Jelena
Kojčinović
a,
Dalibor
Tatar
a,
Stjepan
Šarić
a,
Cora
Bartus Pravda
b,
Andraž
Mavrič
c,
Iztok
Arčon
cd,
Zvonko
Jagličić
ef,
Maximilian
Mellin
g,
Marcus
Einert
g,
Angela
Altomare
h,
Rocco
Caliandro
h,
Ákos
Kukovecz
b,
Jan Philipp
Hofmann
g and
Igor
Djerdj
*a
aDepartment of Chemistry, Josip Juraj Strossmayer University of Osijek, Cara Hadrijana 8/A, 31000 Osijek, Croatia. E-mail: igor.djerdj@kemija.unios.hr
bDepartment of Applied and Environmental Chemistry, University of Szeged, 6720 Szeged, Hungary
cUniversity of Nova Gorica, Vipavska 13, 5000 Nova Gorica, Slovenia
dInstitute Jožef Stefan, Jamova 39, 1000 Ljubljana, Slovenia
eInstitute of Mathematics, Physics, and Mechanics, University of Ljubljana, Jamova 2, 1000 Ljubljana, Slovenia
fFaculty of Civil & Geodetic Engineering, University of Ljubljana, Jadranska 19, 1000 Ljubljana, Slovenia
gSurface Science Laboratory, Department of Materials and Earth Sciences, Technical University of Darmstadt, Otto-Berndt-Strasse 3, 64287 Darmstadt, Germany
hInstitute of Crystallography, CNR, via Amendola 122/o, Bari 70126, Italy
First published on 22nd December 2023
CeNiO3 has been reported in the literature in the last few years as a novel LnNiO3 compound with promising applications in different catalytic fields, but its structure has not been correctly reported so far. In this research, CeNiO3 (RB1), CeO2 and NiO have been synthesized in a nanocrystalline form using a modified citrate aqueous sol–gel route. A direct comparison between the equimolar physical mixture (n(CeO2):
n(NiO) = 1
:
1) and compound RB1 was made. Their structural differences were investigated by laboratory powder X-ray diffraction (PXRD), selected area electron diffraction (SAED), transmission electron microscopy (TEM) with an energy-dispersive X-ray spectroscopy (EDS) detector, and Raman spectroscopy. The surface of the compounds was analyzed by X-ray photoelectron spectroscopy (XPS), while the thermal behaviour was explored by thermogravimetric analysis (TGA). Their magnetic properties were also investigated with the aim of exploring the differences between these two compounds. There were clear differences between the physical mixture of CeO2 + NiO and RB1 presented by all of these employed methods. Synchrotron methods, such as atomic pair distribution function analysis (PDF), X-ray absorption near edge structure (XANES) and extended X-ray absorption fine structure (EXAFS), were used to explore the structure of RB1 in more detail. Three different models for the structural solution of RB1 were proposed. One structural solution proposes that RB1 is a single-phase pyrochlore compound (Ce2Ni2O7) while the other two solutions suggest that RB1 is a two-phase system of either CeO2 + NiO or Ce1−xNixO2 and NiO.
In one study, it was claimed that CeNiO3 was obtained using a citrate–nitrate combustion technique and provided Rietveld refinement data clearly showing a mixture of CeO2 and NiO.31 Detailed literature studies were conducted to find a crystallographic information file (.cif) to directly compare our experimentally obtained XRD pattern and to obtain crystallographic information but without success. Therefore, there is still no crystallographic proof of the existence of CeNiO3 crystallizing in the perovskite structure.
Importantly, we noticed that what the literature claims to be the XRD pattern of the “CeNiO3 perovskite” suspiciously has the same Bragg positions as the mixture of constituent oxides, fluorite CeO2, and cubic NiO. There could be two explanations: either this is just a coincidence or CeNiO3 is a mixture of CeO2 and NiO at the nanoscale level.
Another possibility is that this “CeNiO3” perovskite is a two-phase system containing solid solution Ce1−xNixO2 and NiO. Various authors have investigated the Ni2+ doping effect of the CeO2 structure.32–39 These studies were mostly used for hydrogen production via water splitting35 and as catalysts for the reduction of NO40 or unsaturated organic compounds.34,36,41 A similar application has also been reported for the “CeNiO3” perovskite.20,22–24,26
Therefore, it is important to elucidate the actual structure of what the literature claims to be CeNiO3 so that the structure–property relationship is established. If this is resolved, its properties could be improved by further research and consequently, its application. In this research, a compound that corresponds to CeNiO3 (RB1) and the substituent oxides, CeO2 and NiO, were synthesized in a nanocrystalline form by means of the previously developed modified aqueous sol–gel method.42–48 Ceria and nickel oxide were mixed in an equimolar molar ratio (n(CeO2):
n(NiO) = 1
:
1) after synthesis to obtain the same amounts of constituent oxides as there are in RB1, so the direct comparison between the physical mixture (CeO2 + NiO) and a “compound” (RB1) can be made. Their structural features were thoroughly investigated using laboratory X-ray and electron diffraction, synchrotron, microscopic and spectroscopic methods, with the aim of solving the actual structure of RB1 (CeNiO3). Their magnetic properties have also been explored to point out the differences in the synthesized compounds.
