Next Article in Journal
Novel High-Entropy Aluminide-Silicide Alloy
Previous Article in Journal
Elasto-Plastic Fracture Mechanics Analysis of the Effect of Shot Peening on 300M Steel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Oxidation State and Local Structure of Chromium Ions in LaOCl

Institute of Solid State Physics, University of Latvia, LV-1063 Riga, Latvia
*
Author to whom correspondence should be addressed.
Materials 2021, 14(13), 3539; https://doi.org/10.3390/ma14133539
Submission received: 11 May 2021 / Revised: 22 June 2021 / Accepted: 23 June 2021 / Published: 25 June 2021
(This article belongs to the Section Advanced Materials Characterization)

Abstract

:
LaOCl doped with 0–10 mol% Cr was synthesized by thermal decomposition of chlorides. X-ray diffraction (XRD) analysis revealed that incorporation of chromium results in a decrease of the lattice parameter a and a simultaneous increase of the lattice parameter c. The local structure of chromium ions was studied with X-ray photoelectron (XPS), X-ray absorption (XANES), multifrequency electron paramagnetic resonance (EPR) and electron-nuclear double resonance (ENDOR) spectroscopy techniques. It was determined that synthesis in oxidizing atmosphere promotes the incorporation of chromium ions predominantly in the 5+ oxidation state. Changes of chromium oxidation state and local environment occur after a subsequent treatment in reducing atmosphere. Spin-Hamiltonian (SH) parameters for a Cr5+ and two types of Cr3+ centers in LaOCl were determined from the EPR spectra simulations.

1. Introduction

Lanthanum oxychloride (LaOCl), which is being considered for a variety of applications including gas sensors [1,2,3,4,5,6,7,8], catalysts [9,10,11], and solid electrolytes [12,13,14], is one of the most widely investigated compounds from the class of lanthanide oxyhalides. Historically, LaOCl has been primarily studied as a host matrix for rare-earth ions to establish the systematics of fluorescence spectra transition interrelations with crystal field parameters [15,16,17,18,19]. With appropriate activator combinations, it is an excellent phosphor, which can be tailored for efficient emission in spectral regions covering the whole visible [20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41] and a part of near-infrared regions [42,43].
LaOCl crystallizes in a tetragonal symmetry lattice with space group P4/nmm and cell parameters a = 4.12 Å and c = 6.88 Å [44,45]. As shown in Figure 1, the lattice consists of alternating cationic and anionic layers, which are arranged perpendicular to the c axis direction. La3+ ions occupy sites with C4v symmetry and are coordinated by four O2− ions from one side and five Cl ions from the other (with the fifth being in the next-to-nearest layer). Incorporation of trivalent rare-earth ions in LaOCl occurs via substitution of the single crystallographic La3+ position. Due to the relatively larger ionic radius of La3+, other trivalent rare-earth ions can be efficiently accommodated without distortions in cation site symmetry; however, a decrease in the unit cell parameters is expected [31,34,38,43]. The situation is more complicated for activators that have several stable oxidation states; for example, it is possible to stabilize both Eu3+ [27,28,29,30,31,32,33] and Eu2+ [33,34,35] ions in LaOCl during the synthesis. The larger Eu2+ ions expand the lattice and require charge compensation, possibly in the form of Cl ion vacancies [34]. The introduction of divalent cations, such as Ca2+ and Mg2+, is likely to produce lattice defects in LaOCl which can improve the optical [29] or electrical [13] properties of the material. Studies of Sb [46,47] and Bi [48] -doped LaOCl are especially interesting as the luminescence spectra showed the presence of two emission centers, while there is only one evident crystallographic site for substitution.
Despite the wide research performed on the host, the incorporation mechanisms of multi-valent impurities in the LaOCl lattice are not fully understood. Therefore, in this study we apply local structure and oxidation state sensitive X-ray and magnetic resonance spectroscopy techniques to elucidate the mechanism of multivalent chromium ion incorporation in LaOCl.
It is shown that during synthesis in oxidizing atmosphere, chromium ions enter into the structure of LaOCl predominantly in the 5+ state, producing asymmetric distortions of the lattice parameters. Local structure transformation of Cr upon annealing in reducing atmosphere is also discussed.

2. Experimental Details

LaOCl doped with 0–10 mol% Cr was prepared by thermal decomposition of chlorides. In a typical synthesis 3 mmol of LaOCl:Cr was prepared. An appropriate amount of CrCl3 6H2O (99.5%) was dissolved in deionized water. Afterwards, La2O3 (99.999%) and 36.5% HCl (99.999%) were added. The obtained suspension was heated to boiling point until all LaOCl was dissolved and the excess water and HCl evaporated. The obtained material, consisting of chloride hydrates, was transferred to an Al2O3 crucible and heat-treated at 700 °C for 2 h in air. The heat treatment promoted the thermal decomposition of chlorides and the formation of LaOCl:Cr. Three parallel samples were prepared for all investigated dopant concentrations to test the reproducibility of the synthesis. To induce changes in the oxidation state of chromium ions, selected LaOCl:Cr samples doped with 0.1, 1 and 5% were additionally heat-treated in H2/Ar (5%/95%) at 500–800 °C for 2 h.
X-ray diffraction (XRD) patterns were obtained by Rigaku MiniFlex 600 diffractometer (Rigaku, Tokyo, Japan) with Bragg-Brentano θ-2θ geometry equipped with a 600 W Cu anode (Cu Kα radiation, λ = 1.5406 Å) X-ray tube operated at 40 kV and 15 mA. The Rietveld analysis was performed using the Profex software [50]. The Vesta code [49] was used for the visualization of the crystal structure.
XPS analyses were carried out using the ThermoFisher ESCALAB Xi+ instrument (Waltham, MA, USA) with a monochromatic Al Kα X-ray source. The instrument binding energy scale was calibrated to give binding energy at 932.6 eV for Cu 2p3/2 line of freshly etched metallic copper. For charge compensation, a standard procedure of sample surface irradiation with a flood of electrons was carried out. The spectra were recorded using an X-ray beam size of 900 × 10 microns, pass energy of 20 eV, and a step size of 0.1 eV. Data from all samples are referenced using the main signal of the carbon 1s spectrum assigned to occur at 284.8 eV. The carbon 1s spectrum was collected using high-energy resolution settings.
X-ray absorption spectroscopy study of chromium ion environment in LaOCl:Cr powder samples was performed at P65 Applied XAFS undulator beamline [51] of the PETRA III storage ring. A fixed exit double-crystal Si(111) monochromator was used to scan the energy range from 5300 eV to 6500 eV, and the harmonic rejection was achieved by an uncoated silicon plane mirror. The powder sample was attached to the sticky side of a Kapton tape, which was placed at 45° to the beam propagation direction. X-ray absorption spectra were collected at room temperature in fluorescence mode using an ionization chamber (I0) located before the sample and passivated implanted planar silicon (PIPS) detector (If) placed at 90° to the incident beam. The X-ray absorption coefficient was calculated as µ(E) = If/I0. X-ray absorption near edge structure (XANES) calculations at the Cr K-edge were performed using ab initio real-space FDMNES code [52,53] within the full-multiple-scattering (FMS) approximation. Only dipole transitions were taken into account, and the energy-dependent real Hedin-Lundqvist exchange-correlation potential was used [52,53]. The relativistic FMS calculations were performed with a self-consistent muffin-tin potential. The calculated XANES spectra were broadened to account for the core-hole level width of chromium (Γ(K-Cr) = 1 eV [54]) and other multi-electronic phenomena. The energy origin was set at the Fermi level EF. All calculations were performed for sufficiently large clusters with a radius of 8 Å around the absorbing chromium atom, which were constructed from the crystallographic LaOCl structure [45] by placing the chromium atom at the required place.
EPR spectra measurements were carried out at room temperature on Bruker ELEXSYS-II E500 CW-EPR system (Bruker Biospin, Rheinstetten, Germany) at X (9.83 GHz frequency; 10 mW power) and Q (33.85 GHz; 5.7 mW) microwave bands. Magnetic field modulation amplitude was 0.4 mT for measurements at both bands. Spectra intensities for the X-band EPR measurements have been normalized to sample mass. Electron nuclear double resonance (ENDOR) spectra were acquired with the same CW-EPR spectrometer equipped with DICE-II CW ENDOR measurement system and Bruker EN 901 X-Band CW-ENDOR resonator mounted on Oxford Instruments liquid helium flow cryostat. The temperature during data acquisition was 10 K, the magnetic field was 345.37 mT, and the microwave frequency was 9.5013 GHz at 20 mW power. The radiofrequency modulation type was FM with 100 kHz modulation depth. The resulting spectrum is a sum of 10 ENDOR scans. EPR and ENDOR spectra simulations were performed in EasySpin software [55].

