Next Article in Journal
Plasma-Enhanced Atomic Layer Deposition of Zirconium Oxide Thin Films and Its Application to Solid Oxide Fuel Cells
Next Article in Special Issue
Microstructural, Phase Formation, and Superconducting Properties of Bulk YBa2Cu3Oy Superconductors Grown by Infiltration Growth Process Utilizing the YBa2Cu3Oy + ErBa2Cu3Oy + Ba3Cu5O8 as a Liquid Source
Previous Article in Journal
Control of the Pore Structure of Plasma-Sprayed Thermal Barrier Coatings through the Addition of Unmelted Porous YSZ Particles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Physical Properties of Submicron and Nano-Grained La0.7Sr0.3MnO3 and Nd0.7Sr0.3MnO3 Synthesised by Sol–Gel and Solid-State Reaction Methods

by
Lik Nguong Lau
1,
Kean Pah Lim
1,*,
Amirah Natasha Ishak
1,
Mohd Mustafa Awang Kechik
1,
Soo Kien Chen
1,
Noor Baa’yah Ibrahim
2,
Muralidhar Miryala
3,
Masato Murakami
3 and
Abdul Halim Shaari
1
1
Superconductor and Thin Film Laboratory, Department of Physics, Faculty of Science, Universiti Putra Malaysia, Serdang 43400 UPM, Selangor Darul Ehsan, Malaysia
2
Department of Applied Physics, Faculty of Science and Technology, Universiti Kebangsaan Malaysia, Bangi 43600 UKM, Selangor Darul Ehsan, Malaysia
3
Shibaura Institute of Technology, 3 Chome-7-5 Toyosu, Koto, Tokyo 135-8548, Japan
*
Author to whom correspondence should be addressed.
Coatings 2021, 11(3), 361; https://doi.org/10.3390/coatings11030361
Submission received: 5 March 2021 / Revised: 16 March 2021 / Accepted: 18 March 2021 / Published: 22 March 2021
(This article belongs to the Special Issue New Advance in Superconductor and Superconducting Thin Films)

Abstract

:
La0.7Sr0.3MnO3 (LSMO) and Nd0.7Sr0.3MnO3 (NSMO) possess excellent colossal magnetoresistance (CMR). However, research work on the neodymium-based system is limited to date. A comparative study between LSMO and NSMO prepared by sol–gel and solid-state reaction methods was undertaken to assess their structural, microstructural, magnetic, electrical, and magneto-transport properties. X-ray diffraction and structure refinement showed the formation of a single-phase composition. Sol–gel-synthesised NSMO was revealed to be a sample with single crystallite grains and exhibited intriguing magnetic and electrical transport behaviours. Magnetic characterisation highlighted that Curie temperature (TC) decreases with the grain size. Strong suppression of the metal–insulator transition temperature (TMI) was observed and attributed to the magnetically disordered grain surface and distortion of the MnO6 octahedra. The electrical resistivity in the metallic region was fitted with theoretical models, and the conduction mechanism could be explained by the grain/domain boundary, electron–electron, and electron–magnon scattering process. The increase in the scattering process was ascribed to the morphology changes. Enhancement of low-field magnetoresistance (LFMR) was observed in nano-grained samples. The obtained results show that the grain size and its distribution, as well as the crystallite formation, strongly affect the physical properties of hole-doped manganites.

1. Introduction

Research on perovskite manganites has been carried out since the 1950s and almost reached a saturated point thereafter. Nevertheless, the discovery of the colossal magnetoresistance (CMR) of this material has once again positioned it as an excellent candidate for memory recording applications and magnetic field sensors. The hole-doped perovskite manganites with the general formula of RE1−xAxMnO3, where RE is a rare earth ion (RE = La, Nd, Pr) and A is a divalent alkaline earth metal ion (B = Ba, Sr, Ca), have been investigated extensively due to the correlation between spin, charge, orbital, and lattice degrees of freedom [1,2,3]. Besides CMR, doped manganites have been discovered with some other intriguing properties, such as ferromagnetic–paramagnetic transition, metal–insulator transition, magnetocaloric effect (MCE), and charge order state [4,5,6,7,8]. A double exchange (DE) mechanism between pairs of Mn3+ and Mn4+ has been proposed to understand the interactions of charge and spin in manganites [9]. However, the CMR phenomenon cannot be explained alone by the DE mechanism. Some other interactions should also be considered, such as the Jahn–Teller (JT) polaron and phase separation. The CMR effects can be grouped into two categories, namely intrinsic and extrinsic MR [10]. The intrinsic MR is only expressive at a high magnetic field, while the latter is significantly greater in low magnetic fields (<0.2 T) [11]. The properties of hole-doped manganites can be controlled through the choice of dopants and their quantity. The doping concentration of divalent alkaline earth cation is normally selected in the range of 0.3 ≤ x ≤ 0.4, as this range lead to a high value of Curie temperature (TC) and is more appealing for modern devices. Besides the cation ratio, the oxygen stoichiometric is also crucial in deciding the electrical and magnetic properties of manganites [12,13,14].
At present, there are several well-established methods, such as solid-state reaction [15,16,17], co-precipitation [18], sol–gel [17,19,20,21], microemulsion [22], pyrophoric reaction [23,24,25], hydrothermal [26], and combustion [27,28], that have been utilised to synthesise different grain distributions in manganites. The solid-state reaction method is commonly applied to prepare the manganites. The mixing of starting materials is done by a grinding or ball-milling technique. This process is essential prior to the heat treatment for obtaining a homogeneous composition. The solid-state reaction can be done without a sophisticated setup but it is not efficient. It requires multiple repetitions of the milling and heating cycles to obtain powders with the desired phase and chemical compositions. Therefore, another low-cost method (sol–gel) has been employed by researchers to tackle this shortcoming. Sol–gel synthesis is a bottom-up approach that allows the precursors to mix in liquid medium at the molecular level, thereby resulting in lower processing temperature compared to the solid-state reaction method. Furthermore, the sol–gel method also permits better stoichiometry, shorter heating time, and smaller particle size [29,30].
Ezaami et al. studied the critical behaviour of La0.6Ca0.2Sr0.1MnO3 samples synthesised by sol–gel and solid-state reaction, and the particle variation has been identified as the leading cause of the property changes [31]. Reduction in the particle size was also discovered by Arun et al. when they prepared Nd0.67Sr0.33MnO3 samples by different synthesis routes (sol–gel and solid-state reaction) [17]. They suggested that the particle size reduction had resulted in the increase in effective grain boundaries and surface over volume ratio, as well as the formation of disorders (structural and magnetic) at the grain surfaces. According to the literature, the physical properties, such as magnetic and electrical behaviours, of manganites strongly depend on the particle size. Besides the synthesis route, particle size can also be tuned by varying the sintering temperatures, as demonstrated by Ng et al. [20]. The grain size of Pr0.67Sr0.33MnO3 samples was in the range of 46–245 nm when sintered at different temperatures (600 to 1000 °C). Both grain and crystallite sizes were observed to increase with the sintering temperature due to the congregation effect. Moreover, particle variation could be accomplished by employing different water-to-surfactant ratios in the microemulsion method [22]. Hintze et al. reported that the TC of La1−xSrxMnO3 was reduced significantly (22%) when particle size decreased to ~25 nm with respect to bulk crystals. Due to the unique physical properties of the hole-doped manganites, nano-sized particles often possess physical properties that are dramatically different from the corresponding bulk properties. Hence, manipulation of the particle size of manganites is expected to give rise to rich functionalities.
The large bandwidth manganite La0.7Sr0.3MnO3 (LSMO) has been extensively studied among the hole-doped perovskite manganites because it exhibits high values of MR (%) and TC [32,33]. To date, the study of the neodymium-based system is still limited and less attention has been paid to Nd0.7Sr0.3MnO3 (NSMO), though it is excellent in terms of CMR, accompanied by a TC close to room temperature [34]. Both LSMO and NSMO present a rich phase diagram in the manganite family but there are no studies that report and compare both compounds in detail. Consequently, this paper aims to investigate the structural, microstructural, magnetic, electrical, and magneto-transport properties of LSMO and NSMO. To produce different grain and crystallite sizes, manganite samples in this work were synthesised by sol–gel and solid-state reaction methods.

