Next Article in Journal
Thermal Decomposition Properties of Epoxy Resin in SF6/N2 Mixture
Next Article in Special Issue
Cu-Doped ZnO Electronic Structure and Optical Properties Studied by First-Principles Calculations and Experiments
Previous Article in Journal
Impacts of Defocusing Amount and Molten Pool Boundaries on Mechanical Properties and Microstructure of Selective Laser Melted AlSi10Mg
Previous Article in Special Issue
Enhanced Thermoelectric Properties of Ca3−xAgxCo4O9 by the Sol–Gel Method with Spontaneous Combustion and Cold Isostatic Pressing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Oxygen Nonstoichiometry on Electrical Conductivity and Thermopower of Gd0.2Sr0.8FeO3−δ Ferrite Samples

1
Kirensky Institute of Physics, Federal Research Center Krasnoyarsk Scientific Center, Russian Academy of Sciences, Siberian Branch, 660036 Krasnoyarsk, Russia
2
Institute of Engineering Physics and Radio Electronics, Siberian Federal University, 660041 Krasnoyarsk, Russia
3
Institute of Chemistry and Chemical Technology, Federal Research Center Krasnoyarsk Scientific Center, Russian Academy of Sciences, Siberian Branch, 660036 Krasnoyarsk, Russia
4
Ioffe Institute, 194021 St. Petersburg, Russia
*
Author to whom correspondence should be addressed.
Materials 2019, 12(1), 74; https://doi.org/10.3390/ma12010074
Submission received: 27 November 2018 / Revised: 19 December 2018 / Accepted: 21 December 2018 / Published: 26 December 2018
(This article belongs to the Special Issue Advanced Thermoelectric Materials)

Abstract

:
The behavior of the resistivity and thermopower of the Gd0.2Sr0.8FeO3−δ ferrite samples with a perovskite structure and the sample stability in an inert gas atmosphere in the temperature range of 300–800 K have been examined. It has been established that, in the investigated temperature range, the thermoelectric properties in the heating‒cooling mode are stabilized at δ ≥ 0.21. It is shown that the temperature dependencies of the resistivity obtained at different δ values obey the activation law up to the temperatures corresponding to the intense oxygen removal from a sample. The semiconductor‒semiconductor electronic transitions accompanied by a decrease in the activation energy have been observed with increasing temperature. It is demonstrated that the maximum thermoelectric power factor of 0.1 µW/(cm·K2) corresponds to a temperature of T = 800 K.

1. Introduction

The major characteristics of a thermoelectric material are its thermoelectric efficiency Z = S2σ/ϰ (S is the Seebeck coefficient, also known as thermopower, σ is the electrical conductivity, and ϰ is the thermal conductivity) [1], dimensionless thermoelectric quality factor ZT (T is the absolute temperature), and power factor P = S2σ = S2/ρ (ρ is the resistivity).
Recently, close attention of researchers has been paid to oxide materials [2], since they are stable against high temperatures, nontoxic, and, as a rule, do not contain rare elements. A group of complex oxides of the rare-earth transition metals Ln1−xAxMO3−δ (Ln is a lanthanoid, A is an alkali or alkali-earth metal, and M is a transition metal (Co, Fe, or Mn)) with a perovskite structure holding a special place [3,4,5] among the oxides for promising application. The diverse physicochemical properties of these compounds, which belong to strongly correlated electron systems, are interesting both for application [6] and for fundamental research. Along with their thermoelectric characteristics, the stability of these materials against different environmental factors, including the atmospheric composition and temperature, is of great practical importance. In this study, taking into account the significant differences between the low-temperature properties of the Gd1−xSrxFeO3−δ compounds with different oxygen nonstoichiometry indices δ [7] and the high mobility of oxygen in the Gd1−xSrxСoO3−δ compounds [8], we investigate the effect of the oxygen nonstoichiometry on the thermoelectric properties of the Gd0.2Sr0.8FeO3−δ compound and its stability in the temperature range of 300–800 K.