3. Results and Discussion

Figure 2 shows XRD patterns of LaOCl samples doped with 0–10% Cr. The most intense diffraction peaks can be assigned to tetragonal LaOCl. Trace amounts of orthorhombic LaCrO3 could be detected in the samples doped with 5–10% Cr. Quantitative analysis of phase composition using Rietveld refinement revealed that LaCrO3 content is 4.7 ± 1.4% for the sample doped with 10% Cr and less than 4% in the rest of the samples. LaOCl samples with Cr concentrations below 5%, where impurity phase LaCrO3 was below the detection limit, were selected for further spectroscopic studies.
Gradual changes of lattice parameters with an increase of Cr content were detected (see Figure 3), suggesting the incorporation of Cr ions in the LaOCl lattice. In the case of isovalent substitution, Cr3+ ions are expected to incorporate in La3+ positions. Due to the smaller ionic radius of Cr3+ (0.615 Å) in comparison to La3+ (1.032 Å) [56], a contraction of the LaOCl lattice resulting in a decrease of the lattice parameters a and c is expected.
The calculated values of lattice parameters indicate a linear decrease of the lattice parameter a and a simultaneous increase of the lattice parameter c. This result suggests that the type of substitution is more complex than expected in the case of Cr3+ → La3+ substitution. LaOCl exhibits a layered structure, in which La-O layers alternate with Cl double layers. To prevent repulsion between Cl ions, the electron density is accumulated in the structural cavities between the Cl layers [45]. Similar anisotropic behavior has been reported for interstitial substitution in layered La2(Ni0.9M0.1)O4+δ (M = Fe, Co, and Cu) perovskite [57] and metal alloy [58,59] materials. We propose that in the case of the introduction of chromium in LaOCl, Cr ions might incorporate between the Cl layers, thus increasing the lattice parameter c, or be located in an off-site position in the case of La substitution.
Results of the X-band measurements for LaOCl samples with different concentrations of Cr are presented in Figure 4. No EPR signals could be detected in the undoped sample. EPR spectra of chromium doped samples consist of a broad line at 358 mT (g ≈ 1.96). There is an obvious correlation between its intensity and activator concentration, which is a strong indication that the signal belongs to a Cr-related paramagnetic center.
EPR signals associated with Cr5+ and Cr3+ oxidation states have been reported to be detectable at room temperature; for the measurements of transition metal ions with an even number of electrons, low temperatures are usually required [60]. The more commonly encountered form is Cr3+, which is a spin S = 3/2 system. In a local cubic symmetry, the EPR spectrum consists of a single central line in g ≈ 1.97–1.98 range with a characteristic 53Cr hyperfine (HF) structure [61,62,63]. In the case of lower local symmetry, several transitions in a field range determined by the magnitude of zero-field splitting (ZFS) of the ground state can be expected [63,64,65,66,67,68]. Cr5+ is a S = 1/2 system; therefore, there are no effects related to ZFS, and the EPR signal is not as complicated as for Cr3+. A notable feature occasionally observed in the spectra of Cr5+ is HF structure caused by an interaction with the nearest cations [69,70,71,72]. We can observe a similar effect for the 2 mol% sample; however, due to the high Cr concentration, it is partially obscured. A sample with 0.1 mol% Cr was therefore synthesized for a more detailed analysis of chromium incorporation in LaOCl; the results of combined EPR and ENDOR investigations are shown in Figure 5.
The EPR spectrum exhibits a well-resolved HF structure with approximately 1.15 mT average separation between the lines, which corresponds to a HF coupling value of A ≈ 32 MHz. The number and relative intensities of the lines could be tentatively accounted for in a model with S = 1/2 interacting with four equivalent nuclear spins I = 7/2. From the naturally abundant nuclei in the LaOCl matrix, only 139La have I = 7/2. ENDOR spectra, in general, are determined by the relative magnitude of HF and nuclear Zeeman interactions [73]. The fact that we observe the ENDOR signal at A/2 ≈ 16 MHz implies that our case corresponds to a Larmor-split HF-centered doublet. Consequently, the nucleus responsible for the EPR spectrum HF structure can be identified from half the distance between the ENDOR lines. As a result, we can estimate that the Larmor frequency is 2.1 MHz directly from the experiment, which is close to the value of 139La at B = 345 mT (the magnetic field at which ENDOR was acquired). The asymmetric shape of the ENDOR signal suggests that the HF interaction is anisotropic. A spin-Hamiltonian (SH) with axial HF interaction was used for EPR and ENDOR spectra simulations:
H ^ = g μ B B · S ^ + S ^ · A · I ^
where g is the g-factor; μ B —the Bohr magneton; B —external magnetic field; A —the HF coupling tensor with components A = A x = A y and A = A z [74]. The best simultaneous fit to the experimental EPR and ENDOR spectra was achieved with the following SH parameter values: g = 1.964 ± 0.001, A = 30.72 ± 0.10 MHz, and A = 37.63 ± 0.10 MHz. The lineshape of the EPR signal was modelled with a Gaussian distribution of energy levels (EasySpin HStrain function [55]). The determined full widths at half maximum (FWHM) of the distributions for perpendicular and parallel directions were 33 and 50 MHz respectively. The linewidth used for the ENDOR simulation was 1 MHz. The main result from magnetic resonance spectroscopy analysis can be summarized as follows: the initial incorporation of chromium occurs in the Cr5+ state in a site, which is coordinated by 4 equivalent La3+ ions.
Annealing in reducing (H2/Ar) atmosphere was carried out to induce transformations of the oxidation state of chromium ions. No changes in the phase composition of LaOCl:Cr were detected after the heat treatment at 500–800 °C for 2 h. The results of XPS analysis of chromium 2p peaks for the samples annealed at different temperatures are shown in Figure 6.
The XPS spectrum contains strong signals of La, O, and Cl; however, Cr signal intensity is miniscule due to the relatively low doping content. Nevertheless, two peaks can be resolved in the Cr 2p transition range, which correspond to Cr 2p3/2 and Cr 2p1/2 spin-orbit splitting doublets. The aforementioned peaks monotonously shift toward smaller energies, when the annealing temperature is increased. Complex multiplet splitting of Cr peaks with up to five components has been demonstrated for different compounds [75]. To establish the effect of the annealing temperature on the XPS signal of chromium, the binding energy of the 2p3/2 peak was calculated as a sum of weighed binding energies of the subpeaks for every annealing temperature. The obtained dependency is shown in Figure 6b. A summary of binding energy values for the 2p3/2 peak in chromium oxides is given in Table 1.
The experimentally measured binding energy values for the peak 2p3/2 for different annealing temperatures fall in the set of values characteristic for Crn+, where n varies from n = 5 for the annealing temperature of 400 °C to n = 3 for the annealing temperature of 800 °C.
