Olivia
Aalling-Frederiksen
a,
Rebecca K.
Pittkowski
a,
Andy S.
Anker
a,
Jonathan
Quinson
ab,
Lars
Klemeyer
c,
Benjamin A.
Frandsen
d,
Dorota
Koziej
c and
Kirsten M. Ø.
Jensen
*a
aUniversity of Copenhagen, Department of Chemistry, Nanoscience Center, 2100 Copenhagen Ø, Denmark. E-mail: kirsten@chem.ku.dk
bAarhus University, Department of Biological and Chemical Engineering, 8200 Aarhus, Denmark
cUniversity of Hamburg, Institute for Nanostructure and Solid-State Physics, Center for Hybrid Nanostructures, Luruper Chausse 149, 22761 Hamburg, Germany
dBrighham Young University, Department of Physics and Astronomy, Provo, Utah 84602, USA
First published on 16th September 2024
We here investigate how the synthesis method affects the crystallite size and atomic structure of cobalt iron oxide nanoparticles. By using a simple solvothermal method, we first synthesized cobalt ferrite nanoparticles of ca. 2 and 7 nm, characterized by Transmission Electron Microscopy (TEM), Small Angle X-ray scattering (SAXS), X-ray and neutron total scattering. The smallest particle size corresponds to only a few spinel unit cells. Nevertheless, Pair Distribution Function (PDF) analysis of X-ray and neutron total scattering data shows that the atomic structure, even in the smallest nanoparticles, is well described by the spinel structure, although with significant disorder and a contraction of the unit cell parameter. These effects can be explained by the surface oxidation of the small nanoparticles, which is confirmed by X-ray near edge absorption spectroscopy (XANES). Neutron total scattering data and PDF analysis reveal a higher degree of inversion in the spinel structure of the smallest nanoparticles. Neutron total scattering data also allow magnetic PDF (mPDF) analysis, which shows that the ferrimagnetic domains correspond to ca. 80% of the crystallite size in the larger particles. A similar but less well-defined magnetic ordering was observed for the smallest nanoparticles. Finally, we used a co-precipitation synthesis method at room temperature to synthesize ferrite nanoparticles similar in size to the smallest crystallites synthesized by the solvothermal method. Structural analysis with PDF demonstrates that the ferrite nanoparticles synthesized via this method exhibit a significantly more defective structure compared to those synthesized via a solvothermal method.
It is well established that in general, the properties of spinel-structured ferrites are size dependent.14–16 Such size-dependent changes in material properties are often attributed to the increased surface-to-volume ratio at the nanoscale. However, the small size may also impact the atomic structure of the material, as it has been observed for several other oxide materials.17–19 Nevertheless, the challenges of characterizing the atomic structure in nanostructured materials mean that these effects are not well understood.20,21 Here, we investigate how the solvent used in a solvothermal method influences the resulting crystallite size and the atomic structure in spinel-type CoFe2O4 and reveal local structural features. The solvothermal method, which has also been widely used for ferrites, is simple, low-cost, scalable and highly tuneable, allowing for the attainment of different particle characteristics.22,23
The spinel structure, with the general formula AB2O4 is presented in Fig. 1. The oxygens are arranged in a cubic closed packed structure, while 8 metal ions occupy tetrahedral sites, and 16 metal atoms occupy the octahedral sites.25 In the normal spinel structure, the available tetrahedral sites are occupied by A2+ ions, while the octahedral sites are occupied by B3+ ions, resulting in a [A2+]Td[B3+]2OhO4 structure. However, spinels can also be inverse or partially inverse. The inverse spinel has all A2+ ions in the octahedral sites, while B3+ are distributed between the tetrahedral and octahedral sites, resulting in the formula [B3+]Td[A2+B3+]2OhO4. The partially inverse spinel has an empirical formula of [A1−x2+Bx3+]Td[Ax2+B2−x3+]2OhO4, where x is the inversion degree and represents the fraction of A2+ occupying the octahedral sites.15,16,26–28 Cobalt iron oxide is known to crystallize as an inverse spinel in the bulk state. However, nanosized particles synthesized via solvothermal or microwave synthesis procedures can adopt a partially inverse spinel structure.15,26–28 Previous studies of spinel iron oxide nanoparticles have also shown a large effect of crystallite size on the atomic structure,29–31 where vacancies are observed in the tetrahedral sites in the atomic structure, which influences their surface structure and reactivity.29 It has not yet been uncovered whether similar effects are seen for other spinel ferrites including CoFe2O4. The inversion degree is well known to affect the magnetic properties of the material.15,32–34 It has also been suggested that the cation distribution affects the catalytic properties of cobalt ferrite,35–37 and structural characterization is thus essential for being able to elucidate structure/property relationships.
