.; Khirouni, K.; Dhahri, S. Transport properties of La0.9Sr0.1MnO3 manganite. Eur. Phys. J. Plus 2020, 135, 456. [Google Scholar] [CrossRef]
  52. Choudhary, Y.R.S.; Mangavati, S.; Patil, S.; Rao, A.; Nagaraja, B.S.; Thomas, R.; Kini, S.G. Effect of rare-earth substitution at La-site on structural, electrical and thermoelectric properties of La0.7−xRExSr0.3MnO3 compounds (x = 0, 0.2, 0.3; RE = Eu, Gd, Y). J. Magn. Magn. Mater. 2018, 451, 110–120. [Google Scholar] [CrossRef]
  53. Chen, G.J.; Chang, Y.H.; Hsu, H.W. The effect of microstructure and secondary phase on magnetic and electrical properties of La-deficient La1−xMnO3−δ (0 ≤ x ≤ 0.3) films. J. Magn. Magn. Mater. 2000, 219, 317–324. [Google Scholar] [CrossRef]
  54. Yousefi, S.; Trappen, R.; Mottaghi, N.; Bristow, A.D.; Holcomb, M.B. Oxygen vacancy effect on ultra-fast carrier dynamics of perovskite oxide La0.7Sr0.3MnO3 thin films, Ultrafast Phenomena and Nanophotonics XXIV. Int. Soc. Opt. Photonics 2020, 11278, 112780. [Google Scholar]
  55. Dhahri, R.; Bejar, M.; Hajlaoui, M.; Sdiri, N.; Valente, M.A.; Dhahri, E. Structural properties and electrical behaviour in the polycrystalline lanthanum-deficiency La1−xxMnO3 manganites. J. Magn. Magn. Mater. 2009, 321, 1735–1738. [Google Scholar] [CrossRef]
  56. Ji, D.H.; Tang, G.D.; Li, Z.Z.; Han, Q.J.; Hou, X.; Bian, R.R.; Liu, S.R. Investigation on the maximum content of vacancy at the B sites in La0.75Sr0.25Mn1−xO3 perovskite manganites. J. Appl. Phys. 2012, 111, 113902. [Google Scholar] [CrossRef]
  57. Abdelmoula, N.; Guidara, K.; Cheikh-Rouhou, A.; Dhahri, E. Effects of the Oxygen Nonstoichiometry on the Physical Properties of La0.7Sr0.3MnO3−δδ Manganites (0 ≤ δ ≤ 0.15). J. Solid State Chem. 2000, 151, 139–144. [Google Scholar] [CrossRef]
  58. Malavasi, L.; Mozzati, M.C.; Azzoni, C.B.; Chiodelli, G.; Flor, G. Role of oxygen content on the transport and magnetic properties of La1−xCaxMnO3+δ manganites. Solid State Commun. 2002, 123, 321–326. [Google Scholar] [CrossRef]
  59. Li, Y.; Zhang, H.; Liu, X.; Chen, Q.; Chen, Q. Electrical and magnetic properties of La1−xSrxMnO3 (0.1 ≤ x ≤ 0.25) ceramics prepared by sol–gel technique. Ceram. Int. 2019, 45, 16323–16330. [Google Scholar] [CrossRef]
  60. Anane, A.; Dupas, C.; le Dang, K.; Renard, J.P.; Veillet, P.; Guevara, A.M.d.; Millot, F.; Pinsardi, L.; Revcolevschi, A. Transport properties and magnetic behaviour of Lal-xSrxMnO3 single crystals. J. Phys. Condens. Matter 1995, 7, 7015–7021. [Google Scholar] [CrossRef]
  61. Baaziz, H.; Maaloul, N.K.; Tozri, A.; Rahmouni, H.; Mizouri, S.; Khirouni, K.; Dhahri, E. Effect of sintering temperature and grain size on the electrical transport properties of La0.67Sr0.33MnO3 manganite. Chem. Phys. Lett. 2015, 640, 77–81. [Google Scholar] [CrossRef]
  62. Mleiki, A.; Khlifi, A.; Rahmouni, H.; Guermazi, N.; Khirouni, K.; Hlil, E.K.; Cheikhrouhou, A. Magnetic and dielectric properties of Ba-lacunar La0.5Eu0.2Ba0.3MnO3 manganites synthesized using sol-gel method under different sintering temperatures. J. Magn. Magn. Mater. 2020, 502, 166571. [Google Scholar] [CrossRef]