Experimental XANES spectra recorded in the range of the La L1,2,3-edges and Cr K-edge for 2% Cr LaOCl samples before and after heat-treatment in H2/Ar at 800 °C for 2 h are shown in Figure 7. As one can see, the Cr K-edge is located between the La L1 and L2 edges that complicates its analysis. Moreover, the XANES spectra are distorted by the presence of several “glitches” (Bragg reflections from the crystal monochromator), which occur as vertical spikes. Glitches can be uniquely identified by inspecting the I0 signal measured by the ionization chamber located before the sample. Three of them (one at 5985.8 eV and two at 5998.8 and 6000.1 eV) are, unfortunately, located at the beginning of the Cr K-edge (Figure 7b). Nevertheless, there is a notable difference between the two samples. The as-prepared sample contains a pre-edge peak at 5992.4 eV, which is completely absent in the reduced sample. Besides this difference, there are also other changes in the fine structure above the edge, which distinguish the two samples: in particular, the peak located at 6015 eV is broadened in the reduced sample. The pre-edge peak is attributed to the 1s → 3d(Cr) transition, which is forbidden in a dipole approximation for geometries that possess an inversion center but becomes allowed in a non-centrosymmetric environment, e.g., tetrahedral [81,82]. This means that the treatment in reducing atmosphere strongly affects the local environment of chromium ions.
Both Cr3+ (0.62 Å) and Cr5+ (0.35–0.57 Å) have smaller ionic radii than La3+ (1.16–1.22 Å) [56]; therefore, a substitution of lanthanum ions by chromium ions is structurally possible. Moreover, crystalline CrOCl3 is known [83], where chromium ions are present in the oxidation state 5+ and have a local environment, which is partially close to that of La in LaOCl. In CrOCl3 chromium ions have square pyramidal coordination with four chlorine atoms forming a base of the pyramid with R(Cr-Cl) = 2.16–2.34 Å and one oxygen atom located at the pyramid top with R(Cr-O) = 1.54 Å [83]. The small size of Cr5+ ions also allows us to suggest an alternative placement in LaOCl structure voids present within the layers of chlorine ions. In this case, chromium ions will have tetrahedral coordination by four chlorine atoms.
To discriminate between different structural models, the Cr K-edge XANES spectra were calculated as described above and are shown in Figure 8. In the case of lanthanum substitution (Figure 8a), the influence of chromium atom displacement along the c-axis by Δz = −0.20… + 0.15 Å has been additionally evaluated. The obtained results suggest that the pre-edge peak observed in the experimental XANES for the as-prepared sample can be reproduced using both structural models, i.e., a non-centrosymmetric environment of Cr5+ ions can be associated with their location in tetrahedral [CrCl4] (Figure 8b) or distorted [CrO4Cl4] (Figure 8a) surroundings. In the latter case, the chromium ions should be displaced closer to the oxygen ions by about 0.1 Å, which is favored by the smaller size of Cr5+ ions and stronger Cr5+-O2− Coulomb interaction. Upon reduction to the oxidation state 3+ in the reduced sample, the size of the chromium ions increases and the interaction with oxygen ions becomes weaker, which promotes the displacement of chromium ions toward lanthanum position or even slightly above it. Such displacement leads to a drastic reduction of the pre-edge peak intensity and a lowering of the amplitude of the peak located at 15–20 eV above the Fermi level in Figure 8a (it corresponds to the peak at 6015 eV in the experimental XANES in Figure 7b). Thus, the substitutional model, in which chromium ions are located close to the lanthanum position, can describe the experimentally observed variations in the Cr K-edge XANES in the as-prepared and reduced LaOCl:Cr samples.
X-band EPR spectra of 1% Cr samples annealed at different temperatures in the H2/Ar atmosphere are shown in Figure 9.
A gradual decrease of the Cr5+ signal intensity is observed as the reducing temperature is increased, which is accompanied by an emergence of additional signals over a broader field range. The number of features in the spectrum and the fact that EPR was detected at room temperature suggest the reduction of Cr5+ to Cr3+. EPR measurements at two frequency bands were performed for a more detailed analysis of the Cr3+ local structure; the results are shown in Figure 10.
One of the advantages of measurements at higher microwave frequencies is a simplification of spectra for S > 1/2 systems with a large magnitude of ZFS [84]. Moreover, interpretation of experimental results is more unambiguous, if a simultaneous fit for EPR spectra acquired at different frequencies can be achieved with the same parameter set. Simulations were performed with the following SH:
H ^ = g μ B B · S ^ + S ^ · D · S ^
where D is the ZFS tensor, which can be reduced to two parameters: D and E [85] and S = 3/2. The determined SH parameter values are summarized in Table 2.
The final simulation curves were achieved as a superposition of the initial Cr5+ signal and two Cr3+ centers labelled as I and II respectively. Both Cr3+ centers are axially symmetric (rhombic ZFS parameter E = 0); however, the magnitude of ZFS differs by a factor of two indicating significant variations in the local environment. The relative contribution (EasySpin Weight function) from the Cr3+ I center was determined to be higher by a factor of 30, thereby implying that this site is dominant in LaOCl. There was a feature (530 mT at X-band and 1135 mT at Q-band) in the experimental spectra, which could not be accounted for in the simulations. Its most likely origin is associated with another type of Cr-related defect.
The results of X-ray and magnetic resonance spectroscopy analysis demonstrate that oxidation state and local structure of chromium ions in LaOCl can be controlled during synthesis. The solubility limit of Cr in LaOCl is approximately 5 mol%; at higher concentrations formation of LaCrO3 impurity phase can be detected. In the investigated Cr concentration range, all XRD patterns can be ascribed to the tetragonal LaOCl phase without detectable deviations in crystal symmetry. The incorporation of Cr ions in the crystal structure can be deduced from the shift of XRD peak positions. Rietveld analysis reveals an asymmetric distortion of the lattice, that is, a decrease of the lattice parameter a and simultaneous increase of the lattice parameter c. The relatively small level of Cr doping prevents a direct detection of Cr local environment in LaOCl host matrix by XRD; therefore, local structure sensitive tools were used to probe the local environment around Cr ions in this study. As evidenced by the EPR and XPS data, if synthesized in air, the incorporation of chromium in LaOCl occurs predominantly in the 5+ oxidation state. The EPR and ENDOR spectra of LaOCl:Cr5+ can be simulated in a S = 1/2 model with HF interaction with four equivalent La nuclei. XANES spectra of the corresponding samples contain the forbidden 1s→3d(Cr) pre-edge structure, which is an indication of a non-centrosymmetric environment around Cr5+ ions. The observation can be explained by two different structural models: Cr5+ placement in the voids within the layers of chlorine ions or in an off-site position in the case of La substitution (in both models Cr5+ HF interaction with 4 equivalent La nuclei could be expected).
Annealing in a reducing atmosphere produces changes in the oxidation state and local environment of chromium ions. A gradual Cr5+→Cr3+ transformation is observed in the XPS spectra as the reducing temperature increases. EPR spectra simulations reveal that at least two Cr3+ centers are formed in the LaOCl structure as the result. However, a possibility of other oxidation states cannot be excluded (EPR signal of chromium having oxidation states with an even number of electrons is not expected to be detectable at our experimental conditions). It is likely that the dominant contribution to the EPR spectrum originates from Cr3+ substituting La3+ ions. Such geometry possesses an inversion center, which explains the absence of XANES 1s→3d(Cr) pre-edge structure in the spectrum of the reduced sample. The obtained results show that accommodation of small cations produces distortions to the LaOCl structure, which could be a prospective strategy for tailoring optical, electrical, or catalytical properties of the material.