Chemical name | Manufacturer |
---|---|
Cerium(III) nitrate hexahydrate, 99.5% | Acros Organics, USA |
Nickel nitrate hexahydrate, p.a. | T.T.T., Croatia |
Citric acid monohydrate, Ph Eur | T.T.T., Croatia |
Concentrated ammonia solution (w = 25%), p.a. | Gram-Mol, Croatia |
Chemicals were used as purchased, without further purification. MilliQ ultrapure water was obtained using a PURELAB Flex device for ultrapure water preparation. For pH value adjustment, a pH-meter HANNA pH 211 was used. The reaction solution was heated on a magnetic hotplate stirrer DLAB MS-H-S and dried in a drying oven Instrumentaria ST-01/02 afterwards. Calcination was performed in a muffle furnace Nabertherm LT5/11/B410.
The samples were prepared in the form of homogeneous pellets, pressed from micronized powder mixed with micronized BN, with a total absorption thickness of about 1.5 above the Ni K-edge or Ce L3-edge, and inserted in the monochromatic beam between the first two ionization detectors. The absorption spectra were measured in the energy region from −150 eV to +1000 eV relative to the Ni K-edge (8333 eV), while for the Ce L3-edge (5724 eV) EXAFS scans were stopped at the Ce L2-edge (6165 eV). Equidistant energy steps of 0.3 eV were used in the XANES region, while for the EXAFS region, equidistant k steps of 0.03 Å−1 were adopted, with an integration time of 2 s per step. The exact energy calibration was established with simultaneous absorption measurement on 5 μm thick Ni metal foil for the Ni K-edge, or the CeO2 reference sample (calibrated with vanadium metal foil) for the Ce L3-edge, placed between the second and third ionization chambers. The energy reproducibility of the measured spectra was ±0.03 eV. The quantitative analysis of XANES and EXAFS spectra was performed using the Demeter (IFEFFIT) program package,50 in combination with the FEFF6 program code for ab initio calculations of photoelectron scattering paths in quantitative EXAFS analysis.51
Compound | RB1 (CeNiO3) | CeO2 + NiO (n![]() ![]() ![]() ![]() |
|||||
---|---|---|---|---|---|---|---|
1 | 2 | 3 | |||||
Chemical formula | CeO2 | NiO | Ce0.97Ni0.03O2 | NiO | Ce2Ni2O7 | CeO2 | NiO |
Crystal system | Cubic | Tetragonal | Cubic | Cubic | Cubic | ||
Space group |
Fm![]() |
P42/nmc |
Fm![]() |
Fd![]() |
Fm![]() |
||
Z | 8 | 2 | 8 | 8 | 8 | ||
Calculated density (g cm−3) | 7.20 | 6.76 | 7.07 | 6.74 | 7.83 | 7.19 | 6.80 |
Unit cell parameters (Å) | 5.4165(2) | 4.1785(2) | a = 3.8208(3); c = 5.4492(2) | 4.1903(2) | 10.8406(9) | 5.4162(2) | 4.1861(2) |
Unit cell volume (Å3) | 158.88(6) | 73.36(1) | 79.55(1) | 73.57(1) | 1273.98(8) | 158.91(6) | 72.96(1) |
Phase content (wt%) | 71.0(2) | 29.0(3) | 71.34(1) | 28.66(4) | 100(4) | 66.73(2) | 33.27(2) |
Average crystallite size (nm) | 5.1 | 4.6 | 5.2 | 4.7 | 5.4 | 9.6 | 9.8 |
Average apparent microstrain (×10−4) | 0.95 | 0.95 | 4.47 | 0.95 | |||
R B | 0.91 | 1.02 | 0.88 | 0.75 | 4.40 | 1.45 | 0.68 |
R p, Rwp, Re | 10.9, 9.19, 8.08 | 11.0, 9.21, 8.13 | 19.9, 17.9, 8.59 | 9.87, 8.44, 6.56 | |||
χ 2 | 1.29 | 1.28 | 4.32 | 1.66 |
Based on the literature, the actual structure of RB1 (CeNiO3) is unknown, so we proposed three different possibilities:
(1) RB1 is an equimolar mixture of both cubic Fmm CeO2 and NiO.
