Next Article in Journal
Shaping of Aluminum Foam during Foaming of Precursor Using Steel Mesh with Various Opening Ratios
Previous Article in Journal
Increase in the Mechanical Strength of Mg-8Gd-3Y-1Zn Alloy Containing Long-Period Stacking Ordered Phases Using Equal Channel Angular Pressing Processing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Iron Impurities on Magnetic Properties of Nanosized CeO2 and Ce-Based Compounds

1
CEITEC IPM, Institute of Physics of Materials, AS CR, Zizkova 22, 616 62 Brno, Czech Republic
2
Faculty of the Environment, University of Jan Evangelista Purkyne, Kralova Vysina 7, 400 96 Usti nad Labem, Czech Republic
3
Department of Physics, VSB-Technical University of Ostrava, 17.listopadu 15, 708 33 Ostrava-Poruba, Czech Republic
4
Institute of Physics, University of Silesia in Katowice, Pułku Piechoty 1A, 41-500 Chorzów, Poland
*
Author to whom correspondence should be addressed.
Metals 2019, 9(2), 222; https://doi.org/10.3390/met9020222
Submission received: 17 December 2018 / Revised: 1 February 2019 / Accepted: 9 February 2019 / Published: 13 February 2019
(This article belongs to the Special Issue Rare Earth Compounds)

Abstract

:
CeO2 samples prepared by three technological procedures from the same cerium source, namely cerium (III) nitrate hexahydrate, are studied from the viewpoint of structure, chemical and phase composition, and micro- and macro-magnetic properties. The scanning and transmission electron microscopies completed by energy-dispersive X-ray (EDX) analysis yield nano-structural natures and homogenous chemical compositions of the ceria samples, confirmed also by X-ray diffraction. The diamagnetic, paramagnetic, and ferromagnetic phases in all samples follow from an analysis of the room- and low-temperature measurements of hysteresis loops. Iron impurities in ppm amounts are clearly detected by 57Fe Mössbauer spectrometry not only in the ceria samples but also in the selected input chemicals used for their preparation. This contributes to the explanation of the magnetic behaviour of nanosized ceria.

Graphical Abstract

1. Introduction

Cerium, the chemical element with atomic number 58, is a silver-white metal and the second element in the lanthanide series. It shows +3 and +4 oxidation states. The most widely used compound of cerium is cerium oxide CeO2, known also as ceria. It crystallises in the fluorite crystal structure with space group Fm 3 ¯ m (Figure 1) over the temperature range from room temperature to the melting point. The fluorite structure consists of a face-centred cubic (f.c.c.) unit cell of cations, with anions occupying the tetrahedral interstitial sites. This can also be seen as a superposition of a f.c.c. lattice of cations (Ce4+) with lattice constant a and a simple cubic lattice of anions (O2−) with lattice constant a/2. In this structure, each cerium cation is coordinated by eight nearest-neighbour oxygen anions, while each oxygen anion is coordinated by four nearest-neighbour cerium cations [1]. The colour of CeO2 is pale yellow. The CeO2 nanoparticles can be synthesised by solution-based methods including co-precipitation, micro-emulsion processes, sol-gel techniques, combustion reaction, etc.
Ceria is presently a frequently studied compound because of its unique physical properties, which offer a number of various applications. It is extensively used as a catalyst in various chemical processes, e.g., for the elimination of toxic auto-exhaust gases [2,3], as a glass-polishing material [4], or in biotechnology, environmental chemistry, and medicine [5,6]. The mobility of oxygen vacancies and consequently the high ionic conductivity determines ceria as a promising electrolyte for solid oxide fuel cells [7] and other applications [8]. Just the oxygen vacancies and their effects on the magnetic properties of ceria and some other nonmagnetic oxides, such as Al2O3 or ZnO, call for deeper investigations. The main reason is the puzzling large difference in obtained magnetic moments presented by authors all over the world. The origin of magnetic moments in these oxides, predominantly in their nano-sized form, is ascribed to a formation of localised electron spin moments at the oxygen vacancies particularly present at nano-particle surfaces. In the case of CeO2, another reason for its ferromagnetic behaviour has to be considered, namely the effects of Ce3+ ions on the nano-particle surfaces which, contrary to diamagnetic Ce4+ ions, carry their own magnetic moments. Recently, the magnetic behaviour of nanosized ceria has also been ascribed to a presence of a small amount of magnetic iron-based impurities. Their amount in units of ppm was detected by low-temperature magnetic measurements using a superconducting quantum interference device (SQUID) magnetometer and subsequently confirmed by 57Fe Mössbauer spectrometry, which is highly sensitive to the presence of iron atoms in a compound [9]. Nevertheless, some Ce-based compounds yield a relatively high magnetic moment even if no iron is detected by Mössbauer spectrometry. Therefore, the present study is concentrated on ceria prepared by different technological procedures. It continues our previous investigations [9,10] in order to shed more light on the magnetic behaviours of ceria and Ce-compounds. The nano-sized CeO2 powders were prepared from the same cerium (III) nitrate hexahydrate using three precipitation procedures with different chemicals. Their structural and physical analyses were completed by analysing selected chemicals used for CeO2 production from the viewpoint of possible iron and/or other chemical element contaminations.

