Javier A.
Cuervo Farfán
a,
Críspulo E.
Deluque Toro
b,
Carlos A.
Parra Vargas
c,
David A.
Landínez Téllez
a and
Jairo
Roa-Rojas
*a
aGrupo de Física de Nuevos Materiales, Departamento de Física, Universidad Nacional de Colombia, Bogotá, DC, Colombia. E-mail: jroar@unal.edu.co
bGrupo de Nuevos Materiales, Facultad de Ingeniería, Universidad del Magdalena, Santa Marta, Colombia
cGrupo Física de Materiales, Escuela de Física, Universidad Pedagógica y Tecnológica de Colombia, Tunja, Colombia
First published on 28th September 2020
Polycrystalline samples of Sm2Bi2Fe4O12 were produced by the standard solid state reaction method. The structural characterization revealed that the material crystallizes in an orthorhombic structure (Pnma # 62 space group). Scanning electron microscopic images showed granular and densified characteristics on the surface and inside the samples, with two well-differentiated grain sizes: micrometric and submicrometric. From energy-dispersive X-ray spectra, it was established that the samples contained the expected elements in stoichiometric proportions suggested by the chemical formula of the compound. I–V curves and measurements of electrical permittivity as a function of temperature showed semiconductor-type behaviors, strongly dependent on the polycrystalline character of the material and a Maxwell–Wagner-type dielectric tendency. The optical spectral analysis corroborated the semiconductor behavior with the optical bandgap Eg = 2.62 eV. Temperature-dependent magnetization results have the expected form for ferromagnetic-type materials, with evidence of disorder that gives rise to magnetic irreversibility between the zero-field cooling and field cooling measurement procedures. Magnetization curves as a function of the applied field showed a hysterical response, even at room temperature, revealing the occurrence of ferromagnetic ordering in the material. The ab initio calculations of the electron density of states reveal the appearance of a mean semiconductor band gap whose value is in agreement with that measured experimentally. The effective magnetic moment is attributed to the majority contributions of the 3d-Fe orbitals. At low temperatures, the behaviors of the specific heat at constant volume and pressure are similar, with a tendency of the former towards the Dulong–Petit limit. Strong changes in all the thermophysical properties studied are evident due to the effects of pressure and temperature on the wave behavior of the crystal lattice. At higher temperatures, a strong variation in the Debye temperature induces divergence in thermophysical parameters for different applied pressures. The physical parameters related to the coexistence of the semiconductor and ferromagnetic properties in the material suggest possible technological implications in the spintronics industry.
In order to study the structural, electrical, magnetic and electronic properties of this family of perovskites, in the present work, we propose the Sm2Bi2Fe4O12 novel material, which is the supercell coming from the SmBiFe2O6 double perovskite. The supercell is considered to minimize the partial occupations of Sm and Bi. Considered from simple perovskites, this compound could be a combination in identical proportions of the rare earth orthoferrite SmFeO3 and the bismuth ferrite BiFeO3. The first, which is classified in the family of lanthanide ferrites, also known as orthoferrites, tends to crystallize in a distorted orthorhombic perovskite structure, belonging to the Pbnm space group with a unit cell formed by R3+ cations that occupy the vertices of a straight prism (with rectangular faces), which in its centre contains a Fe3+ cation, octahedrally coordinated with 6 oxygen anions (octahedron FeO6). These compounds have been extensively studied due to their interesting electrical and magnetic properties,18 as well as by its optical19 and dielectric20 responses. However, BiFeO3 is recognized by its rhombohedral structure, R3c space group, and its multiferroic properties.21 It is the reason that bismuth ferrite has been much studied in recent years.22 In the literature, there are some reports regarding the simultaneous presence of Sm and Bi in the perovskite-type ferrite Bi1−xSmxFeO3. A first work presents a structural approach for x = 0.2, 0.4, 0.6, 0.8 and 1.0 in which the obtained diffraction patterns are compared with JCPDS data.23 The most recent report on Bi1−xSmxFeO3, related to the search for a causal analysis conducive to the optimization and design of ferroic materials, was made on thin films for x ≤ 0.2.24 Perhaps the most interesting work, including a complete structural analysis with Rietveld refinement, and characterization of the dielectric and magnetic responses, was done for the Bi1−xSmxFeO3 material with 0.05 ≤ x ≤ 0.2.25 That report cites several works on partial substitution of Sm and other rare earths at the Bi site, but the specific composition x = 0.5, corresponding to a double perovskite26 AA′B2O6, has not been reported. From the experimental point of view, this work aims to carefully analyse the crystal structure, as well as the morphological, magnetic, electrical, and optical properties of the complex perovskite Sm0.5Bi0.5FeO3, conveniently presented as Sm2Bi2Fe4O12, which demonstrate its ferromagnetic semiconductor behaviour as a first approximation to the possible technological implications in the spintronics industry.