(2) RB1 is a mixture of tetragonal P42/nmc Ni-doped CeO2 and cubic Fmm NiO.
(3) RB1 is a novel cubic Fdm compound with a pyrochlore structure and the chemical formula Ce2Ni2O7.
The main challenge in a direct comparison of RB1 and previously reported CeNiO3 is the lack of crystallographic data obtained by the Rietveld refinement in those research studies. The results presented in this study unequivocally indicate that our RB1 sample corresponds to the previously reported CeNiO3. However, variations in synthesis procedures undoubtedly generate visible differences in structural features such as the degree of crystallinity. For instance, Harikrishnan et al.21 employed a co-precipitation method using metal nitrates, NaOH, and potassium carbonate. After prolonged stirring of the solution followed by calcination at 650 °C, the targeted CeNiO3 has been obtained. In another research conducted by the same author,27 CeNiO3 synthesis involved Ni-foam for active material growth, followed by autoclaving at 180 °C for 12 hours and then calcination at 600 °C under an inert (argon) atmosphere. These differences in synthesis approaches affected only the degree of crystallinity, whilst the positions of Bragg reflections of the claimed CeNiO3 and their relative intensities are almost intact. Our initial goal was to directly compare the equimolar physical mixture of CeO2 + NiO (n(CeO2):
n(NiO) = 1
:
1) and RB1 prepared under the same synthesis conditions to identify potential differences in the crystallographic structure. Here, we assumed that RB1 consists of both cubic CeO2 and NiO, so the physical mixture of these two constituent oxides should not differ much from RB1. As for the unit cell parameters and crystal structure, these two samples appear to be the same. However, the phase content slightly differs since there is around 4 wt% more of CeO2 in RB1. Also, there is a large difference in average crystallite size values. The refinement of the physical mixture of CeO2 + NiO reveals crystallite sizes of constituent oxides that are two times larger than those found in RB1. This can be directly seen from the broadened Bragg peaks of RB1 compared to the physical mixture.
Another possibility was that cerium and nickel form a unique single-phase compound. For the proposal of an actual crystal structure of RB1, we have tried to solve it ab initio from the powder XRD pattern. However, this is a complex task since there are only a few broad peaks present, so our attempt was unsuccessful. Therefore, we have tried to find XRD patterns of different compounds in the literature that are similar to the experimental XRD pattern of RB1. The most suitable XRD pattern that was found was that of pyrochlore Ce2Zr2O755 which was used as a starting point. However, the content of the unit cell was changed to refine it with the obtained XRD pattern of RB1. Although this refinement appears to be correct, the plot given in Fig. 1c demonstrates that there are slight differences in the peak positions between the calculated and experimental patterns.
Mahammadunnisa et al.32 investigated the impact of doping CeO2 with 5–30 wt% of Ni2+. In their synthesis procedure, specific amounts of metal nitrates and citric acid were dissolved in distilled water, sonicated, and placed in a preheated furnace at around 450 °C, causing a spark, leading to the formation of a solid product. The samples further underwent calcination at 600 °C for 4 hours to eliminate the carbon content. They have noticed that there is a slight decrease in unit cell parameters due to doping CeO2 with a smaller cation Ni2+ (C.N. = 6, r = 0.69 Å (ref. 56)) in NiO, compared to Ce4+ (C.N. = 8, r = 0.97 Å (ref. 56)) in CeO2, which is also visible in RB1 compared to the physical mixture of constituent oxides (CeO2 + NiO). Two additional peaks in the XRD pattern of Ce–Ni–O systems32 at 2θ = 37° and 2θ = 43° correspond to the NiO phase and only appear when doping of CeO2 is beyond 15 wt%. Zhou et al.57 used three types of synthesis procedures, all including metal nitrates as the starting material: (i) the citrate acid method by adding citric acid into the nitrate aqueous solution, heating until gel was formed and then calcination at 450 °C; (ii) the coprecipitation method with addition of potassium carbonate, followed by the adjustment of the pH value and then calcination at 450 °C, and (iii) the ammonia evaporation method, which included addition of NH3 to aqueous solution of metal nitrates until a specific pH value, followed by calcination at 450 °C. They have shown that Ni-loading in ceria must be beyond 20 mol% for these additional peaks to appear. Therefore, the NiO phase detected by XRD in RB1 might be the NiO that is dispersed on the surface of ceria, while the rest of it could be dissolved in the cerium-nickel-based tetragonal solid solution Ce1−xNixO2−δ.57 Another difference between the physical mixture of CeO2 + NiO and RB1 is in average crystallite size values which can also be estimated from the Bragg peak widths on X-ray diffraction patterns (Fig. 1a compared to Fig. 1b). In the physical mixture CeO2 + NiO, the average crystallite sizes of the constituent oxides are twice as large as those in RB1 (Table 2). Mahammadunisa et al.32 also reported this behaviour as an effect of doping CeO2 with NiO. They attributed it to the incorporation of nickel cations in the initial structure of ceria.58 Even though these authors did not refine the XRD patterns they obtained, it was probably assumed that both Ce1−xNixO2−δ and NiO are cubic, at least according to the given results. Small dopant concentrations can still preserve the cubic crystal structure of ceria.59 However, doping ceria with other elements can also result in symmetry breaking from the cubic to the tetragonal crystal structure, which cannot be detected by XRD.5,60,61
