You are on page 1of 11

Article

pubs.acs.org/JPCC

Structural, Chemical, and Magnetic Investigations of Core−Shell Zinc


Ferrite Nanoparticles
J. A. Gomes,† G. M. Azevedo,‡,# J. Depeyrot,*,†,⊥ J. Mestnik-Filho,§ F. L. O. Paula,† F. A. Tourinho,∥
and R. Perzynski⊥

Complex Fluids Group, Instituto de Física, Universidade de Brasília, Caixa Postal 04455, 70919-970 Brasília (DF) Brazil

Instituto de Física, Universidade Federal do Rio Grande do Sul, Av. Bento Gonçalves 9500, Caixa-Postal 15051, 91501-970 Porto
Alegre (RS) Brazil
§
Complex Fluids Group, Instituto de Química, Universidade de Brasília, Caixa Postal 04478, 70904-970 Brasília (DF), Brazil

Instituto de Pesquisas Energéticas e Nucleares, Av. Prof. Lineu Prestes 2242, 05508-000 São Paulo (SP), Brazil

PECSA, Université Pierre et Marie Curie, UMR 7195, case 51, 4 place Jussieu, 75005 Paris, France

ABSTRACT: We investigate here the internal structure of


zinc ferrite nanoparticles designed and prepared by a soft
chemistry method to elaborate magnetic nanocolloids. The
strategy used to avoid acid dissolution modifies the chemical
composition of the surface of the nanoparticles, which are
described as a core of stoichiometric zinc ferrite surrounded by
a maghemite shell. Measurements of X-ray absorption near-
edge spectroscopy, extended X-ray absorption fine structure,
and X-ray diffraction are undertaken to investigate the local
structure of nontreated nanocrystals and of surface-treated ones as a function of their sizes. The qualitative analysis of X-ray
absorption results indicates a nonequilibrium cation distribution among the interstitial sites of the zinc ferrite nanocrystals core.
Ab-initio calculations of theoretical photoelectron backscattering phases and amplitudes give, by fitting Fourier transformed
EXAFS data at both Zn and Fe K-edges, an average inversion degree of 0.34. This value well matches the result of Rietveld
refinement of X-ray diffraction data. Magnetization measurements performed on dilute aqueous nanocrystal dispersions, liquid at
room temperature and frozen at low temperatures, are carried out in order to test the obtained results.

I. INTRODUCTION With this purpose, the “soft-chemistry” method has been


Magnetic fluids, which are colloidal dispersions of magnetic developed to synthesize nanosized ferrites particles dispersible
nanoparticles in a liquid carrier, constitute a very attractive and in aqueous media.7−10 It leads to spinel-type nanocrystals
promising class of nanomaterials. Thanks to the singular within 10 nm in size, with general molecular formula MFe2O4
combination of their liquid and magnetic properties, these where M is a divalent transition metal, namely Mn, Co, Ni, Cu,
materials develop original responses to a powerful external or Zn. Spinel ferrites crystallize in a face-centered cubic lattice
parameter, the magnetic field. They may thus be confined, (Fd3m) formed by a close-packed arrangement of 32 oxygen
displaced, deformed, and controlled. These unique and striking anions, creating 64 interstices of tetrahedral symmetry (A) and
features make them suited for a large number of applications in 32 interstices of octahedral symmetry (B). These sites are
quite a diverse range of fields from engineering 1 to partially occupied: only 1/8 of the A-sites and 1/2 of the B-sites
biomedicals.2,3 are filled by metallic cations. The following crystallographic
Biological applications basically require nanoparticles func- representation can be used to specify the cation distribution
tionalized with biological effectors to make them interact with among the tetrahedral and octahedral sites:
or bind to biological targets. This aim is usually achieved [M(1−x)2+Fex3+]A[Mx2+Fe(2−x)3+]BO42−. Here x is the inversion
promptly, thanks to the high chemical reactivity of the degree defined as the fraction of A-sites occupied by
nanocrystals surface. Among the different magnetic materials, Fe3+cations or the fraction of B-sites occupied by M2+ cations.
spinel ferrite nanoparticles present several advantages that To disperse properly ferrite nanoparticles in acidic media and
justify their use in biomedical applications. From the chemical obtain chemically stable magnetic nanocolloids, it is necessary
point of view, they are not easily oxidized to nonmagnetic to avoid their dissolution during the synthesis process. Thus, a
species,4 if properly protected. They offer a very reactive core−shell strategy has been developed which consists in
surface, which allows to attach ligands5 or to introduce electric
charges6 in order to tune their interfacial properties. They also Received: June 5, 2012
present low sedimentation rates, high colloidal stability, and Revised: October 8, 2012
facility to be functionalized with biological molecules. Published: October 16, 2012

© 2012 American Chemical Society 24281 dx.doi.org/10.1021/jp3055069 | J. Phys. Chem. C 2012, 116, 24281−24291
The Journal of Physical Chemistry C Article