However, it is advantageous that the calculations made from density functional theory (DFT) have been shown to be potentially powerful tools for predicting physical properties in perovskite-type materials.27 As an additional exercise, in this work, the results of ab initio calculations for the density of states of the Sm2Bi2Fe4O12 material are presented, duly analysed, and correlated with the experimental data. In addition, the macroscopic thermodynamic properties are strongly correlated with the microscopic dynamics of the material's atoms; therefore, considering that the collective vibrations of the crystalline networks in solids take place in discrete energy packages, known as phonons, it is possible to study the fundamental excitations that are associated with these thermodynamic properties. Similarly, the most representative function of phonons takes place in insulators and semiconductors, where they make direct contributions to properties such as specific heat and thermal expansion, indicating their dependence on temperature. Hence, it is expected that their study in the new Sm2Bi2Fe4O12 magnetic semiconductor makes contributions to the understanding of the mechanisms that give rise to their physical microscopic properties. Meanwhile, with reference to the phonons, it can be expected that the vibrations in the crystals have a harmonic character, which would be valid for temperature values lower than the Debye temperature of the solid. In this way, the theoretical methods can constitute a complementary tool for the study of atomic dynamics at relatively high temperatures through approaches such as Debye's quasi-harmonic model.28 The main reason for the application of this method is that its results have been shown to be surprisingly close to those reported experimental values for materials from the olivine family.29 Finally, in this work, the theoretical results of the electronic characteristics of the double perovskite Sm2Bi2Fe4O12 are reported, as a complementary method in search for information about the mechanisms responsible for the magnetic and transport properties evidenced by this material. Thus, the results are presented for the density of electronic states and the behaviour of specific heat, Debye's temperature, entropy, thermal expansion, and the Grüneisen parameter as functions of pressure and temperature.
![]() | ||
Fig. 1 XRD refined pattern for the Sm2Bi2Fe4O12 material. The short vertical lines at the bottom indicate the Bragg positions. |
From the Rietveld analysis it was obtained that the material crystallizes in a complex perovskite-type structure belonging to the Pnma space group (# 62), with octahedral distortions given by the notation a+b−b−, corresponding to the same symmetry of the lanthanide orthoferrite, but with differently defined axes. This is a primitive (P) structure with a glide plane (n) perpendicular to the x-axis, a mirror plane (m) perpendicular to the y-axis, and a glide plane (a) perpendicular to the z-axis. Although the Pnma space group is expected for simple perovskites, it is well known that double perovskites can crystallize in this group when the crystal radii of cations A and A′ have relatively close values.42 In this case the RSm and RBi radii, for coordination 6, are 1.10 Å and 1.17 Å respectively.43 The first important parameters to establish the degree of distortion of the crystalline structure of the complex perovskite with respect to the ideal cubic perovskite is the known tolerance factor τ that for the AA′B2O6 perovskite type can be calculated using the following equation:
![]() | (1) |
The cell parameters obtained from the refinement are presented in Table 1, where a, b and c represent the lattice parameters; x, y and z are the atomic positions of atoms in the unit cell; α, β and γ are the angles between the cell coordinates; χ2; R(F2); Rp and Rwp are the reliability parameters, which guarantee the quality of the refinement procedure. Lattice parameters suggest that this orthorhombic structure corresponds to an elongated cell with axes a and c whose values are close to each other compared to axis b, which are between 28% and 30% greater. Oxygen ions with O1 notation in Table 1 are found along the primitive b-axis of the structure, while O2 anions are located in the basal plane of the cell (ac-plane). As shown in Table 1, there are differences between Wyckoff positions for the same element, which constitutes the second factor indicative of a marked structural distortion.