Another method that is sensitive to phase formation and therefore useful for distinction between cubic and tetragonal crystal structures is Raman spectroscopy. Therefore, the Raman spectra of RB1 and a physical mixture of CeO2 + NiO were recorded in the range from 1200 cm−1 to 80 cm−1 at an excitation wavelength of 532 nm and are shown in Fig. 2. According to Kroumova et al.,62 CeO2 with a fluorite structure should possess only one Raman active mode, T2g, also known in the available literature as the F2g vibration.43 However, cubic NiO has no Raman active modes.62 Therefore, the Raman spectrum of the physical mixture of CeO2 and NiO should contain only one Raman active mode corresponding to fluorite ceria, as shown in Fig. 2 (red). Sole NiO would not be visible in the Raman spectrum. A strong F2g vibration of CeO2 in the Raman spectrum of the physical mixture CeO2 + NiO is found at 450 cm−1. The Raman spectrum of RB1 shows a strong peak slightly shifted to higher wavenumbers, at 455 cm−1. Also, an additional defect peak at 576 cm−1 coincides with the vibrations that are often assigned to oxygen vacancies in CeO2-based compounds.48,59,63–67
![]() | ||
Fig. 2 Raman spectra of the RB1 sample and physical mixture of cerium(IV) and nickel(II) oxide (1![]() ![]() |
Cop et al.59 have shown that doping ceria with different cations results in the appearance of additional defect bands that are a result of increased concentration of oxygen vacancies. Various other authors have shown that these bands are expected when CeO2 is doped with other cations,32,59,68 especially when these cations are aliovalent,59 such as Ni2+. Also, a peak at 220 cm−1 is not active in the fluorite-structured ceria. This phonon mode is often activated in doped ceria, especially if there is a mismatch in ionic radii and ionic charge of dopants.61,69 Therefore, a tetragonal solid solution would explain the symmetry breaking observed in the Raman spectrum of RB1. Popović et al.38 have investigated charge delocalization in ceria upon doping with Fe2+/3+. They observed that the doping of ceria with smaller cations has an impact on the position of F2g mode because of the shrinkage of the unit cell. Also, charge mismatch and the small crystallite size impact oxygen vacancy concentration and the intensity of the vacancy mode.38 Atzori et al.70 have also investigated CeO2–NiO systems synthesized using the surface-templated method with direct synthesis of metal oxides from nitrates, dissolved in water with the template and NaOH, then filtered and calcined to obtain the targeted compound. Another used route was the synthesis of CeO2 while Ni was deposited through incipient wetness impregnation technique. They have shown that this large defect band increases with increased Ni loading. Several other authors have studied Ce1−xNixO2−δ for various catalytic applications.33,41,71,72 Barrio et al.33 even reported that the limit of the solubility of Ni in ceria ranges from 10 to 12 mol%, even though they did not give information on the Ce1−xNixO2−δ unit cell content. These values were calculated from the difference in the NiO content; they assumed that the remaining Ni was dissolved in CeO2. The total Ni-content in RB1 is 30 wt%, which is 50 mol%, but there is only 3 mol% of total Ni-content dissolved in the tetragonal solid solution in our case according to the Rietveld refinement results. It seems that in our case, there is much less Ni that is dissolved in CeO2, but none of the mentioned research papers calculated the actual amount of Ni dissolved in the solid solution directly from the unit cell content. Therefore, we must consider their results carefully.
Additionally, it was necessary to record Raman spectra of CeO2 and NiO separately, to further inspect their structure. Fig. 3a and b shows individual Raman spectra of CeO2 and NiO. The Raman spectrum of pure CeO2 shows typical strong F2g vibration at 462 cm−1 and a wide defect peak from 550–650 cm−1 that occurs due to the presence of oxygen vacancies.73 Now, it can be observed that F2g vibration is red-shifted in both RB1 and the CeO2 + NiO mixture. Also, the NiO spectrum reveals a low-intensity wide peak at 506 cm−1 that cannot be classified as Raman-active mode since its intensity is 150 times lower than the intensity of F2g vibration in CeO2. Also, there are two additional broad peaks at around 190 cm−1 and 1075 cm−1, with even lower intensity. The inset in Fig. 3b shows the Raman spectrum of NiO on the same scale as the CeO2 spectrum in Fig. 3a to visualize the difference in intensities and thus to explain the absence of these peaks in the Raman spectrum of the CeO2 + NiO mixture. According to George et al.74 these peaks appear due to the first- and second-order Raman scattering in NiO-nanostructures because of the structural defects. To conclude, there is a clear difference between RB1 and the CeO2 + NiO mixture in terms of Raman spectroscopy.