protecting the stoichiometric ferrite nanocore with a thin coat section IV, taking into account the chemical core−shell model
of maghemite.10 As various metallic elements M can be used, it while fitting the experimental EXAFS and XRD data. Then, this
leads to nanosized particles with different magnetic properties. crossed analysis provides us with values of the inversion degree
As an example, for magnetic hyperthermia applications, the that are compared with the value deduced from low
conversion of electromagnetic energy into heat can be temperature magnetization of isolated particles.
increased by 1 order of magnitude, taking advantage of the
exchange coupling between a magnetically hard core and II. EXPERIMENTAL METHODS AND COMPUTATIONAL
magnetically soft shell.11 However, on the very small spatial PROCEDURE
scales, new intricate magnetic effects begin to play and modify A. Elaboration of Core−Shell Nanoparticles. Chemical
such properties, like the appearance of superparamagnetism, Synthesis. The investigated samples are ZnFe2O4-based
finite size effects, and surface ones.12 For example, Ni ferrite nanoparticles of aqueous magnetic fluids, which are prepared
nanoparticles are well-known to present magnetic properties as described beforehand.9,10 All the reagents used in this work
associated with their disordered surface spins.13,14 In order to are products of analytical purity. In a first step, the nanocrystals
study specifically this surface magnetism, one could use are chemically synthesized by hydrothermal coprecipitation and
nanoparticles with a ZnFe2O4 core, a ferrite that is not are obtained by alkalinizing 1:2 mixtures of M2+ and Fe3+ salts
magnetic at room temperature. Indeed, in a zinc ferrite crystal with NaOH, at 100 °C and under vigorous stirring. During this
with an ideal normal spinel structure, superexchange inter- process, one can control the mechanisms of nucleation and
actions between Fe3+ cations located at A- and B-sites do not crystalline growth that rule their sizes. For example, it has been
exist since the filled tetrahedral sites are only occupied by shown that acting on the velocity of mixing of the reagents,
nonmagnetic Zn2+ cations. from a quick to a slow dropping procedure, while the
As we will show, the spatial confinement at the nanoscale temperature, reagent concentrations, and stirring rate are
completely modifies the problem since the stoichiometric maintained constant, allows monitoring the nanoparticle size.
ZnFe2O4 cores produced here are magnetic at room temper- Then, in this work, by varying the addition rate of the reagents,
ature. Several works have already shown the existence of a we have obtained a series of precipitates containing nano-
significant magnetic moment of stoichiometric zinc ferrite particles with mean size ranging typically between 8 and 14 nm.
nanoparticles at room temperature, and these findings are At the end of this step, the particles are chemically
usually attributed to a partially inverted structure.9,10,15−19 homogeneous and made of stoichiometric zinc ferrite. As
Thus, it leads us to suspect that cationic redistribution occurs they have to be further dispersed in acidic medium, their
within the nanocrystals cores, making the synthesized ZnFe2O4 surface needs to be protected against acid dissolution. Thus, in
nanoparticles as suited for bioapplications as particles of the a second step, the precipitate is submitted to an acid cleaning
other ferrites. The inversion degree in redistributed spinels that consists in washing twice with water and once in a HNO3
depends on several factors as, for example, synthesis procedure, solution (2 mol/L) in order to reverse the charge of the
spatial confinement on nanoscale, and nature as well as size of nanoparticles and eliminate undesirable less soluble byproducts
the cations.19−24 In stoichiometric zinc ferrite nanocrystals, the formed during the first step. Then, each precipitate is
cation distribution has been extensively determined using hydrothermally treated with a 1 mol/L Fe(NO3)3 solution at
Rietveld refinement of neutrons, diffraction spectra25,26 or X- 100 °C in order to ensure the thermodynamical stability of the
ray ones,27,28 in-field Mö ssbauer spectroscopy measure- nanocrystals, avoiding their degradation in acidic medium. The
ments,17,19,29,30 nuclear magnetic resonance,30 X-ray absorp- surface treatment induces iron enrichment of the nanoparticles,
tion,21,22,32−34 and X-ray circular magnetic dichroism.34 thus creating a superficial layer of maghemite that surrounds
The aim of the present work is to understand these finite size the particle core made of a zinc ferrite.32 In the third and last
effects by means of a systematic study on the local scale of such step, the nanoparticles, now of core−shell type, are dispersed in
ZnFe2O4 nanograins. In particular, we investigate nontreated aqueous medium by monitoring the pH (surface potential and
nanocrystals and surface-treated ones with different mean sizes density of charge) and the ionic strength (screening of the
in order to study the influence on their local structure with the surface potential) of the solution. The nanoparticles powders
surface treatment and the size variation. We check here their analyzed in this work are obtained after evaporation of the
core−shell structure by combining synchrotron X-ray absorp- aqueous liquid. The produced samples are then labeled Z1 to
tion near-edge spectroscopy (XANES), extended X-ray Z5. Moreover, during the elaboration of sample Z4, some
absorption fine structure (EXAFS), and X-ray diffraction amount (called Z4A) of the precipitate is also collected soon
(XRD) experiments. These studies, performed on powders after the coprecipitation step (nontreated nanoparticles) in
obtained at each step of the nanocolloid synthesis, are order to study the influence of the hydrothermal surface
complemented by SQUID magnetization measurements that treatment on their local structure.
are realized with “gas-like” dilute dispersions of independent Chemical Analyses. The chemical composition of the
nanoparticles. For sake of clarity, the paper is organized as nanoparticles is determined using suitable chemical titrations.
follows. In section II, we lay out the details of the experimental Iron(III) titration is performed by dichromatometry. Zinc(II)
methods and the computational procedure we use to titration is quantified using inductively coupled plasma atomic
characterize the structure of the nanoparticles. In the section emission spectroscopy (ICP-AES).
III, we present the chemical, structural, and magnetic (room B. Synchrotron X-ray Studies. Powder Diffraction. Data
temperature) general characteristics of the nanocrystals of X-ray powder diffraction (XPD) are collected at room
combined with qualitative results about their local structure temperature at the D12A-XRD beamline of the Brazilian
as obtained by XANES and EXAFS. The overall results suggest Synchrotron Light Laboratory (LNLS), using a double crystal
a nonequilibrium cation distribution among interstitial sites of Si(111) monochromator with constant offset, a graphite (002)
the structure that enhances the magnetic response of the analyzer, and a scintillation counter as detector. Samples are
nanoparticles. These results are subsequently discussed in placed in a cylindrical sample holder that is spun during data
24282 dx.doi.org/10.1021/jp3055069 | J. Phys. Chem. C 2012, 116, 24281−24291
The Journal of Physical Chemistry C Article

collection in order to minimize effects of an eventual to the values known from crystallographic data. For Fe K-edge
preferential orientation. The measurements are performed (Zn K-edge) only 7 (7) variables have been used during the fit,
with monochromatic X-ray beam, λ = 2.0633 Å (6.01 keV), while 29 (15) independent parameters are available according
with an approximate area of 4 × 1.5 mm2. Diffraction patterns to Nyquist’s criterion. Second, the fits of nanostructures spectra
are obtained typically within 20° ≤ 2θ ≤ 130° interval scanned are performed with the values of E0 and S02 obtained for the
with an angle step of 0.04°. crystalline standard. Debye−Waller factors are modeled as a
The simplified integral breadth method, proposed by sum of structural and dynamical terms, the latter being
Williamson and Hall,35 allows for the deconvolution of the accounted for by a correlated Einstein model with the Einstein
size and strain contributions to line broadening, by using β* = temperature as fitting parameter, cation distribution, intera-
(1/DXR) + 2εd*, where β* = (β cos θ)/λ denotes the integral tomic distances, and coordination numbers being allowed to
breadth in reciprocal space, β being the line broadening, λ the vary. The χ2 and the R-factor express here the quality of all the
X-ray wavelength, 2θ the Bragg diffraction angle, and d* the fits.40
corresponding reciprocal interplane spacing. Using the so-called C. Magnetization Measurements. Magnetization meas-
Williamson−Hall plots, β* versus d*, allows us to determine urements of dilute solutions of magnetic fluid are performed at
the values of the diameter DXR and lattice strain ε. 300 and 5 K using a superconducting quantum interference
In order to complete the analyses, Rietveld structure device (SQUID) in the field range 0−4 × 103 kA/m. Their
refinement of the XPD data is performed using the GSAS volume fraction of nanoparticles ϕp, their pH, and their ionic
software package developed by Larson and Von Dreele.36,37 It strength are adjusted in such a way that the colloidal dispersion
provides a fitting procedure of experimental results and allows can be considered as a “gas of individual particles” at room
determining here the lattice parameter, the oxygen position, temperature.14,41 If the freezing process is sufficiently fast, the
and the cation distribution. The structure refinements are made dispersion state of the magnetic solution is the same at 5 K,
with a peak shape modeled by a pseudo-Voigt function that allowing therefore the determination of the high-field magnet-
includes correction for peak asymmetry and background ization of isolated nanoparticles.
intensity well accounted by a Chebyshev polynomial function.
Each diffractogram is fitted until convergence, the best result III. RESULTS
being selected on the basis of reliability factor of refinement A. Characteristics of Core−Shell Nanoparticles. XRD
(Rwp) and quality of fit (χ2). patterns of samples Z4A and Z4 are presented in Figure 1, and
Absorption Spectroscopy Measurements. XAS spectra are
collected in transmission mode, at 20, 150, and 300 K, at the
LNLS D04B-XAS beamline around the Fe and the Zn K-edges.
Each spectrum corresponds to an average over three
independent scans. The spectra are calibrated in energy by
simultaneous measurements of transmission spectra of Fe and
Zn foils. The energies E0 of the first inflection point at the
absorption edge of these reference samples are respectively
7112 and 9662 eV.
Background removal and isolation of EXAFS oscillations is
performed by means of the AUTOBK algorithm.38 Then, the
extended fine structure appearing above the absorption edge is
isolated and normalized to the edge step height and energy,
thus resulting on a spectrum with a “per-atom” basis scale. The
raw experimental data in energy space is reduced to
photoelectron wave vector and then Fourier transformed to Figure 1. X-ray powder diffraction patterns of zinc ferrite nano-
radial coordinates using a Hanning window (between 2.6 and particles collected before (Z4A) and after (Z4) the protective surface
11 Å−1 for Fe K-edge and 2.5 and 12.0 Å−1 for Zn K-edge). treatment. The intensity of the diffracted beam is plotted as a function
Thus, in this form the data directly reflect the average local of the scattering angle, 2θ, in degrees. The inset shows the
environment around the absorbing atoms. Filtered EXAFS Williamson−Hall plot for both samples; β is the full width at half-
spectrum is obtained by back Fourier transformation between maximum of the diffraction peaks, calculated accounting for additional
source of broadening by using a Si standard monocrystal.
1.0 and 4.6 Å. These processing steps are performed using the
IFEFFIT package,38 and theoretical photoelectron backscatter-
ing phases and amplitudes are calculated ab initio via the FEFF8 the same features are obtained for all others samples. Using
code.39 For each of the cations, theoretical standards are Bragg’s law, the diffraction peaks are indexed and associated
generated with the cation occupying either the octahedral or with only one crystalline phase, which corresponds to the spinel
the tetrahedral sites. Fits of the experimental spectra are structure. Then, the averaged parameter of the cubic lattice ⟨a⟩
performed (in real space) taking into account the weighted sum is determined. Table 1 lists the obtained values from the
of all these contributions. analysis of the eight most intense lines. For all samples, the
The determination of the structural parameters is done as cubic cell size varies between 8.41 and 8.44 Å, to be compared
follows. First, the filtered EXAFS spectra of the ZnFe2O4 with the Zn ferrite bulk value of the Joint Committee on
reference compound taken at different temperatures are fitted Powder Diffraction Standards (JCPDS card No. 82-1049),
simultaneously. Debye−Waller factors σ2, interatomic distances ⟨a⟩ASTM = 8.441 Å. These results also indicate that nano-
of each path, energy threshold E0, and inelastic losses S02 are particles of different sizes and resulting from several chemical
allowed to vary while coordination numbers N for all relevant syntheses crystallize in a structure presenting similar lattice
multiple scattering paths and the cations distribution are fixed parameters. The inset of Figure 1 shows the Williamson−Hall
24283 dx.doi.org/10.1021/jp3055069 | J. Phys. Chem. C 2012, 116, 24281−24291
The Journal of Physical Chemistry C Article