2. Materials and Methods

The LSMO and NSMO samples were prepared using sol–gel (SG) and solid-state reaction (SS) methods to modify the particle size. The samples will hereafter be denoted as LSSG, LSSS, NSSG, and NSSS. A similar sol–gel preparation procedure has been reported in our previous works [20,35]. The stoichiometric amounts of La (NO3)3·6H2O (Alfa Aesar, Heysham, UK; 99.99%), N2O6Sr (Alfa Aesar, Heysham, UK; 99.97%), and Mn (NO3)2·4H2O (Sigma-Aldrich, St. Louis, MO, USA; ≥97%) were used as the precursors to synthesise LSSG. Lanthanum (III) nitrate hexahydrate was replaced with Nd (NO3)3·6H2O (Sigma-Aldrich St. Louis, MO, USA; 99.9%) to prepare NSSG. LSSG and NSSG were sintered at 700 °C in air atmosphere to obtain the nanocrystalline compounds. On the other hand, the starting materials for the solid-state reaction method were La2O3 (Alfa Aesar, Heysham, UK; 99.9%), Nd2O3 (Sigma-Aldrich, St. Louis, MO, USA; 99.9%), SrCO3 (Sigma-Aldrich, St. Louis, MO, USA; ≥99.9%), and MnCO3 (Sigma-Aldrich, St. Louis, MO, USA; ≥99.9%). The stoichiometric ratios of related materials were well mixed by the ball-milling technique with analytical-grade acetone for 24 h before the drying process in the oven at 70 °C (24 h). The dried powder was later calcined at 900 °C for 12 h. Lastly, the powder was pelletised and sintered at 1100 °C for 24 h to produce LSSS and NSSS. Both heat treatment processes were carried out in an air atmosphere.
The structural properties of the samples were characterised by an X-ray diffractometer (XRD, X’Pert Pro PW 3040, Malvern Panalytical, Malvern, UK) and their data were analysed by HighScore Plus software. Field emission scanning electron microscope (FESEM, JEOL 7600 F, JEOL, Tokyo, Japan) and transmission electron microscope (TEM, Talos L120C, Thermo Fisher Scientific, Waltham, MA, USA) were utilised to examine the sample’s morphology and grain size. The magnetic properties were studied using a vibrating sample magnetometer (VSM, Lakeshore 7407, Lake Shore Cryotronics, Westerviller, OH, USA) and AC susceptometer (ACS, CryonBIND T, CryoBIND, Zagred, Croatia). AC susceptibility was measured in a magnetic field of 5 Oe at 219 Hz. The temperature dependence of the electrical resistivity and magneto-transport (variable magnetic field up to 1 T) were assessed by standard four-point probe method using Hall effect measurement system (HMS, Lakeshore 7604, Lake Shore Cryotronics, Westerviller, OH, USA) in the temperature range of 80–300 K.