Nevertheless, additional analyses are necessary to resolve this structural issue. One of them is transmission electron microscopy. Fig. 4 and 5 show STEM, TEM, and SAED images of RB1. STEM-EDS mapping in Fig. 4 shows a uniform distribution of involved cations in RB1. The physical mixture of CeO2 + NiO does not have a uniform distribution of cerium and nickel, as shown in Fig. S1 and S2 in the ESI.† This is a difference that can have an impact on the difference in catalytic activity or some other physical properties.
![]() | ||
Fig. 5 (a) TEM image of RB1 (CeNiO3) nanoparticle agglomerate and (b) a corresponding SAED with assigned d-values in Å. (c) HRTEM of RB1 (CeNiO3) nanoparticles, d-values in Å. |
TEM, HRTEM and SAED images confirm the nanocrystalline nature of synthesized RB1. The crystallite size values nearly correspond to the ones calculated by the Rietveld refinement of the XRD pattern of RB1. Measured interplanar distances (d values) were compared to values obtained from the Rietveld refinement results of RB1 for three different cases. A comparison is given in Table 3.
SAED pattern | Rietveld refinement | |||||
---|---|---|---|---|---|---|
Cubic Fm![]() |
Tetragonal P42/nmc | Cubic pyrochlore Fd![]() |
||||
d (Å) | hkl | d (Å) | hkl | d (Å) | hkl | |
3.17 | 3.13 | 111 | 3.13 | 101 | 3.13 | 222 |
2.79 | 2.71 | 200 | 2.72 | 002 | 2.71 | 400 |
2.48 | 2.41 | 111 | 2.42 | 111 | 2.49 | 331 |
2.13 | 2.09 | 200 | 2.10 | 200 | 2.09 | 511 |
1.96 | 1.92 | 220 | 1.92 | 112 | 1.92 | 440 |
1.67 | — | — | 1.64 | 103 | 1.65 | 533 |
1.61 | 1.63 | 311 | 1.63 | 211 | 1.63 | 622 |
1.52 | 1.56 | 222 | 1.56 | 202 | 1.52 | 551 |
1.39 | 1.35 | 400 | 1.36 | 004 | 1.36 | 800 |
As seen in Table 3, the cubic Fmm crystal structure can be ruled out due to the missing interplanar distance. Therefore, the actual compound present in RB1 could be in the form of a solid solution Ce1−xNixO2, or it could crystalize as a pyrochlore Ce2Ni2O7.
Normalized Ni K-edge XANES spectra recorded for nanocrystalline NiO and RB1 are shown in Fig. 6, together with the spectra of the corresponding crystalline NiO reference compound. The energy position of the Ni K-edge and the edge profile in NiO and RB1 coincide with the energy position and edge profile of crystalline NiO with Ni in the 2+ state. This shows that the valence state and local symmetry of Ni cations is the same as in the crystalline NiO. Ni K-edge EXAFS analysis is used to detect the average local environment of Ni cations in the nanocrystalline NiO and RB1 samples. In the Fourier transform (FT), the magnitude of the k3-weighted EXAFS spectra are plotted in Fig. 7. The contributions of photoelectron scattering on the nearest shells of neighbours around the Ni atoms are observed in the R range of up to about 6 Å. Already by qualitative comparison of the FT EXAFS spectra, it is evident that both samples exhibit the same Ni local structure characteristic of crystalline NiO with an FCC crystal structure, as in the crystalline NiO reference.75 Quantitative EXAFS analysis is used to determine the structural parameters of the average local Ni neighbourhood (the type and average number of neighbours; the radii and Debye–Waller factor of neighbouring shells) in the samples. The structural parameters are quantitatively resolved from the EXAFS spectra by comparing the measured EXAFS signal with the model signal. The FEFF model for the crystalline NiO nanoparticles is based on the cubic crystal structure of NiO with the space group Fmm with the lattice constant a = 4.177 Å,76 where Ni is coordinated with 6 oxygen atoms at a distance of 2.09 Å, 12 Ni atoms at 2.95 Å and 8 oxygen atoms at 3.62 Å.