  63. Venkataiah, G.; Krishna, D.C.; Vithal, M.; Rao, S.S.; Bhat, S.V.; Prasad, V.; Subramanyam, S.V.; Reddy, P.V. Effect of sintering temperature on electrical transport properties of La0.67Ca0.33MnO3. Phys. B 2005, 357, 370–379. [Google Scholar] [CrossRef]
  64. Venkataiah, G.; Lakshmi, Y.K.; Reddy, P.V. Influence of sintering temperature on resistivity, magnetoresistance and thermopower of La0.67Ca0.33MnO3. PMC Phys. B 2008, 1, 7. [Google Scholar] [CrossRef] [Green Version]
  65. Pan, K.Y.; Halim, S.A.; Lim, K.P.; Daud, W.M.; Chen, S.K. Effect of Sintering Temperature on Microstructure, Electrical and Magnetic Properties of La0.85K0.15MnO3 Prepared by Sol-Gel Method. J. Supercond. Nov. Magn. 2012, 25, 1177–1183. [Google Scholar] [CrossRef]
  66. Lakshmi, Y.K.; Reddy, P.V. Influence of sintering temperature and oxygen stoichiometry on electrical transport properties of La0.67Na0.33MnO3 manganite. J. Alloys Compd. 2009, 470, 67–74. [Google Scholar] [CrossRef]
  67. Shannigrahi, S.R.; Tan, S.Y. Effects of processing parameters on the properties of lanthanum manganite. Mater. Chem. Phys. 2011, 129, 15–18. [Google Scholar] [CrossRef]
  68. Li, N.; Mahapatra, M.K.; Singh, P. Sintering of porous strontium doped lanthanum manganite-yttria stabilized zirconia composite in controlled oxygen atmosphere at 1400 °C. J. Power Sources 2013, 221, 57–63. [Google Scholar] [CrossRef]
  69. Othmani, S.; Bejar, M.; Dhahri, E.; Hlil, E.K. The effect of the annealing temperature on the structural and magnetic properties of the manganites compounds. J. Alloys Compd. 2009, 475, 46–50. [Google Scholar] [CrossRef]
  70. Li, C.-X.; Yang, B.; Zhang, S.-T.; Zhang, R.; Cao, W.W. Effects of sintering temperature and poling conditions on the electrical properties of Ba0.70Ca0.30TiO3 diphasic piezoelectric ceramics. Ceram. Int. 2013, 39, 2967–2973. [Google Scholar] [CrossRef]
  71. Worledge, D.C.; Snyder, G.J.; Beasley, M.R.; Geballe, T.H.; Hiskes, R.; DiCarolis, S. Anneal-tunable Curie temperature and transport of La0.67Ca0.33MnO3. J. Appl. Phys. 1996, 80, 5158–5161. [Google Scholar] [CrossRef]
  72. Dey, P.; Nath, T.K. Effect of grain size modulation on the magneto- and electronic-transport properties of La0.7Ca0.3MnO3 nanoparticles: The role of spin-polarized tunneling at the enhanced grain surface. Phys. Rev. B 2006, 73, 214425. [Google Scholar] [CrossRef]
  73. Paredes-Navia, S.A.; Liang, L.; Romo-De-La-Cruz, C.O.; Gemmen, E.; Fernandes, A.; Prucz, J.; Chen, Y.; Song, X. Electrical conductivity increase by order of magnitude through controlling sintering to tune hierarchical structure of oxide ceramics. J. Solid State Chem. 2021, 294, 121831. [Google Scholar] [CrossRef]
  74. Baaziz, H.; Tozri, A.; Dhahri, E.; Hlil, E.K. Size-induced Griffiths phase-like in ferromagnetic metallic La0.67Sr0.33MnO3 nanoparticles. J. Magn. Magn. Mater. 2016, 403, 181–187. [Google Scholar] [CrossRef]
