Next Article in Journal
An Alternative Method for Synthesizing N,2,3-Trimethyl-2H-indazol-6-amine as a Key Component in the Preparation of Pazopanib
Previous Article in Journal
Wettability of a Polymethylmethacrylate Surface by Fluorocarbon Surfactant Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and NEXAFS and XPS Characterization of Pyrochlore-Type Bi1.865Co1/2Fe1/2Ta2O9+Δ

by
Nadezhda A. Zhuk
1,*,
Sergey V. Nekipelov
2,
Olga V. Petrova
2,
Aleksandra V. Koroleva
3,
Aleksey M. Lebedev
4 and
Boris A. Makeev
5
1
Department of Chemistry, Institute of Natural Sciences, Syktyvkar State University, Oktyabrsky Prospect, 55, 167001 Syktyvkar, Russia
2
Institute of Physics and Mathematics of the Komi Science Center UB RAS, Oplesnina St. 4, 167982 Syktyvkar, Russia
3
Resource Center “Physical methods of surface research”, Faculty of Physics, Saint Petersburg State University, University Emb. 7/9, 199034 St. Petersburg, Russia
4
National Research Center—Kurchatov Institute, 1 Akad. Kurchatova Sq., 123182 Moscow, Russia
5
Institute of Geology of the Komi Science Center UB RAS, Pervomaiskaya St. 54, 167982 Syktyvkar, Russia
*
Author to whom correspondence should be addressed.
Chemistry 2024, 6(5), 1078-1088; https://doi.org/10.3390/chemistry6050062
Submission received: 9 August 2024 / Revised: 10 September 2024 / Accepted: 13 September 2024 / Published: 19 September 2024
(This article belongs to the Section Inorganic and Solid State Chemistry)

Abstract

:
A cubic pyrochlore with the composition Bi1.865Co1/2Fe1/2Ta2O9+Δ (space group Fd-3m, a = 10.5013(8) Å) was synthesized from oxide precursors using solid-phase reactions. These ceramics are characterized by a porous microstructure formed by randomly oriented grains of an elongated shape with a longitudinal size of 0.5–1 µm. The electronic state of cobalt and iron ions in oxide ceramics was studied by NEXAFS and XPS spectroscopy. The parameters of the XPS spectra of Bi4f, Bi5d, Ta4f, Co2p, and Fe2p ionization thresholds for a complex pyrochlore were compared with the parameters of the corresponding oxides of the transition elements. The energy position of the XPS-Ta4f and -Ta5p spectra is shifted towards lower energies compared to the binding energy in tantalum(V) oxide by 0.75 eV. According to XPS spectroscopy, bismuth and tantalum cations have the corresponding effective charge of +3 and +(5-δ). The NEXAFS-Fe2p spectrum of ceramics coincides with the spectrum of Fe2O3 in its main spectrum characteristics and indicates the content of iron ions in the oxide ceramics in the form of octahedral Fe(III) ions, and according to the character of the Co2p spectrum, cobalt ions are predominantly in the Co(II) state.