3. Results and Discussion

LSMO and NSMO samples were characterised by X-ray diffraction (XRD) analysis using CuKα radiation at room temperature (300 K). Figure 1 shows the XRD patterns along with Rietveld refinement of LSSG, LSSS, NSSG, and NSSS. All samples were fully crystallised in single-phase by referring to the indexed peaks. LSSG and LSSS possessed a hexagonal crystal structure with the space group of R-3c (167) (ICSD reference code: 98-004-9739). Meanwhile, both NSSG and NSSS samples displayed a strong orientation towards the (121) direction with an orthorhombic crystal structure (ICSD reference code: 98-003-7439).
Goldschmidt’s tolerance factor, tG, is an indicator of the formation of perovskite structures and researchers have been using it to verify the stability and distortion of manganite compounds [7,36,37]. The tolerance factor is defined as:
t G = r A + r O 2 ( r B + r O )
where rA, rB, and rO are the radii of A, B, and O sites in the perovskite structure (ABO3). Hole-doped manganites have a stable perovskite structure if their tG is in the range of 0.89–1.02 [37]. The radii of La3+, Nd3+, Sr2+, Mn3+, Mn4+, and O2− are 1.30, 1.25, 1.40, 0.72, 0.67, and 1.28 Å, respectively. The tolerance factors of LSMO (0.930) and NSMO (0.917) perovskite compounds in this work are within the acceptable range and confirmed the stability of the perovskite structures.
The Rietveld refinement was carried out on the XRD patterns to determine the structural parameters. All diffraction patterns displayed good refinement, fitting with the calculated data by showing the low residual value of the weighted pattern, RWP, and goodness of fit, χ2, as shown in Figure 1. The lattice parameters, bond angles, and lengths of all samples are listed in Table 1. It can be observed that the unit cell volume of LSMO is greater than NSMO. This increase can be linked to the larger ionic radius of La3+ (1.30 Å) compared to Nd3+ (1.25 Å). Moreover, changes were observed in Mn–O bond length and Mn–O–Mn bond angle when the samples were prepared by different methods. These changes indicate the distortion of the MnO6 octahedra and could hamper the DE mechanism as well as the electrical behaviour.
From Figure 1, the peak broadening can be observed in the samples synthesised by the sol–gel method, which suggests the larger value of the full width at half maximum (FWHM). The FWHM and average crystallite size of all samples are summarised in Table 2. The average crystallite size, D, was calculated by Scherrer’s equation [38]:
D = 0.9 λ β cos θ
where β = βsample − βinstrumental, β is the line broadening at half of the maximum intensity (FWHM), λ is the X-ray wavelength (1.5406 Å), and θ is the position of the most intense diffraction peak. NSSG obtained the highest FWHM value and has the smallest crystallite size (21.6 nm). Another sample (LSSG) prepared by the sol–gel method also exhibits much smaller crystallite size (22.6 nm) as compared to the samples synthesised by the solid-state reaction method (LSSS and NSSS). The peak broadening (larger FWHM) is ascribed to the non-uniform lattice strain that comes from the nanocrystalline nature of the sample [39].
Figure 2 depicts the grain size distribution histograms and FESEM micrographs of the internal section for LSMO and NSMO samples. All samples showed a uniform grain size distribution and were composed of grains with an irregular shape. As the solid-state reaction samples were sintered at a higher temperature (1100 °C), LSSS and NSSS possessed a more compact and dense morphology compared to the sol–gel samples (LSSG and NSSG) sintered at lower temperatures. It was found that the grain size of samples prepared by the sol–gel method was in the nanometre range (26–45 nm), and the solid-state reaction method produced grains in submicron size (608–700 nm). These outcomes are in good agreement with the reported works [31,40].
TEM analysis was carried out to further validate the formation of the nanostructured particles in NSSG as it had the smallest grain size (26.46 nm) among all samples. Figure 3 reveals that the grain size of NSSG was in the range of 12–36 nm, with 23.22 nm as the mean grain size. The histogram shown in Figure 3 reflects the local grain distribution and it is comparable to the corresponding FESEM result. The selected area electron diffraction pattern (SAED) shown in Figure 3 confirms the polycrystalline nature of NSSG. The plane orientations such as (101), (121), and (022) were identified to corroborate with the XRD peaks. Furthermore, the lattice fringe observed was applied to the (101) plane, which had a d-spacing of 0.390 nm. A grain normally consists of several crystallites that are arranged in a similar orientation. The grain and crystallite sizes will be increased with the sintering temperature due to the congregation effect [20]. However, single-crystallite grains can be formed when samples are sintered at low temperatures [35] and this was witnessed in our NSSG sample. Furthermore, LSSG was found to have two crystallites in a single grain when referring to the data given in Table 2. As a result, manganite compounds with single (NSSG), double (LSSG), and multiple crystallites (LSSS and NSSS) in a single grain were synthesised and observed in this study.
The magnetic behaviour behind the divalent/monovalent ion-doped perovskite manganites is the existence of Mn3+/Mn4+ pair that leads to magnetic coupling as well as the DE interaction [41]. Vibrating sample magnetometer (VSM) measurement was carried out to characterise the magnetic properties, and the field dependence magnetisation measurement (M-H curve) at room temperature (300 K) for LSSS and NSSS is shown in Figure 4. A soft ferromagnetic hysteresis loop was observed on the LSSS, which indicates that the Curie temperature (TC) of LSMO was above room temperature. Its TC was reported in the range of 350–370 K by Souza et al., who studied the particle reduction (bulk to ~20 nm) of La0.7Sr0.3MnO3 [33]. Nevertheless, NSSS behaves as a paramagnetic curve in the M-H graph. The TC was verified by AC susceptometer measurement in the range of 80–320 K. As the temperature increases, the magnitude of susceptibility demonstrates a rapid drop until it reaches a minimum value (close to 0), and this transition is called ferromagnetic (FM)–paramagnetic (PM) transition. The Tc was identified from the inflection point of dχ’/dT versus temperature plot. There was no drop in AC susceptibility in the LSSG and LSSS measurements (Figure 4), with an χ’ value that was much higher than 0, which further confirms the ferromagnetic behaviour of LSSS observed in VSM analysis. Therefore, the transition of LSMO compounds was clearly above room temperature, which was beyond the range of our ACS measurement. For NSSG and NSSS, the Tc was observed at 255 and 270 K, respectively, as depicted in Figure 4. In Table 2, the results show that TC decreases as the grain size decreases. This behaviour can be attributed to the loss of long-range ferromagnetic ordering and the presence of magnetically disordered grain boundary layers in sol–gel-synthesised samples [20,40,42]. The existence of this layer can be explained by the core-shell model. It is also known as a magnetically dead layer, where the magnetic interactions are modified by the defects, vacancies, stress, and broken bonds, whereas the core is dominated by the DE interaction and promotes the ferromagnetic interaction like a bulk counterpart [43,44]. The thickness/portion of the dead layer with random spin orientation increases with the decrease in grain size. Hence, the NSSG, which is smaller in grain size, has lower TC compared to NSSS. Furthermore, the reduction in TC was also correlated with the changes in the structural parameters of NSMO samples as listed in Table 1. It can be observed that the increase in Mn–O bond length and the decrease in Mn–O–Mn bond angle weakened the DE coupling and hence reduced the TC [22]. The observed broader dχ’/dT peak minimum in NSSS is the result of the broad grain size distribution, as shown by the corresponding FESEM analysis [36]. In other words, the width of the FM–PM transition relies on the grain size distribution. The same outcome was also reported by Ng et al. when investigating the grain size reduction of Pr0.67Sr0.33MnO3 [20]. In conclusion, we can deduce that the TC and FM–PM transition are influenced by the grain distribution and its morphological properties.