4. Conclusions

The oxidation state and local environment of chromium ions were studied in LaOCl samples synthesized in different annealing atmospheres. A strong effect of chromium content on the LaOCl lattice was evidenced by X-ray diffraction. XPS, XANES, multifrequency EPR and ENDOR spectroscopy techniques were used to elucidate how chromium ions are embedded into the matrix.
Surprisingly, we found that chromium ion incorporation in LaOCl occurs predominantly in the 5+ oxidation state, producing asymmetric distortions to the crystal structure: a decrease of the lattice parameter a and an increase of the lattice parameter c. The EPR spectrum of LaOCl:Cr5+ consists of a resonance at g = 1.964 with a well-resolved HF structure from interaction with four equivalent La nuclei. XRD and EPR data, which are validated by XANES calculations, suggest that Cr5+ incorporation occurs either in the voids within the layers of chlorine ions or in an off-site position in the case of La substitution. Annealing in a reducing atmosphere affects the oxidation state and local structure of chromium ions so that at least two types of axial symmetry Cr3+ centers are formed in LaOCl.

Author Contributions

Formal analysis, A.A., G.K., H.O., A.F., A.S. and A.K.; investigation, A.A., G.K., H.O., A.F., A.S. and A.K.; methodology, A.A., G.K., H.O., A.F., A.S. and A.K.; supervision, A.A.; validation, A.A. and G.K.; visualization, A.A., G.K., A.S. and A.K.; writing—original draft, A.A., G.K., A.S. and A.K.; writing—review and editing, A.A., G.K., A.F., A.S. and A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by the Latvian Council of Science, project “Novel transparent nanocomposite oxyfluoride materials for optical applications”, project No. LZP-2018/1-0335. Institute of Solid State Physics, University of Latvia as the Center of Excellence has received funding from the European Union’s Horizon 2020 Framework Programme H2020-WIDESPREAD-01-2016-2017-TeamingPhase2 under grant agreement No. 739508, project CAMART2.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kim, D.H.; Yoon, J.Y.; Park, H.C.; Kim, K.H. CO2-sensing characteristics of SnO2 thick film by coating lanthanum oxide. Sens. Actuators B Chem. 2000, 62, 61–66. [Google Scholar] [CrossRef]
  2. Marsal, A.; Dezanneau, G.; Cornet, A.; Morante, J.R. A new CO2 gas sensing material. Sens. Actuators B Chem. 2003, 95, 266–270. [Google Scholar] [CrossRef]
  3. Marsal, A.; Centeno, M.A.; Odriozola, J.A.; Cornet, A.; Morante, J.R. DRIFTS analysis of the CO2 detection mechanisms using LaOCl sensing material. Sens. Actuators B Chem. 2005, 484–489. [Google Scholar] [CrossRef]
  4. Hwang, D.K.; Kim, S.; Lee, J.H.; Hwang, I.S.; Kim, I.D. Phase evolution of perovskite LaNiO3 nanofibers for supercapacitor application and p-type gas sensing properties of LaOCl-NiO composite nanofibers. J. Mater. Chem. 2001, 21, 1959–1965. [Google Scholar] [CrossRef]
  5. Li, X.; Qian, Q.; Zheng, W.; Wei, W.; Liu, X.; Xiao, L.; Chen, Q.; Chen, Y.; Wang, F. Preparation and Characteristics of LaOCl Nanotubes by Coaxial Electrospinning. Mater. Lett. 2012, 80, 43–45. [Google Scholar] [CrossRef]
  6. Trung, D.D.; Toan, L.D.; Hong, H.S.; Lam, T.D.; Trung, T.; van Hieu, N. Selective detection of carbon dioxide using LaOCl-functionalized SnO2 nanowires for air-quality monitoring. Talanta 2012, 88, 152–159. [Google Scholar] [CrossRef]
  7. van Hieu, N.; Khoang, N.D.; Trung, D.D.; Toan, L.D.; van Duy, N.; Hoa, N.D. Comparative study on CO2 and CO sensing performance of LaOCl-coated ZnO nanowires. J. Hazard. Mater. 2013, 244–245, 209–216. [Google Scholar] [CrossRef] [PubMed]
  8. Xiong, Y.; Xue, Q.; Ling, C.; Lu, W.; Ding, D.; Zhu, L.; Li, X. Effective CO2 detection based on LaOCl-doped SnO2 nanofibers: Insight into the role of oxygen in carrier gas. Sens. Actuators B Chem. 2017, 241, 725–734. [Google Scholar] [CrossRef]
  9. Kijima, N.; Matano, K.; Saito, M.; Oikawa, T.; Konishi, T.; Yasuda, H.; Sato, T.; Yoshimura, Y. Oxidative catalytic cracking of n-butane to lower alkenes over layered BiOCl catalyst. Appl. Catal. A Gen. 2001, 206, 237–244. [Google Scholar] [CrossRef]
  10. Manoilova, O.V.; Podkolzin, S.G.; Tope, B.; Lercher, J.; Stangland, E.E.; Goupil, J.M.; Weckhuysen, B.M. Surface acidity and basicity of La2O3, LaOCl, and LaCl3 characterized by IR spectroscopy, TPD, and DFT calculations. J. Phys. Chem. B 2004, 108, 15770–15781. [Google Scholar] [CrossRef]
  11. Podkolzin, S.G.; Stangland, E.E.; Jones, M.E.; Peringer, E.; Lercher, J.A. Methyl chloride production from methane over lanthanum-based catalysts. J. Am. Chem. Soc. 2007, 129, 2569–2576. [Google Scholar] [CrossRef] [PubMed]
  12. Imanaka, N.; Okamoto, K.; Adachi, G.Y. Water-insoluble lanthanum oxychloride-based solid electrolytes with ultra-high chloride ion conductivity. Angew. Chem. Int. Ed. 2002, 41, 3890–3892. [Google Scholar] [CrossRef]