![]() | ||
Fig. 1 Structural representation of a normal spinel in the space group Fd![]() |
Here, we synthesize spinel cobalt iron oxide nanocrystallites using a simple solvothermal synthesis approach and demonstrate that the spinel crystallite size can be tuned by changing the solvent from water to tert-butanol. We characterize the particles using small-angle X-ray scattering (SAXS), transmission electron microscopy (TEM), and Pair Distribution Function (PDF) analysis of X-ray and neutron total scattering data, which show that the atomic structure in even the smallest nanoparticles can be well described by the spinel structure, albeit with significant disorder. These effects can be explained by surface oxidation of the small nanoparticles, which is confirmed by X-ray near edge absorption spectroscopy (XANES). From the neutron total scattering data, we also use magnetic PDF (mPDF) to analyse the magnetic structure in the nanoparticles.38 We furthermore analyse the structure of cobalt ferrites synthesized by coprecipitation at room temperature. This synthesis leads to particles in the same size range as the small solvothermally synthesized particles; however their structure differs significantly due to the different synthesis method. Here, the PDF shows significant defects, including vacancies in the tetrahedral sites, as previously observed in pure iron oxide nanoparticles.29 The analysis using complementary structure characterization techniques thus provides insight into the structure–synthesis–size relationship of cobalt iron oxide nanocrystallites.
The scattering data were processed using PDFgetX3 software to obtain the PDFs.39 Reciprocal-space Rietveld refinement of the total scattering data was performed using FullProf software.40 Real-space Rietveld refinements of the PDFs were performed using PDFgui.41 The cubic Fe3O4 structure in space group Fdm was used in the refinements for simplification, since Fe and Co scatter X-rays almost equally.42 γ-Fe2O3 in space-group P43212 was furthermore investigated.
Rietveld refinement was performed using FullProf software,43 including both an atomic structural model of CoFe2O4 in the Fdm space group and magnetic structure in the F
space group with antiparallel magnetic moment components on the tetrahedral and octahedral sites. The scattering patterns collected at each of the four detector banks were refined simultaneously.
PDFgetN3 was used for data reduction to obtain the PDFs.44 The experimentally obtained PDFs contain information on both the atomic and magnetic configuration, as the neutrons interact with both nuclei and magnetic moments of the unpaired electrons. Therefore, the atomic PDFs were first modelled using PDFgui software,45 and the residual was treated as the mPDF, which was modelled with the diffpy.mpdf package in DiffPy-CMI.46
Previous studies of this synthesis approach have indicated that the product from the tert-butanol synthesis may be amorphous, which would manifest in the diffraction profile as very broad features.10 To further investigate this hypothesis, and to highlight the impact of different acquisition conditions on the quality of the collected diffraction patterns, both samples were investigated with X-rays under three different conditions: with Ag Kα radiation (λ = 0.56 Å), with Cu Kα radiation (λ = 1.54 Å) as is used in most laboratory diffractometers, and with Cu Kα radiation applying energy discrimination on the detector to suppress fluorescence. For diffraction using Cu Kα radiation, the standard voltage range of the detector discriminator is 0.110–0.250 V. To suppress fluorescence, the voltage range used is instead set to 0.190–0.270 V based on recommendations by the manufacturer.
The data obtained with standard Cu Kα settings, Fig. 2A and B, are dominated by a large background signal arising from fluorescence. Cu Kα radiation causes significant fluorescence from Co and Fe and is thus not well suited for the analysis of cobalt ferrites. Our data show almost flat diffraction profiles, which may lead to the wrong interpretation that completely disordered or amorphous nanoparticles have formed. A clear improvement in the data quality is observed when using energy discrimination on the detector. The best data are obtained using Ag Kα radiation, whose energy (λ = 0.56 Å) does not result in significant fluorescence from Co and Fe. The results emphasize the importance of selecting appropriate instruments for data collection from the samples in question. Special attention should be given when investigating disordered or nanocrystalline materials, as the Bragg peak profile may become smeared and fluorescence can dominate the patterns.