2. Materials and Methods

Sample synthesis. The Gd0.2Sr0.8FeO3−δ samples were obtained using the glycinenitrate-based solution-combustion synthesis [9]. The initial components of the perovskiteoxide were cations of nitrates of each metal component (Gd(NO3)3·6H2O, 99.9% trace metal basis, Rare Metals Plant, Novosibirsk, Russia; Sr(NO3)2, >99.0%, Reachim, Moscow, Russia; Fe(NO3)3·9H2O, >98%, Sigma-Aldrich, St. Louis, MO, USA). The initial gel was obtained via evaporation of the glycinenitrate-based solution at 363 K with subsequent heating until combustion at a low rate. The prepared powder was calcinated in air at 723 K for 5 h, thoroughly grinded in anagate mortar, and sintered in air at 1623 K for 5 h with subsequent cooling to room temperature at a rate of 2 °C/min. Tablets were synthesized at Tc = 1473 K for 24 h and cooled in the same mode.
The oxygen nonstoichiometry of the samples was determined using a NETZSCHSTA449C analyzer equipped with an Aёolos QMS 403C mass spectrometer (NETZSCH, Selb, Germany). The measurements were performed in the 5% H2-Ar mixture flow in a corundum crucible with a perforated cover. The sample mass was 30 mg. The oxygen content in the Gd0.2Sr0.8FeO3−δ samples was determined from the mass decrement ∆m using the equation δ(T) = δ0 + M∆m(T)/1600, where δ(T) is the δ value at temperature T, ∆m(T) (%) is the mass decrement at this temperature, and M is the molecular mass of the Gd0.2Sr0.8FeO2.87 compound. The measurement error was δ = ±0.01. The oxygen nonstoichiometry index of the initial Sr0.8Gd0.2FeO3−δ material was taken to be δ0 = 0.13, according to the data on the Gd-Sr-Fe perovskite synthesized using the same technique [7].
The X-ray diffraction measurements were performed on powder diffractometer (PANalytical X’Pert PRO, Eindhoven, The Netherlands) using Cu Kα radiation in the 2θ-angular range of 20–158°. The crystal structure was refined using the derivative difference minimization (DDM) method [10]. Temperature dependences of the thermopower and resistivity were obtained on an experimental setup developed at the Ioffe Institute, St. Petersburg [11,12].