2. Materials and Methods

Three cerium oxide samples, denoted as S4, S5, and S6, were prepared from the same cerium source (cerium (III) nitrate hexahydrate, 99.0%, Sigma-Aldrich) using different precipitation procedures. The sample S4 was prepared by conventional direct precipitation of the cerous nitrate solution with the saturated solution of oxalic acid, whereas the sample S5 was prepared by homogeneous precipitation with dimethyl oxalate [11]. In both cases, the resulting cerous oxalates were dried and annealed at 500 °C for 2 h in air to convert them into cerium oxides. The sample S6 was prepared by a well-established precipitation/calcination synthetic route [12,13] using ammonium carbonate as a precipitant, with subsequent calcination at 500 °C. Oxalic acid, dimethyl oxalate, and ammonium bicarbonate were obtained from Sigma-Aldrich as reagent-grade chemicals with purity above 99%. All solutions were prepared in deionised water obtained from a Demi Ultra 20 system (Goro, Prague, Czech Republic), in which reverse osmosis and mixed-bed ion-exchange were used for water purification.
The selected chemicals were subjected to supplementary chemical and/or magnetic analysis. Cerous nitrate 5N–N (CeN) of purity 99.999% (Sigma-Aldrich) was analyzed using the Inductively Coupled Plasma Atomic Emission Spectroscopy (ICP-AES), and the content of iron was determined to be 3.61 ppm. This method was also used for cerous chloride 3N–C (denoted CeCl) of purity 99.9% (Sigma-Aldrich), which is similarly frequently used for CeO2 production, and the content of Fe was 2.83 ppm. Both values are comparable with 1.78 ppm obtained for cerium (III) nitrate hexahydrate 99.0% used for the preparation of the S4, S5, and S6 samples.
The morphologies and microstructures of the samples were investigated using a TESCAN LYRA 3XMU FEG/SEM scanning electron microscope (SEM, Brno, Czech Republic) operated at an accelerating voltage of 20 kV. The microscope was also equipped with an XMax80 Oxford Instruments detector (UK) for energy-dispersive X-ray (EDX) analysis. The structure in the nano-size dimension was observed by a transmission electron microscope (TEM) Philips CM12 STEM (Amsterdam, The Netherlands) with a thermoemission electron source operating at 120 kV. The powder samples for SEM observation were prepared by spreading a small amount of powder onto a sticky conductive tape, and those for TEM were treated ultrasonically in an ethanol bath for 5 min and then placed on holey carbon copper grids.
X-ray diffraction (XRD) patterns were measured at room temperature (RT) using an X’PERT PRO diffractometer with Co Kα radiation (λ = 0.17902 nm) in the 2θ range from 20° to 135° in steps of 0.01° and time per step of 5 s. The evaluation of powder patterns was realised by the Rietveld structure refinement method [14] in semi-automatic mode, using the HighScore Plus program Version 4.8.0 (Malvern Panalytical, UK) and the ICSD database (FIZ Karlsruhe, Germany) of inorganic and related structures [15]. The pattern analysis yielded the phase composition, lattice parameters, and the mean crystallite size, representing a size of coherently diffracting domains.
Room- and low-temperature (RT, LT) hysteresis loops were measured using the Quantum Design XL-7 superconducting quantum interference device (SQUID) magnetometer (Quantum Design, CA, USA). Special care was devoted to the sample holder and to the sample handling procedures, to prevent any contamination of the holder as well as the samples. The saturation magnetisation was derived from the hysteresis loops at an accuracy of approximately ±1%.
The Mössbauer spectrometry was used to detect and/or exclude the contamination of the samples with iron impurities. The measurements were performed at RT in standard transmission geometry using 57Co(Rh) source. Calibration of the velocity scale was performed with α-Fe at RT and the isomer shifts are given with respect to its Mössbauer spectrum. All spectra were evaluated using the transmission integral approach in the program CONFIT [16]. The experimental points were analysed by double- and single-line components yielding values of isomer shift (δ) and quadrupole splitting (Δ). The relative representation of these subcomponents is denoted by A.