  13. Nunotani, N.; Misran, M.R.I.B.; Inada, M.; Uchiyama, T.; Uchimoto, Y.; Imanaka, N. Structural environment of chloride ion-conducting solids based on lanthanum oxychloride. J. Am. Ceram. Soc. 2020, 103, 297–303. [Google Scholar] [CrossRef] [Green Version]
  14. Udayakantha, M.; Schofield, P.; Waetzig, G.R.; Banerjee, S. A full palette: Crystal chemistry, polymorphism, synthetic strategies, and functional applications of lanthanide oxyhalides. J. Solid State Chem. 2019, 270, 569–592. [Google Scholar] [CrossRef]
  15. Hölsä, J.; Porcher, P. Crystal field analysis of REOCI:Tb3+. J. Chem. Phys. 1982, 76, 2798–2803. [Google Scholar] [CrossRef]
  16. Yuanbin, C.; Shensin, L.; Wufu, S.; Lizhong, W.; Guangtian, Z. Crystal field analysis for emission spectra of LaOCl:Eu3+ under high pressure. Phys. B+C 1986, 139–140, 555–558. [Google Scholar] [CrossRef]
  17. Reid, M.F. Superposition-model analysis of intensity parameters for Eu3+ luminescence. J. Chem. Phys. 1987, 87, 6388–6392. [Google Scholar] [CrossRef]
  18. Malta, O.L.; Ribeiro, S.J.L.; Faucher, M.; Porcher, P. Theoretical intensities of 4f-4f transitions between stark levels of the Eu3+ ion in crystals. J. Phys. Chem. Solids 1991, 52, 587–593. [Google Scholar] [CrossRef]
  19. Bungenstock, C.; Tröster, T.; Holzapfel, W.B.; Bini, R.; Ulivi, L.; Cavalieri, S. Study of the energy level scheme of Pr3+:LaOCl under pressure. J. Phys. Condens. Matter 1998, 10, 9329–9342. [Google Scholar] [CrossRef]
  20. Rambabu, U.; Annapurna, K.; Balaji, T.; Buddhudu, S. Fluorescence spectra of Er3+: REOCl (RE = La, Gd, Y) powder phosphors. Mater. Lett. 1995, 23, 143–146. [Google Scholar] [CrossRef]
  21. Rambabu, U.; Balaji, T.; Annapurna, K.; Buddhudu, S. Fluorescence spectra of Tm3+-doped rare earth oxychloride powder phosphors. Mater. Chem. Phys. 1996, 43, 195–198. [Google Scholar] [CrossRef]
  22. Konishi, T.; Shimizu, M.; Kameyama, Y.; Soga, K. Fabrication of upconversion emissive LaOCl phosphors doped with rare-earth ions for bioimaging probes. J. Mater. Sci. Mater. Electron. 2007, 18, 183–186. [Google Scholar] [CrossRef]
  23. Lee, S.S.; Joh, C.H.; Byeon, S.H. Highly enhanced blue-emission of LnOCl:Tm prepared by dehydration of Ln(OH)2Cl:Tm (Ln = La and Gd). Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2008, 151, 163–168. [Google Scholar] [CrossRef]
  24. Li, G.; Li, C.; Zhang, C.; Cheng, Z.; Quan, Z.; Peng, C.; Lin, J. Tm3+ and/or Dy3+ doped LaOCl nanocrystalline phosphors for field emission displays. J. Mater. Chem. 2009, 19, 8936–8943. [Google Scholar] [CrossRef]
  25. Li, G.; Li, C.; Hou, Z.; Peng, C.; Cheng, Z.; Lin, J. Nanocrystalline LaOCl:Tb3+/Sm3+ as promising phosphors for full-color field-emission displays. Opt. Lett. 2009, 34, 3833. [Google Scholar] [CrossRef] [PubMed]
  26. Li, G.; Hou, Z.; Peng, C.; Wang, W.; Cheng, Z.; Li, C.; Lian, H.; Lin, J. Electrospinning derived one-dimensional LaOCl: Ln3+ (Ln = Eu/Sm, Tb, Tm) nanofibers, nanotubes and microbelts with multicolor-tunable emission properties. Adv. Funct. Mater. 2010, 20, 3446–3456. [Google Scholar] [CrossRef]
  27. Lee, S.S.; Park, H.I.; Joh, C.H.; Byeon, S.H. Morphology-dependent photoluminescence property of red-emitting LnOCl:Eu (Ln = La and Gd). J. Solid State Chem. 2007, 180, 3529–3534. [Google Scholar] [CrossRef]
  28. Du, Y.P.; Zhang, Y.W.; Sun, L.D.; Yan, C.H. Atomically efficient synthesis of self-assembled monodisperse and ultrathin lanthanide oxychloride nanoplates. J. Am. Chem. Soc. 2009, 131, 3162–3163. [Google Scholar] [CrossRef]
  29. Kim, S.W.; Jyoko, K.; Masui, T.; Imanaka, N. A new type of red-emitting (La,Ca)OCl:Eu3+ phosphors. Chem. Lett. 2010, 39, 604–606. [Google Scholar] [CrossRef]
  30. Kong, Q.; Wang, J.; Dong, X.; Yu, W.; Liu, G. Synthesis and luminescence properties of LaOCl:Eu3+ nanostructures via the combination of electrospinning with chlorination technique. J. Mater. Sci. Mater. Electron. 2013, 24, 4745–4756. [Google Scholar] [CrossRef]
  31. Lv, L.; Wang, T.; Li, S.; Su, Y.; Wang, X. Tuning the optical, electronic and luminescence properties of LaOCl:Eu3+ via structural and lattice strain modulation. Cryst. Eng. Comm. 2016, 18, 907–916. [Google Scholar] [CrossRef]
  32. Xu, Y.; Li, Z.; Liu, X.; Luo, Y.; Qian, Q.; Huang, B.; Xiao, L.; Chen, Q. Electrospun LaOCl:Eu3+, Ce4+ nanofibers with color-tunable fluorescence between red and orange. J. Mater. Sci. Mater. Electron. 2017, 28, 8596–8600. [Google Scholar] [CrossRef]
  33. Waetzig, G.R.; Horrocks, G.A.; Jude, J.W.; Villalpando, G.V.; Zuin, L.; Banerjee, S. Ligand-Mediated Control of Dopant Oxidation State and X-ray Excited Optical Luminescence in Eu-Doped LaOCl. Inorg. Chem. 2018, 57, 5842–5849. [Google Scholar] [CrossRef] [PubMed]
  34. Kim, D.; Jang, J.; Ahn, S.I.; Kim, S.H.; Park, J.C. Novel blue-emitting Eu2+-activated LaOCl:Eu materials. J. Mater. Chem. C 2014, 2, 2799–2805. [Google Scholar] [CrossRef]
  35. Kim, D.; Park, S.; Kim, S.; Kang, S.G.; Park, J.C. Blue-emitting Eu2+-activated LaOX (X = Cl, Br, and I) materials: Crystal field effect. Inorg. Chem. 2014, 53, 11966–11973. [Google Scholar] [CrossRef]
