Next Article in Journal
Dissection of the Factors Affecting Formation of a CH∙∙∙O H-Bond. A Case Study
Previous Article in Journal
Pitch-Length Independent Threshold Voltage of Polymer/Cholesteric Liquid Crystal Nano-Composites
Previous Article in Special Issue
Controlled Deposition of Tin Oxide and Silver Nanoparticles Using Microcontact Printing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nickel-Doped Ceria Nanoparticles: The Effect of Annealing on Room Temperature Ferromagnetism

1
Energy Safety Research Institute (ESRI), College of Engineering, Swansea University, Bay Campus, Fabian Way Swansea SA1 8EN, UK
2
School of Chemistry, University of Manchester, Oxford Road, Manchester M13 9PL, UK
3
UCL Healthcare Biomagnetics Laboratories, Royal Institution of Great Britain, 21 Albemarle Street, London W1S 4BS, UK
4
School of Materials, University of Manchester, Oxford Road, Manchester M13 9PL, UK
*
Author to whom correspondence should be addressed.
Crystals 2015, 5(3), 312-326; https://doi.org/10.3390/cryst5030312
Submission received: 30 June 2015 / Revised: 21 July 2015 / Accepted: 17 August 2015 / Published: 21 August 2015
(This article belongs to the Special Issue Nanostructured Oxide Crystals)

Abstract

:
Nickel-doped cerium dioxide nanoparticles exhibit room temperature ferromagnetism due to high oxygen mobility within the doped CeO2 lattice. CeO2 is an excellent doping matrix as it can lose oxygen whilst retaining its structure. This leads to increased oxygen mobility within the fluorite CeO2 lattice, leading to the formation of Ce3+ and Ce4+ species and hence doped ceria shows a high propensity for numerous catalytic processes. Magnetic ceria are important in several applications from magnetic data storage devices to magnetically recoverable catalysts. We investigate the effect doping nickel into a CeO2 lattice has on the room temperature ferromagnetism in monodisperse cerium dioxide nanoparticles synthesised by the thermal decomposition of cerium(III) and nickel(II) oleate metal organic precursors before and after annealing. The composition of nanoparticles pre- and post-anneal were analysed using: TEM (transmission electron microscopy), XPS (X-ray photoelectron spectroscopy), EDS (energy-dispersive X-ray spectroscopy) and XRD (X-ray diffraction). Optical and magnetic properties were also studied using UV/Visible spectroscopy and SQUID (superconducting interference device) magnetometry respectively.

1. Introduction

Metal oxide nanoparticles have been an area of intense research and rapid development over the last twenty years, because of interest in utilizing their unique properties afforded to them by their small size. Metal oxide nanoparticles straddle the boundary between bulk materials and molecules and are a major driving force behind cutting-edge technological development [1,2,3,4,5]. Metal oxide nanoparticles have seen successful and widespread use in applications as diverse as: energy materials [6,7,8,9], thermochromics [10,11,12], magnetism [13,14,15], medicine [16,17,18,19,20], imaging [21,22,23], communications technology and data storage [24,25,26], catalysts for organic transformations [27] and superhydrophobic surfaces [15,28,29].
Cerium(IV) oxide (ceria) and metal-doped ceria nanoparticles have been developed for their role in catalysis [30,31,32,33,34,35,36]. One of the features that makes ceria attractive for catalysis is its ability to release/uptake oxygen from its lattice without losing its structural integrity [37]. Cerium(IV) oxide has the fluorite structure, in which the oxygen are packed along the same plane. This allows high oxygen mobility through the lattice as the number of oxygen vacancies to increase, with cerium atoms changing oxidation state from Ce4+ to Ce3+ increasing the propensity for redox chemistry. The ability for ceria to readily develop oxygen vacancies is key to its effectiveness as a dopant matrix [38].
Various transition and lanthanide metals have been used to dope ceria to enhance catalytic activity [31,39], performance in solid oxide fuel cells [40] and to develop magnetic properties [41,42]. Transition metals, particularly Mn [43], Fe [44], Co [45] and Ni [46] have all been doped into ceria to affect room temperature ferromagnetism and form dilute magnetic semiconductor oxides. It has been hypothesized by several groups that the room temperature ferromagnetism is due to the creation of oxygen vacancies in the CeO2 lattice. Oxygen vacancies which generate exchange interactions between unpaired electron spins and changes in surface chemical states are both postulated to affect room temperature ferromagnetism [47,48,49]. Oxygen vacancies also effect the band structure of the host CeO2, often dramatically affecting optical properties [50,51,52]. Studies involving the doping of non-magnetic elements into CeO2 such as Ca2+ [53] and Cr3+ [48] support this as they also demonstrate room temperature ferromagnetism.
Thurber et al. demonstrated that room temperature ferromagnetism in Ni-doped CeO2 is not due to the formation/coagulation of ferromagnetic nickel within the material [42]. This was determined by the absence of nickel/nickel oxide diffraction peaks in X-ray diffraction pattern (and observations of shifting of the Curie temperature with different Ni doping levels showing doping throughout the material), and the fact that ferromagnetic properties did not increase with high amounts of nickel doping. Pure phase CeO2 and heavily Ni-doped samples did not show ferromagnetism, indicating that a random impurity could be eliminated as the cause of the ferromagnetism as the magnetic moment would be identical throughout the doped/pure phase samples.
In this paper, we present the synthesis of nickel-doped cerium dioxide nanocrystals by the thermal decomposition of metal-oleate complexes in the presence of surfactants at high temperature; a route which has a high success rate in producing large quantities of highly monodisperse nanocrystals. The synthesis of Ni-doped CeO2 via this route has advantages over Fe, Co or Cr doped CeO2 due to the ease in which the solid precursor, nickel(II) oleate, can be handled and stored for long periods of time in comparison to viscid iron(III) oleate or air-sensitive cobalt(II) oleate. We investigated the effect of annealing in air on the composition, crystallinity and room-temperature ferromagnetic properties using superconducting quantum interference device (SQUID) magnetometry. Nanoparticles were characterised using transmission electron microscopy (TEM) , X-ray photoelectron spectroscopy (XPS) , UV/Visible spectroscopy (UV/Vis), X-ray diffraction (XRD) and energy dispersive X-ray spectroscopy (EDS).