3. Results and Discussion

3.1. Morphology, Phase, and Chemical Composition

The morphologies of samples obtained using SEM are shown in Figure 2 (left panel). The EDX chemical analysis has not detected any other elements besides Ce and oxygen (with a usual EDX inaccuracy in light element quantification). Both integral (area of approx. 1 mm2) and point chemical analyses in all three samples were done from several places, and the mean values summarised in Table 1 do not yield any systematic difference. Similarly, there is no important difference in microstructure among the samples and, therefore, the TEM observation was done for the S6 sample only and the resulting fine nanostructure, including confirmation of CeO2 f.c.c. crystal structure by the diffraction pattern, is seen in Figure 3. The TEM micrograph clearly shows that the particle size ranges between approximately 10 and 50 nm.
The electron microscope observations are in good agreement with X-ray diffraction measurements. The patterns of all samples depicted in Figure 2 (right panel) were analysed by the same ICSD 28753 data sheet for the 225 (Fm 3 ¯ m) space group, and the denoted (hkl) peaks are practically identical. They represent the CeO2 phase. The lattice parameters for S4, S5, and S6 samples: 0.5414(1) nm, 0.5413(1) nm, and 0.5415(1) nm, respectively, are identical in the frame of accuracy of measurement. A slight difference is observed only in the crystallite (microdomain) size, d, calculated using the Scherrer formula employing parameters of (111) peak: d = 15.1 nm for S4, d = 21.8 nm for S5, and d = 19.01 nm for S6. These results point out the nanocrystalline nature of studied samples and confirm the representative TEM result in Figure 3, rather than quite large CeO2 plates (seemingly single crystals) shown by SEM in Figure 2 (left panel).