The electrical transport behaviour of mixed-valence manganites is another fundamental characteristic after magnetism, as these two properties are essential in determining a high-efficiency, workable device. The electrical resistivity is contributed by the grains and inter-grain boundaries as they act as the scattering region for the electron transport [45,46,47,48]. Figure 5 depicts the resistivity as a function of temperature (ρ vs T) for all samples under a magnetic field of 0 and 1 T. The resistivity displayed by NSMO is much higher because it is an intermediate-bandwidth manganite and possesses lower electron mobility compared to LSMO [28]. In addition, there is a noticeable difference in resistivity between the samples produced by sol–gel and solid-state reaction methods. The nanocrystalline samples achieved resistivity in the order of a few magnitudes higher than the sub-micron-sized samples. This significant rise can be attributed to the drastic increase in effective grain boundaries in the samples with nano-sized grains. As reported by Arun et al., the potential barrier developed at the grain boundary layers localised the eg electrons at Mn atom and restricted their flow in the DE mechanism [40]. Besides the influence of the grain boundaries, the loosely packed grains (high porosity) in nanocrystalline samples also notably reduced the electron mobility, where the electron was insufficient to move through the grains.
Most of the perovskite manganites are paramagnetic insulators at room temperature and display an increase in resistivity with a decrease in temperature [7,20,40,46]. The increase in resistivity will reach a peak maximum known as metal–insulator transition temperature (TMI). The TMI was recorded by LSSG, NSSG, and NSSS in this study. The resistivity curves (80–300 K) shown by the LSMO samples agree with the previous study [42] indicating that the nanocrystalline samples (LSSG) exhibit metal–insulator transition below room temperature, which is absent in the bulk compound (LSSS). The strong suppression of the TMI was observed in NSMO as the particle was reduced to nano-size. This could be due to the enhanced scattering of the charge carriers through the higher density of magnetic disorder in effective grain boundary layers for nano-sized samples. Enhancement of grain boundaries in nano-grained samples also reduces the DE interaction between Mn3+ and Mn4+ and leads to an increase in resistivity [17]. In addition, the increase in Mn–O bond length and the decrease in Mn–O–Mn bond angle in the NSSG sample caused a decrease in bandwidth and electron mobility, thereby resulting in a decline in the DE interaction [49,50]. TMI was observed to shift towards high temperatures under the applied magnetic field, as depicted in Figure 5. This is because the magnetic field delocalised the charge carriers (eg electrons at Mn atom) and aligned the magnetic spins in manganite compounds. Hence, the resistivity is decreased when electrons move through DE coupling, resulting in a shift in TMI.
To understand the nature of different scattering mechanisms responsible for the conduction in the metallic region (T < TMI), the electrical resistivity results were fitted with the equation below [42,51,52,53]:
ρ ( T ) = ρ 0 + ρ 2 T 2 + ρ 4.5 T 4.5
where ρ 0 , ρ 2 T 2 , and ρ 4.5 T 4.5 represent the grain/domain boundary effect, electron–electron scattering, and electron–magnon scattering process in the ferromagnetic region, respectively. The experimental data were fitted to the theoretical models (solid lines) as shown in Figure 5. The good match between them is evidenced by the obtained regression coefficient, R2 values (close to 100%). The obtained parameters for all the samples are given in Table 3. The residual resistivity due to the grain/domain boundary ( ρ 0 ) is greatest among all parameters, which indicates that it plays a dominant role in the metallic region conduction. It was observed that ρ 0 decreases under an applied magnetic field and this can be attributed to the increase in domain, size thus reducing the value of ρ 0 [20]. On the other hand, the electron–magnon scattering ( ρ 4.5 ) interaction was small in our study as it originated from the induced spin inhomogeneity [51]. As shown by the tolerance factor, LSMO and NSMO compounds are stable perovskite structures and exhibit less electron spin fluctuation. As seen in Table 3, the fitting parameters ρ 0 and ρ 2 decrease with the increase in particle size and this is associated with the decrease in the grain boundaries [52]. The parameters obtained by NSSG are much higher compared to LSSG, evidencing the increase in Mn–O bond length and the decrease in Mn–O–Mn bond angle in NSSG. Concurrently, this results in an increase in the scattering process and reduces the strength of the DE mechanism.
Figure 6 illustrates the magnetoresistance, MR, as a function of the external applied magnetic field up to 1 T in the variation of temperature for LSSG, LSSS, NSSG, and NSSS. The MR can be calculated using the equation below:
MR ( % ) = ρ H ρ O ρ O × 100
where ρ O and ρ H are the resistivities without and with an applied magnetic field, respectively. It can be observed that the MR curves show two distinct regions as the magnetic field increases. The plot at the lower magnetic field (<0.2 T), better known as low-field magnetoresistance (LFMR), exhibits a steeper slope and it is more pronounced at low temperatures. This effect is attributed to the spin-polarised tunnelling across the grain boundaries. The disorder spins in manganites will align when the magnetic field begins to apply and it leads to the enhancement of electron hopping, thus resulting in a steep decrease in resistivity in the low-field region [54,55]. The other part of the plot is intrinsic MR and it is the result of the suppression from spin fluctuations, where the spins are aligned parallel to the magnetic field and exhibit high MR values near the TMI or TC. The MR (1 T) at 80 K for the samples LSSG, LSSS, NSSG, and NSSS was observed as 21.0%, 19.2%, 25.2%, and 19.5%, respectively. Figure 7 depicts the temperature dependence of MR for manganite samples at 0.2 and 1 T. The MR of LSMO increases monotonically with the decrease in temperature and it is comparable with the results reported earlier [42,56]. Enhancement of LFMR was observed in LSSG and NSSG, as shown in Figure 7a. Moreover, NSSS recorded the highest MR (27.7%) near the TMI and this is attributed to the increase in disordered magnetic spin fluctuations at the high-temperature range [35]. Therefore, we can conclude that NSSS is dominated by intrinsic MR accompanied by weak extrinsic MR.
It is vital to understand the correlation between the microstructural, magnetic, electrical, and magneto-transport properties of manganite compounds. Therefore, the relationship between χ’, ρ, and MR as a function of temperature for NSMO samples is illustrated in Figure 8. The magnitude of susceptibility and resistivity have been normalised for ease of viewing. NSSS recorded TC and TMI values close to room temperature, signifying the strong correlation between magnetic and electrical properties. A significant drop in MR was also observed in the vicinity of TMI with the increase in the applied magnetic field. Nevertheless, this is not seen on the NSSG sample, where it is mainly attributed to its single crystallite grains and extrinsic MR arising from the spin-polarised tunnelling. The DE interaction is also greatly suppressed and caused a shift in TMI towards lower temperatures compared to NSSS. The values of TC and TMI are usually close to each other, as demonstrated by the bulk sample (NSSS), but these temperatures might behave differently in the nanocrystalline sample (NSSG). This is because the electrical properties greatly depend on the morphological properties (grain size and grain boundary) while the magnetic behaviour is a cumulative result of the intrinsic properties [28,35]. All characterisation results gathered here clearly indicate that LSMO and NSMO are promising candidates for magnetic field sensors and can be tuned easily by varying the particle size to suit different applications.