![]() | ||
Fig. 6 Ni K-edge XANES spectrum of nanocrystalline NiO and RB1 samples with the crystalline NiO reference compound. Spectra are shifted vertically for clarity. |
The FEFF model comprised three single scattering and two significant multiple scattering paths up to 3.6 Å, with 8 variable parameters: coordination shell distance (R) and Debye–Waller factors (σ2) of all single scattering paths, and the amplitude reduction factor (S02) and shift of the energy origin of the photoelectron (ΔEo), common to all scattering paths, are introduced. The structural parameters of multiple scattering paths are constrained to those of the corresponding single scattering paths. The model was tested on the EXAFS spectrum measured for the crystalline NiO.75 The same FEFF model is used in the fit of the EXAFS spectra of the nanocrystalline NiO and RB1 samples, but here, the amplitude reduction factor (S02) was fixed to a value of 0.90 obtained for the crystalline NiO, and the shell coordination numbers were used as variable parameters. A very good EXAFS fit (Fig. 7) is obtained in the k range of 3–12 Å and the R-range of 1.2–3.3 Å. The best-fit structural parameters are given in Table 4. The results show that the local structure around Ni cations in the nanocrystalline RB1 sample is the same as that in the nanocrystalline NiO compound. The structural parameters of the RB1 sample are the same as those of the nanocrystalline NiO sample, except that the Debye–Waller factors are slightly larger in RB1, indicating a larger structural disorder in the crystalline structure.
Neighbour | N | R (Å) | σ 2 (Å2) | R-factor |
---|---|---|---|---|
NiO | ||||
O | 6.0(5) | 2.079(4) | 0.0052(7) | 0.00086 |
Ni | 12.0(5) | 2.953(2) | 0.0058(2) | |
O | 8.0(5) | 3.46(3) | 0.010(7) | |
RB1 | ||||
O | 6.1(5) | 2.074(4) | 0.0061(8) | 0.0019 |
Ni | 11.0(7) | 2.957(3) | 0.0067(3) | |
O | 7.3(4) | 3.46(2) | 0.010(3) |
To test for the eventual presence of the RB1 nanocrystalline structure one single scattering path from Ce neighbours is added to the FEFF model at the distance characteristic for Ni–Ce in the RB1 crystal structure. The presence of Ce neighbour in the local neighbourhood around Ni cations is excluded by the fit, showing there is no presence of Ce in Ni neighbourhood or it is below the detection limit.
Normalized Ce L3-edge XANES spectra measured on CeO2 and RB1 are shown in Fig. 8, together with the spectra of the corresponding Ce3+ reference compounds (crystalline CeVO4).77
The energy position of the Ce L3-edge and the edge profile in RB1 completely coincide with the energy position and edge profile of CeO2, indicating that all Ce cations in the sample are in the Ce4+ valence state. If the sample contains a mixture of two or more compounds of the same cation with different local structures and valence states, the measured XANES spectrum is a linear combination of individual XANES spectra of different cation sites. In such cases, the relative amounts of the cation at each site and the average valence state of the cation in the sample can be determined by the linear combination fit (LCF) with XANES spectra. LCF analysis of the XANES spectra on RB1 with the XANES spectra of the reference compounds (CeO2 as reference for Ce4+, and CeVO4 as reference for Ce3+) shows that there is no significant presence of Ce3+ species. Quantitative EXAFS analysis to determine the structural parameters of the average local Ce neighbourhood was carried out for CeO2 and RB1 samples (see ESI, Fig. S3 and Table S2†). The structural parameters for RB1 are quantitatively and equally well resolved from the EXAFS spectra using the FEFF models for CeO2 with a cubic (Fmm) and tetragonal (P42/nmc) structure. This shows that EXAFS cannot resolve between cubic and tetragonal structures for the RB1 sample. The obtained structural parameters for RB1 are in agreement with those obtained for pure CeO2 (Table S2†).
The atomic pair distribution function (PDF) profile of RB1 has been fitted to test the three modelling hypotheses. The crystal cell parameters and atomic displacement parameters and positions not constrained by symmetry have been refined for all the crystal phases. Results are shown in Fig. 9, while the main fitting parameters are reported in Table 5. It can be noted that the hypothesized pyrochlore structure is ruled out by the poorer PDF fitting, as the corresponding calculated profile shows significant differences compared to the observed one even at small interatomic distances (Fig. 9c). For the remaining structural hypotheses (1) and (2) there is an overall agreement with reciprocal space determinations (Rietveld analysis) for crystal cell and weight fraction parameters, and the average crystallite size is in good agreement with the particle diameter estimated by PDF (SPdiameter). However, PDF data definitively resolves the question of which structural description is the best among the two linear combinations of crystal phases. A better fit is achieved when the cubic NiO crystal phase is coupled with the tetragonal Ni-doped CeO2 rather than the cubic CeO2 one. In the latter case, a weight agreement factor Rw = 0.163 is obtained and the observed-calculated difference profile in Fig. 9b only contains only high-frequency noise throughout the profile.