3.2. Macroscopic Magnetic Properties

The hysteresis curves of the S4, S5, and S6 samples measured at RT are presented in Figure 4 (left panel) and those obtained at a low temperature of 2 K are summarised in the middle panel. They consist of diamagnetic, paramagnetic, and ferromagnetic contributions. The diamagnetic contribution is reflected at all hysteresis loops by a decrease of magnetisation at higher magnetic fields. It is described by a negative magnetic susceptibility, χD, practically independent on the applied temperature. Estimated values of χD of the studied samples are summarised in Table 2. A sharp magnetisation reversal at lower magnetic fields seen at the RT hysteresis loops in Figure 4 (left panel) is indicative of the ferromagnetic contribution. This can originate in a large number of structural defects, particularly surface oxygen vacancies mentioned by some authors [17,18,19,20,21], and/or magnetic impurities, namely iron [9,10] contributing to small values of RT ferromagnetic saturation magnetisation of an order of 10−3 Am2/kg, detected also in other usually nonmagnetic compounds [22]. The analysis of the present magnetic measurements of the S4, S5, and S6 samples has resulted in the detection of a small amount of iron impurities, in units of ppm as seen in Table 2.
The paramagnetic contribution is practically hidden in the RT measurements, but it can be well determined from the LT curves. Contrary to the diamagnetic contribution, the paramagnetic contribution following the Curie–Weiss law is strongly dependent on temperature. It could be determined from the LT hysteresis loops in Figure 4 (middle panel) after a subtraction of the diamagnetic and ferromagnetic contributions. This has resulted in magnetisation curves in Figure 4 (right panel) that are compared with the Brillouin function (dash-dotted line) calculated using a total electronic angular momentum J = 5/2 and the Lande’s g-factor g = 6/7. The estimated saturation paramagnetic moment Mspm (Table 2) can be converted to the Bohr magneton per formula unit (Mspm/12), and then compared with the theoretical saturation moment of Ce3+ ions equal to 2.54 μB. This resulted in a low concentration of the Ce3+ ions in all samples, reaching about 10−3, which is substantially less compared to the diamagnetic Ce4+ ions.
As it was mentioned above, the selected chemicals, namely those at which the iron impurities were evidenced by ICP-AES analysis, were studied by magnetic techniques as well. The small amounts of iron impurities (2.83 ppm in CeN and 3.61 ppm in CeCl) have caused a weak magnetisation reversal at the room temperature and at the low applied fields. This was similar to ferromagnetic contributions detected in the CeO2 samples. Moreover, the RT hysteresis loops (not presented here) exhibited paramagnetic responses with mass magnetisation of about 0.3–0.5 Am2/kg at magnetic fields of 5.6 MA/m (7 T), and mass susceptibilities of 5 × 10−8 m3/kg for CeN and of 6 × 10−8 m3/kg for CeCl. The much stronger paramagnetic moment of about 15–20 Am2/kg was obtained from the hysteresis loops measured at the low temperature of 2 K. Based on the certificates mentioning cerium occurring in combination with other lanthanide impurity elements, the strong paramagnetic moment was successfully analyzed by the Brillouin function of neodymium (J = 4, g = 0.603) and/or praseodymium (J = 4.5, g = 0.732) overlapping the magnetically poor response of cerium.

3.3. Mössbauer Spectrometry

The Mössbauer spectrometry was used to verify a presence of iron in the samples and selected chemicals frequently used in the production of cerium oxide. The measured and analysed Mössbauer spectra in Figure 5a clearly prove a presence of iron in all three samples. The lower effect of measurements of coarser powders of the chemicals (Figure 5b) has evoked extended times of data reading. Nevertheless, the spectra confirm a presence of iron in the chemicals as well. A slightly better effect, observed at CeN, corresponds to the higher Fe content followed from the ICP-AES. The spectra of the samples were evaluated using one double-line component with hyperfine parameters (isomer shift, δ, and quadrupole slitting, Δ) not too different in the frame of experimental errors: S4: δ = 0.229(10) mm/s, Δ = 0.316(15) mm/s; S5: δ = 0.251(19) mm/s, Δ = 0.303(34) mm/s; and S6: δ = 0.232(12) mm/s, Δ = 0.333(18) mm/s. The spectra of chemicals were analysed using the singlet of δ = 0.218(68) mm/s for CeCl and δ = 0.374(31) mm/s for CeC, and a doublet for CeN of δ = 0.397(43) mm/s and Δ = 0.118(106) mm/s. Though the measurements are limited by the sensitivity of the Mössbauer effect, they support the magnetic results and confirm well the chemical analysis yielding small amounts of iron (see Section 2). They are not substantially different from the previous findings presented in Refs. [9,10].
It is also evident that the preparation of the “iron-free” cerium oxide would require more sophisticated synthetic routes and/or a utilisation of the iron-free starting materials. Although some of the used procedures are recommended for the separation of rare earths from common metals (precipitation of oxalates) [23], they are obviously unable to remove the residual amounts of iron from cerium at the ppm level completely.