2. Results and Discussion

The thermal decomposition of metal-oleate complexes has been shown to be an excellent method for the synthesis of large quantities of monodisperse metal oxide nanoparticles. The method involves the synthesis of the metal-oleate complex by a salt metathesis reaction between the metal chloride and sodium oleate, to give the metal-oleate and sodium chloride. Aside from the formation of sodium chloride, phase transfer in the two-phase solvent system (a mixture of water, ethanol and n-hexane) provides a major driving force.
Once isolated, cerium(III) and nickel(II) oleate complexes are gold and bright green waxy solids respectively, and are easier to manipulate than their iron(III) or cobalt(II) counterparts due to their higher viscosity. Ni-doped ceria nanoparticles were synthesised by decomposing different ratios of cerium(III) and nickel(II) oleate in the presence of a surfactant (oleic acid) and a high boiling point solvent (1-octadecene) at 320 °C. The ratios for different nickel doping levels are given in Table 1.
Table 1. Samples synthesized and used in this study with the amounts of starting reagents listed.
Table 1. Samples synthesized and used in this study with the amounts of starting reagents listed.
SampleCe[oleate]3 used/g, mmolNi[oleate]2 used/g, mmolSample Number
CeO24, 4.060, 01
NiO0, 02.52, 4.062
50% Ni-CeO22, 2.030.662, 2.033
10% Ni-CeO23.59, 3.650.25, 0.4064
8% Ni-CeO23.62, 3.680.2, 0.3255
5% Ni-CeO23.86, 3.80.125, 0.2036
4% Ni-CeO23.84, 3.90.1, 0.1627
3% Ni-CeO23.88, 3.940.075, 0.1228
2% Ni-CeO23.92, 3.980.05, 0.089
1% Ni-CeO24.02, 3.960.025, 0.040610
Post-decomposition, nanoparticles were isolated as black sticky residues, several ethanol washes/centrifuge cycles were used to remove excess oleate-type species and 1-octadecene. Once washed, the nanoparticles were readily dispersible in organic solvents. At this stage, the samples were divided in two, with one half kept for analysis and the other half annealed at 450 °C in air for 12 h. The annealed powders were free flowing and different in color, with the most marked example being Sample 1 changing from dark brown to yellow. The composition of nanoparticles pre- and post-anneal were characterised using: TEM, XPS, EDS and XRD. Their optical properties were also analysed using UV/Visible spectroscopy.
TEM analysis demonstrated the effectiveness of the metal-oleate decomposition method for the synthesis of large quantities of monodisperse nanoparticles (Figure 1).
Figure 1. TEM images of pre- and post annealed samples. (a) Sample 1 (CeO2 nanoparticles); (b) Sample 1 (annealed); (c) Sample 2 (NiO nanoparticles); (d) Sample 2 (annealed) showing a d-spacing of 0.238 nm assigned to the <111> plane of NiO; (e) Sample 3 (50% Ni-CeO2); (f) Sample 7 (4% Ni-CeO2); (g) Sample 9 (2% Ni-CeO2, annealed) and (h) Sample 5 (8% Ni-CeO2, annealed) showing showing a d-spacing of 0.238 nm assigned to the <111> plane of CeO2.
Figure 1. TEM images of pre- and post annealed samples. (a) Sample 1 (CeO2 nanoparticles); (b) Sample 1 (annealed); (c) Sample 2 (NiO nanoparticles); (d) Sample 2 (annealed) showing a d-spacing of 0.238 nm assigned to the <111> plane of NiO; (e) Sample 3 (50% Ni-CeO2); (f) Sample 7 (4% Ni-CeO2); (g) Sample 9 (2% Ni-CeO2, annealed) and (h) Sample 5 (8% Ni-CeO2, annealed) showing showing a d-spacing of 0.238 nm assigned to the <111> plane of CeO2.
Crystals 05 00312 g001
Sample 1 (CeO2) showed spherical nanoparticles of 1.83 ± 0.47 nm in diameter with very little shape anisotropy (on annealing 7.41 ± 2.1 nm). However, on nickel doping, shape anisotropy increased, as did overall sample polydispersity. Sample 5 gave a mean size of 2.63 nm with a standard deviation of 1.18 nm (Figure 1). TEM and therefore EDS analysis was exacting due to the amount of carbon contamination on the TEM grids from the blanket coverage of oleic acid on nanoparticle surfaces/excess oleic acid in solution. Once annealed, high resolution images were taken, as was a more accurate EDS analysis with little carbon contamination (Figure 2).
On annealing, nanoparticle size increased. Sample 5 increased from 2.63 ± 1.18 nm to 7.11 ± 3.37 nm. Annealing had a particularly marked effect on Sample 2 (NiO) which showed a very large increase in particle size (3.77 ± 0.90 nm to 21.49 ± 5.28 nm) and crystallinity, supported by peak narrowing in its XRD pattern (Figure 4j).
Characterisation by TEM was supported by compositional analysis by XPS and EDS spectroscopy. Results from composition analysis by quantitative XPS are fully summarised in Table 2. EDS spectra were effective in identifying the elements present in pre-annealed samples, although a quantitative comparison of compositions was achieved by analysing XPS spectra.
EDS spectra showed the presence of Ni, Ce and O in the pre-annealed samples, but the high levels of surfactants present (oleic acid) introduced increased background interference. As evidenced in Figure 2a, c and d, this was solved by examining the annealed samples, as annealing entirely removed this background.