4. Conclusions

In summary, we have investigated the grain size effect on the structural, microstructural, magnetic, electrical, and magneto-transport properties of LSMO and NSMO. The XRD analysis showed that all samples are fully crystallised in single-phase. According to Rietveld refinement, the unit cell volume of LSMO is greater than NSMO. This is because the ionic radius of La3+ (1.30 Å) is larger than Nd3+ (1.25 Å). The differences in Mn–O bond length and Mn–O–Mn bond angle indicate the distortion of the MnO6 octahedra when the samples were prepared by different methods. All samples exhibited a uniform grain size distribution with irregularly shaped grains. NSSG had the smallest grain size (26.46 nm) and appeared to be a sample with single crystallite grains. The TC of LSMO was observed above room temperature, as evident from the magnetic studies. For NSSG and NSSS, the TC was observed at 255 and 270 K, respectively. From the results obtained by ACS, it can be deduced that the TC and FM–PM transition are influenced by the grain size and its distribution. The metal–insulator transition was recorded below 300 K for LSSG but it was absent in LSSS. The strong suppression of the TMI was observed in NSMO as the particle size declined. The MR of LSMO samples increased monotonically with the decrease in temperature. NSSS recorded the highest MR (27.7%) near the TMI, which could be attributed to the increase in disordered magnetic spin fluctuations in the high-temperature range. Overall, this work has shown that the physical properties of manganites are strongly influenced by their grain and crystallite sizes.

Author Contributions

Conceptualization, L.N.L. and K.P.L.; methodology, L.N.L. and K.P.L.; validation, L.N.L. and A.N.I.; formal analysis, L.N.L. and K.P.L.; investigation, L.N.L. and A.N.I.; resources, K.P.L., M.M.A.K., S.K.C., N.B.I., M.M. (Muralidhar Miryala), M.M. (Masato Murakami) and A.H.S.; writing—original draft preparation, L.N.L.; writing—review and editing, L.N.L. and K.P.L.; visualization, L.N.L. and K.P.L.; supervision, K.P.L., M.M.A.K., S.K.C., N.B.I., M.M. (Muralidhar Miryala), M.M. (Masato Murakami) and A.H.S.; project administration, K.P.L., M.M.A.K., S.K.C. and A.H.S.; funding acquisition, K.P.L., M.M.A.K., S.K.C., N.B.I. and A.H.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was fully funded and supported by the Ministry of Higher Education, Malaysia (MOHE), through the Fundamental Research Grant Scheme (FRGS/1/2019/STG07/UPM/02/4) and a Universiti Putra Malaysia (UPM) research Grant (GP-IPS/2018/9663900).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are openly available in Zenodo at http://doi.org/10.5281/zenodo.4621283 (accessed on 19 March 2021).