4. Conclusions

The study is devoted to a clarification of the magnetic behaviour of ceria in its nano- up to micro-sized forms. The room- and low-temperature magnetic measurements of the samples prepared from the cerium (III) nitrate hexahydrate source and using three different precipitation reactions clearly prove the presence of the diamagnetic, paramagnetic, and ferromagnetic phases. While the effect of Ce3+ ions as carriers of their own magnetic moments can be excluded due to their insignificant concentration compared to diamagnetic Ce4+ ions in ceria, the iron impurities detected by chemical analysis, reflected in hysteresis loops, and confirmed by Mössbauer spectrometry play a significant role despite their small amount. Their well-established presence in ceria excludes the effect of oxygen vacancies as a unique source of ceria magnetism.
As follows from the present ICP-AES analyses and from certificates provided by the supplier, the raw materials (Ce sources) for the ceria preparation contained trace amounts of iron, typically at the level of several ppm, that was also confirmed by other experimental methods used here. Similarly, the precipitating agents were also not completely iron-free. For example, the content of iron in oxalic acid and ammonium carbonate ranges between 2 and 5 ppm according to specifications, and a similar value may be expected also in dimethyl oxalate. This raises a question of whether it is in general possible to produce ceria free of iron, in spite of the synthetic routes (especially those utilising oxalic acid and dimethyl oxalate) which are declared as highly selective separation methods for the rare earths purification. The effect of neodymium and/or praseodymium, detected in the chemicals, on the magnetic behaviour of ceria also remains unclear.

Author Contributions

Sample preparation, P.J.; electron microscope observations. J.B.; SQUID magnetic measurements, A.C.; room-temperature magnetic measurements, O.Z.; magnetic curve analysis, O.Z. and J.L.; XRD and Mössbauer measurements, Y.J.; manuscript preparation, Y.J.; manuscript editing, J.B., O.Z., and J.L.