  36. Kong, Q.; Wang, J.; Dong, X.; Yu, W.; Liu, G. Synthesis and luminescence properties of Yb3+-Er3+ co-doped LaOCl nanostructures. J. Mater. Sci. 2014, 49, 2919–2931. [Google Scholar] [CrossRef]
  37. Park, S.; Cho, S.H. Spectral-converting study of La1-m-nErmYbnOCl (m = 0.001–0.2, n = 0–0.1) phosphors. J. Lumin. 2014, 153, 90–95. [Google Scholar] [CrossRef]
  38. Yu, W.; Kong, Q.; Wang, J.; Dong, X.; Liu, G. Fabrication of Er3+-doped LaOCl nanostructures with upconversion and near-infrared luminescence performances. J. Mater. Sci. Mater. Electron. 2014, 25, 46–56. [Google Scholar] [CrossRef]
  39. Yu, H.; Yu, A.; Li, Y.; Song, Y.; Wu, Y.; Sheng, C.; Chen, B. Energy transfer processes in electrospun LaOCl:Ce/Tb nanofibers. J. Alloys Compd. 2016, 683, 256–262. [Google Scholar] [CrossRef]
  40. Bai, H.; Song, Y.; Li, D.; Ma, Q.; Dong, X.; Yu, W.; Yang, Y.; Wang, J.; Liu, G.; Wang, T. Realizing white light emitting in single phased LaOCl based on energy transfer from Tm3+ to Eu3+. Ceram. Int. 2018, 44, 6754–6761. [Google Scholar] [CrossRef]
  41. Waetzig, G.R.; Horrocks, G.A.; Davidson, R.D.; Jude, J.W.; Villalpando, G.V.; Zuin, L.; Banerjee, S. In a Different Light: Deciphering Optical and X-ray Sensitization Mechanisms in an Expanded Palette of LaOCl Phosphors. J. Phys. Chem. C 2018, 122, 16412–16423. [Google Scholar] [CrossRef]
  42. Guan, M.; Mei, L.; Huang, Z.; Yang, C.; Guo, Q.; Xia, Z. Synthesis and near-infrared luminescence properties of LaOCl:Nd3+/Yb3+. Infrared Phys. Technol. 2013, 60, 98–102. [Google Scholar] [CrossRef]
  43. Kong, Q.; Wang, J.; Dong, X.; Yu, W.; Liu, G. Synthesis and luminescence properties of LaOCl:Nd3+ nanostructures via combination of electrospinning with chlorination technique. Mater. Express 2014, 4, 13–22. [Google Scholar] [CrossRef]
  44. Brixner, L.H.; Moore, E.P. Single-crystal refinement of the structure of LaOCl. Acta Crystallogr. Sect. C 1983, 39, 1316. [Google Scholar] [CrossRef]
  45. Maslen, E.N.; Streltsov, V.A.; Streltsova, N.R.; Ishizawa, N. Synchrotron X-ray Electron Density in the Layered LaOCI Structure. Acta Crystallogr. Sect. B Struct. Sci. 1996, 52, 576–579. [Google Scholar] [CrossRef]
  46. Swindells, F.E. Lanthanum Oxychloride Phosphors. J. Electrochem. Soc. 1954, 101, 415. [Google Scholar] [CrossRef]
  47. van Steensel, L.I.; Blasse, G. The luminescence of Sb3+ in LaOCl. J. Alloys Compd. 1996, 232, 60–62. [Google Scholar] [CrossRef]
  48. Wolfert, A.; Blasse, G. Luminescence of the Bi3+ ion in compounds LnOCl (Ln = La, Y, Gd). Mater. Res. Bull. 1984, 19, 67–75. [Google Scholar] [CrossRef]
  49. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  50. Doebelin, N.; Kleeberg, R. Profex: A graphical user interface for the Rietveld refinement program BGMN. J. Appl. Crystallogr. 2015, 48, 1573–1580. [Google Scholar] [CrossRef] [Green Version]
  51. Welter, E.; Chernikov, R.; Herrmann, M.; Nemausat, R. A beamline for bulk sample x-ray absorption spectroscopy at the high brilliance storage ring PETRA III. AIP Conf. Proc. 2019. [Google Scholar] [CrossRef]
  52. Joly, Y. X-ray absorption near-edge structure calculations beyond the muffin-tin approximation. Phys. Rev. B—Condens. Matter Mater. Phys. 2001, 63, 125120. [Google Scholar] [CrossRef]
  53. Bunau, O.; Joly, Y. Self-consistent aspects of x-ray absorption calculations. J. Phys. Condens. Matter 2009, 21, 345510. [Google Scholar] [CrossRef] [PubMed]
  54. Keski-Rahkonen, O.; Krause, M.O. Total and partial atomic-level widths. At. Data Nucl. Data Tables 1974, 14, 139–146. [Google Scholar] [CrossRef]
  55. Stoll, S.; Schweiger, A. EasySpin, a comprehensive software package for spectral simulation and analysis in EPR. J. Magn. Reson. 2006, 178, 42–55. [Google Scholar] [CrossRef]
  56. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  57. Nakamura, T.; Ling, Y.; Amezawa, K. The effect of interstitial oxygen formation on the crystal lattice deformation in layered perovskite oxides for electrochemical devices. J. Mater. Chem. A 2015, 3, 10471–10479. [Google Scholar] [CrossRef]
  58. Sherby, O.D.; Wadsworth, J.; Lesuer, D.R.; Syn, C.K. Revisiting the structure of martensite in iron-carbon steels. Mater. Trans. 2008, 49, 2016–2027. [Google Scholar] [CrossRef] [Green Version]
  59. Hasegawa, T.; Niibori, T.; Takemasa, Y.; Oikawa, M. Stabilisation of tetragonal FeCo structure with high magnetic anisotropy by the addition of V and N elements. Sci. Rep. 2019, 9, 1–9. [Google Scholar] [CrossRef] [PubMed]
  60. Al’tshulter, S.A.; Kozyrev, B.M. Electron Paramagnetic Resonance in Compounds of Transition Elements; Wiley: Hoboken, NJ, USA, 1974. [Google Scholar]
  61. Low, W.; Rubins, R.S. Electron spin resonance in the cubic crystalline field of calcium oxide. Phys. Lett. 1962, 1, 316–318. [Google Scholar] [CrossRef]
  62. Woonton, G.A.; Dyer, G.L. On the hyperfine structure of trivalent chromium in a cubic environment. Can. J. Phys. 1967, 45, 2265–2279. [Google Scholar] [CrossRef]