The use of Ag Kα radiation also allows data collection over a wide Q-range. Here, we acquired data up to a Qmax of 17 Å−1. PDFs from the samples are shown in Fig. 2C, which demonstrate that the water-synthesized nanoparticles exhibit extended structural order compared to the tert-butanol-synthesized nanoparticles. However, as visible in the inset of Fig. 2C, the two PDFs are locally quite similar, differing only in a slight shift of the PDF peaks towards lower r-values for the PDF of the tert-butanol-synthesized sample. In the inset of Fig. 2C, the first inter-atomic correlations are highlighted. The closest metal-to-oxygen distance (M–O, with M = Fe and Co) appears in the PDF at 2.0 Å. The edge-sharing octahedra give rise to a metal-to-metal (MOh–MOh) PDF-peak at 3.0 Å, while the corner-sharing octahedra–tetrahedra metal-to-metal distance (MOh–MTd) is evident in a peak at 3.5 Å.
![]() | ||
Fig. 3 Real-space Rietveld refinement of the (A) water and (B) tert-butanol-synthesized nanoparticles. In both cases, the spinel structure in space-group Fd![]() |
The data thus show that even the smallest particles retain the spinel structure. However, the refinement parameters presented in Fig. 3A and B show a contraction of the unit cell parameters, a, when exchanging water with tert-butanol. It is also observed that the atomic displacement parameters, (ADPs) Uiso, for the tetrahedral and octahedral metal sites are more than twice as large for the tert-butanol-synthesized nanoparticles compared with the water-synthesized nanoparticles. Similar results were obtained for Rietveld refinements of the XRD patterns in Q-space (Table S3 and Fig. S7 in ESI†). Besides describing the thermal vibrations of the atoms in the samples, the ADP-values also account for displacements resulting from static disorder.51,52 This points towards the formation of ultra-small spinel cobalt iron oxide nanoparticles with some structural disorder compared to the larger, more ordered nanoparticles formed in water. We note that the structural parameters of the particles formed in water are consistent with earlier studies on similar cobalt iron oxide materials.15,53
Previous in situ formation studies of CoFe2O4 and the closely related maghemite γ-Fe3O4 nanoparticles have shown an expansion of the unit cell for small nanoparticles compared to the bulk value.53–55 This effect has been explained by surface-restructuring in small nanoparticles, which results in an expansion of the crystallographic unit cell size. However, we observe the opposite trend here, as the refined unit cell parameter for the 2 nm particles is smaller than that of the 9 nm particles (8.30 Å vs. 8.40 Å). This effect could be related to the amount of cobalt in the sample. The smaller ionic radius of Co compared with Fe means that the unit cell parameter, a, of Co3O4 spinel is smaller than that for CoFe2O4.56 However, from SEM-EDS measurements, presented in Fig. S8–S9 and Tables S4 and S5,† we obtain a Co:
Fe atomic ratio of 1
:
2 for both samples, ruling out the first possible explanation. Another explanation might be that the larger relative surface area of the smallest crystallites leads to more oxidized nanoparticles, which effectively decreases the ionic radii and therefore, results in a smaller refined unit cell for the smallest particles.
To further confirm this, we use X-ray absorption spectroscopy, XAS, to investigate the electronic state of the samples. Fig. 3C and D show the Fe K-edge XAS spectra from the samples, along with two reference spectra from γ-Fe2O3 (maghemite)50 and Fe3O4 (magnetite).24 Both iron oxide references have the spinel structure; however maghemite has vacancies in the octahedral sites to achieve charge balance.
The Fe K-edge XANES spectra of the nanoparticles synthesized in water align with the magnetite reference spectra. The XANES spectra of nanoparticles synthesized in tert-butanol exhibit a 1 eV shift of the white line to higher energies, indicating an increase in the Fe oxidation state of Fe3+, consistent with the reference spectra of maghemite. The derivatives of the normalized μ(E) are presented in Fig. S10 in the ESI.† The two samples exhibit a similar Co K-edge shape, as shown in their XAS spectra in Fig. 3E; however, an absorption edge shift for the tert-butanol-synthesized nanoparticles to higher energies also suggests a more oxidized state of Co, Fig. 3F. The observations from our XAS analysis thus indicate that more oxidized nanoparticles form in tert-butanol, which means that vacancies are most likely present in the small spinel nanoparticles to accommodate charge compensation. We investigate this explanation further below.