Acknowledgments

The authors are grateful to the support staff who assisted in the characterisation measurements and for the facilities provided by UPM. L.N. is thankful to the Shibaura Institute of Technology, SIT as the host of aPBL under the Sakura Science Plan and facilitated some of the sample characterisation in this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mishra, D.; Roul, B.; Singh, S.; Srinivasu, V. Possible observation of Griffith phase over large temperature range in plasma sintered La0.67Ca0.33MnO3. J. Magn. Magn. Mater. 2018, 448, 287–291. [Google Scholar] [CrossRef]
  2. Panchal, G.; Choudhary, R.J.; Phase, D.M. Magnetic properties of La0.7Sr0.3MnO3 film on ferroelectric BaTiO3 substrate. J. Magn. Magn. Mater. 2018, 448, 262–265. [Google Scholar] [CrossRef]
  3. Gong, J.; Zheng, D.; Li, D.; Jin, C.; Bai, H. Lattice distortion modified anisotropic magnetoresistance in epitaxial La0.67Sr0.33MnO3 thin films. J. Alloys Compd. 2018, 735, 1152–1157. [Google Scholar] [CrossRef]
  4. Jonker, G.; Van Santen, J. Ferromagnetic compounds of manganese with perovskite structure. Physica 1950, 16, 337–349. [Google Scholar] [CrossRef]
  5. Yang, H.; Hua, S.; Zhang, P.; Zhang, S.; Ge, H.; Pan, M. Effect of Fe substitution on charge order state and magnetocaloric effect in Pr0.5Sr0.5Mn1−xFexO3 (0.00 ≤ x ≤ 0.08) manganites. J. Alloys Compd. 2016, 680, 768–772. [Google Scholar] [CrossRef]
  6. Von Helmolt, R.; Wecker, J.; Holzapfel, B.; Schultz, L.; Samwer, K. Giant negative magnetoresistance in perovskitelike La2/3Ba1/3Mnx ferromagnetic films. Phys. Rev. Lett. 1993, 71, 2331. [Google Scholar] [CrossRef]
  7. Rozilah, R.; Ibrahim, N.; Yahya, A.K. Inducement of ferromagnetic–metallic phase and magnetoresistance behavior in charged ordered monovalent-doped Pr0.75Na0.25MnO3 manganite by Ni substitution. Solid State Sci. 2019, 87, 64–80. [Google Scholar] [CrossRef]
  8. Ahmed, A.; Mohamed, H.; Paixão, J.; Mohamed, S.A. Thermopower and magnetocaloric properties in NdSrMnO/CrO3 composites. J. Magn. Magn. Mater. 2018, 456, 217–222. [Google Scholar] [CrossRef]
  9. Zener, C. Interaction between the d-shells in the transition metals. II. Ferromagnetic compounds of manganese with perovskite structure. Phys. Rev. 1951, 82, 403. [Google Scholar] [CrossRef]
  10. Tang, T.; Tien, C.; Hou, B. Low-field magnetoresistance of Ag-substituted perovskite-type manganites. Phys. B Condens. Matter 2008, 403, 2111–2115. [Google Scholar] [CrossRef]
  11. Nagaev, E.L. Colossal-magnetoresistance materials: Manganites and conventional ferromagnetic semiconductors. Phys. Rep. 2001, 346, 387–531. [Google Scholar] [CrossRef]
  12. Trukhanov, S.V.; Trukhanov, A.V.; Vasil’ev, A.N.; Maignan, A.; Szymczak, H. Critical behavior of La0.825Sr0.175MnO2.912 anion-deficient manganite in the magnetic phase transition region. JETP Lett. 2007, 85, 507–512. [Google Scholar] [CrossRef]
  13. Trukhanov, S.V.; Trukhanov, A.V.; Szymczak, H.; Botez, C.E.; Adair, A. Magnetotransport properties and mechanism of the a-site ordering in the Nd–Ba optimal-doped manganites. J. Low Temp. Phys. 2007, 149, 185–199. [Google Scholar] [CrossRef]
  14. Trukhanov, S.V.; Khomchenko, V.A.; Lobanovski, L.S.; Bushinsky, M.V.; Karpinsky, D.V.; Fedotova, V.V.; Troyanchuk, I.O.; Trukhanov, A.V.; Stepin, S.G.; Szymczak, R.; et al. Crystal structure and magnetic properties of Ba-ordered manganites Ln0.70Ba0.30MnO3−δ (Ln = Pr, Nd). J. Exp. Theor. Phys. 2006, 103, 398–410. [Google Scholar] [CrossRef]
  15. Ahsan, M.Z.; Ahsan, P.M.A.; Islam, M.A.; Asif, F.C. Structural and electrical properties of copper doped lanthanum manganite NPs. Results Phys. 2019, 15, 102600. [Google Scholar] [CrossRef]
  16. Mohamed, Z.; Ibrahim, N.; Ghani, M.A.; Safian, S.D.; Mohamed, S.N. Structural and electrical transport properties of (La0.7−xYx)Ca0.3MnO3 manganites. Results Phys. 2019, 12, 861–866. [Google Scholar] [CrossRef]
  17. Arun, B.; Suneesh, M.V.; Vasundhara, M. Comparative study of magnetic ordering and electrical transport in bulk and nano-grained Nd0.67Sr0.33MnO3 manganites. J. Magn. Magn. Mater. 2016, 418, 265–272. [Google Scholar] [CrossRef]
  18. Abdel-Khalek, E.K.; Salem, A.F.; Mohamed, E.A. Study on the influence of magnetic phase transitions on the magnetocaloric effect in Sm0.7Sr0.3Mn0.95Fe0.05O3 manganite. J. Alloys Compd. 2014, 608, 180–184. [Google Scholar] [CrossRef]
  19. Xia, W.; Li, L.; Wu, H.; Xue, P.; Zhu, X. Structural, morphological, and magnetic properties of sol-gel derived La0.7Ca0.3MnO3 manganite nanoparticles. Ceram. Int. 2017, 43, 3274–3283. [Google Scholar] [CrossRef]
  20. Ng, S.; Lim, K.; Halim, S.; Jumiah, H. Grain size effect on the electrical and magneto-transport properties of nanosized Pr0.67Sr0.33MnO3. Results Phys. 2018, 9, 1192–1200. [Google Scholar] [CrossRef]
  21. Shokr, F.S.; Hussein, M. Physical properties of La0.67Ca0.13Sr0.2Mn1−xFexO3 compounds doped Fe. Results Phys. 2018, 8, 186–189. [Google Scholar] [CrossRef]
  22. Hintze, C.E.; Fuchs, D.; Merz, M.; Amari, H.; Kübel, C.; Huang, M.-J.; Powell, A.; v. Löhneysen, H. Size-induced changes of structural and ferromagnetic properties in La1−xSrxMnO3 nanoparticles. J. Appl. Phys. 2017, 121, 214303. [Google Scholar] [CrossRef]
  23. Paul, S.; Nath, T. Microstructural, electrical-, magneto-transport properties of grain size modulated nanocrystalline Nd0.67Sr0.33MnO3 CMR manganites. MRS Online Proc. Libr. 2008, 1118, 3–8. [Google Scholar] [CrossRef]
  24. Paul, S.; Nath, T.K. Magneto-transport and magnetic properties of Fe doped nanometric polycrystalline La0.7Sr0.3MnO3 CMR manganites. MRS Online Proc. Libr. (OPL) 2009, 1183. [Google Scholar] [CrossRef]
  25. Dey, P.; Nath, T. Tunable room temperature low-field spin polarized tunneling magnetoresistance of La0.7Sr0.3MnO3 nanoparticles. Appl. Phys. Lett. 2006, 89, 163102. [Google Scholar] [CrossRef]