1. Introduction

Synthetic pyrochlores based on bismuth tantalate attract close attention from scientists due to their promising dielectric properties. Showing low values of dielectric losses and high permittivity, an adjustable temperature of the coefficient of capacitance, and chemical compatibility with low-melting metal conductors, materials based on bismuth-containing pyrochlores are promising as multilayer ceramic capacitors and tunable microwave dielectric components [1,2]. The cubic crystal structure of the pyrochlores A2B2O6O’ are formed by the weakly interacting cationic sublattices A2O’ and B2O6 (Figure 1). The three-dimensional framework of the octahedral sublattice B2O6 ensures the stability of the pyrochlore structure with respect to iso- and hetero-valent substitutions of ions in the cation sublattices and to oxygen vacancies in A2O’ [3], which makes it possible to significantly vary the chemical composition and study the effect of chemical modification of the composition of pyrochlores on the physicochemical properties of pyrochlores. As a rule, the octahedral positions B of the B2O6 sublattice are occupied by relatively small and electronegative cations (Ti4+, Nb5+, Ta5+); larger ions (Bi3+, Ca2+) are distributed in the eight-coordinated A positions. As numerous studies have shown, doping of pyrochlores with 3d-element ions often leads to the formation of so-called mixed pyrochlores [4,5,6,7,8]. A feature of bismuth-containing pyrochlores doped with 3d-element ions is the partial vacancy of the bismuth sublattice and the distribution of dopants in the cation sublattices of both bismuth and tantalum [4,5,6,7,9].
As studies show, the reason for the dual distribution of transition element ions in the bismuth sublattice is the strain of the octahedral framework caused by the placement of heterovalent and incommensurate 3d-element ions in the tantalum(V)/niobium(V) sublattice. This conclusion is confirmed by the percentage ratio of cations occupying positions in the bismuth/tantalum sublattices. According to numerous experiments [4,5,6,7], this ratio does not exceed 1/3. The paper [10] shows the possibility of the formation of iron-containing pyrochlores of the general composition Bi3.36Fe2.08+xTa2.56−xO14.56−x (−0.32 ≤ x ≤ 0.48). Using the data of magnetic susceptibility, and Mössbauer and electron energy loss spectroscopy (EELS) analysis of iron-containing pyrochlores, it was found that iron ions are in the high spin state Fe(III), occupying mainly octahedral positions Nb/Ta/Sb and Nb/Ta/Sb [10,11,12,13,14,15]. It was also found that some of the Fe(III) ions can be placed in the bismuth positions. In particular, depending on the composition of the ceramics, only 4–15% of the A-sites are occupied by Fe3+ ions in the pyrochlores of the Bi2O3-Fe2O3-Nb2O5 system [11]. For the Bi2O3–Fe2O3–Sb2Ox and Bi1.8Fe0.2(FeSb)O7 systems the occupation value equals to 7–25% [12] and 10% [15], respectively. At the same time, it was indicated in [15] that all compositions of Bi2−xFex(FeSb)O7 (x = 0.1, 0.2, 0.3) demonstrate the state of spin glass due to the presence of some Fe(III) ions in crystallographic positions A. In the work [16], a sample of Bi1.6Co0.8Ta1.6O7.2 (sp. gr. Fd-3m:2, 10.5526(2) Å, Z = 8 (No. 227)) containing an admixture of BiTaO4 was obtained and characterized for the first time by the solid-phase synthesis method. The sample is characterized by a dense microstructure with a small number of pores formed by partially intergrown grains. The electronic state of the atoms was studied by the XPS and NEXAFS methods. According to spectroscopy data, bismuth atoms have an effective charge of +3, cobalt atoms +2, and tantalum ions +(5-δ). Co-doped bismuth magnesium tantalate with a pyrochlore structure (space group Fd-3m) was synthesized in the work [17]. Using X-ray phase analysis, it was established that single-phase Bi2Mg1−xCoxTa2O9 samples are formed at x < 0.7. At a higher cobalt content in the samples, the impurity phase β-BiTaO4 is detected, the amount of which is proportional to the degree of cobalt doping. The unit cell parameter of the Co,Mg co-doped bismuth tantalate phase increases with increasing content of cobalt ions in the samples from 10.5412(8) (x = 0.3) to 10.5499(8) Å (x = 0.7). The electronic state of ions in Bi2Mg1−xCoxTa2O9 was studied using X-ray spectroscopy. According to NEXAFS and XPS data, it was established that doping with cobalt and magnesium does not change the oxidation state of bismuth and tantalum in the pyrochlore. The ions have the oxidation states Bi(+3), Mg(+2), and Ta(+5). Meanwhile, in the Ta4f-, Ta5p-, and Ta4d-spectra of the Bi2Mg1−xCoxTa2O9 samples, an energy shift of the absorption bands towards lower energies is observed, which is characteristic of tantalum ions with an effective charge (+5-δ). In the XPS-Bi4f and -Bi5d spectra, a shift of bands to lower energies is detected, due to the distribution of some low-charge Co(II) ions in the bismuth position. According to NEXAFS spectroscopy data, the oxidation state of cobalt ions is predominantly 2+ and 3+. Similar cobalt pyrochlores based on bismuth niobate have been studied in detail in a wide concentration range for the Bi2O3–CoO–Nb2O5 system [7,18,19]. The single-phase pyrochlores were found to have Bi-deficient stoichiometry, with 8 to 25% of the A positions occupied by Co ions. In addition, the Bi and O atoms are shifted from ideal positions to the 96g and 32e positions by 0.34 Å and 0.45 Å, respectively. It was noted that a similar disordering of the structure is observed for pyrochlores in the Bi–M–Nb–O system (M = Zn, Fe, or Mn). It was also shown that a weak antiferromagnetic exchange is realized between paramagnetic atoms, and cobalt atoms have an effective charge of +(2.2).
New studies of pyrochlores based on bismuth tantalate/niobate have shown that pyrochlores are successfully synthesized in which the octahedral sublattice can simultaneously accommodate ions of seven different transition elements. The possibility of synthesizing phase-pure pyrochlores doped with Cr, Mn, Fe, Co, Ni, Cu, and Zn ions was described for the first time in the literature [20,21,22]. Several studies of complex oxide tantalates have shown [20,22] that they are low-porosity ceramics with moderate values (compared to complex niobates) of relative permittivity and low dielectric losses. As studies have shown [22], the permittivity of the complex oxide Bi2-1/3Cr1/6Mn1/6Fe1/6Co1/6Ni1/6Cu1/6Zn1/6Ta2O9+Δ is 49, and the dielectric losses are 0.004, showing average values of dielectric parameters compared to monodoped pyrochlores. The authors of the work suggested that the individual effect of transition element ions in multi-element pyrochlores is leveled by the combined effect of all dopants. Despite the growing interest of researchers in multi-element oxide systems, the degree and nature of the interaction of dissimilar ions with each other and their effect on the properties of oxide ceramics as a whole have not been finally clarified. It should be noted that the properties of two- and three-component (by the number of transition elements as dopants) oxide systems are currently being actively studied as promising catalysts and anodes for lithium batteries [23,24,25,26]. In [27], it was shown that the two-component pyrochlore Fe,Co-doped bismuth tantalate exhibits tunable dielectric relaxation due to variable oxidation states of the Fe3+/Co2+ ions. In addition, the samples exhibit photocatalytic activity with respect to organic dyes under the influence of visible radiation. It was found that a pyrochlore with a low cobalt content showed better photocatalytic activity, and a pyrochlore with a high cobalt content demonstrated a higher photoinduced current density. In our work, the possibility of solid-phase synthesis of the phase-pure pyrochlore Bi2−xFe1/2Co1/2Ta2O9+Δ with an equivalent content of cobalt and iron was demonstrated; oxidation states of transition element ions in the ceramics were analyzed based on X-ray spectroscopy data (Near Edge X-ray Absorption Fine Structure (NEXAFS) and X-ray Photoelectron Spectroscopy (XPS)).