Table 1. Characteristics of ZnFe2O4-Based Nanoparticlesa The chemical characteristics and the values of the room
temperature magnetization of the nanoparticles are also given
⟨a⟩ ε tsh mp
samples DXR (nm) (Å) (10−3) χM ϕs/ϕp (nm) (kA/m) in Table 1. The content of metallic cations is measured by
Z4A 9.6 8.42 0 0.33 0 0 208
chemical titrations giving the molar fraction of divalent metal,
Z1 7.9 8.41 2 0.20 0.42 0.71 171
χM. The chemical core−shell model,10 based on a core of
Z2 8.5 8.42 2 0.19 0.45 0.83 163
stoichiometric ferrite and a shell of maghemite, allows one to
Z3 11.7 8.42 2 0.23 0.32 0.79 188
deduce the shell volume fraction ϕs/ϕp and the shell thickness
Z4 12.4 8.43 2 0.26 0.23 0.58 207
tsh, calculated by using the crystalline size DXR. In Table 1 one
Z5 13.9 8.44 1 0.27 0.20 0.55 187
can see that χM increases, while ϕs/ϕp decreases, as the particle
size increases. Indeed, the surface-to-volume ratio largely
a
DXR is the crystalline size and ε the strain, both deduced from the
Williamson−Hall method, excepted for sample Z4A, deduced from
increases with decreasing size, therefore offering more contact
Rietveld analysis;44 ⟨a⟩ is the average lattice parameter deduced using between the particles and the Fe(NO3)3 solution during the
the Bragg’s law for the eight most intense diffracted line; χM is the hydrothermal treatment performed to protect them against acid
molar fraction of divalent metal calculated using the molar dissolution. The maghemite shell thickness is always of the
concentrations of each metal measured by chemical titrations; ϕs/ϕp order of the lattice parameter of the spinel cell.
is the volume fraction of the maghemite shell and tsh its thickness, both The nanoparticle magnetization mp obtained at 300 K is
determined10 using the values of DXR; mp is the particle magnetization listed in the last column of Table 1. The values show that our
determined9 for samples Z1 to Z5 from the Langevin analysis of ZnFe2O4-based nanoparticles present a magnetization that is, at
magnetization curves, obtained at 300 K with dilute liquid dispersions, room temperature, comparable to the values obtained for the
as a function of the applied field. For sample Z4A, the magnetization is
measured on powder sample, and mp is the value of the magnetization
others ferrite nanoparticles9,10 and very different from the zero
at 4 × 103 kA/m. value expected for the bulk material. It therefore suggests cation
inversion in the ZnFe2O4 core of the nanoparticles with the
presence of Zn2+ ions at octahedral sites.
B. XANES. Figure 2a compares the experimental XANES
plot where β cos θ is plotted against sin θ for all the indexed spectra at Zn K-edge for bulk zinc ferrite and both
lines of the diffractograms. Here, β is calculated accounting for nanocrystalline samples Z4A and Z4. The spectra exhibit the
additional source of broadening by using a Si standard same three peaks, with the central one (9667 eV) more intense
monocrystal. A linear extrapolation is applied to this plot, in the nanoparticles spectra than in the bulk spectrum, and the
and then the intercept gives λ/DXR and the slope gives 4ε. The
values of the nanocrystals mean size are collected in Table 1.
They show that as the velocity of mixing of the reagents varies
from a quick to a slow dropping procedure, the nanoparticles
average diameter continuously increases from 7.9 nm for
sample Z1 to 13.9 nm for sample Z5. The size difference
between nontreated nanoparticles (sample Z4A, DXR = 9.6 nm)
and surface-treated ones (sample Z4, DXR = 12.4 nm) is due to
the acid cleaning step performed soon after the coprecipitation,
which leads to small particles dissolution and therefore shifts
the average size toward larger value. Recently, such a method
has been successfully employed on the X-ray peak profile to
determine the lattice strain in maganites 42 and zinc
ferrite43nanocrystals. The inset of Figure 1 shows that particles
obtained soon after coprecipitation step (sample Z4A) do not
present any strain. On the contrary, for samples Z1 to Z5 (only
shown for sample Z4), a nonzero slope is observed, indicating
the existence of internal stresses. The obtained strain values are
collected in Table 1. Then, one can conclude here that the
enrichment of the surface with iron induced by the surface
treatment is accompanied by strain of the nanocrystal lattice,
this effect being more pronounced as the nanocrystals size
decreases. One would wonder if any nanosized particles do not
suffer from the same kind of stress and always present nonzero
strain. However, several studies, based on the Williamson−Hall
plot to separate the line broadenings relative to the grains size
effect and the strain effects have indicated that this is not always
the case. In cerium oxide nanoparticles, it suggests no
heterogeneous strain in nanoparticles smaller than 15 nm.44 Figure 2. (a) Experimental XANES spectra, obtained at Zn K-edge, of
zinc ferrite bulk material and nanoparticles samples Z4A and Z4. (b)
In cobalt ferrite nanoparticles, the milling-induced high FEFF calculations of the Zn K-edge in ZnFe2O4 nanoparticles as a
coercivity has been associated with the microstructural strain function of the degree of site substitution. The solid black and red
deduced from the Williamson−Hall plot.45 In this same study, lines correspond to Fe absorbers at the Fe and Zn sites, respectively.
the coprecipitated particle submitted to a low temperature The dashed lines correspond to the Zn Ferrite presenting and
annealing and mechanical milling remains in the 10−20 nm size increasingly higher site inversion. The raw FEFF energy scale was
and does not present strain before and after mechanical milling. offset by 10 eV for better comparison with the experimental results.