Figure 2. Exemplar EDS spectra of non-annealed and annealed samples. (a) CeO2 nanoparticles (annealed); (b) NiO nanoparticles (non-annealed), showing the high carbon content from the nanoparticle ligands; (c) Ni-CeO2 (8% doping, Sample 5, annealed) and (d) Ni-CeO2 (2% doping, Sample 9, annealed).
Figure 2. Exemplar EDS spectra of non-annealed and annealed samples. (a) CeO2 nanoparticles (annealed); (b) NiO nanoparticles (non-annealed), showing the high carbon content from the nanoparticle ligands; (c) Ni-CeO2 (8% doping, Sample 5, annealed) and (d) Ni-CeO2 (2% doping, Sample 9, annealed).
Crystals 05 00312 g002
XPS was used to determine the presence and composition of elements in the doped and undoped samples. Quantitative XPS results are listed in Table 2, with all samples showing the desired ratios of Ce to Ni in accordance with the amount of precursors used (Table 1). The survey scan revealed a degree of sodium in the samples and is a possible contaminant derived from sodium oleate used in the precursor synthesis.
High resolution Ce 3d XPS spectra showed spectra typical of a Ce(IV) species, specifically for cerium(IV) oxide (CeO2) [54,55,56]. Ce4+ exhibits six peaks comprised of 3 pin-orbit doublets due to different Ce 4f occupancies post-excitation. The peak positions and doublet separation (Δ) for Sample 3 are as follows: peak positions: 880.6, 886.5, 896.3, 899.2, 905.8 and 914.7 eV, Δ = 18.5 eV).
One of the difficulties in analysing Ni-doped cerium dioxide is that the most intense signals for Ni and Ce (Ni 2p and Ce 3d) have a degree of overlap in Figure 3c. Figure 3d is the Ni 2p high resolution scan from Sample 2 (pure phase NiO) and it is clear that a substantial portion of the Ni 2p1/2 is masked by the Ce 3d3/2 peak. The full Ni2p scan is evident in Figure 3d, which is indicative of NiO (peak positions: 853.5, 855.2, 860.7, 872.0 and 879.3 eV, Δ = 18.5 eV) [54,57].
Figure 3. (a) XPS survey scan of Sample 3 (50% Ni-CeO2) showing elements present; (b) Fitted Ce 3d high resolution XPS spectrum showing the 3 doublets that are indicative of Ce(IV) and indeed CeO2; (c) Ni 2p region of Sample 3, showing the overlapping Ce 3d region which masks the Ni 2p1/2 region; (d) is a high resolution scan of the Ni 2p region in Sample 2 (pure phase NiO).
Figure 3. (a) XPS survey scan of Sample 3 (50% Ni-CeO2) showing elements present; (b) Fitted Ce 3d high resolution XPS spectrum showing the 3 doublets that are indicative of Ce(IV) and indeed CeO2; (c) Ni 2p region of Sample 3, showing the overlapping Ce 3d region which masks the Ni 2p1/2 region; (d) is a high resolution scan of the Ni 2p region in Sample 2 (pure phase NiO).
Crystals 05 00312 g003
Table 2. Nanoparticle compositional analysis by quantitative XPS spectroscopy.
Table 2. Nanoparticle compositional analysis by quantitative XPS spectroscopy.
SampleCe Composition by XPS/at.%O Composition by XPS/at.%Ni Composition by XPS/at.%
137.7962.21
249.5250.48
322.0254.5723.40
436.5159.643.86
532.7264.802.48
634.2663.881.86
732.9265.811.27
836.9862.130.90
937.9561.470.58
1033.4366.320.25
XRD patterns were easily obtained on powder samples post-annealing and demonstrated the propensity of the ceria lattice to receive metal ion dopants without significant structural re-arrangement or collapse (Figure 4). All patterns (bar NiO, Sample 2) were shown to exhibit the halite structure of cerium(IV) oxide (COD: 4343161) [50,58], with Sample 2 showing the cubic structure of NiO (COD: 1010093).
Figure 4. Powder X-ray diffraction patterns of: (a) Sample 3 (50% Ni-CeO2, annealed); (b) Sample 4 (10% Ni-CeO2, annealed); (c) Sample 5 (8% Ni-CeO2, annealed); (d) Sample 6 (5% Ni-CeO2, annealed); (e) Sample 7 (4% Ni-CeO2, annealed); (f) Sample 8 (3% Ni-CeO2, annealed); (g) Sample 9 (2% Ni-CeO2, annealed); (h) Sample 10 (1% Ni-CeO2, annealed); (i) Sample 1 (CeO2, annealed) and (j) Sample 2 (NiO, annealed). Circles represent the diffraction peaks from the standard pattern of CeO2 (COD: 4343161), squares standard data from NiO (COD: 1010093) and triangles standard data from Ni metal (COD: 1512526).
Figure 4. Powder X-ray diffraction patterns of: (a) Sample 3 (50% Ni-CeO2, annealed); (b) Sample 4 (10% Ni-CeO2, annealed); (c) Sample 5 (8% Ni-CeO2, annealed); (d) Sample 6 (5% Ni-CeO2, annealed); (e) Sample 7 (4% Ni-CeO2, annealed); (f) Sample 8 (3% Ni-CeO2, annealed); (g) Sample 9 (2% Ni-CeO2, annealed); (h) Sample 10 (1% Ni-CeO2, annealed); (i) Sample 1 (CeO2, annealed) and (j) Sample 2 (NiO, annealed). Circles represent the diffraction peaks from the standard pattern of CeO2 (COD: 4343161), squares standard data from NiO (COD: 1010093) and triangles standard data from Ni metal (COD: 1512526).
Crystals 05 00312 g004
When nickel doping was increased beyond 8%, Figure 3b,c shows the emergence of the nickel oxide pattern. In Sample 3 (1:1 molar ratio of cerium and nickel precursors), it is evident that a mixed phase system exists, supported by TEM micrographs which show two distinct nanoparticle populations. On annealing, Sample 2 (NiO) exhibited narrowed peaks compared to those of the non-annealed sample (Figure S1) consistent with nanoparticle sintering and increased crystallised size seen by TEM micrographs (Figure 2c,d). The diffraction pattern for Sample 2 also revealed that even after annealing, the NiO was not phase-pure; indeed there was a degree of Ni present as shown by reflections at 44.7° and 51.7° assigned as the <111> and <200> of Ni metal (COD: 1512526). However, this was not seen by high resolution Ni2p XPS spectra of Sample 2, which showed solely NiO. This is indirect evidence of a Ni/NiO core/shell type system as XPS is surface sensitive, whereas XRD analyses the entire sample.