2. Experimental Part

For solid-phase synthesis of Bi2−xFe1/2Co1/2Ta2O9+Δ (x = 0; 0.135) samples, stoichiometric amounts of bismuth (III), tantalum (V), iron (III), and cobalt (II, III) oxides of analytical grade were used (Acros Organics BVBA, Geel, Belgium). The stoichiometric mixture of oxides was thoroughly homogenized in an agate mortar for one hour, then pressed into disks using a manual plexiglass press (pressure 5 MPa). The samples were successively calcined in 4 stages at temperatures of 650, 850, 950, and 1050 °C for 15 h at each stage of heat treatment. The temperature regime was selected based on the results of a study of the phase formation of pyrochlores based on bismuth tantalate [9,20]. After each calcination stage, the sample was thoroughly homogenized and pressed again in the form of disks to ensure tight contact of the ceramic grains. X-ray data were obtained using a Shimadzu 6000 X-ray diffractometer (Kyoto, Japan) (Cu radiation; 2θ = 10–80°; Δ2θ = 0.05°; scanning rate 2.0°/min). Analysis of the experimental the X-ray diffraction pattern was performed using the Powder Cell program V. 2.3, Berlin, Germany [28]. To calculate the unit cell parameter, the WinCSD software package, V. 4.0, Russia, was used [29]. The analysis of the impurity content in the samples was carried out by the Rietveld method using the Topas V. 5.0 software package [30].
The microstructure, local chemical analysis, and elemental mapping of the sample surface were studied using scanning electron microscopy and energy-dispersive X-ray spectroscopy (Tescan VEGA 3LMN scanning electron microscope, INCA Energy 450 energy-dispersive spectrometer, Tescan, Czech Republic). NEXAFS spectroscopy studies were carried out at the NanoPES beamline of the KISI synchrotron radiation source at the Kurchatov Institute (Moscow, Russia) [31]. NEXAFS spectra were obtained by total electron yield (TEY) recording with an energy resolution of 0.5 eV and 0.7 eV in the regions of the Fe2p and Co2p absorption edges, respectively. To construct NEXAFS spectra we used the built-in software (Atom V. 1.0, Russia) on the NanoPES experimental station. The X-ray Photoelectron Spectroscopy (XPS) study was performed on a Thermo Scientific ESCALAB 250Xi X-ray spectrometer (Thermo Fisher Scientific, Great Britain, UK). An Al Kα X-ray tube (1486.6 eV) was used as a source of ionizing radiation. In the experiments, an ion-electron charge compensation system was used to neutralize the charge of the sample. All peaks were calibrated with respect to the C1s peak at 284.6 eV. The experimental data were processed using the ESCALAB 250 Xi software (Thermo Fisher Scientific, Great Britain, UK). Standards for XPS are precursors for solid-phase synthesis; purchased oxides were α-Bi2O3 (monoclinic, sp. gr. P21/c), β-Ta2O5 (orthorhombic, sp. gr. Pna2), and Co3O4 (sp. gr. Fd-3m), α-Fe2O3 (sp. gr. R-3c), grade pure for analysis, from Acros Organics BVBA, Belgium.