  63. O’Donnell, K.P.; Henderson, B.; O’Connell, D.; Henry, M.O. Axial Cr3+ centres in MgO: EPR and fluorescence studies. J. Phys. C Solid State Phys. 1977, 10, 3877–3884. [Google Scholar] [CrossRef]
  64. Zverev, G.M.; Prokhorov, A.M. Fine structure and hyperfine structure of paramagnetic resonance of Cr+++ in synthetic ruby. Sov. Phys. JETP 1958, 7, 354. [Google Scholar]
  65. Glynn, T.J.; Kelleher, L.; Imbusch, G.F.; Larkin, D.M.; Merritt, F.R.; Berggren, M.J. EPR study of LiGa5O8: Cr3+. J. Chem. Phys. 1971, 55, 2925–2930. [Google Scholar] [CrossRef]
  66. Patel, J.L.; Davies, J.J.; Cavenett, B.C.; Takeuchi, H.; Horai, K. Electron spin resonance of axially symmetric Cr3+ centres in KMgF3 and KZnF3. J. Phys. C Solid State Phys. 1976, 9, 129–138. [Google Scholar] [CrossRef]
  67. Takeuchi, H.; Arakawa, M.; Aoki, H.; Yosida, T.; Horai, K. EPR and 19F-ENDOR of Cr3+ Impurity Centres in K2ZnF4 and K2MgF4. J. Phys. Soc. Japan 1982, 51, 3166–3172. [Google Scholar] [CrossRef]
  68. Takeuchi, H.; Arakawa, M. EPR Study of Cr3+-Li+ Centres in Several Perovskite Fluorides. J. Phys. Soc. Jpn 1984, 53, 376–380. [Google Scholar] [CrossRef]
  69. Grunin, V.S.; Patrina, I.B.; Zonn, Z.N. Temperature Dependence of the EPR Spectra of Cr3+ and Cr5+ Ions in CaV2O6 and BaV2O6 Crystals. Phys. Status Solidi 1980, 98, 765–771. [Google Scholar] [CrossRef]
  70. Greenblatt, M.; Pifer, J.H.; McGarvey, B.R.; Wanklyn, B.M. Electron spin resonance of Cr5+ in YPO4 and YVO4. J. Chem. Phys. 1981, 74, 6014–6017. [Google Scholar] [CrossRef]
  71. Aboukaïs, A.; Zhilinskaya, E.A.; Filimonov, I.N.; Nesterenko, N.S.; Timoshin, S.E.; Ivanova, I.I. EPR investigation, before and after adsorption of naphtalene, of mordenite containing Fe3+ and Cr5+ ions as impurities. Catal. Lett. 2006, 111, 97–102. [Google Scholar] [CrossRef]
  72. Böttcher, R.; Pöppl, A.; Hoentsch, J.; Rakhmatullin, R.M. The Jahn-Teller effect in Cr5+-doped PbTiO3: A multi-frequency electron paramagnetic resonance study. J. Phys. Condens. Matter 2010, 22, 65902. [Google Scholar] [CrossRef]
  73. Telser, J. Electron-Nuclear Double Resonance (ENDOR) Spectroscopy. Encycl. Inorg. Chem. 2008. [Google Scholar] [CrossRef]
  74. Bennati, M. EPR interactions-hyperfne couplings. Emagres 2017, 271–282. [Google Scholar] [CrossRef]
  75. Biesinger, M.C.; Brown, C.; Mycroft, J.R.; Davidson, R.D.; McIntyre, N.S. X-ray photoelectron spectroscopy studies of chromium compounds. Surf. Interface Anal. 2004, 36, 1550–1563. [Google Scholar] [CrossRef]
  76. Si, P.Z.; Wang, H.X.; Jiang, W.; Lee, J.G.; Choi, C.J.; Liu, J.J. Synthesis, structure and exchange bias in Cr2O3/CrO2/Cr2O5 particles. Thin Solid Film. 2011, 519, 8423–8425. [Google Scholar] [CrossRef]
  77. Weckhuysen, B.M.; Wachs, I.E.; Schoonheydt, R.A. Surface chemistry and spectroscopy of chromium in inorganic oxides. Chem. Rev. 1996, 96, 3327–3350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Brand, E.; Kellett, D.; Enever, M.D.; Fellows, J.T.; Egdell, R.G. Magnetic properties of thin CrO2 layers supported on polycrystalline TiO2. J. Mater. Chem. 2005, 15, 1141–1147. [Google Scholar] [CrossRef]
  79. Heinig, N.F.; Jalili, H.; Leung, K.T. Fabrication of epitaxial CrO2 nanostructures directly on MgO(100) by pulsed laser deposition. Appl. Phys. Lett. 2007, 91, 253102. [Google Scholar] [CrossRef] [Green Version]
  80. Harrison, P.G.; Lloyd, N.C.; Daniell, W. The nature of the chromium species formed during the thermal activation of chromium-promoted tin(iv) oxide catalysts: An epr and xps study. J. Phys. Chem. B 1998, 52, 10672–10679. [Google Scholar] [CrossRef]
  81. Pantelouris, A.; Modrow, H.; Pantelouris, M.; Hormes, J.; Reinen, D. The influence of coordination geometry and valency on the K-edge absorption near edge spectra of selected chromium compounds. Chem. Phys. 2004, 300, 13–22. [Google Scholar] [CrossRef]
  82. Farges, F. Chromium speciation in oxide-type compounds: Application to minerals, gems, aqueous solutions and silicate glasses. Phys. Chem. Miner. 2009, 36, 463–481. [Google Scholar] [CrossRef]
  83. Hector, A.L.; Levason, W.; Light, M.E.; Reid, G.; Sardar, K.; Zhang, W. Chromium(V) oxide trichloride, and some pentachlorido-oxido-chromate(V) salts: Structures and spectroscopic characterization. Z. Anorg. Allg. Chem. 2013, 639, 906–910. [Google Scholar] [CrossRef]
  84. Misra, S.K. Multifrequency Electron Paramagnetic Resonance: Theory and Applications; Wiley: Hoboken, NJ, USA, 2011; p. 1022. [Google Scholar] [CrossRef]
  85. Telser, J. EPR interactions-zero-field splittings. Emagres 2017, 6, 207–234. [Google Scholar] [CrossRef]
Figure 1. Crystal structure of LaOCl. Visualized in VESTA [49] using atom coordinates from [45].
Figure 1. Crystal structure of LaOCl. Visualized in VESTA [49] using atom coordinates from [45].
Materials 14 03539 g001
Figure 2. XRD patterns of LaOCl samples with different concentrations of Cr.
Figure 2. XRD patterns of LaOCl samples with different concentrations of Cr.