As expected, the presence of Ni metal decreases when comparing the non-annealed and annealed states in sample 2 (Figure S1).XRD patterns were further analysed for unit cell parameter and unit cell volume (Table S1) using QualX software with standard data taken from the Crystallography Open Database (COD). However, there was little correlation between the cell parameter values or the unit cell volumes with nickel doping. Reports have suggested that cell parameter decreases with nickel doping due to the smaller size of Ce3+ and Ce4+ compared to Ni, Ni+ or Ni2+ [42], or that nanoscale effects overtake this, and an expansion in unit cell parameters is reported with decreasing nanoparticle size due to Ce4+ replacement with smaller Ce3+ [59,60,61].
One of the marked effects of introducing oxygen vacancies into CeO2 by transition metal doping is the alteration of the band structure. This can be seen using UV/visible spectroscopy, specifically diffuse reflectance spectroscopy on powder samples. A sample of the results are summarised in Figure 5.
Figure 5 shows the change in diffuse reflectance optical spectra with increased nickel doping. Using these data, it is possible to determine the optical band-gap of the samples after applying the Kubelka-Munk function (Equation (1)):
F ( R ) =   (   ( 1 R ) 2 R ) 2
The band gap of Sample 1 (pure CeO2) was found to be 3.81 eV, in agreement with the literature value of 3.78 eV for cerium(IV) oxide [62]. With increasing nickel doping, the band gap increases to 3.7, 3.5, 3.81 eV and 3.93 eV for 1%, 2%, 3% and 8% doping respectively. This is consistent with reports that show that ceria nanoparticles of decreasing size show an increase in bad gap energy [59]. TEM image analysis post annealing showed that particle size decreased slightly on Ni-doping, with the exception of sample 2 (pure NiO) which was dramatically different to any of the samples containing cerium. Thurber et al. reported changes in band gap energy of nickel doped ceria systems as a combination of particle size effects and structural changes brought about by interstitial Ni incorporation, even at the 1% level [42]. The authors also show a decrease in the band gap up to x ≤ 0.04 (Ce1−xNixO2) but an increase thereafter, and correlated their findings with lattice strain induced by nickel doping. Our samples did exhibit this trend, but only up to x ≤ 0.02 after which band gap energy values increased, possibly due to smaller particle sizes obtained from our syntheses.
Figure 5. Diffuse-reflectance spectra of nickel-doped CeO2 samples, with reflectance intensity decreasing markedly with increased nickel doping. “QS” is the reflectance spectrum of blank quartz slides.
Figure 5. Diffuse-reflectance spectra of nickel-doped CeO2 samples, with reflectance intensity decreasing markedly with increased nickel doping. “QS” is the reflectance spectrum of blank quartz slides.
Crystals 05 00312 g005
Along with changes in the optical properties common in materials of interest in spintronics, the magnetic properties of the Ni-doped CeO2 samples were investigated using SQUID magnetometry (Figure 6). Both pre- and post-annealed NiO samples (Sample 2) have a coercive field that increase from 14 Oe to 70 Oe post-annealing. The annealing process also impacts the saturation magnetisation with a 17% increase in saturation post-annealing, commensurate with the increase in crystallinity and particle size.
Figure 6. Magnetic hysteresis curves at 300 K for: (a) NiO annealed and non-annealed “slurry” samples (Sample 2) and (b) 5% Ni-CeO2 (Sample 6). Magnetic saturation values were calculated by removing the high field linear component (χ.H) from the data.
Figure 6. Magnetic hysteresis curves at 300 K for: (a) NiO annealed and non-annealed “slurry” samples (Sample 2) and (b) 5% Ni-CeO2 (Sample 6). Magnetic saturation values were calculated by removing the high field linear component (χ.H) from the data.
Crystals 05 00312 g006
It is noteworthy that the non-annealed CeO2 sample (Sample 1) showed no magnetism at room temperature, with the annealed sample giving a weak magnetic moment of 1.9 × 10−3 emu/g. Sample 6, 5% Ni-CeO2, showed weak ferromagnetic behaviour, however there was an absence of a coercive field in both annealed and non-annealed samples, presumably due to the small size of the nanoparticles, and the absence of ripening in the doped samples compared with NiO (Sample 2). Curiously there is however a decrease in the magnetic saturation post-annealing, which is contrary to expectations, however previous studies [42] noted a linear increase of ferromagnetic moment in dopant concentrations below 4%, with a dramatic decrease thereafter. Therefore, annealing in the case of Sample 6 could be detrimental to the overall moment due to increased crystallinity of the annealed product.