24284 dx.doi.org/10.1021/jp3055069 | J. Phys. Chem. C 2012, 116, 24281−24291


The Journal of Physical Chemistry C Article

shoulder, centered in the bulk spectrum at 9675 eV, absent in


the nanoparticles spectra. Figure 2b shows the absorption
spectra calculated for several values of the inversion degree
using the FEFF 8.4 code39 for Zn ferrite structures. Because
XAFS probes the average environment around the absorbing
atom, the calculations described in more detail elsewhere46 here
include contributions due to Zn atoms located in both
tetrahedral and octahedral sites (solid, black, and red curves,
respectively). The evolution of the XANES spectra with the
degree of site inversion is depicted in Figure 2b as dashed lines.
They are obtained from weighted averages of the spectra
corresponding to Zn in tetrahedral and octahedral sites (solid
lines). Our calculations show that as the inversion degree
increases, the intensities of the central peak, located at ∼9667
eV, increases and the intensity of the shoulder, located at 9675
eV, diminishes, in good qualitative agreement with our
experimental results. Two comments are worth to be made.
First, our observations are in good agreement with both the
experimental and theoretical results reported in ref 22, in which
similar FEFF calculations of bulk ZnFe2O4 were performed
with the same SCF and FMS parameters, in the so-called “Z + 1
approximation”. Second, this qualitative behavior with increas-
ing inversion degrees is the same as the one observed in the
experimental spectra of Z4A and Z4 samples. Such results show
unambiguously the presence of Zn2+ ions at octahedral sites of
nanoparticles both before and after the surface treatment step
of the synthesis. Figure 3. (a) Experimental XANES spectra, at Fe K-edge, of zinc
Room temperature XANES spectra at Fe K-edge, of ferrite and Fe3O4 bulk materials and nanoparticles samples Z4A and
references compounds, of sample Z4A and sample Z4 are Z4. The inset displays the distinctive pre-edge resonances. (b) FEFF
presented in Figure 3. The absorption edge energy is taken as calculations of the Fe K-edge in ZnFe2O4 nanoparticles as a function
the maximum of the first derivative of the normalized of the degree of site substitution. The solid black and red lines
correspond to Fe absorbers at the Fe and Zn sites, respectively. The
absorbance. The values obtained for nanoparticles and bulk dashed, dotted, and dot-dashed lines correspond to the Zn ferrite
ZnFe2O4 samples are equal, whereas the absorption edge of presenting and increasingly higher site inversion. The inset exhibits a
other reference compounds, such as Fe3O4, FeO, and Fe, is magnification of the pre-edge region.
shifted by a few electronvolts to lower energy due to their
smaller average iron valence. The energy edges deduced from region. It is readily noticeable that the same trends observed for
the spectra of Figure 3 shows, as expected, that the average the pre-edge peaks in the experimental data are reproduced by
valence of iron ions in the synthesized nanoparticles and in bulk the FEFF calculations. It thus corroborates the existence of
samples is always equal to +3. Similar observations hold for the cation inversion in our nanoscaled samples, with iron ions in
spectra taken at the K-edge of the divalent metal and indicate both tetrahedral and octahedral sites, differently from bulk zinc
that the average oxidation state of zinc cations is equal to +2. ferrite, where Fe3+ ions are not present in tetrahedral sites.
Moreover, we observe that neither the surface treatment with C. EXAFS. Figure 4 presents Fourier-transformed EXAFS
ferric nitrate nor the nanograins size influences the valence state data taken at 20 K at Zn K-edge, for samples of nanoparticles
significantly. Z4A and Z4 and of bulk material. In all the cases, the Fourier
The inset of Figure 3 exhibits a magnification of the pre-edge
region, which shows a small peak associated with the electronic
1s → 3d quadrupole and 1s → 3d/4p dipole transition. The
prepeak observed for magnetite spectrum is more intense than
for bulk zinc ferrite as its intensity is larger for cations localized
at tetrahedral sites than for cations at octahedral sites.22,46 Both
Z4A and Z4 nanoparticles samples have intermediate prepeak
intensities, then corroborating the existence of iron ions both in
tetrahedral and octahedral sites, differently from bulk zinc
ferrite, where no Fe3+ ions occupy tetrahedral sites. Our Fe K-
edge XANES experimental results are in good qualitative
agreement with Fe K-edge FEFF calculations presented in
Figure 3b, where the XANES spectra due to Fe ions located at
the Fe and Zn crystallographic sites (solid red and black lines,
respectively) are singled out. The dashed, dotted, and dot-
dashed lines correspond to the Zn Ferrite presenting
increasingly higher degree of inversion and are obtained, as in Figure 4. Magnitude of the k3-weighted Fourier-transformed EXAFS
Figure 2b, by suitable weighted average of the solid lines. As in signals taken at 20 K, at Zn K-edge, for zinc ferrite bulk material and
Figure 3, the inset shows a magnification of the pre-edge nanoparticles samples Z4A and Z4.

24285 dx.doi.org/10.1021/jp3055069 | J. Phys. Chem. C 2012, 116, 24281−24291


The Journal of Physical Chemistry C Article

transform amplitude is extracted from EXAFS signal using the


same parameters. Then the data can be represented in the same
coordinate system and allow a direct comparison. All spectra
exhibit two major contributions always located around the same
radial distances, indicating that both nanocrystals and bulk
compound present very similar local structures, a result
consistent with XRD patterns of Figure 1. For radial
coordinates larger than 4 Å, the observed peaks have small
amplitudes showing that EXAFS data are overall mainly
sensitive to the first coordination shells. The amplitudes of
the spectra of nanoparticles are systematically lower than bulk
counterpart due to structural disorder and/or reduced
coordination number. Indeed, the atoms localized in the
surface shell are likely under-coordinated, and their proportion
largely increases with the size reduction. The first peak of bulk
Figure 5. Average coordination number of zinc ions in the ith
spectrum, centered close to 1.5 Å, is associated with the first coordination shell normalized to its bulk value NNi (i = 1−9), as a
coordination shell of oxygen atoms around the zinc cations. function of the nanoparticles diameter. The inset shows the size
The second one, located around 3 Å, is assigned mainly to the dependence of the ratio between the coordination number of the
second shell of neighboring atoms and includes simple and second and first shells. The vertical line around 5 nm indicates the
multiple scattering paths involving zinc ions located at threshold at which the variations of the coordination numbers occur in
tetrahedral sites. The nanocrystals spectra present the same the same proportion.
two main contributions. Nevertheless, one peak develops
around 2.6 Å as a consequence of additional scattering paths
involving zinc ions located at octahedral sites. An analogous behavior is observed for Fe ions. Thus, these
This behavior therefore confirms the previous analyses of calculations demonstrate that, in our samples of nanoparticles,
XANES data, and the whole of these results strongly support a which are all well larger than 5 nm in diameter, our hypothesis
cationic distribution of metallic cations inside the nanocrystals on the variations of the coordination number is sound.
different from that of ideal ZnFe2O4 material. Moreover, this Figure 6a shows, for samples Z4A and Z4, the Fourier-
cation inversion occurs during the coprecipitation step of the transformed EXAFS data taken at 20 K at Zn K-edge (open
synthesis as indicated by the spectra obtained for sample Z4A symbols) and their respective best fit (solid line) obtained as
from Figures 2−4. described in section II. Filtered contributions of the first