3. Results

In contrast to the complex cobalt oxides Re1−xSrxCoO3−δ (Re is the rare-earth ion), which, depending on the cooling rate and synthesis temperature, can form compounds ordered or disordered over А sites of the perovskite structure [13,14,15,16,17,18], the Gd0.2Sr0.8FeO3−δ compound synthesized by us is disordered and characterized by the random distribution of oxygen vacancies and cations over the А sites. The X-ray diffraction study revealed no foreign phases. According to the X-ray powder diffraction data (Figure 1), the synthesized Gd0.2Sr0.8FeO3−δ samples, similar to the heavily-doped La1−xSrxFeO3−δ (0.8 ≤ x ≤ 1.0) compounds [19], have a form of a disordered perovskite with the cubic symmetry (sp. gr.Pm3m (a = 3.8688(1) Å), in which all the iron sites were identical, as in the mixed-valence iron compounds Sr2LaFe3O8.94 [20]. The structural formula derived from the refined occupancies of atomic positions (Table 1) was Sr0.8Gd0.2FeO2.79. Similar perovskites Sr0.75Gd0.25FeO2.87 and Sr0.75Gd0.25FeO2.94 described in [7] have smaller lattice parameters (3.8665(1) and 3.8582(1) Å), which is consistent with the smaller oxygen stoichiometry estimated from the XRD refinement for our sample.
Atomic positions, and their occupancies and thermal parameters in the crystal structure of Sr0.8Gd0.2FeO2.79 are given in Table 1.
Figure 2 shows the temperature-programmed reduction data. The change in the mass caused by removal of oxygen from the sample upon heating in the 5% H2-Ar reducing atmosphere started at ≈500 K and depended on the heating rate, which is related to the solid-state reduction kinetics. As the heating rate increased, the thermogravimetric (TG) and derivative thermogravimetric (DTG) curves shifted to the high-temperature region and the maximum reduction rate was observed at 640 and 725 K at heating rates of 2 and 20 K/min, respectively.
The electrical resistivity and thermopower were measured on the rectangular bar samples 0.9 × 4.5 × 9 mm in size at temperatures from 300 to 800 K in the inert He (99.999%) atmosphere. To examine the changes in the temperature dependences of the resistivity and thermopower, the measurements were performed during the heating–cooling cycles and the courses of the obtained temperature curves were compared. Three thermal cycles were performed. The heating and cooling rates were 5 °C/min. The results obtained are shown in Figure 3. It can be seen that the temperature dependences of the electrical resistivity and thermopower of the investigated samples in the heating‒cooling cycles were essentially different, which is indicative of the strong effect of oxygen on the physical properties of the Gd0.2Sr0.8FeO3-δ compound. In the first thermocycle, when oxygen was actively removed from the sample, the behavior of the thermoelectric parameters of the compound was especially anomalous (Figure 4а). For clarity, the electrical resistivity and Seebeck coefficient measured during the first heating with the active oxygen removal (Figure 4а) and the last cooling (Figure 4b), when the sample was almost stable, are plotted.
The anomalies in the behavior of the Seebeck coefficient correlated with the data of thermogravimetric analysis of the sample mass loss caused by the removal of oxygen and are observed at much lower temperatures than in the behavior of the resistivity (Figure 4a).
The oxygen nonstoichiometry index δ increased from one heating‒cooling cycle to another and the deviation from the semiconductor-type conductivity shifted toward higher temperatures. At δ = 0.21, the temperature dependency of the resistivity was qualitatively consistent with the semiconductor-type conductivity dρ/dT < 0 over the entire temperature range of interest (Figure 4b). The conductivity in the regions corresponding to the semiconductor type obeyed the thermal activation law:
ρ ( T ) = ρ 0 × e x p ( E a / k B T )
where ρ 0 is the coefficient weakly dependent on temperature, E a is the activation energy, and k B is the Boltzmann constant. The dependences of the logarithmic resistivity of the sample on the reciprocal temperature in the corresponding temperature ranges for several heating and cooling processes are presented in Figure 5. For all the thermal cycles, the ln(ρ)(1/T) curves contain two portions that are described well by the thermal activation law with different activation energies E a . At each specific δ value, the activation energy decreased with increasing temperature. At the same time, the activation energy grew with the oxygen nonstoichiometric index. The absolute values of the oxygen nonstoichiometry indices and activation energies for different temperature ranges and electronic transition temperatures were determined by the crossing points of the approximation curves are given in Table 2.
Figure 6 presents the plots of activation energy E a in the low-temperature (below the temperature of the electronic transition, T < Tp-p) and high-temperature (above the transition temperature, T > Tp-p) regions and semiconductor‒semiconductor transition temperature Tp-p as functions of the oxygen content in the sample. It can be seen that, in the investigated temperature and δ ranges, all the dependencies are linear.
The E a value in the low-temperature range was several times higher than in the high-temperature range; the difference slightly decreased with increasing δ. The electrical conductivity of the materials under study was implemented using impurity carriers induced by different-valence Gd+3/Sr+2 cation substitutions and holes related to oxygen vacancies. At high temperatures, we observed the properties of an almost degenerate semiconductor, while at low temperatures, the activation energy increased. The temperature Tp-p of the asymptotic change weakly depended on the vacancy concentration. The fact that the activation energy depended on the vacancy concentration is evidence for the shift of the impurity level from the allowed state band extremum. It is worth noting that the concentration dependences of the activation energies (straights 1 and 2 in Figure 6) are almost parallel.
Figure 7 shows the temperature dependencies of the power factor P for the first and last temperature cycles. It can be seen that, as the temperature increased, the power factor of the samples grew, but did not attain its maximum, in the investigated temperature range. In the sample stability region, the increase was monotonic, without jumps, and almost linear for the sample with δ = 0.21. At a temperature of T = 500 K, the power factor of the sample with δ = 0.21 exceeded almost fourfold the power factor of the samples with δ = 0.13 (0.052 µW/(K2·cm) and 0.013 µW/(K2·cm)) and the maximum P value obtained at T = 800 K was 0.1 µW/(K2·cm).
The power factors obtained in this work for the sample Gd0.2Sr0.8FeO3-δ (P = 0.1 μW/(cm·K2)) are consistent with those presented in Reference [21] for the LaCo1−xNixO3 (P = 0,12 μW/(cm·K2), LaCo1−xTixO3−δ (P = 0.28 μW/(cm·K2)) [22], La1−xSrxCo0.8Ni0.1Fe0.1O3 (P = 0.76 μW/(cm·K2)) [5], and La1−xNaxCoO3 (P = 0.1 μW/(cm·K2)) polycrystalline samples [23], although our data are inferior to the value of P ≈ 3 μW/(cm·K2) for the Ca3−xBixCo4O9+δ samples from study [24].