3. Experimental Section

3.1. Materials

Cerium(III) chloride (purum p.a., ≥ 98.0%), nickel(II) chloride hexahydrate (ReagentPlus), 1-octadecene (90%) oleic acid (90% technical grade) and sodium oleate (≥82% fatty acids) were purchased from Sigma-Aldrich Ltd., UK Ethanol (≥99.8% (GC)) and n-hexane (Laboratory Reagent, ≥95%) were purchased from VWR limited, UK. UHQ deionized water with a resistivity of not less than 18.2 MΩ·cm−1 (Millipore, UK) was used for aqueous solutions.

3.2. Methods

Metal-oleate syntheses: Nickel(II) and cerium(III) oleate complexes were synthesized according to a modified procedure by Park et al. [63]. Briefly, nickel(II) chloride hexahydrate (7.13 g, 30 mmol) was dissolved in deionized water (60 mL) and added to a 250 mL three-necked flask fitted with a condenser and charged with sodium oleate (18.23 g, 60 mmol), n-hexane (105 mL), deionized water (45 mL) and ethanol (60 mL). The mixture was refluxed under stirring (~70 °C) for 4 h, giving a dark green suspension in hexane. On cooling, the organic (hexane) layer was isolated and washed with deionized water (2 × 100 mL) and evaporated in vacuo to give solid green nickel(II) oleate.
Cerium(III) oleate was synthesized using the method above, with cerium(III) chloride heptahydrate (11.2 g, 30 mmol) and 3 equivalents of sodium oleate (27.4 g, 90 mmol) used to give cerium(III) oleate as a waxy solid.
Nanoparticle synthesis: Cerium(III) oleate, nickel(II) oleate or a mixture of both (see Table 1) was added to a three-necked 250 mL flask containing oleic acid (0.64 mL, 2 mmol) and 1-octadecene (30 mL). The mixture was then heated to 320 °C at a constant rate of 3.3 °C·min−1 under stirring and kept at 320 °C for 1 h, developing a vigorous reflux. On cooling, the black suspension was subjected to several ethanol washings/centrifuge cycles (10 min at 3000× g) until a black solid was isolated.
Half the product was then placed into a ceramic crucible before being annealed in air for 12 h at 450 °C at a heating rate of 10 °C·min−1, yielding free flowing powders free of ligands.
The remainder of the product was re-dispersed in n-hexane (ca. 20 mL) giving a colloidally stable black dispersion.

3.3. Instrumentation

TEM images were recorded using a FEI Tecnai G2 20 with a LaB6 source at an acceleration voltage of 200 kV. EDS spectra were taken with an Oxford XMax 80 TLE detector running AZTEC software. UV-Vis absorption spectra were recorded using a Shimadzu UV-1800 UV/Vis spectrometer (Japan) in the wavelength range 200–800 nm. X-ray photoelectron spectra (XPS) were recorded on a Kratos Axis Supra instrument (Kratos Analytical, Manchester, UK) using a monochromated Al Kα source. All spectra were recorded using a charge neutralizer to limit differential charging and subsequently calibrated to the main CxHy carbon peak at a binding energy of 284.8 eV. Survey scans were recorded at a pass energy of 160 eV and high resolution data at a pass energy of 20 eV. Data was fitted using CASA XPS with Shirley backgrounds. XRD diffraction patterns were acquired using a Bruker D8 Advance diffractometer (Germany), using a Cu Kα source (λ = 0.154018 nm) with a Gobel mirror optic and Soller slits on the detector side. Scans were acquired at a grazing incidence of 3° over a 2θ range of 20°–80° with 0.05° steps and 3 s per step. Samples for magnetometery were prepared in polycarbonate sample holders and weighed using a 6 d.p. microbalance before mounting them in a rigid brass tube specifically designed for SQUID magnetometry measurements. A Quantum Design MPMS SQUID-VSM (San Diego, CA, USA) was used to measure M vs. H curves at room temperature (300 K) with a field range of ±7T.

4. Conclusions

Ni-CeO2 nanoparticles were successfully synthesised using the thermal decomposition of metal-oleate precursors. A portion of the obtained nanoparticles were annealed at 450 °C leading to some sintering (in the case of NiO) and increased crystallinity. Nanoparticle structure and composition were characterised using: TEM, XPS, EDS and XRD, giving expected levels of nickel doping and structural deviation. Optical properties were analysed using reflectance spectroscopy, with even low levels of Ni doping (ca. 1%) having a marked effect on overall reflectance. SQUID magnetometry was used to compare annealed and non-annealed samples, with an increase in ferromagnetic moment observed on annealing of the pure NiO sample and indeed the pure CeO2 sample. The 5% Ni-CeO2 sample demonstrated a decrease on annealing, perhaps due to the drop-off of ferromagnetism in Ni-CeO2 samples with nickel doping levels of above 4%.
Magnetic ceria are important in several applications from magnetic data storage devices to magnetically recoverable catalysts, and weak room temperature ferromagnetism induced by transition metal doping is a facile method to realise the aforementioned applications at the nanoscale.

Supplementary Files

Supplementary File 1

Acknowledgments

C.W.D. and J.C.B. thank the Ramsay trust for past and current Ramsay Fellowships. John Warren (School of Materials, University of Manchester) is thanked for his generous help with XRD. James McGettrick, Tristan Watson (SPECIFIC), Danny Ovens and Jonathan Counsell (Kratos Analytical Ltd.) are thanked for XPS measurements. Andreas Kafizas (Imperial College London) is thanked for his help in interpretation of UV-Visible spectra.