  26. Ngida, R.E.A.; Zawrah, M.F.; Khattab, R.M.; Heikal, E. Hydrothermal synthesis, sintering and characterization of nano La-manganite perovskite doped with Ca or Sr. Ceram. Int. 2019, 45, 4894–4901. [Google Scholar] [CrossRef]
  27. Kumar, D.; Singh, A.K. Investigation of structural and magnetic properties of Nd0.7Ba0.3Mn1−xTixO3 (x = 0.05, 0.15 and 0.25) manganites synthesized through a single-step process. J. Magn. Magn. Mater. 2019, 469, 264–273. [Google Scholar] [CrossRef]
  28. Thombare, B.; Dusane, P.; Kekade, S.; Salunkhe, A.; Choudhary, R.J.; Phase, D.M.; Devan, R.S.; Patil, S.I. Influence of nano-dimensionality on magnetotransport, magnetic and electrical properties of Nd1−xSrxMnO3−δ (0.3 ≤ x ≤ 0.7). J. Alloys Compd. 2019, 770, 257–266. [Google Scholar] [CrossRef]
  29. Brinker, C.J.; Scherer, G.W. Sol-Gel Science: The Physics and Chemistry of Sol-Gel Processing; Academic Press: London, UK, 2013. [Google Scholar]
  30. Danks, A.E.; Hall, S.R.; Schnepp, Z. The evolution of ‘sol–gel’ chemistry as a technique for materials synthesis. Mater. Horiz. 2016, 3, 91–112. [Google Scholar] [CrossRef] [Green Version]
  31. Ezaami, A.; Sfifir, I.; Cheikhrouhou-Koubaa, W.; Koubaa, M.; Cheikhrouhou, A. Critical properties in La0.7Ca0.2Sr0.1MnO3 manganite: A comparison between sol-gel and solid state process. J. Alloys Compd. 2017, 693, 658–666. [Google Scholar] [CrossRef]
  32. Ren, Q.; Zhang, Y.; Chen, Y.; Wang, G.; Dong, X.; Tang, X. Structure and magnetic properties of La0.67Sr0.33MnO3 thin films prepared by sol–gel method. J. Sol-Gel Sci. Technol. 2013, 67, 170–174. [Google Scholar] [CrossRef]
  33. Souza, A.D.; Babu, P.; Rayaprol, S.; Murari, M.; Mendonca, L.D.; Daivajna, M. Size control on the magnetism of La0.7Sr0.3MnO3. J. Alloys Compd. 2019, 797, 874–882. [Google Scholar] [CrossRef]
  34. Feng, X.; Wen, H.; Shen, Y. Microwave absorption properties of Nd0.7Sr0.3MnO3 prepared using high-energy ball milling. J. Alloys Compd. 2013, 555, 145–149. [Google Scholar] [CrossRef]
  35. Lim, K.P.; Ng, S.W.; Lau, L.N.; Kechik, M.M.A.; Chen, S.K.; Halim, S.A. Unusual electrical behaviour in sol–gel-synthesised PKMO nano-sized manganite. Appl. Phys. A 2019, 125, 745. [Google Scholar] [CrossRef]
  36. Jeddi, M.; Gharsallah, H.; Bekri, M.; Dhahri, E.; Hlil, E. Structural characterization and ZFC/FC magnetization study of La0.6Ca0.4−xSrxMnO3 nanoparticle compounds. Appl. Phys. A 2020, 126, 6. [Google Scholar] [CrossRef]
  37. Bouzidi, S.; Gdaiem, M.A.; Rebaoui, S.; Dhahri, J.; Hlil, E. Large magnetocaloric effect in La0.75Ca0.25−xNaxMnO3 (0 ≤ x ≤ 0.10) manganites. Appl. Phys. A 2020, 126, 60. [Google Scholar] [CrossRef]
  38. Lau, L.N.; Ibrahim, N.B.; Baqiah, H. Influence of precursor concentration on the structural, optical and electrical properties of indium oxide thin film prepared by a sol–gel method. Appl. Surf. Sci. 2015, 345, 355–359. [Google Scholar] [CrossRef]
  39. Kurchania, R.; Rathore, D.; Pandey, R. Size dependent strain and nanomagnetism in CoFe2O4 nanoparticles. J. Mater. Sci. Mater. Electron. 2015, 26, 9355–9365. [Google Scholar] [CrossRef]
  40. Arun, B.; Akshay, V.R.; Chandrasekhar, K.D.; Mutta, G.R.; Vasundhara, M. Comparison of structural, magnetic and electrical transport behavior in bulk and nanocrystalline Nd-lacunar Nd0.67Sr0.33MnO3 manganites. J. Magn. Magn. Mater. 2019, 472, 74–85. [Google Scholar] [CrossRef]
  41. Verma, M.K.; Sharma, N.D.; Sharma, S.; Choudhary, N.; Singh, D. High magnetoresistance in La0.5Nd0.15Ca0.25A0.1MnO3 (A = Ca, Li, Na, K) CMR manganites: Correlation between their magnetic and electrical properties. Mater. Res. Bull. 2020, 125, 110813. [Google Scholar] [CrossRef]
  42. Navin, K.; Kurchania, R. The effect of particle size on structural, magnetic and transport properties of La0.7Sr0.3MnO3 nanoparticles. Ceram. Int. 2018, 44, 4973–4980. [Google Scholar] [CrossRef]
  43. 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]
  44. Giri, S.K.; Yusuf, S.M.; Mukadam, M.D.; Nath, T.K. Emergence of Griffiths phase and glassy mixed phase in Sm0.5Ca0.5MnO3 nanomanganites. J. Alloys Compd. 2014, 591, 181–187. [Google Scholar] [CrossRef]
  45. Mandal, S.; Nath, T.; Rao, V. Effect of nanometric grain size on electronic-transport, magneto-transport and magnetic properties of La0.7Ba0.3MnO3 nanoparticles. J. Phys. Condens. Matter 2008, 20, 385203. [Google Scholar] [CrossRef] [PubMed]
  46. Manh, D.H.; Phong, P.T.; Thanh, T.D.; Hong, L.V.; Phuc, N.X. Low-field magnetoresistance of La0.7Ca0.3MnO3 perovskite synthesized by reactive milling method. J. Alloys Compd. 2010, 499, 131–134. [Google Scholar] [CrossRef]
  47. Anshul, A.; Amritphale, S.S.; Kaur, S.; Hada, R. Nanomaterials synthesis and effect of granular disorders on electrical behavior of lanthanum manganites. J. Mater. Sci. Technol. 2011, 27, 691–695. [Google Scholar] [CrossRef]
  48. Bhatt, R.C.; Singh, S.K.; Srivastava, P.C.; Agarwal, S.K.; Awana, V.P.S. Impact of sintering temperature on room temperature magneto-resistive and magneto-caloric properties of Pr2/3Sr1/3MnO3. J. Alloys Compd. 2013, 580, 377–381. [Google Scholar] [CrossRef] [Green Version]
  49. Sakthipandi, K.; Arunachalam, M.; Thamilmaran, P.; Sivabharathy, M.; Sankarrajan, S. Influence of ionic radius on magnetic phase transition in R1−xSrxMnO3 perovskites. AIP Conf. Proc. 2017, 1832, 030016. [Google Scholar]
  50. Sakthipandi, K.; Rajendran, V.; Jayakumar, T. Phase transitions of bulk and nanocrystalline La1−xSrxMnO3 (x = 0.35 and 0.37) perovskite manganite materials using in situ ultrasonic studies. Mater. Res. Bull. 2013, 48, 1651–1659. [Google Scholar] [CrossRef]
  51. Li, Y.; Zhang, H.; Chen, Q.; Li, D.; Li, Z.; Zhang, Y. Effects of A-site cationic radius and cationic disorder on the electromagnetic properties of La0.7Ca0.3MnO3 ceramic with added Sr, Pb, and Ba. Ceram. Int. 2018, 44, 5378–5384. [Google Scholar] [CrossRef]
  52. Mohamed, H. Influence of sodium doping on the electrical and magnetic properties of La0.90Li0.10MnO3 manganites. J. Magn. Magn. Mater. 2017, 424, 44–52. [Google Scholar] [CrossRef]