1 | 2 | 3 | |||
---|---|---|---|---|---|
Chemical formula | CeO2 | NiO | Ce0.97Ni0.03O2 | NiO | Ce2Ni2O7 |
Crystal system | Cubic | Tetragonal | Cubic | Cubic | |
Space group |
Fm![]() |
P42/nmc |
Fm![]() |
Fd![]() |
|
Fit range | 1.4–40.0 Å | 1.4–40.0 Å | 1.4–40.0 Å | ||
R w | 0.193 | 0.163 | 0.261 | ||
Q broad | 0.040 | 0.028 | 0.040 | ||
Scale | 0.136 | 0.141 | 0.08 | ||
Phase content (wt%) | 76.9 | 23.1 | 77.0 | 23.0 | 100 |
δ 1 | 2.07 | 1.50 | 2.10 | 1.54 | 2.44 |
Unit cell parameters (Å) | 5.402 | 4.175 | a = 3.812; c = 5.417 | 4.175 | 10.804 |
SPdiameter (nm) | 5.5 | 5.2 | 5.4 | 5.0 | 5.7 |
Additionally, X-ray photoelectron spectroscopy (XPS) was conducted to investigate the surface of the materials. XPS spectra were recorded for NiO and CeO2 separately and for all synthesized compounds. Fig. 10 shows high-resolution Ce 3d spectra for CeO2 and RB1, respectively. The Ce 3d spectrum of CeO2 shows the typical 3d5/2 at 883.1 eV and the 3d3/2 at 901.7 eV with a spin–orbit splitting of 18.6 eV. The spectra exhibit strong charge-transfer satellites78,79 and additional overlapping peaks from multiplet effects.80 The position and intensity ratio between those peaks depends on the oxidation state, the ligand type, and the next-nearest neighbour.80 This makes analysing core spectra of mixed transition metal and lanthanide oxides difficult. It was pointed out81 that satellites 3 and 4 can be directly related to cerium 4+ (CeO2), which shows that both materials in Fig. 10 mainly consist of cerium 4+. However, the CeO2 deviates from PLD-deposited films shown in ref. 81 indicating that cerium 3+ is present on the surface, which could be due to the formation of cerium hydroxide, clearly seen in the O 1s spectra (Fig. S4b†).
In the Ce 3d spectra of RB1, the left shoulder of the 3d5/2 main peak broadens while the 3d3/2 main peak increases in intensity, which indicates the higher cerium 3+ content in RB1 compared to the CeO2 sample. This could hint at a mixed oxidation state of cerium and nickel or an oxygen-deficient structure. However, in the O 1s spectra of RB1 (Fig. S4c†), the cerium hydroxide content on the surface is high, which would result in cerium in the 3+ state. Reference spectra without the influence of hydroxide could be used in the fitting procedure to analyze the content of different cerium components. Peak position analysis is summarized in Table S3.†
Similar analysis was performed for the high-resolution Ni 2p spectra, in order to investigate the stoichiometry of the compounds and the oxidation state of nickel cations in both samples. The Ni 2p spectra are shown in Fig. 11 for NiO and RB1 and the peak position analysis is summarized in Table S4.†
The Ni 2p spectrum of NiO shows 2p3/2 (853.0 eV) and 2p1/2 (871.5 eV) doublets with a spin–orbit splitting of 17.5 eV.82 The core level peaks at 856.1 and 873.2 eV are a sign of the presence of non-stoichiometry in NiO and the existence of Ni3+, respectively.82,83 It has been previously reported that satellite peaks 2 and 3 also point out towards the existence of Ni3+.83,84 Nickel oxide prepared at lower temperatures (>700 °C) has a higher concentration of Ni3+ than Ni2+.82,85 Interestingly, a fully stoichiometric NiO compound is formed after calcination at 1100 °C.