Author Contributions

J.C.B. performed the experimental work and wrote the bulk of the manuscript. P.O.B. and P.D.M. provided TEM, UV/Visible and XRD characterization, P.S. performed SQUID measurements and wrote the magnetism section and C.W.D. supervised the work and conceived the idea.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ozin, G.A.; Arsenault, A.; Cademartiri, L. Nanochemistry: A Chemical Approach to Nanomaterials, 2nd ed.; Royal Society of Chemistry: London, UK, 2008. [Google Scholar]
  2. Patzke, G.R.; Zhou, Y.; Kontic, R.; Conrad, F. Oxide nanomaterials: Synthetic developments, mechanistic studies, and technological innovations. Angew. Chem. Int. Ed. 2011, 50, 826–859. [Google Scholar] [CrossRef] [PubMed]
  3. Bear, J.; Charron, G.; Fernández-Argüelles, M.T.; Massadeh, S.; McNaughter, P.; Nann, T. In Vivo Applications of Inorganic Nanoparticles. In BetaSys: Systems Biology of Regulated Exocytosis in Pancreatic β-Cells; Systems Biology Series; Springer: New York, NY, USA, 2011; Volume 2, pp. 185–220. [Google Scholar]
  4. Franke, M.E.; Koplin, T.J.; Simon, U. Metal and metal oxide nanoparticles in chemiresistors: Does the nanoscale matter? Small 2006, 2, 36–50. [Google Scholar] [CrossRef] [PubMed]
  5. Chen, X.; Mao, S.S. Titanium Dioxide Nanomaterials: Synthesis, Properties, Modifications, and Applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef] [PubMed]
  6. Zoolfakar, A.S.; Rani, R.A.; Morfa, A.J.; O’Mullane, A.P.; Kalantar-zadeh, K. Nanostructured copper oxide semiconductors: A perspective on materials, synthesis methods and applications. J. Mater. Chem. C 2014, 2, 5247–5270. [Google Scholar] [CrossRef]
  7. Hu, Y.H. A Highly Efficient Photocatalyst—Hydrogenated Black TiO2 for the Photocatalytic Splitting of Water. Angew. Chem. Int. Ed. 2012, 51, 12410–12412. [Google Scholar] [CrossRef] [PubMed]
  8. Ni, M.; Leung, M.K.H.; Leung, D.Y.C.; Sumathy, K. A review and recent developments in photocatalytic water-splitting using for hydrogen production. Renew. Sustain. Energy Rev. 2007, 11, 401–425. [Google Scholar] [CrossRef]
  9. Scanlon, D.O.; Dunnill, C.W.; Buckeridge, J.; Shevlin, S.A.; Logsdail, A.J.; Woodley, S.M.; Catlow, C.R.A.; Powell, M.J.; Palgrave, R.G.; Parkin, I.P.; et al. Band alignment of rutile and anatase TiO2. Nat. Mater. 2013, 12, 798–801. [Google Scholar] [CrossRef] [PubMed]
  10. Li, S.-Y.; Niklasson, G.A.; Granqvist, C.G. Nanothermochromics with VO2-based core-shell structures: Calculated luminous and solar optical properties. J. Appl. Phys. 2011, 109, 113515. [Google Scholar] [CrossRef]
  11. Chen, Z.; Gao, Y.; Kang, L.; Cao, C.; Chen, S.; Luo, H. Fine crystalline VO2 nanoparticles: Synthesis, abnormal phase transition temperatures and excellent optical properties of a derived VO2 nanocomposite foil. J. Mater. Chem. A 2014, 2, 2718–2727. [Google Scholar] [CrossRef]
  12. Son, J.-H.; Wei, J.; Cobden, D.; Cao, G.; Xia, Y. Hydrothermal Synthesis of Monoclinic VO2 Micro- and Nanocrystals in One Step and Their Use in Fabricating Inverse Opals. Chem. Mater. 2010, 22, 3043–3050. [Google Scholar] [CrossRef]
  13. Zhao, Y.; Dunnill, C.W.; Zhu, Y.; Gregory, D.H.; Kockenberger, W.; Li, Y.; Hu, W.; Ahmad, I.; McCartney, D.G. Low-Temperature Magnetic Properties of Hematite Nanorods. Chem. Mater. 2007, 19, 916–921. [Google Scholar] [CrossRef]
  14. Peveler, W.J.; Bear, J.C.; Southern, P.; Parkin, I.P. Organic–inorganic hybrid materials: Nanoparticle containing organogels with myriad applications. Chem. Commun. 2014, 50, 14418–14420. [Google Scholar] [CrossRef] [PubMed]
  15. Crick, C.R.; Bear, J.C.; Southern, P.; Parkin, I.P. A general method for the incorporation of nanoparticles into superhydrophobic films by aerosol assisted chemical vapour deposition. J. Mater. Chem. A 2013, 1, 4336–4344. [Google Scholar] [CrossRef]
  16. Bear, J.C.; Yu, B.; Blanco-Andujar, C.; McNaughter, P.D.; Southern, P.; Mafina, M.-K.; Pankhurst, Q.A.; Parkin, I.P. A low cost synthesis method for functionalised iron oxide nanoparticles for magnetic hyperthermia from readily available materials. Faraday Discuss. 2014, 175, 83–95. [Google Scholar] [CrossRef] [PubMed]
  17. Jordan, A.; Scholz, R.; Wust, P.; Fähling, H.; Felix, R. Magnetic fluid hyperthermia (MFH): Cancer treatment with AC magnetic field induced excitation of biocompatible superparamagnetic nanoparticles. J. Magn. Magn. Mater. 1999, 201, 413–419. [Google Scholar] [CrossRef]
  18. Dobson, J. Magnetic nanoparticles for drug delivery. Drug Dev. Res. 2006, 67, 55–60. [Google Scholar] [CrossRef]
  19. Dunnill, C.W.; Ansari, Z.; Kafizas, A.; Perni, S.; Morgan, D.J.; Wilson, M.; Parkin, I.P. Visible light photocatalysts—N-doped TiO2 by sol-gel, enhanced with surface bound silver nanoparticle islands. J. Mater. Chem. 2011, 21, 11854–11861. [Google Scholar] [CrossRef]
  20. Dunnill, C.W.; Parkin, I.P. Nitrogen-doped TiO2 thin films: Photocatalytic applications for healthcare environments. Dalton Trans. 2011, 40, 1635–1640. [Google Scholar] [CrossRef] [PubMed]