  53. Venkataiah, G.; Venugopal Reddy, P. Electrical behavior of sol–gel prepared Nd0.67Sr0.33MnO3 manganite system. J. Magn. Magn. Mater. 2005, 285, 343–352. [Google Scholar] [CrossRef]
  54. Zhou, Y.; Zhu, X.; Li, S. Structure, magnetic, electrical transport and magnetoresistance properties of La0.67Sr0.33Mn1−xFexO3 (x = 0–0.15) doped manganite coatings. Ceram. Int. 2017, 43, 3679–3687. [Google Scholar] [CrossRef]
  55. Modi, A.; Bhat, M.A.; Pandey, D.K.; Tarachand; Bhattacharya, S.; Gaur, N.K.; Okram, G.S. Structural, magnetotransport and thermal properties of Sm substituted La0.7−xSmxBa0.3MnO3 (0 ≤ x ≤ 0.2) manganites. J. Magn. Magn. Mater. 2017, 424, 459–466. [Google Scholar] [CrossRef]
  56. Zhou, Y.; Zhu, X.; Li, S. Effect of particle size on magnetic and electric transport properties of La0.67Sr0.33MnO3 coatings. PCCP 2015, 17, 31161–31169. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Rietveld refinement of the XRD patterns from LSSG, LSSS, NSSG, and NSSS.
Figure 1. Rietveld refinement of the XRD patterns from LSSG, LSSS, NSSG, and NSSS.
Coatings 11 00361 g001
Figure 2. Grain size distribution histograms and FESEM micrographs of manganite samples.
Figure 2. Grain size distribution histograms and FESEM micrographs of manganite samples.
Coatings 11 00361 g002
Figure 3. TEM micrographs, grain size distribution histogram, and SAED of NSSG.
Figure 3. TEM micrographs, grain size distribution histogram, and SAED of NSSG.
Coatings 11 00361 g003
Figure 4. VSM measurement at room temperature and temperature dependence of AC susceptibility for LSMO and NSMO compounds.
Figure 4. VSM measurement at room temperature and temperature dependence of AC susceptibility for LSMO and NSMO compounds.
Coatings 11 00361 g004
Figure 5. Electrical resistivity as a function of temperature for LSMO and NSMO manganites under applied magnetic field of 0 and 1 T. The solid line represents the best fit of experimental data.
Figure 5. Electrical resistivity as a function of temperature for LSMO and NSMO manganites under applied magnetic field of 0 and 1 T. The solid line represents the best fit of experimental data.
Coatings 11 00361 g005
Figure 6. MR as a function of external applied magnetic field curves from 80–300 K for (a) LSSG, (b) LSSS, (c) NSSG, and (d) NSSS.
Figure 6. MR as a function of external applied magnetic field curves from 80–300 K for (a) LSSG, (b) LSSS, (c) NSSG, and (d) NSSS.
Coatings 11 00361 g006
Figure 7. Temperature dependence of MR for LSMO and NSMO manganite samples at (a) 0.2 T and (b) 1 T.
Figure 7. Temperature dependence of MR for LSMO and NSMO manganite samples at (a) 0.2 T and (b) 1 T.
Coatings 11 00361 g007
Figure 8. Relationship between χ’, ρ, and MR as a function of temperature for NSSG and NSSS.
Figure 8. Relationship between χ’, ρ, and MR as a function of temperature for NSSG and NSSS.
Coatings 11 00361 g008
Table 1. Tolerance factor and structural parameters obtained from the Rietveld refinement of the XRD patterns for LSSG, LSSS, NSSG, and NSSS.
Table 1. Tolerance factor and structural parameters obtained from the Rietveld refinement of the XRD patterns for LSSG, LSSS, NSSG, and NSSS.
Sample CodeLSSGLSSSNSSGNSSS
SampleLa0.7Sr0.3MnO3Nd0.7Sr0.3MnO3
ICSD Reference Code98-004-973998-003-7493
Crystal StructureHexagonalOrthorhombic
Space GroupR-3c (167)Pnma (62)
Tolerance Factor, tG0.9300.917
Lattice Parameter-
a (Å)5.4865.5055.4375.450
b (Å)5.4865.5057.6707.700
c (Å)13.38113.3665.4805.466
Volume (Å3)348.704350.757228.516229.398
Bond Angle and Length----
∠Mn-O1-Mn (°)166.496161.681159.880165.804
∠Mn-O2-Mn (°)--158.591160.057
Mn-O1 (Å)1.9501.8601.9671.943
Mn-O2 (Å)--1.9511.955
Table 2. FWHM, crystallite size, grain size, TC, TMI, and magnetoresistance (MR) at 1 T for LSMO and NSMO samples.
Table 2. FWHM, crystallite size, grain size, TC, TMI, and magnetoresistance (MR) at 1 T for LSMO and NSMO samples.
SampleFWHM (°)Average Crystallite Size D (nm)Average Grain Size GS (nm)Curie Temperature, TC (K)Metal–Insulator Temperature, TMI (K)MR at 1 T (%)
80 KTMI
LSSG0.41722.644.4>30023821.07.1
LSSS0.115127.4699.9>300>30019.2
NSSG0.43421.626.525515025.217.8
NSSS0.16373.3607.527024419.527.7
Table 3. Parameters obtained corresponding to the best fit of the experimental data based on Equation (3) from LSMO and NSMO compounds.
Table 3. Parameters obtained corresponding to the best fit of the experimental data based on Equation (3) from LSMO and NSMO compounds.
SampleH (T)ρο (Ω cm)ρ2 (Ω cm K−2)ρ4.5 (Ω cm K−4.5)R2
LSSG02.4637.12 × 10−5−4.00 × 10−10.999
11.8547.53× 10−5−3.89 × 10−10.999
LSSS04.79 × 10−31.91 × 10−7−5.47 × 10−10.999
13.77 × 10−31.95 × 10−7−5.21 × 10−10.999
NSSG018190.232−2.85 × 10−70.999
112930.192−2.08 × 10−70.999
NSSS08.30 × 10−28.18 × 10−63.79 × 10−120.999
15.33 × 10−28.43 × 10−61.89 × 10−120.999
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lau, L.N.; Lim, K.P.; Ishak, A.N.; Awang Kechik, M.M.; Chen, S.K.; Ibrahim, N.B.; Miryala, M.; Murakami, M.; Shaari, A.H. The Physical Properties of Submicron and Nano-Grained La0.7Sr0.3MnO3 and Nd0.7Sr0.3MnO3 Synthesised by Sol–Gel and Solid-State Reaction Methods. Coatings 2021, 11, 361. https://doi.org/10.3390/coatings11030361

AMA Style

Lau LN, Lim KP, Ishak AN, Awang Kechik MM, Chen SK, Ibrahim NB, Miryala M, Murakami M, Shaari AH. The Physical Properties of Submicron and Nano-Grained La0.7Sr0.3MnO3 and Nd0.7Sr0.3MnO3 Synthesised by Sol–Gel and Solid-State Reaction Methods. Coatings. 2021; 11(3):361. https://doi.org/10.3390/coatings11030361

Chicago/Turabian Style

Lau, Lik Nguong, Kean Pah Lim, Amirah Natasha Ishak, Mohd Mustafa Awang Kechik, Soo Kien Chen, Noor Baa’yah Ibrahim, Muralidhar Miryala, Masato Murakami, and Abdul Halim Shaari. 2021. "The Physical Properties of Submicron and Nano-Grained La0.7Sr0.3MnO3 and Nd0.7Sr0.3MnO3 Synthesised by Sol–Gel and Solid-State Reaction Methods" Coatings 11, no. 3: 361. https://doi.org/10.3390/coatings11030361

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