4. Conclusions

The behavior of electrical resistivity and thermopower of the disordered Gd0.2Sr0.8FeO3−δ perovskite in the temperature range of 300–800 K and its stability in the helium atmosphere were investigated. It was established that the stability of thermoelectric parameters strongly depends on the oxygen content in a sample and is obtained at temperatures from 300 to 800 K at an oxygen nonstoichiometric index of δ ≥ 0.21.
In the sample stability region, the temperature dependences of electrical resistivity for all δ values contained portions that were described well by the thermal activation law; as the temperature increased, spread semiconductor‒semiconductor transitions were observed, which are accompanied by a decrease in the activation energy. The latter grows with decreasing oxygen content in a sample. In this case, the temperature of electronic transition depends linearly on the oxygen nonstoichiometry.
We explained the nature of the semiconductor transitions as the result of occupancy/deoccupancy by the oxygen ions of different positions in the nonstoichiometry lattice. The electron bands near the Fermi level substantially changed and led to significant changes in the band gap and the activation energy. The activation energies in the low- and high-temperature regions depended similarly on the δ value. The increase in the δ value led to the significant enhancement of the power factor P. In particular, at a temperature of T = 500 K, the power factor of the sample with δ = 0.21 was almost fourfold higher than the power factor of the samples with δ = 0.13 (0.052 µW/(K2·cm) and 0.013 µW/(K2·cm). In the sample with δ = 0.21, the power factor increased with temperature almost linearly and attained its maximum value of 0.1 µW/(cm·K2) at T = 800 K. Thus, varying the oxygen content in a sample, one can control the temperature ranges of stability and the power factor of the complex transition metal oxides.

Author Contributions

Conceptualization & Methodology, V.D. and S.O.; Investigation, A.B., S.V., L.S. and S.N.; Writing-Original Draft Preparation, Y.O.; Writing-Review & Editing, A.F.; Visualization, V.D.; Supervision, Project Administration & Funding Acquisition, A.F.

Funding

This research was funded by the Russian Science Foundation, project No. 16-13-00060.