The explanation of the nonstoichiometry in nanocrystalline nickel oxide lies in the formation of Ni2+ vacancies85,86 because of excess oxygen. Positively charged nickel-vacant holes (h++) are typically neutralized by receiving two electrons from the neighbouring Ni2+ species, causing them to increase their oxidation state to Ni3+. In addition, there is another possible way of neutralizing the charge mismatch by neighbouring O2− species, thus creating reactive electrophilic O− species.85,87 These reactions are summarized in eqn (1) and (2):
2Ni2+ + h++ → 2Ni3+ | (1) |
2O2− + h++ → 2O−. | (2) |
The electrophilic oxygen species are ideal candidates for nucleophilic attack by hydroxide or water thus making them highly active in the formation of O–O bonds.87 Therefore, they are very important for the oxygen evolution reaction (OER).88–90 However, electrophilic oxygen can also be used for the reduction of unsaturated organic compounds.34,36,41
When comparing the Ni 2p spectrum of NiO and RB1, one can see that the RB1 spectrum is very noisy and cannot be interpreted easily. This is probably due to the interference of the charge-transfer satellite peaks in the Ni 2p region with the Ce 3d region in RB1.91 However, there is a sign of core level peaks at 854 and 856 eV, where the peak at 856 is higher in intensity and broadened compared to the peak at 854 which could indicate a higher content of Ni3+ in RB1 than in pure NiO. This could be due to the higher hydroxide content on the surface of RB1 and the formation of NiOOH, as seen in the high-resolution O 1s spectra shown in Fig. S4.†
All O 1 s spectra in Fig. S4† show a lattice oxygen (OL) peak at 530 eV. While the NiO and CeO2 spectra show two additional peaks at 531 eV and 533 eV. The peak at 531 eV can be related to loosely bound/adsorbed surface oxygen species (Oads), while the peak at around 533 eV is ascribed to adsorbed OH-groups or water molecules (OOH), in good agreement with the literature.85,88,92 The deconvolution parameters of O 1s spectra are summarized in Table S5 (ESI).† NiO and CeO2 have a similar concentration of adsorbed oxygen species (around 30%), while RB1 contains almost 67% of adsorbed hydroxide groups. This could be attributed to the almost two times smaller crystallite size of RB1, providing substantially more surface area for oxygen species to adsorb at.
The results of thermogravimetric analysis (Fig. S5†) can also confirm the presence of higher content of Ni3+ and Ce3+ in RB1 than in a physical mixture of constituent oxides since there is a higher oxygen uptake in RB1. During heating in an oxidative atmosphere, Ce3+ is oxidized to Ce4+, while the surface hydroxides are removed. However, heating non-stoichiometric Ni-based oxides increases stoichiometry and thus decreases the concentration of nickel vacancies.82
To further corroborate our structural findings, aiming to clarify the earlier assumption that RB1 is CeNiO3 or not, we have synthesized an additional compound Ce0.97Ni0.03O2 and conducted several additional measurements (PXRD and Raman spectroscopy). These additional measurements prove that, indeed, the equimolar introduction of Ni into the CeO2 lattice does not yield the single CeNiO3 compound. Instead, it likely forms a mixture with two distinct phases: Ce1−xNixO2 and NiO. The synthesis of 3% Ni-doped CeO2 yielded phase-pure, tetragonal Ce0.97Ni0.03O2, which agrees well with crystallographic data reported for RB1. Rietveld refinement results presented in Fig. 12a and Table S6† confirmed almost identical values for this compound compared to the phases reported for RB1 in Table 2.
In addition to that, Raman spectroscopy measurements were employed to provide a further proof (Fig. 12b). The defect peak at 576 cm−1 decreased proportionally with the reduction in Ni-dopant concentration, aligning with the 50% dopant level in RB1 and the 3% Ni-doped sample. Moreover, a consistent increase in the CeO2 Raman active mode intensity (F2g at 450 cm−1) occurred with decreasing dopant content. Additionally, the intensity evident at 220 cm−1 suggests Ni integration into CeO2 as in the RB1 compound.
![]() | ||
Fig. 13 Temperature dependent susceptibility (a) and isothermal magnetization (b) of CeO2, NiO, RB1 and 3% Ni doped CeO2. |
The magnetic behaviour of the newly synthesised compounds RB1 and CeO2 (3% doped Ni) differ significantly from the magnetic properties of CeO2 and NiO nanoparticles. A much larger splitting between ZFC and FC susceptibility and hysteresis with a coercive magnetic field of 600 Oe and a remanent magnetization of 0.02 emu g−1, are clear evidence of a ferromagnetic order in CeNiO3 (RB1). The local maximum of ZFC susceptibility at about 100 K and the temperature at which the ZFC and FC curves diverge (about 220 K) are quite different, indicating a rather broad size distribution of RB1 nanoparticles. The temperature where ZFC and FC curves diverge shifts to lower temperatures in CeO2 (3% doped Ni) while the isothermal magnetization at 2 K of only 3% doped CeO2 is already quite similar to the M(H) curve of RB1. These results support the finding that RB1 is the mixture of Ce0.97Ni0.03O2 and NiO. Several publications have reported on the ferromagnetism of Ce1−xNixO2 nanoparticles.97–101 In our final product (RB1) there is a lot of NiO and only a small portion of Ce1−xNixO2, which makes the ferromagnet even more interesting as this small portion of Ce1−xNixO2 makes a huge difference in the magnetic properties of the system.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3dt03280a |
This journal is © The Royal Society of Chemistry 2024 |