Atom | Wyckoff site | Site symmetry | Atomic coordinates | Pnma space group | ||
---|---|---|---|---|---|---|
X | Y | z | Cell parameters | |||
Reliability factors: χ2 = 2.696, R(F2) = 4.79%, Rp = 2.65% and Rwp = 3.69%. XRD-density: 7.94 g cm−3. | ||||||
Sm, Bi | 4c | −1 | 0.5497(1) | 0.2500 | 0.0075(5) | a = 5.6197(1) Å |
Fe | 4a | −1 | 0.0000 | 0.0000 | 0.0000 | b = 7.7775(2) Å |
O1 | 4c | m | 0.9781(19) | 0.2500 | 0.8953(20) | c = 5.4266(1) Å |
O2 | 8d | 1 | 0.3056(18) | 0.0373(16) | 0.1984(19) | V = 237.18(3) Å3 |
The density reported in Table 1 corresponds to the mass of the cell divided by the volume that was calculated from the lattice parameters. In the notation of the table, a, c and d are the letters of Wyckoff, which determine all the x points for which the possible symmetry groups of each cell site are conjugated subgroups of the Pnma spatial group.44 The space group obtained is consistent with that reported in ref. 25 for substitutions greater than 12.5% of Sm instead of Bi. Meanwhile, the volume presented in Table 1 reveals that the cell is 0.8% more compact for our material with 50% Sm at the Bi site than that reported in ref. 25 for 20% Sm at the Bi crystallographic site. This result suggests that the unit cell of the Sm2Bi2Fe4O12 perovskite is more stable than in the case of substitutions that represent different percentages of Sm and Bi in the structure.
Fig. 2 is the graphical representation of the expected structure from the Rietveld analysis. In the figure, it is possible to observe the distortion deviation of the cell with respect to that expected for a cubic lattice, which gives rise to the obtained value of the tolerance factor. Moreover, the rotations and tilting of the octahedra due to the Glazer notation (a+b−b−) for this non-symmetric space group is another relevant source of distortion, as shown in Fig. 2a and b. However, Fig. 2c and d allow us to establish the differences between the octahedral distances of the O1–Fe–O2 bonds (1.99 Å and 2.05 Å) with each other and with respect to the O1–Fe–O1 (2.03 Å) bonds. Similarly, the octahedral bond angles Fe–O2–Fe and Fe–O1–Fe also evidence substantial differences with 150.77° and 146.71° values, respectively.
The valence of the bond has the property that its sum around each atom in a compound is equal to the oxidation state (valence) of that atom. The valence of the bond is equivalent to the number of bond electrons distributed within the bond. For the case of FeO6 octahedra in the material Sm2Bi2Fe4O12, the bond valence sum is 2.883, which is very close to the oxidation state of the cation (3+).
Fig. 3 exemplifies the surface characteristics of the material in micrographs produced by secondary and backscattered electrons (3a) and (3b), respectively, under the application of 10 kV and for a magnification of 20 kx. The figure shows a distribution of grains of various sizes and shapes, which make up very compact surface, with a relatively less porosity and very well-defined grain boundaries, that is, low intergranular diffusivity. Despite the variety in granular forms, no evidence of grains corresponding to impurities or precursor oxides that have not reacted is perceived, so it can be said that all the grains that appear in the image are part of the material.