  75. Baaziz, H.; Tozri, A.; Dhahri, E.; Hlil, E.K. Influence of grain size and sintering temperature grain size on the critical behavior near the paramagnetic to ferromagnetic phase transition temperature in La0.67Sr0.33MnO3 nanoparticles. J. Magn. Magn. Mater. 2018, 449, 207–213. [Google Scholar] [CrossRef]
  76. Baaziz, H.; Tozri, A.; Dhahri, E.; Hlil, E.K. Magnetocaloric properties of La0.67Sr0.33MnO3 tunable by particle size and dimensionality. Chem. Phys. Lett. 2018, 691, 355–359. [Google Scholar] [CrossRef]
  77. Baaziz, H.; Tozri, A.; Dhahri, E.; Hlil, E.K. Effect of particle size reduction on the structural, magnetic properties and the spin excitations in ferromagnetic insulator La0.9Sr0.1MnO3 nanoparticles. Ceram. Int. 2015, 41, 2955–2962. [Google Scholar] [CrossRef]
  78. Baaziz, H.; Tozri, A.; Dhahri, E.; Hlil, E.K. Size reduction effect on the critical behavior near the paramagnetic to ferromagnetic phase transition temperature in La0.9Sr0.1MnO3 nanoparticles. Solid State Commun. 2015, 208, 45–52. [Google Scholar] [CrossRef]
  79. Baaziz, H.; Tozri, A.; Dhahri, E.; Hlil, E.K. Effect of particle size reduction on the magnetic phase transition and the magnetocaloric properties in ferromagnetic insulator La0.9Sr0.1MnO3 nanoparticles. Chem. Phys. Lett. 2015, 625, 168–173. [Google Scholar] [CrossRef]
  80. Moualhi, Y.; Rahmouni, H.; Khirouni, K. Usefulness of theoretical approaches and experiential conductivity measurements for understanding manganite-transport mechanisms. Results Phys. 2020, 19, 103570. [Google Scholar] [CrossRef]
  81. Moualhi, Y.; Rahmouni, H.; Gassoumi, M.; Khirouni, K. Summerfield scaling model and conduction processes defining the transport properties of silver substituted half doped (La–Ca) MnO3 ceramic. Ceram. Int. 2020, 46, 24710–24717. [Google Scholar] [CrossRef]
  82. Banerjee, A.; Pal, S.; Bhattecharya, S.; Chaudhuri, B.K.; Yang, H.D. Particle size and magnetic field dependent resistivity and thermoelectric power of La0.5Pb0.5MnO3 above and below metal–insulator transition. J. Appl. Phys. 2002, 91, 5125–5134. [Google Scholar] [CrossRef]
  83. Venkataiah, G.; Lakshmi, Y.K.; Prasad, V.; Reddy, P.V. Influence of Particle Size on Electrical Transport Properties of La0.67Sr0.33MnO3 Manganite System. J. Nanosci. Nanotechnol. 2007, 7, 2000–2004. [Google Scholar] [CrossRef] [PubMed]
  84. Huang, Y.-H.; Yan, C.-H.; Wang, Z.-M.; Liao, C.-S.; Xu, G.-X. Variation of structural, magnetic and transport properties in RE0.7Sr0.3MnO3 granular perovskites. Solid State Commun. 2001, 118, 541–546. [Google Scholar] [CrossRef]
  85. Venkataiah, G.; Huang, J.C.A.; Reddy, P.V. Low temperature resistivity minimum and its correlation with magnetoresistance in La0.67Ba0.33MnO3 nanomanganites. J. Magn. Magn. Mater. 2010, 322, 417–423. [Google Scholar] [CrossRef]
  86. Mleiki, A.; Hanen, R.; Rahmouni, H.; Guermazi, N.; Khirouni, K.; Hli, E.K.; Cheikhrouhou, A. Study of magnetic and electrical properties of Pr0.65Ca0.25Ba0.1MnO3 manganite. RSC Adv. 2018, 8, 31755–31763. [Google Scholar] [CrossRef] [Green Version]