3. Results and Discussion

3.1. Synthesis and Microstructure

As shown by X-ray phase analysis, as a result of solid-phase synthesis of a sample of the Bi2Fe1/2Co1/2Ta2O9+Δ composition, a two-phase preparation was reproducibly formed, containing, in addition to the pyrochlore, an impurity phase of bismuth orthotantalate of the triclinic modification β-BiTaO4 (sp. gr. P-1) in an amount of 6.56 mol.% (Figure 2). As shown by early studies [22], the formation of bismuth orthotantalate is associated with the placement of a part of the dopants—ions of transition elements in the cationic sublattice of bismuth(III). Low-charge cations, the ionic radius of which is significantly larger than the ionic radius of tantalum(V), preferring tetrahedral coordination (Zn2+, Co2+, Cu+, and Mn2+), are distributed into the crystallographic positions of bismuth. In works [9,20] it was demonstrated that the synthesis of a single-phase sample is possible if additional vacancies are created in the bismuth sublattice by an amount proportional to the amount of the β-BiTaO4 impurity. For our case, based on the exact value of the mole fraction of the impurity, 6.56 mol.% β-BiTaO4, the defective bismuth composition of the single-phase ceramics Bi2−xFe1/2Co1/2Ta2O9+Δ were calculated. Solid-phase synthesis of the Bi2−xFe1/2Co1/2Ta2O9+Δ sample (x = 0.135) was carried out under the same conditions as Bi2Fe1/2Co1/2Ta2O9+Δ, described in the experimental part. The X-ray diffraction pattern of the sample is shown in Figure 1. As shown by the X-ray phase analysis, the synthesized sample Bi1.865Fe1/2Co1/2Ta2O9+Δ does not contain impurity phases. It is interesting to note that the synthesis of the pyrochlore of the Bi2CoTa2O9 composition also led to the formation of a two-phase preparation. Taking into account the amount of the β-BiTaO4 impurity, it was possible to synthesize a single-phase pyrochlore Bi1.862CoTa2O9 in [9]. Comparing these two compositions, we see that the coefficients for the bismuth atom absolutely coincide within the error limits of calculating the impurity content. In this case, several conclusions can be made. Obviously, Co(II) ions are distributed in the bismuth position in the Bi1.865Fe1/2Co1/2Ta2O9+Δ and Bi1.862CoTa2O9 samples; these ions have a large ionic radius compared to the ionic radius of tantalum (V) (R(Co(II))c.n-6 = 0.745 Å and R(Ta(V))c.n-6 = 0.64 Å)) [29]. Cobalt(II) ions are distributed in the bismuth position in equal amounts for the two compositions Bi1.865Fe1/2Co1/2Ta2O9+Δ and Bi1.862CoTa2O9. This indicates a non-random (non-statistical) distribution of ions over the cation sublattices of bismuth/tantalum and the presence of a pattern in the distribution of cations. It should be noted that the data are highly reproducible, since the samples were not synthesized simultaneously, but with a difference of several years. Similar behavior of the cobalt(II) ions in the two compositions can be explained by the proximity of the ionic radii of Ta(V) and Fe(III) (R(Fe(III))c.n-6 = 0.645 Å and R(Ta(V))c.n-6 = 0.64 Å)) in octahedral coordination. Moreover, it can be assumed that the ionic and spin composition (Co(II),Co(III)) of cobalt is almost the same in both samples.
Calculation of the unit cell parameter of the pyrochlore Bi1.865Fe1/2Co1/2Ta2O9+Δ showed a value of 10.5013(8) Å, which is significantly less than the unit cell parameter of 10.54051(3) Å for the cobalt-containing pyrochlore based on bismuth tantalate Bi1.49Co0.8Ta1.6O7.0 (Bi1.862CoTa2O9) [9], which, with the same bismuth content, is associated with a significant difference in the radii of iron (III) and cobalt (II) ions (R(Fe(III))c.n.-6 = 0.645 Å and R(Co(II))c.n.-6 = 0.645 Å) [32]. The unit cell parameter of 10.4969 Å for the iron-containing pyrochlore Bi2FeTa2O9+Δ is comparable with that calculated for Bi1.865Fe1/2Co1/2Ta2O9+Δ [33]. This is due to the fact that the lack of bismuth(III) ions in the composition of the studied sample is compensated by the difference in the radii of the iron(III) and cobalt(II) ions. The unit cell parameter of the pyrochlore Bi1.865Fe1/2Co1/2Ta2O9+Δ calculated by us corresponds to the values given in the article [10] for iron-containing pyrochlores based on bismuth tantalate (10.4979–10.5033 Å). Elemental mapping of the Bi1.865Fe1/2Co1/2Ta2O9+Δ sample showed a uniform distribution of element atoms over the sample surface (Figure 3), and EDS analysis confirmed that the experimental sample composition corresponded to the specified one. According to the EDS spectrum, the chemical composition of the sample corresponds to the formula Bi1.74Fe0.47Co0.48Ta2O9+Δ and corresponds to the nominal composition.
Microphotographs of the surface of the sample synthesized at 1050 °C are shown in Figure 4. The porous, loose microstructure is formed by partially intergrown grains. Individual small crystallites and large agglomerates up to 4 μm are recorded as a result of grain intergrowth. The average crystallite size determined by X-ray diffraction using the Scherrer formula is ~67 nm, while larger crystallites in the range of 0.5–1 μm were determined using a scanning electron microscope (SEM). Apparently, the crystallites in the micrographs are aggregated ceramic grains of significantly smaller sizes.
It is interesting to note that the microstructure of the cobalt pyrochlore of the same bismuth stoichiometry is less porous, coarse grains are fused with each other, and the grain size is 1–3 μm [9]. Meanwhile, the microstructure of iron-containing preparations is porous [33] and formed by partially fused weakly faceted grains. The sizes of round crystallites, determined by the Scherrer method, are ∼49 nm. Grains with a longitudinal size of 0.2–2 μm were recorded by scanning electron microscopy. Taking into account the above, the microstructure of the ceramics is affected by the relative content of bismuth, iron, and cobalt ions. Apparently, cobalt and bismuth ions contribute to the production of dense, low-porosity ceramics, while iron ions contribute to the production of loose, fine-grained ceramics. This important observation will allow targeted regulation of the microstructure of ceramics by chemical modification of the pyrochlore composition.