Materials 14 03539 g002
Figure 3. LaOCl lattice parameter dependence on the level of Cr content.
Figure 3. LaOCl lattice parameter dependence on the level of Cr content.
Materials 14 03539 g003
Figure 4. EPR spectra of LaOCl samples with different concentrations of Cr.
Figure 4. EPR spectra of LaOCl samples with different concentrations of Cr.
Materials 14 03539 g004
Figure 5. (a) EPR and (b) ENDOR spectra simulations of LaOCl sample with 0.1 mol% Cr.
Figure 5. (a) EPR and (b) ENDOR spectra simulations of LaOCl sample with 0.1 mol% Cr.
Materials 14 03539 g005
Figure 6. (a) XPS spectrum of the 5% Cr LaOCl sample prior annealing in a reducing atmosphere; inset: background corrected XPS spectra in Cr 2p peak range of 5% Cr LaOCl samples annealed at different temperatures in H2/Ar atmosphere and (b) binding energy of Cr 2p3/2 to the annealing temperature of the samples.
Figure 6. (a) XPS spectrum of the 5% Cr LaOCl sample prior annealing in a reducing atmosphere; inset: background corrected XPS spectra in Cr 2p peak range of 5% Cr LaOCl samples annealed at different temperatures in H2/Ar atmosphere and (b) binding energy of Cr 2p3/2 to the annealing temperature of the samples.
Materials 14 03539 g006
Figure 7. Experimental XANES spectra recorded in the range of the La L1,2,3-edges and Cr K-edge for 2% Cr LaOCl samples before and after annealing in reducing atmosphere. The I0 signal measured by the ionization chamber located before the sample is shown at the top in (a). Several “glitches” are observed and labelled with the letter “G” in (b).
Figure 7. Experimental XANES spectra recorded in the range of the La L1,2,3-edges and Cr K-edge for 2% Cr LaOCl samples before and after annealing in reducing atmosphere. The I0 signal measured by the ionization chamber located before the sample is shown at the top in (a). Several “glitches” are observed and labelled with the letter “G” in (b).
Materials 14 03539 g007
Figure 8. (a) A structural model of the lanthanum ion substitution by chromium (left). Calculated Cr K-edge XANES spectra for different positions of chromium atoms displaced along the c-axis by Δz relative to the lanthanum position (right). (b) A fragment of the structural model of the chromium environment with the tetrahedral coordination by chlorine atoms (left). The calculated Cr K-edge XANES spectrum (right). The pre-edge peak is indicated by arrows.
Figure 8. (a) A structural model of the lanthanum ion substitution by chromium (left). Calculated Cr K-edge XANES spectra for different positions of chromium atoms displaced along the c-axis by Δz relative to the lanthanum position (right). (b) A fragment of the structural model of the chromium environment with the tetrahedral coordination by chlorine atoms (left). The calculated Cr K-edge XANES spectrum (right). The pre-edge peak is indicated by arrows.
Materials 14 03539 g008
Figure 9. EPR spectra of 1% Cr LaOCl samples after annealing at different temperatures in reducing atmosphere. Spectra intensities in 0–300 and 400–800 mT ranges have been magnified 30-fold.
Figure 9. EPR spectra of 1% Cr LaOCl samples after annealing at different temperatures in reducing atmosphere. Spectra intensities in 0–300 and 400–800 mT ranges have been magnified 30-fold.
Materials 14 03539 g009
Figure 10. (a) X-band and (b) Q-band EPR spectra simulations of 0.1% Cr LaOCl sample after annealing at 800 °C in reducing atmosphere.
Figure 10. (a) X-band and (b) Q-band EPR spectra simulations of 0.1% Cr LaOCl sample after annealing at 800 °C in reducing atmosphere.
Materials 14 03539 g010
Table 1. A summary of 2p3/2 binding energies in chromium oxides.
Table 1. A summary of 2p3/2 binding energies in chromium oxides.
Binding Energy, eVCompoundReference
576.7–577.2Cr2O3[75]
577.1Cr2O3[76]
577.0Cr2O3[77]
577.0CrO2[78]
576.6CrO2[79]
579.3Cr2O5[76]
579.0Cr2O5[77]
579.3CrO3[78]
580.3CrO3[80]
580.0CrO3[77]
578.5Annealed at 400 °CCurrent work
576.9Annealed at 800 °CCurrent work
Table 2. Fitted SH parameter values of Cr3+ centers in LaOCl.
Table 2. Fitted SH parameter values of Cr3+ centers in LaOCl.
Cr3+ CentergD, MHzHStrain, MHz
I1.973 ± 0.00112063 ± 200253 ± 25
II1.977 ± 0.0016865 ± 100115 ± 15
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Antuzevics, A.; Krieke, G.; Ozols, H.; Fedotovs, A.; Sarakovskis, A.; Kuzmin, A. Oxidation State and Local Structure of Chromium Ions in LaOCl. Materials 2021, 14, 3539. https://doi.org/10.3390/ma14133539

AMA Style

Antuzevics A, Krieke G, Ozols H, Fedotovs A, Sarakovskis A, Kuzmin A. Oxidation State and Local Structure of Chromium Ions in LaOCl. Materials. 2021; 14(13):3539. https://doi.org/10.3390/ma14133539

Chicago/Turabian Style

Antuzevics, Andris, Guna Krieke, Haralds Ozols, Andris Fedotovs, Anatolijs Sarakovskis, and Alexei Kuzmin. 2021. "Oxidation State and Local Structure of Chromium Ions in LaOCl" Materials 14, no. 13: 3539. https://doi.org/10.3390/ma14133539

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