IV. DISCUSSION
In the following we propose a crystalline core−shell model of
nanoparticles, probed here by combining the quantitative
analyses of EXAFS spectra with Rietveld refinement of XRD
data. In order to account for the modifications of nanocrystal
structure induced by the surface treatment during the synthesis
procedure, we first examine sample Z4A. In that sample, the
nanoparticles have not been surface-treated and thus do not
present any maghemite shell. Then, the associated surface-
treated sample Z4 is examined using the core−shell scheme for
Rietveld analysis of XRD spectra, and the same method is
employed for nanoparticles with different sizes. Finally, we
show that magnetization results obtained at low temperature
corroborate our structural analysis.
A. EXAFS. Fitted Fourier-transformed EXAFS data are
obtained as detailed in part B of section II. Measurements
performed at various temperatures are fitted simultaneously in
order to reduce the correlation between the Debye−Waller
factor and the coordination number. Moreover, we impose that
the variations of the coordination numbers occur in the same
proportion for all coordination shells.
To check the validity of this hypothesis, we calculated using a
FORTRAN code, and considering an ideal zinc ferrite
structure, the average coordination number of each coordina-
tion shell as a function of the nanoparticle diameter. The
variations of NNi, the average coordination number of the ith
shell normalized to its bulk value, are obtained for Zn ions (1 ≤
i ≤ 9) and presented in Figure 5. Its inset depicts the size Figure 6. (a) Fourier transform amplitude of the experimental EXAFS
dependence of the ratio NN2/NN1, and the whole Figure 5 spectrum (dots) and its fit (line) obtained for samples Z4A and Z4at
typically shows that above a diameter of 5 nm, the relative Zn K-edge at 20 K. (b) Filtered contributions of the first shells of
variations of the NNi are approximately the same whatever i. coordination for samples Z4A and Z4.

24286 dx.doi.org/10.1021/jp3055069 | J. Phys. Chem. C 2012, 116, 24281−24291


The Journal of Physical Chemistry C Article

Table 2. Interatomic Distances, in Å, Obtained by Fitting the Fourier-Transformed EXAFS Data, at Both Fe and Zn K-Edgesa
sample Fe(A)−O Fe(B)−O Fe(B)−MC Fe(A)−MC Zn(A)−O Zn(B)−O Zn(B)−MC Zn(A)−MC
bulk 2.05 2.96 1.92 3.54
Z4A 1.85 2.02 2.91 3.41 1.92 2.09 3.01 3.53
Z4 1.85 2.01 2.90 3.40 1.92 2.09 3.01 3.53
Z5 1.85 2.01 2.90 3.40 1.92 2.09 3.01 3.53
Z1 1.85 2.01 2.90 3.40 1.92 2.09 3.01 3.53
a
The associated error is ±0.01 Å.

coordination shell for both samples are presented in Figure 6b. 65% of its ideal bulk value. A similar reduction of the average
At both Fe and Zn K-edges, the fit of the spectra includes coordination of zinc ions is observed after analysis at the Zn K-
contributions of single and multiple scattering. As expected for edge. Such reduced average coordination is ascribed to surface
ferrite nanocrystals and as we observed here, only simple effects, since atoms sitting near the surface of the particles have
scattering paths are relevant to describe the signal of the first smaller coordination numbers as compared to atoms sitting
two coordination shells around the absorbing atoms.47 That is within the core. In the case of surface-treated nanoparticles
not the case for more distant coordination shells, where (sample Z4), the average coordination of iron atoms decreases
multiple scattering contributions should necessarily be taken even more and reaches only 48% of the bulk coordination
into account. As seen in Figures 6a and 6b, the agreement number. This value being much smaller than the one
between the experimental data and the fit is good in the 1−3.5 corresponding to the particles obtained soon after the
Å range, for both nontreated and surface treated nanocrystals. coprecipitation suggests that the iron atoms, incorporated to
Similar agreement is observed for all the other samples analyzed the nanocrystal shell during the protective surface treatment,
in this work. are largely under-coordinated. On the contrary, at Zn K-edge,
Table 2 summarizes the results of interatomic distances, in NN varies from 69% in sample Z4A to 78% in sample Z4. The
angstroms, obtained at both edges for all the samples. As an larger value obtained with surface-treated nanoparticles is
example, Fe(A)−O stands for the distance between iron ions in attributed to zinc ions, which, after the surface treatment, are
tetrahedral sites and the first layer of oxygen anions, Fe(B)− mainly localized within the nanocrystal core and therefore more
MC stands for the distance between the iron ions in octahedral coordinated. At both edges, size-sorted samples Z1 and Z5 also
sites and the first layer of metallic cation (iron or zinc, present approximately the same reduction of the average
depending on the cation distribution in the nanocrystal), and coordination when compared to sample Z4.
Zn(B)−O corresponds to the distance between the inverted Table 3 also lists the filling rate of tetrahedral sites
zinc ions in octahedral sites and the first layer of oxygen ions. determined at both Zn and Fe K-edges and the inversion
The values relative to nanoscaled samples present interatomic degree x deduced from the analysis at Zn K-edge and associated
distances very close to the values obtained for the reference with the zinc ferrite core of the nanoparticles in surface treated
bulk material, therefore confirming that all the nanomaterials samples. As mentioned in the Introduction, ideal zinc ferrite
investigated here crystallize in a spinel-type arrangement. The crystallizes in a normal spinel structure, with zinc ions located
comparison of the interatomic distances between samples Z4A at tetrahedral sites and iron ions located at octahedral sites (x =
and Z4 shows that the ion−ion distances are not affected by the 0). Then, for bulk sample, the percentage of zinc ions at
protective surface treatment. Moreover, it can also be seen that tetrahedral sites is found to be 100, as expected. For sample
the collected values do not depend on the nanocrystal size. Z4A, the results obtained at Fe K-edge indicate that 34% of the
The values of the relative average number of coordination tetrahedral sites are occupied by iron ions. Furthermore, the
NN, obtained now by considering the contributions of all the fraction of A-site filled by zinc ions, derived from EXAFS fits at
coordination shells, are presented in Table 3. At Fe K-edge, the Zn K-edge, is 64%. The agreement between both amounts,
average coordination of iron atoms, for nanoparticles obtained derived from a collection of different experiments, can be
soon after the coprecipitation step (sample Z4A), is found to be considered as excellent since their sum almost matches 100. It
suggests that nanoparticles obtained just after the coprecipita-
Table 3. Results of Fourier-Transformed EXAFS Data at tion step already present a different cation distribution as
Both Zn and Fe K-Edgesa compared to the normal spinel structure. It is characterized by
an inversion degree deduced at Zn K-edge equal to 0.36(2),
filling rate at A-sites which corresponds to the migration, on average, of 2.88 zinc
(%)
ions from A-sites to B-sites, per unit cell. In the literature, the
NN Fe K-edge NN Zn K-edge Fe Zn occurrence of such a partial inversion appears as strongly
samples (%) (%) K-edge K-edge x
dependent on the method used to prepare the ZnFe2O4
bulk 100 100 00 100 0
nanomaterial. Methods that use high-temperature treatment
Z4A 65 69 34 64 0.36
generally lead to the ideal normal spinel structure,20,25,48
Z4 48 78 52 64 0.36
whereas high energy ball milling,17,26 mechanochemistry,49,50
Z5 51 82 53 67 0.33
and soft chemical routes15,16,22,27−29 induce a nonequilibrium
Z1 54 81 55 68 0.32
cation distribution among the interstitial sites. The inversion
a
NN is the relative average number of coordination, in percentage of degree, found here for our coprecipitated nanoparticles,
the ideal bulk value; x is the inversion degree deduced from the compares well with the values relative to others nanocrystals
analysis at Zn K-edge (associated with the zinc ferrite core of the also obtained by soft chemical routes.
nanoparticles in surface-treated samples). The associated errors are After the protective treatment of the nanoparticles surface,
±3%, ±5%, and ±5%, respectively. the fraction of A-sites occupied by zinc ions remains constant,
24287 dx.doi.org/10.1021/jp3055069 | J. Phys. Chem. C 2012, 116, 24281−24291
The Journal of Physical Chemistry C Article