3.2. XPS and NEXAFS Spectroscopy

The XPS spectra of bismuth tantalate co-doped with cobalt and iron atoms Bi1.865Fe1/2Co1/2Ta2O9+Δ are shown in Figure 5a–e. The spectra of the initial oxides, the precursors for the synthesis, are also shown for comparison. Figure 5a shows the XPS spectrum in a wide energy range, and Figure 5b–e shows the spectral dependences in the region of the Bi4f, Bi5d, Ta4f, Co2p, and Fe2p ionization thresholds of the studied pyrochlore. Table 1 shows the energy positions of the XPS spectra components. The figures also show the results of decomposition of spectral dependences into individual peaks modeled by Gauss–Lorentz curves, and the background lines were modeled by the Shirley or smart approximation. The Survey XPS spectrum contains a C1s peak, which is associated with the presence of random surface organic contaminants on the surface. The presence of organic impurities contributes to the intensity of the O1s peak and distorts it. In this regard, the analysis of the surface composition of the samples was carried out based on the analysis of the spectra of metal ions.
Comparison of XPS-Bi4f and -Bi5d spectra of the studied sample and Bi2O3 oxide (Figure 5b) showed that the energy position and peak width in the spectrum of the sample almost completely coincide with the spectrum of Bi2O3 oxide. This suggests that the bismuth cation in the pyrochlore has a charge state of Bi3+. The shape of the peaks in the spectrum of tantalum ions (Figure 5c) clearly indicates that all cations are in the same charge state (there is no splitting or distortion of the peaks), while the energy position of the peaks has a characteristic shift toward lower energies compared to the binding energy in pentavalent tantalum oxide Ta2O5. The shift toward lower energies is characteristic of a decrease in the effective positive charge; in particular, for the Ta4f and Ta5p spectra presented by us, this energy shift is ΔE = 0.75 eV. This, in turn, allows us to assume that tantalum atoms have the same effective charge +(5-δ).
The energy range of the Co2p spectrum (Figure 5d) of the sample includes a peak responsible for the binding energy of the Ta4p1/2 level, which somewhat complicates the perception of the cobalt spectrum. Meanwhile, when comparing the pyrochlore spectrum with the Co3O4 spectrum we obtained and the CoO spectrum known from the literature [34], it can be noted that the energy position of the main peaks in all the spectra presented is almost the same. Bands characteristic of Co(II) are observed at 781 and 797 eV. At the same time, the spectrum of the pyrochlore and cobalt oxides contains satellite peaks, the energy position of which in the oxides is different. In the CoO spectrum, they are shifted toward lower energies and are more intense. Based on the position and intensity of the satellite lines in the pyrochlore spectrum, one can conclude that the spectra of CoO and pyrochlore correspond. This means that the charge state of cobalt ions in the pyrochlore is predominantly +2. In the XPS-Fe2p spectra (Figure 5e) the situation is different from that in the cobalt spectra. Comparison of the Fe2p spectrum in the pyrochlore with the spectra of FeO [35,36] and Fe2O3 oxides shows that the binding energy of the Fe2p3/2 and Fe2p1/2 levels (711 and 724 eV) and the energy position of the satellite peaks (719 and 733 eV) in the pyrochlore coincide with the spectrum of Fe2O3. This allows us to unambiguously determine the charge state of iron atoms in the composites as Fe3+. NEXAFS spectra of the Bi1.865Fe1/2Co1/2Ta2O9+Δ sample are shown in Figure 6.
When analyzing NEXAFS spectra, it is necessary to take into account that absorption spectra have a richer structure compared to XPS spectra, since they are associated with transitions of internal electrons of atoms (in this case, 2p electrons of iron and cobalt atoms) to free valence states of these atoms [37]. In this case, in accordance with the dipole selection rules, these states should be formed from 3d-orbitals. As is known, it is 3d electrons that participate in the formation of covalent bonds in the compounds under consideration, and accordingly, NEXAFS 2p spectra are most sensitive to changes in the environment of 3d-atoms.
When comparing the NEXAFS Fe 2p3/2 spectrum of the pyrochlore and the spectra of iron oxides, it is evident (Figure 6a) that the energy position and intensity ratio of the absorption bands at 709 and 708 eV in the pyrochlore coincide with the characteristics of the bands in the Fe 2p3/2 spectrum of Fe2O3 oxide, which also confirms the state of iron oxides as Fe(III). In the Co 2p3/2 spectrum of the pyrochlore (Figure 6b), a peak at 779 and a shoulder at 781 eV and 778 eV appear, which in terms of energy position and intensity ratio correlate well with the main features of the CoO spectrum. Based on this, it can be concluded that cobalt ions in pyrochlore are predominantly in the Co(II) state.

4. Conclusions

The possibility of forming a Fe/Co co-doped pyrochlore with equivalent iron and cobalt content was investigated. As a result, two samples of Bi2−xCo1/2Fe1/2Ta2O9+Δ (x = 0; 0.135) were synthesized by the solid-phase reaction method. According to the XRD data, the Bi2Co1/2Fe1/2Ta2O9+Δ sample is not single-phase and contains an impurity of bismuth orthotantalate in the amount of 6.56%. We associate the formation of the impurity with the distribution of a part of the cobalt(II) cations in the bismuth sublattice. In our work, we were able to prevent the formation of the impurity by creating vacancies in the bismuth sublattice proportional to the amount of BiTaO4 (x = 0.135). In this regard, the Bi1.865Co1/2Fe1/2Ta2O9+Δ sample (sp. gr. Fd-3m) turned out to be phase-pure. According to the XPS data, a characteristic shift of the Ta4f and Ta5p spectra to lower energies by 0.75 eV is observed, which may be due to the distribution of low-charge cobalt and iron cations in the octahedral positions of tantalum (V). The effective charge of tantalum cations is +(5-δ) and that of bismuth cations is +3. According to the character of the Co2p XPS spectrum of the complex oxide, cobalt ions are predominantly in the charge state of +2. According to the XPS data, the effective charge of iron cations is +3.

Author Contributions

Conceptualization, N.A.Z.; Funding acquisition, S.V.N. Investigation, N.A.Z., B.A.M., A.M.L., A.V.K. and S.V.N.; Resources, B.A.M., O.V.P., A.M.L. and A.V.K.; Validation, N.A.Z.; Visualization, N.A.Z. and S.V.N.; Writing—original draft, N.A.Z. and S.V.N.; Formal analysis, O.V.P. All authors have read and agreed to the published version of the manuscript.