The volumetric densification of the material plays an important role in the physical properties and the ferroic or magnetodielectric properties, on which many of the potential applications of the material depend. For this reason, the samples were fractured to analyse the granular morphology in the bulk of the material through SEM images such as those presented in Fig. 3c and d. This was observed that in the internal part of the samples the grains appear more diffused with each other and that the average size of the grains is dominated by the smaller grains, unlike the surface, where a majority of grains are observed of micrometric order. The internal grain size analysis of the samples carried out using the ImageJ software45 evidenced the domain of nanometric grains. This result will be useful later in explaining behaviours observed in the magnetic and electric responses. By comparing the granular density of the sample reported in Section 2 (6.36 g cm−3) with the density obtained by XRD (7.94 g cm−3), it can be established that the granular density of the samples is approximately 80.1% of the theoretical value expected for a sample made up of a perfect crystal.
As shown in Fig. 4, semiquantitative analysis of the composition was carried out by the deconvolution of the EDS spectra obtained through the incidence of the X-ray microprobe on different regions of the surface and inside the samples (using the fracture areas presented in Fig. 3c and d). The results reveal that the material does not contain impurities or elements other than those expected from the composition of the formula.
![]() | ||
Fig. 4 Composition analysis from EDS for the Sm2Bi2Fe4O12 material. The inset represents a comparison between the experimental and expected weight percentages. |
The inset of Fig. 4 contains a table with theoretical and experimental percentage weights, which were calculated from the stoichiometry of the samples and through the compositional analysis by means of EDS, respectively. The experimental and theoretical percentages are in good agreement, considering that light elements such as oxygen can introduce errors that affect the entire set of obtained values.
![]() | ||
Fig. 5 Characteristic response of current–voltage (a), and real and complex contribution to the electrical permittivity (b) measured in samples of Sm2Bi2Fe4O12. The inset in Fig. 6a corresponds to the quasi-linear behaviour that takes place under the application of high current values. |
The occurrence of Schottky barriers formed by grain boundaries is responsible for the nonlinear characteristic of the observed I–V curve.46 These barriers are idealized as micro junctions where the electrical responses for a grain pair behave as two consecutive Zener diodes. The intergranular limits act as resistances that produce currents like that expected in varistor diodes. When high-voltage values are applied, the resistance adopts an ohmic trend, as shown in the inset of Fig. 5a. This behaviour has place due to the presence of intragranular boundaries,47 but in this regime the resistivity decreases because the intragranular contributions are more relevant than the intergranular ones. Two-dimensional models have been proposed to understand this type of response, where the varistor microstructure is formulated in the form of a geometric network, representing the ceramic grains and boundaries.47 The three-dimensional models consider that the polycrystalline system can be simulated as a Voronoi-like tessellation.48 In this way, the microstructural electrical transport is modelled using equivalent electrical circuits with resistive and capacitive elements, which are related to the intragranular and intergranular transport currents inside the material.47 The I–V merit figure of the varistor follows the characteristics given by the following equation:
I = cVm, | (2) |
Fig. 5b exemplifies complex impedance measurements as a function of temperature under the application of frequencies from 100 Hz up to 10 MHz. The resulting curves reveal the strongly dispersive feature of complex electrical permittivity, with a shift of the imaginary permittivity (ε′′) towards the region of higher temperatures when frequency values are increased. By contrast, it is observed that the real permittivity (ε′) diminishes with the increase in applied frequency. These behaviours are correlated with non-homogeneous Maxwell–Wagner-type polarisation processes that involve conduction effects on the dielectric response that usually appear in rare-earth-based perovskites, where the conductive response is governed by the Arrhenius law.14 Undoubtedly, the results observed in the electrical permittivity curves are influenced by the crystal growth in the form of grains of different sizes and shapes, which make up a relatively compact distribution, with slight porosity and the occurrence of intergranular diffusion processes and grain boundaries observed in the SEM images of Fig. 3.