We note here that data collection strategies should again be carefully considered regarding characterization of nanoscale materials and the information that can be extracted from PDF data. As mentioned above, X-ray total scattering data for PDF analysis were also collected using synchrotron radiation in the RA-PDF set-up,57 where a wide Q-range is obtained by using high energy X-rays (here λ = 0.1619 Å) and positioning the detector close to the sample (here 283 mm). This results in lower Q-resolution than in the Ag Kα diffractometer, and thus a lower limit for which crystallite sizes can reliably be quantified from PDF refinements. The refinement is shown in Fig. S6† and refined parameters are listed in Table S2 in the ESI.† PDF modelling of the synchrotron data shows an average crystallite size of only 4 nm for the water-synthesized nanoparticles (obtained using the one phase model). The crystallite size is thus clearly underestimated, again illustrating the need for a critical approach in data collection methods. The different Q-resolutions for the two data collection strategies are expressed in the Qdamp parameters, which describe instrumental broadening. While this parameter refines to 0.011 Å−1 for the Ag Kα diffractometer, it is 0.034 Å−1 for the synchrotron data.
TEM micrographs from the samples are shown in Fig. 4A and B. These confirm that close-to-spherical particles on the nanometer scale have been synthesized. A high level of polydispersity is observed for the water-synthesized nanoparticles, consistent with the findings of the PDF refinements. The high degree of agglomeration in the sample challenges the size determination from TEM; however, the nanoparticles synthesized in water have size in the range of 4–20 nm, comparable to the crystallite size of 7 nm obtained from PDF analysis. For the tert-butanol-synthesized sample, size estimation is difficult from the TEM images due to the agglomeration of the small nanoparticles. Additional TEM data are shown in Fig. S11 and S12 in the ESI.†
![]() | ||
Fig. 4 TEM micrographs of the (A) water-synthesized and (B) tert-butanol synthesized nanoparticles. SAXS refinements of the (C) water-synthesized and (D) tert-butanol synthesized nanoparticles. |
Complementary SAXS measurements were also conducted. The SAXS data were modelled with a polydisperse sphere model. The results from the SAXS refinements shown in Fig. 4C and D, reveal sizes of 7.7 and 4.1 nm for the water and tert-butanol synthesized nanoparticles, respectively. Our refinements reveal a wide size distribution for both samples as seen from the lognormal distribution presented in the insets of Fig. 4C and D. The refined SAXS parameters are listed in Table S6 in the ESI.† The full Q-range of the data collected is also presented in Fig. S13 in the ESI.†
Together, PDF, TEM and SAXS thus show that significantly larger particles are formed in water compared to tert-butanol. A broad size distribution is observed especially for the water synthesized samples, and particles in the size range of ca. 4 to 20 nm are obtained. The nanoparticles synthesized in tert-butanol falls in the size range of around 2 nm.
We use [Co1−xFex]tet[CoxFe2−x]octO4 in space group Fdm as the structural starting model for the refinements, constraining the Co
:
Fe atomic ratio to 1
:
2 as determined by SEM-EDS analysis (Fig. S7, S8, Tables S4 and S5 in ESI†). The nPDF fits are shown in Fig. 5, and refined parameters are given in Table S7 in ESI.† The full fitting range (1.5–60 Å) of the water synthesized sample is presented in Fig. S15 in the ESI.† From the nPDF refinements, we obtain inversion degrees of 0.57 and 0.46 for the spinel structures of the water- and tert-butanol-synthesized particles, respectively. This gives the formulas [Co0.43Fe0.57]Td[Co0.57Fe1.43]OhO4 for the water-synthesized nanoparticles, and [Co0.54Fe0.46]Td[Co0.46Fe1.54]OhO4 for the tert-butanol synthesized particles. Bulk CoFe2O4 has an inversion degree of 1, and our data thus confirm that nanoparticles have a more mixed character than bulk materials, as it has been demonstrated in related studies.28,53 Possibly, a decreasing size results in more mixing; however, analysis of data from a wider range of different sizes and data points would be needed to confirm this trend.