Funding

The NEXAFS study was carried out within the framework of the state budget topic 122040400069-8 and also with the financial support of the Ministry of Science and Higher Education of Russia under agreement No. 075-15-2021-1351.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The XPS studies were performed on the equipment of the Resource Center “Physical methods of surface investigation” of the Scientific Park of St. Petersburg University. The NEXAFS studies were performed on the synchrotron radiation from the station “NanoPES” storage ring (National Research Center “Kurchatov Institute”). XRD studies were conducted at the Center for X-Ray Diffraction Studies of the Research Park of St. Petersburg State University within the project AAAA-A19-119091190094-6. The authors thank M.G. Krzhizhanovskaya for analyzing the X-ray powder diffraction using the Rietveld method.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Giampaoli, G.; Siritanon, T.; Day, B.; Li, J.; Subramanian, M.A. Temperature in-dependent low loss dielectrics based on quaternary pyrochlore oxides. Prog. Solid State Chem. 2018, 50, 16–23. [Google Scholar] [CrossRef]
  2. Du, H.; Yao, X. Structural trends and dielectric properties of Bi-based pyrochlores. J. Mater. Sci. Mater. Electron. 2004, 15, 613–616. [Google Scholar]
  3. Subramanian, M.A.; Aravamudan, G.; Rao Subba, G.V. Oxide pyrochlores—A review. Prog. Solid State Chem. 1983, 15, 55–143. [Google Scholar] [CrossRef]
  4. Lufaso, M.W.; Vanderah, T.A.; Pazos, I.M.; Levin, I.; Roth, R.S.; Nino, J.C.; Provenzano, V.; Schenck, P.K. Phase formation, crystal chemistry, and properties in the system Bi2O3–Fe2O3–Nb2O5. J. Solid State Chem. 2006, 179, 3900–3910. [Google Scholar] [CrossRef]
  5. Vanderah, T.A.; Lufaso, M.W.; Adler, A.U.; Levin, I.; Nino, J.C.; Provenzano, V.; Schenck, P.K. Subsolidus phase equilibria and properties in the system Bi2O3:Mn2O3±x:Nb2O5. J. Solid State Chem. 2006, 179, 3467–3477. [Google Scholar] [CrossRef]
  6. Nguyen, H.B.; Noren, L.; Liu, Y.; Withers, R.L.; Wei, X.R.; Elcombe, M.M. The disordered structures and low temperature dielectric relaxation properties of two misplaced-displacive cubic pyrochlores found in the Bi2O3-MO-Nb2O5 (M = Mg, Ni) systems. J. Solid State Chem. 2007, 180, 2558–2565. [Google Scholar] [CrossRef]
  7. Vanderah, T.A.; Siegrist, T.; Lufaso, M.W.; Yeager, M.C.; Roth, R.S.; Nino, J.C.; Yates, S. Phase Formation and Properties in the System Bi2O3:2CoO1+x:Nb2O5. Eur. J. Inorgan. Chem. 2006, 2006, 4908–4914. [Google Scholar] [CrossRef]
  8. Zhuk, N.A.; Sekushin, N.A.; Krzhizhanovskaya, M.G.; Kharton, V.V. Multiple relaxation, reversible electrical breakdown and bipolar conductivity of pyrochlore–type Bi2Cu0.5Zn0.5Ta2O9 ceramics. Solid State Ion. 2022, 377, 115868. [Google Scholar] [CrossRef]
  9. Zhuk, N.A.; Krzhizhanovskaya, M.G.; Sekushin, N.A.; Sivkov, D.V.; Abdurakhmanov, I.E. Crystal structure, dielectric and thermal properties of cobalt doped bismuth tantalate pyrochlore. J. Mater. Res. Technol. 2023, 22, 1791–1799. [Google Scholar] [CrossRef]
  10. Jusoh, F.A.; Tan, K.B.; Zainal, Z.; Chen, S.K.; Khaw, C.C.; Lee, O.J. Novel pyrochlores in the Bi2O3-Fe2O3-Ta2O5 (BFT) ternary system: Synthesis, structural and electrical properties. J. Mater. Res. Techn. 2020, 9, 11022–11034. [Google Scholar] [CrossRef]
  11. Zhuk, N.A.; Krzhizhanovskaya, M.G.; Koroleva, A.V.; Reveguk, A.A.; Sivkov, D.V.; Nekipelov, S.V. Thermal expansion, crystal structure, XPS and NEXAFS spectra of Fe-doped bismuth tantalate pyrochlore. Ceram. Intern. 2022, 48, 14849–14855. [Google Scholar] [CrossRef]
  12. Egorysheva, A.V.; Ellert, O.G.; Maksimov, Y.V.; Volodin, V.D.; Efimov, N.N.; Novotortsev, V.M. Subsolidus phase equilibria and magnetic characterization of the pyrochlore in the Bi2O3–Fe2O3–Sb2Ox system. J. Alloys Compd. 2013, 579, 311–314. [Google Scholar] [CrossRef]
  13. Matsuda, C.K.; Barco, R.; Sharma, P.; Biondo, V.; Paesano, A.; da Cunha, J.B.M.; Hallouche, B. Iron-containing pyrochlores: Structural and magnetic characterization. Hyperfine Interact. 2007, 175, 55–61. [Google Scholar] [CrossRef]
  14. Filoti, G.; Rosenberg, M.; Kuncser, V.; Seling, B.; Fries, T.; Spies, A.; KemmlerSack, S. Magnetic properties and cation distribution in iron containing pyrochlores. J. Alloys Comp. 1998, 268, 16–21. [Google Scholar] [CrossRef]
  15. Whitaker, M.J.; Marco, J.F.; Berry, F.J.; Raith, C.; Blackburn, E.; Greaves, C. Structural and magnetic characterisation of the pyrochlores Bi2−xFex(FeSb)O7, (x = 0.1, 0.2, 0.3), Nd1.8Fe0.2(FeSb)O7 and Pr2(FeSb)O7. J. Solid State Chem. 2013, 198, 316–322. [Google Scholar] [CrossRef]
  16. Zhuk, N.A.; Krzhizhanovskaya, M.G.; Koroleva, A.V.; Nekipelov, S.V.; Sivkov, D.V.; Sivkov, V.N.; Lebedev, A.M.; Chumakov, R.G.; Makeev, B.A.; Kharton, V.V.; et al. Spectroscopic characterization of cobalt doped bismuth tantalate pyrochlore. Solid State Sci. 2022, 125, 106820. [Google Scholar] [CrossRef]
  17. Zhuk, N.A.; Makeev, B.A.; Koroleva, A.V.; Nekipelov, S.V.; Petrova, O.V. NEXAFS and XPS Studies of Co Doped Bismuth Magnesium Tantalate Pyrochlores. Chemistry 2024, 6, 323–332. [Google Scholar] [CrossRef]
  18. Piir, I.V.; Prikhodko, D.A.; Ignatchenko, S.V.; Schukariov, A.V. Preparation and structural investigations of the mixed bismuth niobates, containing transition metals. Solid State Ion. 1997, 101–103, 1141–1146. [Google Scholar] [CrossRef]
  19. Smolenskii, G.A.; Isupov, V.A.; Golovshchifova, G.I.; Tutov, A.G. New compounds with pyrochlore-type structure and their dielectric properties. Izv. Akad. Nauk. SSSR Neorg. Mater. 1976, 12, 255–258. [Google Scholar]
  20. Rylchenko, E.P.; Makeev, B.A.; Sivkov, D.V.; Korolev, R.I.; Zhuk, N.A. Features of phase formation of pyrochlore-type Bi2Cr1/6Mn1/6Fe1/6Co1/6Ni1/6Cu1/6Ta2O9+Δ. Lett. Mater. 2022, 12, 486–492. [Google Scholar] [CrossRef]
  21. Zhuk, N.A.; Makeev, B.A.; Krzhizhanovskaya, M.G.; Nekipelov, S.V.; Sivkov, D.V.; Badanina, K.A. Features of the Phase Formation of Cr/Mn/Fe/Co/Ni/Cu Codoped Bismuth Niobate Pyrochlore. Crystals 2023, 13, 1202. [Google Scholar] [CrossRef]