  21. Na, H.; Song, I.; Hyeon, T. Inorganic Nanoparticles for MRI Contrast Agents. Adv. Mater. 2009, 21, 2133–2148. [Google Scholar] [CrossRef]
  22. Bridot, J.; Dayde, D.; Riviere, C.; Mandon, C.; Billotey, C.; Lerondel, S. Hybrid gadolinium oxide nanoparticles combining imaging and therapy. J. Mater. Chem. 2009, 19, 2328–2335. [Google Scholar] [CrossRef]
  23. Na, H.B.; Lee, J.H.; An, K.; Park, Y.I.; Park, M.; Lee, I.S.; Nam, D.-H.; Kim, S.T.; Kim, S.-H.; Kim, S.-W.; et al. Development of a T1 Contrast Agent for Magnetic Resonance Imaging Using MnO Nanoparticles. Angew. Chem. Int. Ed. 2007, 46, 5397–5401. [Google Scholar] [CrossRef] [PubMed]
  24. Cornell, R.M.; Schwertmann, U. The Iron Oxides: Structure, Properties, Reactions, Occurrences and Uses; Wiley: Hoboken, NJ, USA, 2006. [Google Scholar]
  25. Terris, B.D.; Thomson, T. Nanofabricated and self-assembled magnetic structures as data storage media. J. Phys. Appl. Phys. 2005, 38, R199–R222. [Google Scholar] [CrossRef]
  26. Reiss, G.; Hütten, A. Magnetic nanoparticles: Applications beyond data storage. Nat. Mater. 2005, 4, 725–726. [Google Scholar] [CrossRef] [PubMed]
  27. Chng, L.L.; Erathodiyil, N.; Ying, J.Y. Nanostructured Catalysts for Organic Transformations. Acc. Chem. Res. 2013, 46, 1825–1837. [Google Scholar] [CrossRef] [PubMed]
  28. Crick, C.R.; Bear, J.C.; Kafizas, A.; Parkin, I.P. Superhydrophobic Photocatalytic Surfaces through Direct Incorporation of Titania Nanoparticles into a Polymer Matrix by Aerosol Assisted Chemical Vapor Deposition. Adv. Mater. 2012, 24, 3505–3508. [Google Scholar] [CrossRef] [PubMed]
  29. Xu, L.; Karunakaran, R.G.; Guo, J.; Yang, S. Transparent, Superhydrophobic Surfaces from One-Step Spin Coating of Hydrophobic Nanoparticles. ACS Appl. Mater. Interfaces 2012, 4, 1118–1125. [Google Scholar] [CrossRef] [PubMed]
  30. Bhatta, U.M.; Ross, I.M.; Saghi, Z.; Stringfellow, A.; Sayle, D.; Sayle, T.X.T.; Karakoti, A.; Reid, D.; Seal, S.; Möbus, G. Atomic motion on various surfaces of ceria nanoparticles in comparison. J. Phys. Conf. Ser. 2012, 371. [Google Scholar] [CrossRef]
  31. Chattopadhyay, J.; Son, J.E.; Pak, D. Preparation and characterizations of Ni-alumina, Ni-ceria and Ni-alumina/ceria catalysts and their performance in biomass pyrolysis. Korean J. Chem. Eng. 2011, 28, 1677–1683. [Google Scholar] [CrossRef]
  32. Zhang, S.; Muratsugu, S.; Ishiguro, N.; Tada, M. Ceria-Doped Ni/SBA-16 Catalysts for Dry Reforming of Methane. ACS Catal. 2013, 3, 1855–1864. [Google Scholar] [CrossRef]
  33. Pino, L.; Vita, A.; Cipitì, F.; Laganà, M.; Recupero, V. Hydrogen production by methane tri-reforming process over Ni-ceria catalysts: Effect of La-doping. Appl. Catal. B Environ. 2011, 104, 64–73. [Google Scholar] [CrossRef]
  34. Xu, J.; Xue, B.; Liu, Y.-M.; Li, Y.-X.; Cao, Y.; Fan, K.-N. Mesostructured Ni-doped ceria as an efficient catalyst for styrene synthesis by oxidative dehydrogenation of ethylbenzene. Appl. Catal. Gen. 2011, 405, 142–148. [Google Scholar] [CrossRef]
  35. Qureshi, U.; Dunnill, C.W.; Parkin, I.P. Nanoparticulate cerium dioxide and cerium dioxide–titanium dioxide composite thin films on glass by aerosol assisted chemical vapour deposition. Appl. Surf. Sci. 2009, 256, 852–856. [Google Scholar] [CrossRef]
  36. Gomez, V.; Clemente, A.; Irusta, S.; Balas, F.; Santamaria, J. Identification of TiO2 nanoparticles using La and Ce as labels: Application to the evaluation of surface contamination during the handling of nanosized matter. Environ. Sci. Nano 2014, 1, 496–503. [Google Scholar] [CrossRef]
  37. Naganuma, T.; Traversa, E. Stability of the Ce3+ valence state in cerium oxide nanoparticle layers. Nanoscale 2012, 4, 4950–4953. [Google Scholar] [CrossRef] [PubMed]
  38. Trovarelli, A. Catalytic Properties of Ceria and CeO2-Containing Materials. Catal. Rev. 1996, 38, 439–520. [Google Scholar] [CrossRef]
  39. Tao, Z.; Hou, G.; Xu, N.; Zhang, Q. A highly coking-resistant solid oxide fuel cell with a nickel doped ceria: Ce1−xNixO2−y reformation layer. Int. J. Hydrog. Energy 2014, 39, 5113–5120. [Google Scholar] [CrossRef]
  40. Sun, C.; Li, H.; Chen, L. Nanostructured ceria-based materials: Synthesis, properties, and applications. Energy Environ. Sci. 2012, 5, 8475–8505. [Google Scholar] [CrossRef]
  41. Dohčević-Mitrović, Z.D.; Paunović, N.; Radović, M.; Popović, Z.V.; Matović, B.; Cekić, B.; Ivanovski, V. Valence state dependent room-temperature ferromagnetism in Fe-doped ceria nanocrystals. Appl. Phys. Lett. 2010, 96, 203104. [Google Scholar] [CrossRef]
  42. Thurber, A.; Reddy, K.M.; Shutthanandan, V.; Engelhard, M.H.; Wang, C.; Hays, J.; Punnoose, A. Ferromagnetism in chemically synthesized CeO2 nanoparticles by Ni doping. Phys. Rev. B 2007, 76, 165206. [Google Scholar] [CrossRef]
  43. Xia, C.; Hu, C.; Chen, P.; Wan, B.; He, X.; Tian, Y. Magnetic properties and photoabsorption of the Mn-doped CeO2 nanorods. Mater. Res. Bull. 2010, 45, 794–798. [Google Scholar] [CrossRef]
  44. Brito, P.C.A.; Santos, D.A.A.; Duque, J.G.S.; Macêdo, M.A. Structural and magnetic study of Fe-doped CeO2. Phys. B Condens. Matter 2010, 405, 1821–1825. [Google Scholar] [CrossRef]