However, two well-differentiated behaviours are observed in the regimes 116 K < T < 225 K and T > 225 K of complex permittivity presented in Fig. 5b. This behaviour can also be observed in ref. 25 for Sm substitutions greater than 17.5% at the crystallographic Bi sites. In order to analyse the mechanisms that give rise to these differences in transport properties, an electrical stimulus analysis was performed by applying the Arrhenius-type equation for the frequency:
f = f0e−EA/kBT, | (3) |
![]() | ||
Fig. 6 Arrhenius analysis of the resonance frequency as a function of the inverse of the temperature in the low-frequency (a) and high-frequency (b) regimes for the Sm2Bi2Fe4O12 perovskite. |
The activation energies EA for the low- and high-frequency regimes are 0.34 eV and 0.58 eV, with pre-exponential factors f0 of 1013.77 Hz and 1012.55 Hz, respectively, suggesting the occurrence of ionic-type polarisation.54 The value of the activation energy in the temperature regime close to 116 K could be related to a ferroelectric-paraelectric transition as has been predicted for BaTiO3,55 which would be in agreement with the indirect evidence reported for Eu2Bi2Fe4O12 where this transition is observed in pyroelectric current measurements.14
![]() | ||
Fig. 7 (a) Diffuse reflectance spectrum and (b) the Kubelka–Munk analysis of the experimental determination of the optical band gap for the Sm2Bi2Fe4O12 material. |
This type of excitation is in accordance with the irreducible representation of the vibrations given by the following expression:
Γ = 5RSM + 7B1u + 9B2u + 9B3u, | (4) |
When the applied magnetic field is increased up to H = 2000 Oe, its intensity causes the maximum magnetic order value to increase (M = 0.50 emu g−1) but occurring at lower temperatures (T = 150 K), while the anomaly is now it manifests in a temperature regime above the maximum magnetic order value (at T = 175 K). For higher field strengths (H = 10000 Oe), the maximum magnetization reaches a value of 0.69 emu g−1 at T = 61.6 K and the anomaly tends to disappear (T = 156 K) in the presence of the applied magnetic field, since intergranular magnetic domains begin to obtain a greater long range order. In the case of the FC curves, the results reveal that the material shows a magnetic ordering behaviour that is pronounced with the decrease in temperature and with the increase in applied magnetic field intensity. The trend suggests that the material has a ferromagnetic type character, with a small decrease in the magnetic moment at low temperatures, due to the occurrence of magnetic domains, whose spins are oriented in the opposite direction to the applied field, which gives it a slight ferrimagnetic tendency. These antiparallel domains can be caused by disorientations due to canting of the Fe3+ electronic spins, which takes place as a consequence of the octahedral distortions observed in Fig. 2. However, due to the occurrence of magnetic ordering in the entire temperature regime in which the magnetic response was experimentally observed, it is not possible to make an adjustment that allows setting the Curie temperature value of the Sm2Bi2Fe4O12 perovskite. In the meantime, there is no evidence of the anomaly observed in ZFC measurements, because the anisotropies that took place at low temperatures for that measurement recipe now are dominated by the magnetic field that has been applied at high temperatures, so that, by decreasing the temperature, the correlation potential between magnetic domains has already been established and remains while the temperature changes. E. Gil-González et al.25 showed in differential scanning calorimetry measurements that TC is above 1123 K for Sm substitutions in the Bi sites greater than 12.5%. For this reason, it is not possible to establish the TC value or the effective magnetic moment through fittings following the Curie–Weiss law. Another interesting feature of the curves presented in Fig. 8 is the remarkable irreversibility that takes place for the three values of applied magnetic field strength. The irreversibility temperature value must be above the maximum measurement temperature, even for moderately high fields such as H = 10 kOe. This result is associated with the obviously disorganized nature of the material, which contains grains of various sizes and shapes, as well as inhomogeneous porosities and grain boundaries that hinder the orientation of the magnetic domains during the ZFC procedure, forming regions or clusters of domains with ferromagnetic short-range orders.