  22. Parshukova, K.N.; Sekushin, N.A.; Makeev, B.A.; Krzhizhanovskaya, M.G.; Koroleva, A.V.; Zhuk, N.A. Synthesis and dielectric properties, XPS spectroscopy study of high-entropy pyrochlore. Lett. Mater. 2022, 12, 469–474. [Google Scholar] [CrossRef]
  23. Sukhanov, K.S.; Gilev, A.R.; Kiselev, E.A.; Cherepanov, V.A. Functional properties and structure-size factor in La1.4A0.6Ni0.6Fe0.4O4+δ (A = Ca, Sr, Ba). J. Alloys Comp. 2024, 990, 174369. [Google Scholar] [CrossRef]
  24. Aksenova, T.V.; Mysik, D.K.; Cherepanov, V.A. Crystal Structure and Properties of Gd1−xSrxCo1-yFeyO3-δ Oxides as Promising Materials for Catalytic and SOFC Application. Catalysts 2022, 12, 1344. [Google Scholar] [CrossRef]
  25. Gilev, A.R.; Kiselev, E.A.; Sukhanov, K.S.; Korona, D.V.; Cherepanov, V.A. Evaluation of La2−x(Ca/Sr)xNi1-yFeyO4+δ (x = 0.5, 0.6; y = 0.4, 0.5) as cathodes for proton-conducting SOFC based on lanthanum tungstate. Electrochim. Acta 2022, 421, 140479. [Google Scholar] [CrossRef]
  26. Shevchenko, V.A.; Komayko, A.I.; Sivenkova, E.V.; Samigullin, R.R.; Skvortsova, I.A.; Abakumov, A.M.; Nikitina, V.A.; Drozhzhin, O.A.; Antipov, E.V. Effect of Ni/Fe/Mn ratio on electrochemical properties of the O3–NaNi1-x-yFexMnyO2 (0.25 ≤ x, y ≤ 0.75) cathode materials for Na-ion batteries. J. Power Sources 2024, 596, 234092. [Google Scholar] [CrossRef]
  27. Sun, S.H.; Xue, Y.; Yang, D.; Pei, Z.; Fang, L.; Xia, Y.; Ti, R.; Wang, C.; Liu, C.; Xiong, B.; et al. Bismuth pyrochlores with varying Fe/Co ratio for efficient Multi-Functional Catalysis: Structure evolution versus Photo- and Electro-catalytic activities. Chem. Eng. J. 2022, 448, 137580. [Google Scholar] [CrossRef]
  28. Kraus, W.; Nolze, D.G. POWDER CELL—A program for the representation and manipulation of crystal structures and calculation of the resulting X-ray powder patterns. J. Appl. Cryst. 1996, 29, 301–303. [Google Scholar] [CrossRef]
  29. Akselrud, L.G.; Grin, Y.N.; Zavalij, Y.P. CSD-universal program package for single crystal or powder structure data treatment. In Proceedings of the 12th European Crystallographic Meeting, Moscow, Russia, 20–29 August 1989; p. 155. [Google Scholar]
  30. Bruker, A.X.S. Topas, V.5.0 Software. General Profile and Structure Analysis Software for Powder Diffraction Data; Bruker AXS: Karlsruhe, Germany, 2014.
  31. Lebedev, A.M.; Menshikov, K.A.; Nazin, V.G.; Stankevich, V.G.; Tsetlin, M.B.; Chumakov, R.G. NanoPES Photoelectron Beamline of the Kurchatov Synchrotron Radiation Source. J. Surf. Investig. X-ray Synchrotron Neutron Tech. 2021, 15, 1039–1044. [Google Scholar] [CrossRef]
  32. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  33. Zhuk, N.A.; Sekushin, N.A.; Semenov, V.G.; Fedorova, A.V.; Selyutin, A.A.; Krzhizhanovskaya, M.G.; Lutoev, V.P.; Makeev, B.A.; Kharton, V.V.; Sivkov, D.N.; et al. Dielectric properties, Mössbauer study, ESR spectra of Bi2FeTa2O9.5 with pyrochlore structure. J. Alloys Comps. 2022, 903, 163928. [Google Scholar] [CrossRef]
  34. Hassel, M.; Freund, H.-J. High Resolution XPS Study of a Thin CoO(111) Film Grown on Co(0001). Surf. Sci. Spectr. 1996, 4, 273–278. [Google Scholar] [CrossRef]
  35. Regan, T.J.; Ohldag, H.; Stamm, C.; Nolting, F.; Luning, J.; Stöhr, J.; White, R.L. Chemical effects at metal/oxide interfaces studied by x-ray-absorption spectroscopy. Phys. Rev. B 2001, 64, 214422. [Google Scholar] [CrossRef]
  36. Yamashita, T.; Hayes, P. Analysis of XPS spectra of Fe2+ and Fe3+ ions in oxide materials. Appl. Surf. Sci. 2008, 4, 2441–2449. [Google Scholar] [CrossRef]
  37. Stohr, J. NEXAFS Spectroscopy; Springer: Berlin/Heidelberg, Germany, 1992; 662p. [Google Scholar]
Figure 1. Unit cell of the A2B2O7 pyrochlore (sp.gr. Fd-3m).
Figure 1. Unit cell of the A2B2O7 pyrochlore (sp.gr. Fd-3m).
Chemistry 06 00062 g001
Figure 2. X-ray diffraction powder patterns of Bi1.865Fe1/2Co1/2Ta2O9+Δ and Bi2Fe1/2Co1/2Ta2O9+Δ. * denotes the reflections of the impurity phase—bismuth orthotantalate.
Figure 2. X-ray diffraction powder patterns of Bi1.865Fe1/2Co1/2Ta2O9+Δ and Bi2Fe1/2Co1/2Ta2O9+Δ. * denotes the reflections of the impurity phase—bismuth orthotantalate.
Chemistry 06 00062 g002
Figure 3. EDS spectrum and elemental maps of the Bi1.865Fe1/2Co1/2Ta2O9+Δ samples.
Figure 3. EDS spectrum and elemental maps of the Bi1.865Fe1/2Co1/2Ta2O9+Δ samples.
Chemistry 06 00062 g003
Figure 4. Microphotograph of the Bi1.865Fe1/2Co1/2Ta2O9+Δ surface in the secondary and elastically reflected electron modes.
Figure 4. Microphotograph of the Bi1.865Fe1/2Co1/2Ta2O9+Δ surface in the secondary and elastically reflected electron modes.
Chemistry 06 00062 g004
Figure 5. Survey XPS spectrum of the Bi1.865Co1/2Fe1/2Ta2O9+Δ (CoFe) (a); XPS-Bi4f spectra of Bi2O3 and CoFe (b); XPS spectra of tantalum and bismuth cations in Bi2O3,Ta2O5, and CoFe (c); XPS-Co2p spectra of CoO, Co3O4, and CoFe (d); and XPS-Fe2p spectra of FeO, Fe2O3, and CoFe (e).
Figure 5. Survey XPS spectrum of the Bi1.865Co1/2Fe1/2Ta2O9+Δ (CoFe) (a); XPS-Bi4f spectra of Bi2O3 and CoFe (b); XPS spectra of tantalum and bismuth cations in Bi2O3,Ta2O5, and CoFe (c); XPS-Co2p spectra of CoO, Co3O4, and CoFe (d); and XPS-Fe2p spectra of FeO, Fe2O3, and CoFe (e).
Chemistry 06 00062 g005
Figure 6. NEXAFS-Fe2p (a) and -Co2p spectra (b) of the Bi1.865Co1/2Fe1/2Ta2O9+Δ.
Figure 6. NEXAFS-Fe2p (a) and -Co2p spectra (b) of the Bi1.865Co1/2Fe1/2Ta2O9+Δ.
Chemistry 06 00062 g006
Table 1. Energy positions of the components of the XPS spectra for Bi1.865Co1/2Fe1/2Ta2O9+Δ.
Table 1. Energy positions of the components of the XPS spectra for Bi1.865Co1/2Fe1/2Ta2O9+Δ.
PeakEnergy (eV)PeakEnergy (eV)
Bi4f7/2158.78Co2p3/2780.16
Bi4f5/2164.10Co2p1/2795.93
Bi5d5/225.84Co2p sat785.27
Bi5d3/228.84Fe2p3/2710.56
Ta4f7/225.41Fe2p1/2724.32
Ta4f5/227.31Fe2p sat718.60
Ta5p3/235.78Fe2p sat733.22
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhuk, N.A.; Nekipelov, S.V.; Petrova, O.V.; Koroleva, A.V.; Lebedev, A.M.; Makeev, B.A. Synthesis and NEXAFS and XPS Characterization of Pyrochlore-Type Bi1.865Co1/2Fe1/2Ta2O9+Δ. Chemistry 2024, 6, 1078-1088. https://doi.org/10.3390/chemistry6050062