The nPDF refinements presented in Fig. 5A and B show a wave-like feature in the fit-residual shown in green. This underlying systematic feature in the fit-residual appears too regular to be mere noise and can be attributed to a contribution from the magnetic structure.38,59 Magnetic PDF (mPDF) analysis were therefore performed on the residuals of the nuclear PDF analysis and the fits are shown in brown. The refined parameters are listed in Table S8 in the ESI.†
The magnetic structure is described with a collinear model, where the magnetic moments on the tetrahedral and octahedral sites are oriented antiparallel. This has been described before for CoFe2O4 by Andersen et al.15,60 The magnetic model was constructed from the refined nuclear structure by adding magnetic species with antiparallel alignment on the tetrahedral and octahedral sites. The spin direction was 100 in the coordinate system of the cubic structure. The average magnetic moment size and the ratio of the moment size on the octahedral site to the moment on the tetrahedral site were refined, together with a parameter representing a spherical magnetic domain size. The average ordered moment for the water-synthesized sample was 3.83(13) μB, consistent with the expectations for mixed Co and Fe moments. The magnetic correlation length for the water-synthesized sample was determined to be 6.5 nm from refinements, which is 80% of the crystallite size refined from the neutron scattering data. The data thus reveal that the magnetic coherence is smaller than the crystallite size itself. This finding is consistent with a study of the magnetic structure of non-stoichiometric spinel iron oxide nanocrystallites, where the magnetic domain size was found to be 60–70% of the nanocrystallites.61Q-space Rietveld refinement is shown in Fig. S16† and refined parameters are listed in Table S9 in the ESI† for the water-synthesized nanoparticles. Here we obtain similar results, which are comparable to the refined values from the PDF-analysis. Hence, the lattice parameter was refined to 8.38 Å and an inversion degree of 0.63 was obtained.
We use the same approach to describe the contribution from a magnetic structure in the fit-residual of the nuclear nPDF refinement for tert-butanol synthesized nanoparticle, Fig. 5B. Here, the magnetic correlations length was determined to be 2.0 nm, which is close to the average crystallite domain size.
The Q-space data and PDF for the nanoparticles synthesized at RT are plotted with the data from the particles of the solvothermal synthesis in Fig. S17 in the ESI† for comparison. At first glance, the Q-space data show features at positions expected for the spinel structure. However, the PDFs show significant differences. When performing a PDF refinement using the Fe3O4 spinel structural model in space-group Fdm (Fig. 6A), the spinel Fd
m structure cannot fully describe the MOh–MOh to MOh–MTd ratio highlighted in purple in Fig. 6A. A similar effect has previously been observed in PDF analysis of small nanoparticles of iron oxide,29,55 where it was shown that vacancies in the tetrahedral sites appear with decreasing nanoparticle size. In these studies, the particles were found to adopt the maghemite structure, γ-Fe2O3, in space-group P43212.24,50 As described above, charge-balance in maghemite is achieved through ordered vacancies on the MOh-site. The MOh-sites (Fe4) which have occupancies of 0.33 are highlighted in orange in the structure presented in Fig. 6B. Here, we perform a similar analysis of the cobalt iron oxide nanoparticles at room temperature.
![]() | ||
Fig. 6 PDF analysis of the precipitate in water collected at RT. (A) PDF refinement using the magnetite model in space-group Fd![]() |
Using maghemite γ-Fe2O3 as the structural model and refining the MTd site occupancies, we improve the description of the experimental PDF, as shown in Fig. 6B. The MTd site occupancy is refined to 0.59, and the spherical crystallite size is refined to 1.8 nm. All refined parameters are listed in Table S10 in the ESI.† The fit quality is further improved if we allow refinement of all metal positions. This fit is presented in Fig. S18 in the ESI.†
We note that when using the γ-Fe2O3 structural model to describe the PDF of the tert-butanol-synthesized nanoparticles discussed above, no significant improvement in the fit quality is achieved compared to the magnetite structure used in Fig. 3B. The fit is shown in Fig. S19 in the ESI.† The MTd site occupancy is refined to 0.81.
Our analysis thus shows that the ultra-small cobalt ferrite oxide nanoparticles that form at RT in water are significantly more defect-rich compared to the spinel-structured tert-butanol synthesized particles of similar size. When synthesized in water at RT, the particles show similarities with spinel nanoparticles of iron oxide, where a certain size is needed before the ideal spinel structure can describe the structure.29 The results thus illustrate a clear effect of the synthesis method and particle size on the atomic structure of the materials.
To further address size/structure/synthesis relationships, we synthesized ultra-small nanoparticles from co-precipitation at room temperature in water. Unlike the particles synthesized by a solvothermal method, we were not able to describe these with the simple spinel structure due to a high concentration of vacancies observed on the tetrahedral sites in the structure, altering the local atomic structure. The observation of vacancies aligns well with the size-dependent structure observed for iron oxide nanoparticles.29,55
Our study thus shows that size–structure relationships are not always specific for a given material, but are also highly influenced by the synthesis method. Apart from the structural analysis, we have provided data showing how appropriate data collection strategies are required, as the use of Cu Kα radiation can lead to misleading conclusions regarding the crystal structure of Co and Fe containing nanomaterials.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4na00590b |
This journal is © The Royal Society of Chemistry 2024 |