  87. Hanen, R.; Mleiki, A.; Rahmouni, H.; Guermazi, N.; Khirouni, K.; Hlil, E.K.; Cheikhrouhou, A. Effect of the nature of the dopant element on the physical properties of X-PrCaMnO system (X = Cd, Sr, and Pb). J. Magn. Magn. Mater. 2020, 508, 166810. [Google Scholar] [CrossRef]
  88. Hayat, K.; Nadeem, M.; Iqbal, M.J.; Rafiq, M.A.; Hasan, M.M. Analysis of electro-active regions and conductivity of BaMnO3 ceramic by impedance spectroscopy. Appl. Phys. A 2014, 115, 1281–1289. [Google Scholar] [CrossRef]
  89. Li, M.; Sinclair, D.C. The extrinsic origins of high permittivity and its temperature and frequency dependence in Y0.5Ca0.5MnO3 and La1.5Sr0.5NiO4 ceramics. J. Appl. Phys. 2012, 111, 054106. [Google Scholar] [CrossRef]
  90. Moualhi, Y.; Nofal, M.M.; M’nassri, R.; Rahmouni, H.; Selmi, A.; Gassoumi, M.; Khirouni, K.; Cheikrouhou, A. Double Jonscher response and contribution of multiple mechanisms in electrical conductivity processes of Fe-PrCaMnO ceramic. Ceram. Int. 2020, 46, 1601–1608. [Google Scholar] [CrossRef]
  91. Moualhi, Y.; M’nassri, R.; Nofal, M.M.; Rahmouni, H.; Selmi, A.; Gassoumi, M.; Chniba-Boudjada, N.; Khirouni, K.; Cheikhrouhou, A. Magnetic properties and impedance spectroscopic analysis in Pr0.7Ca0.3Mn0.95Fe0.05O3 perovskite ceramic. J. Mater. Sci. Mater. Electron. 2020, 31, 21046–21058. [Google Scholar] [CrossRef]
  92. Lahouli, R.; Massoudi, J.; Smari, M.; Rahmouni, H.; Khirouni, K.; Dhahri, E.; Bessais, L. Investigation of annealing effects on the physical properties of Ni0.6Zn0.4Fe1.5Al0.5O4 ferrite. RSC Adv. 2019, 9, 19949–19964. [Google Scholar] [CrossRef] [Green Version]
Figure 1. X-ray diffraction patterns and Rietveld refinement of the La0.9Sr0.1MnO3 system sintered at 600 (a), 800 (b), and 1000 °C (c). TEM micrographs of the sintered samples (TS = 600 °C (d), 800 °C (e), and 1000 °C (f)) for the La0.9Sr0.1MnO3 system.
Figure 1. X-ray diffraction patterns and Rietveld refinement of the La0.9Sr0.1MnO3 system sintered at 600 (a), 800 (b), and 1000 °C (c). TEM micrographs of the sintered samples (TS = 600 °C (d), 800 °C (e), and 1000 °C (f)) for the La0.9Sr0.1MnO3 system.
Catalysts 12 00340 g001aCatalysts 12 00340 g001b
Figure 2. Temperature dependence of the electrical DC-resistivity for the La0.9Sr0.1MnO3 compound at different sintering temperatures (TS = 600 °C (a), 800 °C (b), and 1000 °C (c)). Evolution of TM-SC with sintering temperature TS (d).
Figure 2. Temperature dependence of the electrical DC-resistivity for the La0.9Sr0.1MnO3 compound at different sintering temperatures (TS = 600 °C (a), 800 °C (b), and 1000 °C (c)). Evolution of TM-SC with sintering temperature TS (d).
Catalysts 12 00340 g002
Figure 3. Frequency dependence of the normalized imaginary part of the impedance (Z″) at different temperatures for La0.9Sr0.1MnO3 compound sintered at 600 °C (a), 800 °C (b), and 1000 °C (c). Plot of ln(τ) versus the inverse of temperature (1/kB ×T) of the La0.9Sr0.1MnO3 compound at different sintering temperatures (d).