With the aim to analyse the characteristics of the magnetic ordering in the Sm2Bi2Fe4O12 material, magnetization curves as a function of applied magnetic fields up to H = 30000 Oe were performed, as exemplified in Fig. 9 for T = 50, 200 and 300 K. The hysteresis curves are narrow, with very small coercive fields (HC) and remnant magnetization (MR), but with relatively high saturation magnetizations (MS), as presented in Table 2. This can be inferred from the results presented in Table 2 that both HC and MR decay almost linearly with the increase in temperature. However, between the temperature values 50 K and 200 K, the MS remains approximately constant, while between the temperature values 200 K and 300 K, it shows a decreasing trend. The decay of HC and MR with the increase in temperature is expected from the M(T) curves observed in Fig. 9, because, although the temperature region between 50 K and 300 K corresponds to an ordering state of the ferromagnetic type, the magnetic moment produced by the domains oriented parallel to the applied field is not constant with the temperature.
T (K) | H C (Oe) | M R (±0.001 emu g−1) | M S (±0.001 emu g−1) |
---|---|---|---|
50 | 41.1 | 0.287 | 0.555 |
200 | 27.4 | 0.235 | 0.554 |
300 | 19.2 | 0.200 | 0.472 |
Meanwhile, the saturation magnetization behaviour is very interesting. Strictly speaking, it is not possible to affirm that there is saturation, because under the application of magnetic fields greater than H = 10 kOe, the hysteresis curves M(H) adopt an increasing linear characteristic, at least up to H = 30 kOe. Similar results were reported by E. Gil-González et al. for substitutions up to 20% of Sm in the Bi sites,25 where the indeterminacy of saturation magnetization prevents the experimental obtaining of the material's effective magnetic moment. This type of response could be associated with the nanometric character of a good volumetric portion of the Sm2Bi2Fe4O12 samples that were examined. This is known that the presence of grains with submicron size usually introduces border and surface effects that give rise to the formation of magnetic domains whose behaviour resembles that of paramagnetic materials, because each of them assumes the character of an independent magnetic entity.59 One of the macroscopic manifestations of this type of response takes place in the form of an apparent increasing linearity of the saturation magnetization regime in the hysteresis curves, as shown in Fig. 9. Similarly, the occurrence of these independent domains could significantly influence the frustrated feature of the system, which manifests itself through the irreversibility observed in the M(T) curves for the ZFC and FC recipes in Fig. 8. For this reason, the hysteresis curve observed is the result of the contributions of the ferromagnetic response originating in the micrometric grains plus the superparamagnetic character due to the submicrometric grains.
The distribution calculated for the electronic orbitals around the Fermi level in the material Sm2Bi2Fe4O12 is exemplified in Fig. 11, where the density of states with spin polarisation is presented, defining E = 0 eV as the value of the Fermi energy (EF). One of the observable peculiarities in the density of states is the asymmetry between the orbitals corresponding to the spin-up and spin-down electronic orientations. This behaviour allows us to obtain an effective magnetic moment of 20 μB in the unit cell that is largely due to the 3d-Fe orbitals, for which a majority orientation of spin-up electrons is observed with respect to the spin-down electrons. Another very important phenomenological characteristic that emerges from Fig. 11 is the quasi-insulating character of the electrons in the spin up configuration, which show an Eg value of 3.35 eV, compared to a semiconductor feature of the electrons in the spin-down configuration, for which Eg = 1.40 eV. Interestingly, the average between these two values (∼2.38 eV) corresponds 90% to the macroscopic value experimentally determined at room temperature, as presented in Section 4.3 of this document. As described in Section 3, the Hubbard U potential was determined by considering that the 4f-Sm electrons are frozen as core states, which can be corroborated in Fig. 11, where clearly the contributions of these electrons to the density of states occur far from the Fermi level (below −17 eV in the valence band and above 7 eV in the conduction band). It is important to note that the value of the Hubbard U potential obtained by the method described by Liechtenstein et al.,36 from the Fk parameters required in the consideration of 3d-Fe orbitals, led to a band gap value that is in agreement with the experimental results reported in this document.