Acknowledgments

This study was supported by the Russian Science Foundation, project No. 16-13-00060.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ioffe, A.F. Semiconductor Thermoelements, and Thermoelectric Cooling (Infosearch, London 1957); Rowe, D.M., Ed.; CRC Handbook of Thermoelectrics; CRC Press: Boca Raton, FL, USA, 1995. [Google Scholar]
  2. Shiand, X.; Chenand, L.; Uher, C. Recent advances in high-performance bulk thermoelectric materials. Int. Mater. Rev. 2016, 61, 379–415. [Google Scholar]
  3. Terasaki, I.; Sasago, Y.; Uchinokura, K. Large thermoelectric power in NaCo2O4 single crystals. Phys. Rev. B 1997, 56, R12685. [Google Scholar] [CrossRef]
  4. Weidenkaff, A.; Aguirre, M.H.; Bocher, L.; Trottmann, M.; Tomes, P.; Robert, R. Development of Perovskite-type Cobaltates and Manganates for Thermoelectric Oxide Modules. J. Korean Ceram. Soc. 2010, 47, 47–53. [Google Scholar] [CrossRef]
  5. Harizanova, S.G.; Zhecheva, E.N.; Valchev, V.D.; Khristova, M.G.; Stoyanova, R.K. Improving the thermoelectric efficiency of Co based ceramics. Mater. Today Proc. 2015, 2, 4256–4261. [Google Scholar] [CrossRef]
  6. Zlati´c, V.; Boyd, G.R.; Freericks, J.K. Universal thermopower of bad metals. Phys. Rev. B 2014, 89, 155101. [Google Scholar] [CrossRef]
  7. Blasco, J.; Stankiewicz, J.; García, J. Phase segregation in the Gd1−xSrxFeO3−δ series. J. Solid State Chem. 2006, 179, 898–908. [Google Scholar] [CrossRef]
  8. Vereshchagin, S.N.; Dudnikov, V.A.; Solovyov, L.A. Study of Mobile Oxygen in Ordered/Disordered Nonstoichiometric Sr-Gd-Cobaltate by Simultaneous Thermal Analysis. J. Sib. Fed. Univ. Chem. 2017, 10, 346–357. [Google Scholar] [CrossRef] [Green Version]
  9. Chick, L.A.; Pederson, L.R.; Maupin, G.D.; Bates, J.L.; Thomas, L.E.; Exarhos, G. Glycine-nitrate combustion synthesis of oxide ceramic powders. J. Mamep Lett. 1990, 10, 6. [Google Scholar] [CrossRef]
  10. Solovyov, L.A. Full-Profile Refinement by Derivative Difference Minimization. J. Appl. Crystallogr. 2004, 37, 743–749. [Google Scholar] [CrossRef]
  11. Burkov, A.T.; Heinrich, A.; Konstantinov, P.P.; Nakama, T.; Yagasaki, K. Experimental set-up for thermopower and resistivity measurements at 100–1300 K. Meas. Sci. Technol. 2001, 12, 264–272. [Google Scholar] [CrossRef]
  12. Burkov, A.T.; Fedotov, A.I.; Kasyanov, A.A.; Panteleev, R.I.; Nakama, T. Methods and technique of thermopower and electricalconductivity measurements of thermoelectric materials at high temperatures. Sci. Tech. J. Inf. Technol. Mech. Opt. 2015, 15, 173–195. [Google Scholar] [CrossRef]
  13. James, M.; Tedesco, T.; Cassidy, D.J.; Withers, R.L. Oxygen vacancy ordering in strontium doped rare earth cobaltate perovskites Ln1−xSrxCoO3−δ (Ln = La, Pr and Nd; x > 0.60). Mater. Res. Bull. 2005, 40, 990. [Google Scholar] [CrossRef]
  14. James, M.; Morales, L.; Wallwork, K. Structure and Magnetism in Rare Eath Strontium-Doped Cobaltates. Phys. B 2006, 199, 385–386. [Google Scholar]
  15. Vereshchagin, S.N.; Dudnikov, V.A.; Shishkina, N.N.; Solovyov, L.A. Phase transformation behavior of Sr0.8Gd0.2CoO3−δ perovskite in the vicinity of order-disorder transition. Thermochim. Acta 2017, 655, 34–41. [Google Scholar] [CrossRef]
  16. Vereshchagin, S.N.; Solovyov, L.A.; Rabchevskii, E.V.; Dudnikov, V.A.; Ovchinnikov, S.G.; Anshits, A.G. Methane oxidation over A-site ordered and disordered Sr0.8Gd0.2CoO3−δ perovskites. Chem. Commun. 2014, 50, 6112–6115. [Google Scholar] [CrossRef]
  17. Yoshida, S.; Kobayashi, W.; Nakano, T.; Terasaki, I.; Matsubayashi, K.; Uwatoko, Y.; Grigoraviciute, I.; Karppinen, M.; Yamauchi, H. Chemical and physical pressure effects on the magnetic and transport properties of the A-site ordered perovskite Sr3YCo4O10.5. J. Phys. Soc. Jpn. 2009, 78, 094711. [Google Scholar] [CrossRef]
  18. Fukushima, S.; Sato, T.; Akahoshi, D.; Kuwahara, H. Comparative study of ordered and disordered Y1−xSrxCoO3−δ. J. Appl. Phys. 2008, 103, 07F705. [Google Scholar] [CrossRef]
  19. Dann, S.E.; Currie, D.B.; Weller, M.T.; Thomas, M.F.; Al-Rawwas, A.D. The Effect of Oxygen Stoichiometry on Phase Relations and Structure in the System La1−xSrxFeO3−δ (0 <= x <= 1, 0 <= δ <= 0.5). J. Solid State Chem. 1994, 109, 134–144. [Google Scholar] [CrossRef]
  20. Battle, P.D.; Gibb, T.C.; Lightfoot, P. The structural consequences of charge disproportionation in mixed-valence iron oxides. The crystal structure of Sr2LaFe3O8.94 at room temperature and 50 K. J. Solid State Chem. 1990, 84, 271–279. [Google Scholar] [CrossRef]