  45. Ferrari, V.; Llois, A.M.; Vildosola, V. Co-doped ceria: Tendency towards ferromagnetism driven by oxygen vacancies. J. Phys. Condens. Matter 2010, 22. [Google Scholar] [CrossRef] [PubMed]
  46. Thurber, A.; Reddy, K.M. High-temperature magnetic-field-induced activation of room-temperature ferromagnetism in Ce1−xNixO2. J. Appl. Phys. 2007, 101, 09N506. [Google Scholar] [CrossRef]
  47. Fernandes, V.; Mossanek, R.J.O.; Schio, P.; Klein, J.J.; de Oliveira, A.J.A.; Ortiz, W.A.; Mattoso, N.; Varalda, J.; Schreiner, W.H.; Abbate, M.; et al. Dilute-defect magnetism: Origin of magnetism in nanocrystalline CeO2. Phys. Rev. B 2009, 80. [Google Scholar] [CrossRef]
  48. Chen, S.-Y.; Fong, K.-W.; Peng, T.-T.; Dong, C.-L.; Gloter, A.; Yan, D.-C.; Chen, C.-L.; Lin, H.-J.; Chen, C.-T. Enhancement of Ferromagnetism in CeO2 Nanoparticles by Nonmagnetic Cr3+ Doping. J. Phys. Chem. C 2012, 116, 26570–26576. [Google Scholar] [CrossRef]
  49. 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. [Google Scholar] [CrossRef]
  50. Ranjith, K.S.; Saravanan, P.; Chen, S.-H.; Dong, C.-L.; Chen, C.L.; Chen, S.-Y.; Asokan, K.; Rajendra Kumar, R.T. Enhanced Room-Temperature Ferromagnetism on Co-Doped CeO2 Nanoparticles: Mechanism and Electronic and Optical Properties. J. Phys. Chem. C 2014, 118, 27039–27047. [Google Scholar] [CrossRef]
  51. Wang, Z.; Quan, Z.; Lin, J. Remarkable Changes in the Optical Properties of CeO2 Nanocrystals Induced by Lanthanide Ions Doping. Inorg. Chem. 2007, 46, 5237–5242. [Google Scholar] [CrossRef] [PubMed]
  52. Kumar, A.; Babu, S.; Karakoti, A.S.; Schulte, A.; Seal, S. Luminescence Properties of Europium-Doped Cerium Oxide Nanoparticles: Role of Vacancy and Oxidation States. Langmuir 2009, 25, 10998–11007. [Google Scholar] [CrossRef] [PubMed]
  53. Chen, X.; Li, G.; Su, Y.; Qiu, X.; Li, L.; Zou, Z. Synthesis and room-temperature ferromagnetism of CeO2 nanocrystals with nonmagnetic Ca2+ doping. Nanotechnology 2009, 20. [Google Scholar] [CrossRef] [PubMed]
  54. NIST X-ray Photoelectron Spectroscopy (XPS) Database, Version 3.5. Available online: http://srdata.nist.gov/xps/ (accessed on 23 February 2015).
  55. Burroughs, P.; Hamnett, A.; Orchard, A.F.; Thornton, G. Satellite structure in the X-ray photoelectron spectra of some binary and mixed oxides of lanthanum and cerium. J. Chem. Soc. Dalton Trans. 1976, 1686–1698. [Google Scholar] [CrossRef]
  56. Mullins, D.R.; Overbury, S.H.; Huntley, D.R. Electron spectroscopy of single crystal and polycrystalline cerium oxide surfaces. Surf. Sci. 1998, 409, 307–319. [Google Scholar] [CrossRef]
  57. Moulder, J.F.; Stickle, W.F.; Sobol, P.E.; Bomben, K.D. Handbook of X-ray Photoelectron Spectroscopy: A Reference Book of Standard Spectra for Identification and Interpretation of Xps Data; Physical Electronics: Eden Prairie, MN, USA, 1995; pp. 142–143. [Google Scholar]
  58. Taguchi, M.; Takami, S.; Adschiri, T.; Nakane, T.; Sato, K.; Naka, T. Supercritical hydrothermal synthesis of hydrophilic polymer-modified water-dispersible CeO2 nanoparticles. CrystEngComm 2011, 13, 2841–2848. [Google Scholar] [CrossRef]
  59. Zhang, F.; Chan, S.-W.; Spanier, J.E.; Apak, E.; Jin, Q.; Robinson, R.D.; Herman, I.P. Cerium oxide nanoparticles: Size-selective formation and structure analysis. Appl. Phys. Lett. 2002, 80, 127–129. [Google Scholar] [CrossRef]
  60. Zhou, X.D.; Huebner, W. Size-induced lattice relaxation in CeO2 nanoparticles. Appl. Phys. Lett. 2001, 79, 3512–3514. [Google Scholar] [CrossRef]
  61. Mogensen, M.; Sammes, N.M.; Tompsett, G.A. Physical, chemical and electrochemical properties of pure and doped ceria. Solid State Ion. 2000, 129, 63–94. [Google Scholar] [CrossRef]
  62. Griffiths, T.R.; Davies, M.J.; Hubbard, H.V.S.A. Spectroscopic studies on single crystals having the fluorite lattice. Part 1—The fundamental absorption edge; Urbach’s rule and the Debye temperature in CeO2. J. Chem. Soc. Faraday Trans. 2 Mol. Chem. Phys. 1976, 72, 765–772. [Google Scholar] [CrossRef]
  63. Park, J.; An, K.; Hwang, Y.; Park, J.-G.; Noh, H.-J.; Kim, J.-Y.; Park, J.-H.; Hwang, N.-M.; Hyeon, T. Ultra-large-scale syntheses of monodisperse nanocrystals. Nat. Mater. 2004, 3, 891–895. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Bear, J.C.; McNaughter, P.D.; Southern, P.; O’Brien, P.; Dunnill, C.W. Nickel-Doped Ceria Nanoparticles: The Effect of Annealing on Room Temperature Ferromagnetism. Crystals 2015, 5, 312-326. https://doi.org/10.3390/cryst5030312

AMA Style

Bear JC, McNaughter PD, Southern P, O’Brien P, Dunnill CW. Nickel-Doped Ceria Nanoparticles: The Effect of Annealing on Room Temperature Ferromagnetism. Crystals. 2015; 5(3):312-326. https://doi.org/10.3390/cryst5030312

Chicago/Turabian Style

Bear, Joseph C., Paul D. McNaughter, Paul Southern, Paul O’Brien, and Charles W. Dunnill. 2015. "Nickel-Doped Ceria Nanoparticles: The Effect of Annealing on Room Temperature Ferromagnetism" Crystals 5, no. 3: 312-326. https://doi.org/10.3390/cryst5030312

Article Metrics

Back to TopTop