![]() | ||
Fig. 11 Density of states for the spin-up and -down orientations with the contributions due to the electronic orbitals of Sm, Bi, Fe and O indicated in the picture. |
In the innermost region of the valence band, from −20 eV to −18.3 eV, there are representative contributions to the density of states due to orbitals of 5p-Sm with a minor contribution from the 2s-O electrons. Another hybridization has been observed in the region from −18.3 eV and −15.2 eV, where the highest contribution corresponds to the 2s-O states and the lowest to the 5p-Sm states, which have place for the two electron spin orientations. In the energy range from −10.2 eV to −8.7 eV, a significant contribution from the 6s-Bi orbitals and a smaller one due to the 2p-O orbitals are observed.
Between −8.7 eV and −6.2 eV, the 3d-Fe electronic states are definitely relevant for spin-up configuration, whereas in the energy regime from −6.2 eV up to the Fermi level, there are hybridizations of the 2p-O orbitals (major contributor), 4d-Sm and 6p-Bi for the two spin configurations. Above the band gap, from 1,387 eV to 3.4 eV, the states that are responsible for the majority contribution are 3d-Fe, which occur for the spin-down configuration. At last, starting at 3.4 eV to 11.48 eV, the major contributors to the density of states are the 6p-Bi, 4d-Sm, 2p-O and 4s-Fe orbitals, which appear asymmetrically for the two spin configurations, with majority hybridizations due to the 2p-O electrons. From the interpretation of the electronic states close to the Fermi level, it is possible to affirm that the semiconductor nature of the material is essentially due to the hybridization of the 2p-O orbitals with 4d-Sm and 6p-Bi, which confine the contributions of the 3d-Fe electronic states at lower levels in the valence band, giving rise to the band gap that energetically interposes between these states with majority spin-up orientation, and the 3d-Fe orbitals of the conduction band, in which dominates the spin-down orientation.
![]() | ||
Fig. 12 Specific heat at constant volume (a) and constant pressure (b) as a function of temperature for the Sm2Bi2Fe4O12 material calculated from Debye's quasi-harmonic model. |
It is observed in Fig. 12a and b that for temperature values below 300 K, the specific heats at constant volume and constant pressure follow the same feature for all the applied pressures. Meanwhile, Fig. 11a shows a clear tendency of the specific heat at a constant volume towards an independent behaviour of temperature, known as the Dulong–Petit limit (LDP),28 which is not observed in Fig. 12b, where, at high temperatures, the specific heat curves change as a result of the applied pressure, showing that each of the atoms in the material absorbs a different amount of energy with the applied temperature gradient. However, there seems to be a relationship between structural symmetry and the LDP value, since for perovskite-type materials that tend to adopt an ideal cubic unit cell, the LDP value (241.11 J mol−1 K−1)40 practically corresponds to half of the value reported for less symmetrical structures (476.36 J mol−1 K−1).61 Therefore, the further the structure moves away from that cubic considered ideal, the greater the thermal energy required to achieve the excitation of the atoms under the application of a temperature gradient, although it is also feasible that the process of absorption of thermal energy is less efficient than that in the case of structurally symmetrical cells. This is the probable reason why in the present work, for the Pnma space group of the Sm2Bi2Fe4O12 orthorhombic structure, a high value LDP = 490.22 J mol−1 K−1 was obtained.
The results presented here suggest that at low temperatures the purely electronic character of specific heat may be predictable for the complex perovskite studied, while the response to high temperatures must be associated with mechanisms of a phononic nature, so that in this regime the model loses reliability. By correlating the results of specific heat with the density of states presented in Fig. 10, it can be affirmed that the 2p-O, 4d-Sm and 6p-Bi orbitals that are very close to the Fermi level provide relevant electronic contributions to the specific heat, while the contribution of the 3d-Fe orbitals is very incipient. For this reason, at high temperature values, the electronic contributions to the specific heat coexist with important contributions from the vibration of the bonds between cations and anions, which move around their equilibrium positions as a consequence of the absorption of thermal energy, thereby increasing the value of the total specific heat. Additionally, in the case of ceramic compounds, the granular characteristics can introduce variations in possible experimental results, so it can be expected that the specific heat also depends on the porosity of the material, because the thermal energy necessary to increase the temperature is expected. It is less in porous materials than in those that are denser.