Acknowledgments

This work was supported from ERDF/ESF New Composite Materials for Environmental Applications (No. CZ.02.1.01/0.0/0.0/17_048/0007399), CEITEC 2020 - National Sustainability Programme II (No. LQ1601), and project SP2018/43, all funded by Ministry of Education, Youth and Sports of the Czech Republic, and from the National Science Centre of Poland via project 2015/19/B/ST8/02636.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schwarz, K. Materials design of solid electrolytes. Proc. Natl. Acad. Sci. USA 2006, 103, 3497. [Google Scholar] [CrossRef] [PubMed]
  2. Alikin, E.A.; Vedyagin, A.A. High Temperature Interaction of Rhodium with Oxygen Storage Component in Three-Way Catalysts. Top. Catal. 2016, 59, 1033–1038. [Google Scholar] [CrossRef]
  3. Trovarelli, A. Catalytic properties of ceria and CeO2-containing materials. Catal. Rev. 1996, 38, 439–520. [Google Scholar] [CrossRef]
  4. Armini, S.; De Messenmaeker, J.; Whelan, C.M.; Moinpour, M.; Maex, K. Composite polymer core-ceria shell abrasive particles during oxide CMP: A defectivity study. J. Electrochem. Soc. 2008, 155, H653–H660. [Google Scholar] [CrossRef]
  5. Tarnuzzer, R.W.; Colon, J.; Patil, S.; Seal, S. Vacancy engineered ceria nanostructures for protection from radiation-induced cellular damage. Nano Lett. 2005, 5, 2573–2577. [Google Scholar] [CrossRef] [PubMed]
  6. Asati, A.; Santra, S.; Kaittanis, C.; Nath, S.; Perez, J.M. Oxidase-Like Activity of Polymer-Coated Cerium Oxide Nanoparticles. Angew. Chem. 2009, 48, 2308–2312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Park, S.D.; Vohs, J.M.; Gorte, R.J. Direct oxidation of hydrocarbons in a solid-oxide fuel cell. Nature 2000, 404, 265–267. [Google Scholar] [CrossRef] [PubMed]
  8. Younis, A.; Chu, D.; Li, S. Cerium Oxide Nanostructures and their Applications. In Functionalized Nanomaterials; Farrukh, M.A., Ed.; IntechOpen: Rijeka, Croatia, 2016; pp. 53–68. [Google Scholar] [CrossRef]
  9. Lunacek, J.; Zivotsky, O.; Janos, P.; Dosek, M.; Chrobak, A.; Marysko, M.; Jiraskova, Y. Structure and magnetic properties of synthesized fine cerium dioxide nanoparticles. J. Alloys Comp. 2018, 753, 167–175. [Google Scholar] [CrossRef]
  10. Jiraskova, Y.; Bursik, J.; Zivotsky, O.; Lunacek, J.; Janos, P. Magnetic and Mossbauer Study of Cerium-Based Reactive Sorbent. Acta Phys. Pol. A 2017, 131, 1096–1098. [Google Scholar] [CrossRef]
  11. Salutsky, M.L.; Kirby, H.W. Precipitation of actinium oxalate from homogeneous solution. Anal. Chem. 1956, 28, 1780–1782. [Google Scholar] [CrossRef]
  12. Janos, P.; Kuran, P.; Kormunda, M.; Stengl, V.; Grygar, T.M.; Dosek, M.; Stastny, M.; Ederer, J.; Pilarova, V.; Vrtoch, L. Cerium dioxide as a new reactive sorbent for fast degradation of parathion methyl and some other organophosphates. J. Rare Earths 2014, 32, 360–370. [Google Scholar] [CrossRef]
  13. Janos, P.; Henych, J.; Pelant, O.; Pilarova, V.; Vrtoch, L.; Kormunda, M.; Mazanec, K.; Stengl, V. Cerium oxide for the destruction of chemical warfare agents: A comparison of synthetic routes. J. Hazard. Mater. 2016, 304, 259–268. [Google Scholar] [CrossRef] [PubMed]
  14. Young, R.A. (Ed.) The Rietveld Method; International Union of Crystallography, Oxford University Press Inc.: New York, NY, USA, 1993. [Google Scholar]
  15. ICSD Database, Version 1.9.4; NIST/FIZ; FIZ Karlsruhe: Karlsruhe, Germany, 2014; Volume 1.
  16. Zak, T.; Jiraskova, Y. CONFIT: Mossbauer spectra fitting program. Surf. Interface Anal. 2006, 38, 710–714. [Google Scholar] [CrossRef]
  17. Il’ves, V.G.; Sokovnin, S.Y. Production and studies of properties of nanopowders on the basis of CeO2. Nanotechnol. Rus. 2012, 7, 213–226. [Google Scholar] [CrossRef]
  18. Ge, M.Y.; Wang, H.; Liu, E.Z.; Liu, J.F.; Jiang, J.Z.; Li, Y.K.; Xu, Z.A.; Li, H.Y. On the origin of ferromagnetism in CeO(2) nanocubes. Appl. Phys. Lett. 2008, 93, 062505. [Google Scholar] [CrossRef]
  19. Chen, S.Y.; Lu, Y.H.; Huang, T.W.; Yan, D.C.; Dong, C.L. Oxygen Vacancy Dependent Magnetism of CeO2 Nanoparticles Prepared by Thermal Decomposition Method. J. Phys. Chem. C 2010, 114, 19576–19581. [Google Scholar] [CrossRef]
  20. Sundaresan, A.; Bhargavi, R.; Rangarajan, N.; Siddesh, U.; Rao, C.N.R. Ferromagnetism as a universal feature of nanoparticles of the otherwise nonmagnetic oxides. Phys. Rev. B 2006, 74, 161306. [Google Scholar] [CrossRef]