equal to 64%, as shown from the analysis at Zn K-edge for distribution in the ZnFe2O4 core is fixed to the value obtained
sample Z4. It suggests that the cation distribution of the from the Rietveld refinement of sample Z4A, a value which
ZnFe2O4 nanoparticle core is not affected by the surface agrees well with EXAFS determinations and is independent of
treatment. Table 3 also presents the cation distributions the nanocrystal size.
obtained for samples Z1 and Z5. Both show an inversion One can wonder if such an analysis would work since the
degree different from that of the bulk material but very similar surface layer of maghemite could be too thin in order to allow
to the value determined for sample Z4, equal to 0.36. for a reliable result. However, all of the fits presented better
At Fe K-edge, the value found for the filling rate of reliability factors whenever the two phases are taken into
tetrahedral sites by iron ions increases from 34% to 52% and account. Figure 7a,b well illustrates this result for sample Z4.
the sum of the filling rates determined at both edges is no more
around 100. This is not surprising as the tetrahedral sites now
seen at Fe K-edge are no longer related to the zinc ferrite
particle core but associated with the whole particle. It therefore
means that the iron ions incorporated to the maghemite surface
shell mainly localizes at tetrahedral sites. Indeed, the hydro-
thermal surface treatment induces, first, the release of zinc ions
from the nanocrystal shell and, second, the enrichment of this
shell by iron ions.10 This is clearly seen with the crystallo-
graphic representation deduced from our results. Using the
filling rate of tetrahedral site derived at Fe K-edge, the formula
that corresponds to the particle structure obtained just after the
coprecipitation step writes [Zn5.32+Fe2.73+]A[Zn2.72+Fe13.33+]B-
O322−. After the surface treatment, this representation does not
stand anymore for the whole particle structure but only for that
of the particle core. The formula associated with the crystalline
structure of the shell is now the maghemite formula that writes
[Fe83+]A[Fe13.33+□2.7]BO322−. Thus, one can deduce that at
octahedral sites, the vacancies represented by □, correspond to
the release of zinc ions from the outer layer and at tetrahedral
sites, the released zinc ions are substituted by iron ions.
B. Rietveld Refinement of XRD. We have already fitted X-
ray diffraction data of sample Z4A by using the Rietveld analysis
performed in a one-phase (spinel-type) model, and the
procedure is detailed elsewhere.46 The nanoparticles mean
Figure 7. Rietveld refinement of XRD patterns. (a) Sample Z4: the
size deduced from the refinement of the diffracted intensity refinement is performed using only one spinel-type phase (χ2 = 7.24,
profile is 9.6 nm. The lattice parameter is found to be equal to Rwp = 8.34%). (b) Sample Z4: the refinement is performed using two
8.430(1) Å, in good agreement with the standard JCPDS value spinel-type phases (χ2 = 4.86, Rwp = 5.25%); X-ray data are shown as
of 8.441 Å (JCPDS card No. 82-1049). Note that the value of plus marks; the solid lines are the best fit to the data and the tic marks
Table 1 equal to 8.42 Å, obtained by using Bragg’s law applied show the positions of the allowed reflections of the used phases. The
to the eight most intense diffracted lines, well agrees with that lower curves represent the difference between observed and calculated
given by the Rietveld approach. The value found for the profiles. The inset of (b) displays the density of the nanoparticles dp
inversion degree, x = 0.33(2), matches excellently well our deduced from the Rietveld refinement, as a function of the
EXAFS determination, x = 0.36(2). nanoparticle diameter DXR.
Let us now focus on the X-ray diffraction data corresponding
to the samples obtained after the surface treatment. Because of
the enrichment of the nanocrystal surface with iron atoms, a Figure 7a presents the Rietveld refinement of sample Z4
more realistic modeling is obtained in terms of a crystalline obtained using only one spinel-type phase: the reliability factors
core−shell model. Rietveld refinements of the diffraction are χ2 = 7.24 and Rwp = 8.34%. When two spinel-type phases
patterns of samples Z1 to Z5 are therefore performed using are used, these parameters become χ2 = 4.86 and Rwp = 5.25%.
two phases of spinel type, namely, the zinc ferrite core and the Their values are significantly lower, and Figure 7b also shows
protective surface layer of maghemite.10 The percentage of each for sample Z4 the better agreement between calculated and
phase corresponds to the values determined by chemical experimental data. In particular, it can easily be seen that the
analysis and collected in Table 1. Both the nanoparticles contribution of the [440] peak is better fitted. All the other
diameter and the strain parameter are fixed by the results samples present similar fitting quality by using the two spinel-
obtained from the Williamson−Hall method presented in type phases, which is confirmed by the low values of χ2 and Rwp
section III. Indeed, in our case, there is no way to obtain the full collected in Table 4. In each sample, we have also tested the
width at half-maximum separately for each phase. However, the refinement results for slight variations of the nanoparticles size
Williamson−Hall is a good approximation to determine the and strain. No improvement of the reliability factors is obtained
average line width. We use it in the refinement with two phases, when considering variations of about 10%. The fit is for all the
and we see that the result is better than refining with a single samples better when using two spinel-type phases, and this
phase. Thus, in our fit, the refined structural parameters are the result can be probably attributed, first, to the large relative
cubic cell sizes, acore and ashell, and the oxygen positions, ucore volume of the shell, which, for samples Z1 to Z5, varies
and ushell, of each phase. Then, for each sample, the cation between 20% and 45% of the whole particle, respectively.
24288 dx.doi.org/10.1021/jp3055069 | J. Phys. Chem. C 2012, 116, 24281−24291
The Journal of Physical Chemistry C Article

Table 4. Crystallographic Parameters of Core−Shell Nanoparticles Deduced from Rietveld Analysis of XRD Data Using Two
Spinel Ferrite Phasesa
samples acore (Å) ashell (Å) ucore ushell dp (g/cm3) χ2 Rwp (%)
Z1 8.445 (1) 8.395 (1) 0.253 (1) 0.259 (1) 5.09 2.23 3.96
Z2 8.458 (1) 8.413 (1) 0.251 (1) 0.258 (1) 5.05 2.93 5.07
Z3 8.431 (1) 8.396 (1) 0.255 (1) 0.259 (1) 5.16 4.36 4.95
Z4 8.431 (1) 8.396 (1) 0.254 (1) 0.258 (1) 5.21 4.86 5.25
Z5 8.425 (1) 8.432 (1) 0.256 (1) 0.253 (1) 5.21 2.77 5.04
a
acore and ashell are the cubic cell size of the zinc ferrite core and the maghemite shell, ucore and ushell are the oxygen parameter of these two phases, dp
is the density of the core−shell nanoparticles, χ2 characterizes the quality of fit, and Rwp is the reliability factor of refinement.