Fig. 13a exemplifies the entropy response as a function of temperature under the application of several pressures from 0 GPa up to 12 GPa. As Debye's quasi-harmonic model predicts,40 when the thermal energy of the system is increased, the randomness in the behaviour of their physical properties is magnified, thereby increasing the effects of inter- and intramolecular vibrations, structural distortions and thermal expansion (Fig. 13b), which grow rapidly in the temperature interval 0 K < T < 300 K, with a significant decrease in the slope for T > 300 K due to a decreasing trend in Debye's temperature. The temperature dependence of the coefficient of thermal expansion, α(T), for different values of applied pressure, is shown up to 12 GPa in Fig. 13b, where a systematic decrease in α(T) is shown when the pressure is increased. This effect is more dramatic in the high temperature regime, where it can be seen that α falls from 7.15 × 10−5 K−1 for P = 0 GPa to 4.25 × 10−5 K−1 for P = 12 GPa at T = 1000 K.
![]() | ||
Fig. 13 Behaviour of the entropy (a) and thermal expansion coefficient (b) as a function of temperature for the Sm2Bi2Fe4O12 ferrite. |
However, in the low temperature range 0 K < T < 300 K, the coefficient of thermal expansion increases rapidly with the increase in temperature for all applied pressure values. Due to the orthorhombic characteristic of the structure in Sm2Bi2Fe4O12, with the three network parameters in its unit cell, marked differences could be expected between the values of the coefficient of thermal expansion along the three crystallographic directions. This behaviour has to do with the octahedral distortions that were mentioned in Section 4.1, such that changes in temperature and pressure cause rotations and inclinations of the FeO6 octahedra, as well as the elongation or compression of the crystallographic cell along the crystallographic axes. Then, the behaviours observed in α(T) for the different applied pressures have a correlation with the distorting nature of the cell, as well as with the possible structural transitions that take place due to the changes in temperature and pressure to which the material is being subjected. Therefore, from the experimental point of view, the occurrence of deviations with respect to the theoretical results presented in Fig. 13b is possible. Regarding the relatively low absolute value of the coefficient of thermal expansion in this family of magnetic semiconductor, its potential applicability in microelectronic circuits for computers and similar devices could be suggested.62
A graph of Debye's temperature as a function of temperature, ΘD(T), under the application of pressures up to 12 GPa, is presented in Fig. 14a, where it is observed that the increase in the system pressure gives rise to a systematic increase in the elastic waves of the system, producing a ΘD growth. This increase is considerably more pronounced than that observed in the case of cubic perovskites.61 Likewise, the dependence with the increase in temperature shows a continuous non-linear decreasing behaviour of ΘD for all the applied pressures, which has been reported by other authors for materials of double perovskite-type A2BB′O6.63,64 From the results, it can be interpreted that the pressure exerts an effect of increasing the vibration frequencies of the cation–anion bonds in the cell, while the increase in temperature expands the structure, increasing the wavelength of the vibrations, therefore the frequency decreases, as does ΘD.
![]() | ||
Fig. 14 Debye temperature (a) and Grüneisen parameter (b) as a function of temperature for the Sm2Bi2Fe4O12 perovskite-like material. |
Finally, Fig. 14b shows that an increase in the applied pressure produces a gradual decrease in the Grüneisen parameter γ, which occurs for the entire temperature regime analysed. In the same way, it is evident that γ grows continuously and not linearly with the increase in temperature. As discussed above, these behaviours have to do with the changes produced in the vibration frequency of the crystal lattice,65 since γ was obtained by calculating the logarithmic derivative of ΘD with respect to volume.40
Footnotes |
† The raw/processed data necessary to reproduce these findings may be viewed at DOI: http://10.17632/nny9w27zb3.1. |
‡ CCDC 2011265. For crystallographic data in CIF or other electronic format see DOI: 10.1039/d0tc02935a |
This journal is © The Royal Society of Chemistry 2020 |