Figure 3. Frequency dependence of the normalized imaginary part of the impedance (Z″) at different temperatures for La0.9Sr0.1MnO3 compound sintered at 600 °C (a), 800 °C (b), and 1000 °C (c). Plot of ln(τ) versus the inverse of temperature (1/kB ×T) of the La0.9Sr0.1MnO3 compound at different sintering temperatures (d).
Catalysts 12 00340 g003aCatalysts 12 00340 g003b
Figure 4. The Nyquist diagrams of La0.9Sr0.1MnO3 system sintered at 600 °C (a), 800 °C (b), and 1000 °C (c).
Figure 4. The Nyquist diagrams of La0.9Sr0.1MnO3 system sintered at 600 °C (a), 800 °C (b), and 1000 °C (c).
Catalysts 12 00340 g004aCatalysts 12 00340 g004b
Figure 5. Temperature dependence of the grain boundary resistance for the La0.9Sr0.1MnO3 compound at different sintering temperatures (a). Evolution of TM-SC with sintering temperature TS (b). Temperature dependence of the electrical resistivity for the sample sintered at 600 °C performed at different frequencies (c). Plot of ln(Rgb/T) against (1/kB × T) at different sintering temperatures (d).
Figure 5. Temperature dependence of the grain boundary resistance for the La0.9Sr0.1MnO3 compound at different sintering temperatures (a). Evolution of TM-SC with sintering temperature TS (b). Temperature dependence of the electrical resistivity for the sample sintered at 600 °C performed at different frequencies (c). Plot of ln(Rgb/T) against (1/kB × T) at different sintering temperatures (d).
Catalysts 12 00340 g005
Table 1. Structural parameters of the La0.9Sr0.1MnO3 system sintered at 600, 800, and 1000 °C.
Table 1. Structural parameters of the La0.9Sr0.1MnO3 system sintered at 600, 800, and 1000 °C.
TS (°C)6008001000
a(Å)5.50(3)5.51(1)5.51(8)
c(Å)13.35(0)13.34(6)13.34(7)
v(Å3)350.14(7)351.11(4)351.92(8)
d(XRD) (nm)457585
d(TEM) (nm)437284
Table 2. Fitted values of grains, grain boundaries, and electrodes resistances extracted from the electrical equivalent circuit model.
Table 2. Fitted values of grains, grain boundaries, and electrodes resistances extracted from the electrical equivalent circuit model.
TS (°C)6008001000
Resistance (Ω)
T (K)RgRgbReRgRgbReRgRgbRe
80632390272682483541551150
1208030654738537049221800300
17080315057012061777293450460
180125300212012163377293587470
200105215145511058263273415400
21088180126810859266224057505
220741509649555063233985489
230651207809049256213685480
240631026167941243233000160
28037452575221422131717151
30029301805014819131390130
34065207161987111269784
380701226196161359055
400701234176551456822
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hizi, W.; Rahmouni, H.; Gorji, N.E.; Guesmi, A.; Ben Hamadi, N.; Khezami, L.; Dhahri, E.; Khirouni, K.; Gassoumi, M. Impact of Sintering Temperature on the Electrical Properties of La0.9Sr0.1MnO3 Manganite. Catalysts 2022, 12, 340. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12030340

AMA Style

Hizi W, Rahmouni H, Gorji NE, Guesmi A, Ben Hamadi N, Khezami L, Dhahri E, Khirouni K, Gassoumi M. Impact of Sintering Temperature on the Electrical Properties of La0.9Sr0.1MnO3 Manganite. Catalysts. 2022; 12(3):340. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12030340

Chicago/Turabian Style

Hizi, Wided, Hedi Rahmouni, Nima E. Gorji, Ahlem Guesmi, Naoufel Ben Hamadi, Lotfi Khezami, Essebti Dhahri, Kamel Khirouni, and Malek Gassoumi. 2022. "Impact of Sintering Temperature on the Electrical Properties of La0.9Sr0.1MnO3 Manganite" Catalysts 12, no. 3: 340. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12030340

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