Second, it has to be considered the relatively high contrast in the case of zinc ferrite, reduces to mS = 10xNdμB/MM. We
between the Zn and Fe X-ray form factors. therefore performed magnetization measurements at low
Table 4 also lists the cubic lattice parameter determined for temperature in order to evaluate the inversion degree associated
the ferrite core, which matches well with the value of standard with the ZnFe2O4 core of investigated nanoparticles.
bulk material. Note that the cell size obtained for the Figure 8 shows the magnetization curves performed at 5 K
maghemite shell is larger than the maghemite reference value, on dilute magnetic solutions of samples Z1, based on smaller
equal to 8.346 Å (JCPDS card No. 39-1346), and smaller than
the lattice parameter of the core phase. Indeed, the maghemite
layer can be seen as an epitaxial continuation of the core
structure, and this last result illustrates the reduction of the cell
mismatch at the core−shell interface. Moreover, as already
pointed out for sample Z4A, the values of the oxygen positions,
determined for both the spinel phases of the core−shell model,
do not correspond to a perfect compact packing. The value
obtained for the maghemite shell is larger than the value
corresponding to the zinc ferrite core. It is probably related to
the protective surface treatment used to avoid the acid
dissolution of the particles that enhances the distortions of
oxygen atoms mainly at the surface layer.
Finally, the density dp of the core−shell nanoparticles can be
determined by using a weighted sum of core and shell Figure 8. Magnetization curves recorded at 5 K for samples Z1, Z3
contributions where dcore = 5.33 ± 0.04 g/cm3 and dshell = 4.76 and Z4.
± 0.03 g/cm3 are the densities, directly obtained from the
Rietveld refinement, of the core and of the shell, respectively. particles, and Z3 and Z4, based on larger ones. All show a good
Both values well compare with generally used ones around 4.9 saturation performance and the high field magnetization that
g/cm3 for maghemite51,52 and 5.3 g/cm3 for ZnFe2O4.53,54 As can be measured is 368, 391, and 404 kA/m for particles of 7.9,
shown in the inset of Figure 7b, the density of the nanoparticles 11.7, and 12.4 nm, respectively. These large and unexpected
increases as the nanocrystal size increases, since the proportion values are related to the presence of Fe3+ at A-sites of the spinel
of the maghemite shell becomes smaller. A linear extrapolation structure associated with the zinc ferrite core. Indeed, strong
of the nanoparticles density as a function of the nanoparticles superexchange interactions between Fe3+ ions located at both
diameter DXR would lead to the density of ZnFe2O4 bulk sites of the spinel nanostructure can develop. From this high
material for particle sizes larger than 20 nm. field values at low temperature, the cation inversion parameter
C. Magnetization Measurements. In spinel ferrites, the x related to the redistributed zinc ferrite core was determined,
metallic cations located at A- and B-sites occupy the nodes of considering the saturation magnetization of the nanocrystal as a
two subnetworks of spins and the superexchange interaction sum of weighted contributions of a zinc ferrite core and a
between them favors antiparallel alignment of spins, leading to maghemite shell. Then, using, the volume fraction of core and
antiferromagnetic ordering. However, thanks to the difference shell of Table 1 and a magnetization of the maghemite shell
between the numbers of A- and B-sites and to the way they are equal to 350 kA/m,56 it is found x = 0.295 for sample Z1, x =
occupied, the overall behavior is ferrimagnetic. Then, if the 0.33 for sample Z3, and x = 0.34 for sample Z4. These values
distribution of the metallic cations inside A- and B-sites and the are in excellent agreement with those deduced fitting EXAFS
magnetic moment of each ion are known, the saturation spectra (see Table 3) and with the value found by Rietveld
magnetization of the ferrite at T = 0 K writes55 mS = (Nd/ analysis of X-ray diffractogram for nontreated nanoparticles,
MM)[∑BnB,B − ∑AnB,A]μB, where nB,i is the number of Böhr which is also relative to the ZnFe2O4 core of surface-treated
magnetons, μB, associated with the i site of the unit cell, MM is nanoparticles.
the molar mass of ferrite, d its density, and N is Avogadro’s
number. The summation extends over the A and B sites of the V. CONCLUSION
unit cell. In nanocrystals based on spinel ferrites, cation We have investigated at the atomic level the structure of ferrite
redistribution can modify significantly the properties of nanoparticles especially designed to chemically synthesize
saturation magnetization when compared to bulk material. aqueous magnetic nanocolloids. These particles are made of a
However, it would be theoretically possible to deduce the zinc ferrite core surrounded by a protective maghemite shell.
inversion degree x from the value of the magnetization, which, Because of the close connection between magnetic properties,
24289 dx.doi.org/10.1021/jp3055069 | J. Phys. Chem. C 2012, 116, 24281−24291
The Journal of Physical Chemistry C