  21. Robert, R.; Aguirre, M.H.; Bocher, L.; Trottmann, M.; Heiroth, S.; Lippert, T.; Döbeli, M.; Weidenkaff, A. Thermoelectric Properties of LaCo1−xNixO3 Polycrystalline Samples and Epitaxial Thin Films. Solid State Sci. 2008, 10, 502–507. [Google Scholar] [CrossRef]
  22. Robert, R.; Bocher, L.; Trottmann, M.; Reller, A.; Weidenkaff, A. Synthesis and High-Temperature Thermoelectric Properties of Ni and Ti Substituted LaCoO3. J. Solid State Chem. 2006, 179, 3893–3899. [Google Scholar] [CrossRef]
  23. Behera, S.; Kamble, V.B.; Vitta, S.; Umarji, A.M.; Shivakumara, C. Synthesis Structure and Thermoelectric Properties of La1−xNaxCoO3 Perovskite Oxides. Bull. Mater. Sci. 2017, 40, 1291–1299. [Google Scholar] [CrossRef]
  24. Moser, D.; Karvonen, L.; Populoh, S.; Trottmann, M.; Weidenkaff, A. Influence of the Oxygen Content on Thermoelectric Properties of Ca3−xBixCo4O9+δ System. Solid State Sci. 2011, 13, 2160–2164. [Google Scholar] [CrossRef]
Figure 1. Experimental (1), calculated (2), and difference (3) X-ray diffraction patterns of the Gd0.2Sr0.8FeO2.79 compound after DDM refinement.
Figure 1. Experimental (1), calculated (2), and difference (3) X-ray diffraction patterns of the Gd0.2Sr0.8FeO2.79 compound after DDM refinement.
Materials 12 00074 g001
Figure 2. (1, 2) TG and (3, 4) DTG curves of temperature-programmed reduction of the Gd0.2Sr0.8FeO2.87 compound in the 5% H2-Ar mixture flow at heating rates of 2 (black solid curve) and 20 K/min (red dotted curve).
Figure 2. (1, 2) TG and (3, 4) DTG curves of temperature-programmed reduction of the Gd0.2Sr0.8FeO2.87 compound in the 5% H2-Ar mixture flow at heating rates of 2 (black solid curve) and 20 K/min (red dotted curve).
Materials 12 00074 g002
Figure 3. Temperature dependences of (а) the resistivity and (b) thermopower for three continuous heating‒cooling cycles.
Figure 3. Temperature dependences of (а) the resistivity and (b) thermopower for three continuous heating‒cooling cycles.
Materials 12 00074 g003
Figure 4. Temperature dependences of the resistivity and thermopower (insets) during (a) the first heating and (b) the third cooling.
Figure 4. Temperature dependences of the resistivity and thermopower (insets) during (a) the first heating and (b) the third cooling.
Materials 12 00074 g004
Figure 5. Reciprocal temperature dependence of the Gd0.2Sr0.8FeO3−δ logarithmic resistivity. Black straight lines show the thermal activation law processing: (1) first heating, (2) first cooling, and (3) third cooling.
Figure 5. Reciprocal temperature dependence of the Gd0.2Sr0.8FeO3−δ logarithmic resistivity. Black straight lines show the thermal activation law processing: (1) first heating, (2) first cooling, and (3) third cooling.
Materials 12 00074 g005
Figure 6. Dependencies of the activation energy E a : (1) is the temperature region below the electronic transition and (2) is the high-temperature region, and (3) semiconductor–semiconductor transition temperature Tp-p on the oxygen nonstoichiometry index.
Figure 6. Dependencies of the activation energy E a : (1) is the temperature region below the electronic transition and (2) is the high-temperature region, and (3) semiconductor–semiconductor transition temperature Tp-p on the oxygen nonstoichiometry index.
Materials 12 00074 g006
Figure 7. Temperature dependences of the power factor P for the first heating (δ = 0.13 at T = 300 K) (1) and the third cooling (δ = 0.21) (2).
Figure 7. Temperature dependences of the power factor P for the first heating (δ = 0.13 at T = 300 K) (1) and the third cooling (δ = 0.21) (2).
Materials 12 00074 g007
Table 1. Atomic positions, and their occupancies and thermal parameters in the crystal structure of Sr0.8Gd0.2FeO2.79. Space group is Pm3m, with lattice parameter a = 3.8688(1) Å.
Table 1. Atomic positions, and their occupancies and thermal parameters in the crystal structure of Sr0.8Gd0.2FeO2.79. Space group is Pm3m, with lattice parameter a = 3.8688(1) Å.
AtomPositionOccupancyxyzUiso2]
Sr1a0.800(4)0000.0092(7)
Gd6e0.0334(7)0.080(4)000.0092(7)
Fe1b11/21/21/20.0082(5)
O12h0.232(1)0.452(2)1/200.0211(14)
Table 2. Oxygen nonstoichiometry indices, activation energies, and electronic transition temperatures for the Gd 0.2Sr0.8FeO3−δ ferrite sample.
Table 2. Oxygen nonstoichiometry indices, activation energies, and electronic transition temperatures for the Gd 0.2Sr0.8FeO3−δ ferrite sample.
RegimeδEa (T < Tp-p) (eV)Tp-p (K)Ea (T > Tp-p) (eV)
First heating0.13 ± 0.010.16 ± 0.01461 ± 10.04 ± 0.01
First cooling0.19 ± 0.010.20 ± 0.01495 ± 10.07 ± 0.01
Second cooling0.20 ± 0.010.21 ± 0.01502 ± 10.08 ± 0.01
Third cooling0.21 ± 0.010.22 ± 0.01507 ± 10.09 ± 0.01