  21. Sundaresan, A.; Rao, C.N.R. Ferromagnetism as a universal feature of inorganic nanoparticles. Nano Today 2009, 4, 96–106. [Google Scholar] [CrossRef]
  22. Esquinazi, P.; Setzer, A.; Höhne, R.; Semmelhack, C.; Kopelevich, Y.; Spemann, D.; Butz, T.; Kohlstrunk, B.; Losche, M. Ferromagnetism in oriented graphite samples. Phys. Rev. B 2002, 66, 024429. [Google Scholar] [CrossRef]
  23. Abreu, R.D.; Morais, C.A. Purification of rare earth elements from monazite sulphuric acid leach liquor and the production of high-purity ceric oxide. Miner. Eng. 2010, 23, 536–540. [Google Scholar] [CrossRef]
Figure 1. Fluorite-type crystal structure of CeO2 (Ce atoms are in dark grey and O atoms are in light grey).
Figure 1. Fluorite-type crystal structure of CeO2 (Ce atoms are in dark grey and O atoms are in light grey).
Metals 09 00222 g001
Figure 2. Left panel: SEM morphology of the cerium oxide samples S4 (a), S5 (b), and S6 (c). Right panel: X-ray diffraction patterns.
Figure 2. Left panel: SEM morphology of the cerium oxide samples S4 (a), S5 (b), and S6 (c). Right panel: X-ray diffraction patterns.
Metals 09 00222 g002
Figure 3. TEM micrographs of the S6 sample nanostructure (a,b) and diffraction with analysed face-centred cubic (f.c.c.) CeO2 phase (c).
Figure 3. TEM micrographs of the S6 sample nanostructure (a,b) and diffraction with analysed face-centred cubic (f.c.c.) CeO2 phase (c).
Metals 09 00222 g003
Figure 4. Measured hysteresis loops at 300 K (left panel) and 2 K (middle panel) of the S4 (a), S5 (b), and S6 (c) samples. Paramagnetic contribution is obtained after diamagnetic and ferromagnetic correction and compared to the Brillouin function (right panel).
Figure 4. Measured hysteresis loops at 300 K (left panel) and 2 K (middle panel) of the S4 (a), S5 (b), and S6 (c) samples. Paramagnetic contribution is obtained after diamagnetic and ferromagnetic correction and compared to the Brillouin function (right panel).
Metals 09 00222 g004
Figure 5. Room temperature Mössbauer spectra (a) of the S4, S5, and S6 samples prepared from different chemicals and (b) of selected chemicals, cerous chloride (CeCl) and cerous nitrate (CeN).
Figure 5. Room temperature Mössbauer spectra (a) of the S4, S5, and S6 samples prepared from different chemicals and (b) of selected chemicals, cerous chloride (CeCl) and cerous nitrate (CeN).
Metals 09 00222 g005
Table 1. Mean chemical composition (at. %) of the S4, S5, and S6 samples obtained by energy-dispersive X-ray (EDX) analysis.
Table 1. Mean chemical composition (at. %) of the S4, S5, and S6 samples obtained by energy-dispersive X-ray (EDX) analysis.
Sample/ElementAreaPoint
S4Ce28.4±4.619.1±4.1
O71.680.9
S5Ce28.6±0.224.8±5.0
O71.475.2
S6Ce27.9±0.426.2±2.7
O72.173.8
Table 2. Parameters determined from the magnetic measurements and from the Brillouin function: χD—diamagnetic susceptibility established from 300 K and 2 K hysteresis loops; Msfm 300 K—saturation magnetisation of ferromagnetic contribution at 300 K; Mspm 2 K—saturation magnetisation of paramagnetic contribution at 2 K.
Table 2. Parameters determined from the magnetic measurements and from the Brillouin function: χD—diamagnetic susceptibility established from 300 K and 2 K hysteresis loops; Msfm 300 K—saturation magnetisation of ferromagnetic contribution at 300 K; Mspm 2 K—saturation magnetisation of paramagnetic contribution at 2 K.
SampleχD
(m3/kg)
Msfm 300 K (Am2/kg)Fe Content (ppm)Mspm 2 K
(Am2/kg)
Ce3+ Concentration (-)
S4−4.5·10−90.002090.0280.0009
S5−1.1·10−80.000830.0450.0015
S6−2.1·10−80.001460.1600.0052

Share and Cite

MDPI and ACS Style

Jiraskova, Y.; Bursik, J.; Janos, P.; Lunacek, J.; Chrobak, A.; Zivotsky, O. Effect of Iron Impurities on Magnetic Properties of Nanosized CeO2 and Ce-Based Compounds. Metals 2019, 9, 222. https://doi.org/10.3390/met9020222

AMA Style

Jiraskova Y, Bursik J, Janos P, Lunacek J, Chrobak A, Zivotsky O. Effect of Iron Impurities on Magnetic Properties of Nanosized CeO2 and Ce-Based Compounds. Metals. 2019; 9(2):222. https://doi.org/10.3390/met9020222

Chicago/Turabian Style

Jiraskova, Yvonna, Jiri Bursik, Pavel Janos, Jiri Lunacek, Artur Chrobak, and Ondrej Zivotsky. 2019. "Effect of Iron Impurities on Magnetic Properties of Nanosized CeO2 and Ce-Based Compounds" Metals 9, no. 2: 222. https://doi.org/10.3390/met9020222

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