AMA Style

Zhuk NA, Nekipelov SV, Petrova OV, Koroleva AV, Lebedev AM, Makeev BA. Synthesis and NEXAFS and XPS Characterization of Pyrochlore-Type Bi1.865Co1/2Fe1/2Ta2O9+Δ. Chemistry. 2024; 6(5):1078-1088. https://doi.org/10.3390/chemistry6050062

Chicago/Turabian Style

Zhuk, Nadezhda A., Sergey V. Nekipelov, Olga V. Petrova, Aleksandra V. Koroleva, Aleksey M. Lebedev, and Boris A. Makeev. 2024. "Synthesis and NEXAFS and XPS Characterization of Pyrochlore-Type Bi1.865Co1/2Fe1/2Ta2O9+Δ" Chemistry 6, no. 5: 1078-1088. https://doi.org/10.3390/chemistry6050062

APA Style

Zhuk, N. A., Nekipelov, S. V., Petrova, O. V., Koroleva, A. V., Lebedev, A. M., & Makeev, B. A. (2024). Synthesis and NEXAFS and XPS Characterization of Pyrochlore-Type Bi1.865Co1/2Fe1/2Ta2O9+Δ. Chemistry, 6(5), 1078-1088. https://doi.org/10.3390/chemistry6050062

Article Metrics

Back to TopTop