Share and Cite

MDPI and ACS Style

Dudnikov, V.; Orlov, Y.; Fedorov, A.; Solovyov, L.; Vereshchagin, S.; Burkov, A.; Novikov, S.; Ovchinnikov, S. Effect of Oxygen Nonstoichiometry on Electrical Conductivity and Thermopower of Gd0.2Sr0.8FeO3−δ Ferrite Samples. Materials 2019, 12, 74. https://doi.org/10.3390/ma12010074

AMA Style

Dudnikov V, Orlov Y, Fedorov A, Solovyov L, Vereshchagin S, Burkov A, Novikov S, Ovchinnikov S. Effect of Oxygen Nonstoichiometry on Electrical Conductivity and Thermopower of Gd0.2Sr0.8FeO3−δ Ferrite Samples. Materials. 2019; 12(1):74. https://doi.org/10.3390/ma12010074

Chicago/Turabian Style

Dudnikov, Vyacheslav, Yury Orlov, Aleksandr Fedorov, Leonid Solovyov, Sergey Vereshchagin, Alexander Burkov, Sergey Novikov, and Sergey Ovchinnikov. 2019. "Effect of Oxygen Nonstoichiometry on Electrical Conductivity and Thermopower of Gd0.2Sr0.8FeO3−δ Ferrite Samples" Materials 12, no. 1: 74. https://doi.org/10.3390/ma12010074

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