Article

nanocrystal chemistry and structure, we have shown that a clear ACKNOWLEDGMENTS


understanding of the internal nanoparticles atomic arrangement
The authors thank the Brazilian agencies CNPq, FAPESP,
at each step of the chemical synthesis is essential to account for
FAPDF, and FINATEC and are greatly indebted to LNLS for
their magnetic properties. Although the heterogeneity of the
beam-time obtained on D12A-XRD1 beamline (X-ray
chemical composition modifies the problem, it has been
possible to study the nonequilibrium site occupancy in the zinc diffraction) and on D04B-XAFS1 beamline (X-ray absorption).
ferrite core of the particles, by crossing the analyses of Fe and J. Depeyrot is very grateful for the CAPES grant contract BEX
Zn K-edge X-ray absorption near-edge spectroscopy and 1439/11-1, and we acknowledge the exchange program
extended X-ray absorption fine structure measurements, CAPES-COFECUB no. 714/11 and PICS no. 5939. We also
Rietveld refinement of X-ray diffraction patterns, and magnet- thank M. H. Sousa for technical assistance.
ization curves.
The very good agreement between the qualitative features of
experimental Zn K-edge XANES spectra and their ab initio
■ REFERENCES
(1) Bayat, N.; Nethe, A.; Guldbakke, J. M.; Hesselbach, J.; Naletova,
calculated counterparts clearly shows the large transference of J. A.; Stahlmann, H. D.; Uhlmann, E.; Zimmermann, K. Colloidal
zinc ions from tetrahedral sites to octahedral ones. Moreover, Magnetic Fluids − Basics, Development and Applications of Ferrofluids;
the intensity profile of the Fe K-edge prepeak associated with Springer Press: Heidelberg, 1995; Vol. 763, pp 359−427.
the electronic 1s → 3d quadrupole and 1s → 3d/4p dipole (2) Maier-Hauff, K.; Ulrich, F.; Nestler, D.; Niehoff, H.; Wust, P.;
transition indicates the consequent Fe3+ inversion from B-sites Thiesen, B.; Orawa, H.; Budach, V.; Jordan, A. J. Neurooncol. 2011,
to A-ones. This cation redistribution takes place during the 103, 317−324.
coprecipitation step without altering the long-range structural (3) de Rosales, R. T. M.; Tavaré, R.; Galaria, A.; Varma, G.; Protti, A.;
order of the whole nanocrystals. These results are also Blower, P. J. Bioconjugate Chem. 2011, 22, 455−465.
confirmed by the qualitative analysis of Fourier-transformed (4) Tartaj, P.; Morales, M.; del, P.; Verdaguer, S. V.; Carreño, T. G.;
Serna, C. J. J. Phys. D: Appl. Phys. 2003, 36, R182−R197.
EXAFS data obtained at Zn−K edge. Fourier-transformed (5) Connolly, J.; St. Pierre, T. G.; Rutnakornpituk, M.; Riffle, J. S.
EXAFS data obtained at both edges are fitted using ab initio Eur. Cells Mater. 2002, 3, 106−109.
calculations of theoretical photoelectron backscattering phases (6) Campos, A. F. C.; Tourinho, F. A.; da Silva, G. J.; Lara, M. C. F.
and amplitudes. The hypothesis made on the variations of the L.; Depeyrot. J. Eur. Phys. E 2001, 6, 29−35.
coordination numbers, occurring in the same proportion for all (7) Massart, R. IEEE Trans. Magn. 1981, 17, 1247−1248.
coordination shells, is carefully checked. The good agreement (8) Tourinho., F. A.; Franck, R.; Massart, R. J. Mater. Sci. 1990, 25,
between experimental and calculated spectra allows an accurate 3249−3254.
quantitative analysis. Our results show that the interatomic (9) Sousa, M. H.; Tourinho, F. A.; Depeyrot, J.; da Silva, G. J.; Lara,
distances are very close to those of standard bulk ferrites and M. C. F. L. J. Phys. Chem. B 2001, 105, 1168−1175.
affected neither by the protective surface treatment nor by the (10) Gomes, J. A.; Sousa, M. H.; Tourinho, F. A.; Aquino, R.; da
size reduction. As a consequence of interface effects, the Silva, G. J.; Depeyrot, J.; Dubois, E.; Perzynski, R. J. Phys. Chem. C
average coordination number of surface atoms decreases and 2008, 112, 6220−6227.
(11) Lee, J. H.; Jang, J. T.; Choi, J. S.; Moon, S. H.; Noh, S. H.; Kim,
this reduction is enhanced after the surface treatment. It is also
J. W.; Kim, J. G.; Kim, I. S.; Park, K. I.; Cheon, J. Nat. Nanotechnol.
concluded that the averaged percentage of Zn2+ ions located at 2011, 6, 418−422.
B sites of the zinc ferrite core is around 34%, a value that does (12) Fiorani, D. Surface Effects in Magnetic Nanoparticles; Springer
not change after the surface treatment and is independent of Press: New York, 2005.
the size of the particles. This degree of cationic exchange is also (13) Kodama, R. H.; Berkowitz, A. E.; McNiff, E. J.; Foner, S. Phys.
found by Rietveld refinement of X-ray diffractogram using a Rev. Lett. 1996, 77, 394−397.
one-phase modeling for nontreated particles. Finally, high field (14) Sousa, E. C.; Rechenberg, H. R.; Depeyrot, J.; Gomes, J. A.;
magnetization measurements performed at low temperatures Aquino, R.; Tourinho, F. A.; Dupuis, V.; Perzynski, R. J. Appl. Phys.
are strongly consistent with the presence of Fe3+ at A sites of 2009, 106, 093901.1−093901.7.
the spinel nanostructure cores and the consequent develop- (15) Sato, T.; Haneda, K.; Seki, M.; Iijima, T. Appl. Phys. A: Mater.
ment of superexchange interactions between iron ions at both Sci. Process. 1990, 50, 13−16.
interstitial sites. (16) Anantharaman, M. R.; Jagatheesan, S.; Malini, K. A.; Sindhu, S.;
Narayanasamy, A.; Chinnasamy, C. N.; Jacobs, J. P.; Reijne, S.; Seshan,
In the future, it would be interesting to perform in-field
K.; Smits, R. H. H.; Brongersma, H. H. J. Magn. Magn. Mater. 1998,
Mössbauer experiments which allow resolving the sextet 189, 83−88.
contributions corresponding to Fe3+ spins at A and B sites (17) Chinnasamy, C. N.; Narayanasamy, A.; Ponpandian, N.;
due to their antiparallel ordering. We also plan to realize Chattopadhyay, K.; Guérault, H.; Grenèche, J. M. J. Phys.: Condens.
neutron diffraction in order to get a better contrast between the Matter 2000, 12, 7795−7805.
contributions of both metallic cations. This would probably be (18) Roy, M. K.; Haldar, B.; Verma, H. C. Nanotechnology 2006, 17,
very helpful in order to develop an actual core−shell model as 232−237.
well for the diffraction analysis as for the theoretical EXAFS (19) Ammar, S.; Jouini, N.; Fiévet, F.; Beji, Z.; Smiri, L.; Moliné, P.;
calculations. Danot, M.; Grenèche, J. M. J. Phys.: Condens. Matter 2006, 18, 9055−

■ AUTHOR INFORMATION
Corresponding Author
9069.
(20) Schiessl, W.; Potzel, W.; Karzel, H.; Steiner, M.; Kalvius, G. M.;
Martin, A.; Krause, M. K.; Halevy, I.; Gal, J.; Schäfer, W.; Will, G.;
Hillberg, M.; Wäppling, R. Phys. Rev. B 1996, 53, 9143−9152.
*E-mail depeyrot@fis.unb.br.
(21) Nakashima, S.; Fujita, K.; Tanaka, K.; Hirao, K.; Yamamoto, T.;
Notes Tanaka, I. Phys. Rev. B 2007, 75, 174443.1−174443.5.
The authors declare no competing financial interest. (22) Stewart, S. J.; Figueroa, S. J. A.; RamalloLópez, J. M.; Marchetti,
# ́
Formerly at Laboratório Nacional de Luz Sincrotron - LNLS, S. G.; Bengoa, J. F.; Prado, R. J.; Requejo, F. G. Phys. Rev. B 2007, 75,
Caixa Postal 6192, 13084-971 Campinas (SP), Brazil. 073408.1−073408.4.

24290 dx.doi.org/10.1021/jp3055069 | J. Phys. Chem. C 2012, 116, 24281−24291


The Journal of Physical Chemistry C Article

(23) Chinnasamy, C. N.; Narayanasamy, A.; Ponpandian, N.; (55) Taillades, P.; Villette, C.; Rousset, A.; Kulkarni, G. U.; Kannan,
Chattopadhyay, K.; Shinoda, K.; Jeyadevan, B.; Tohji, K.; Nakatsuka, K, R.; Rao, C. N. R.; Lenglet, M. J. Solid State Chem. 1998, 141, 56−63.
K.; Furubayashi, T.; Nakatani, I. Phys. Rev. B 2001, 63, 184108.1− (56) Delahaye, E.; Escax, V.; El Hassan, N.; Davidson, A.; Aquino, R.;
184108.6. Dupuis, V.; Perzynski, R. J. Phys. Chem. B 2006, 110, 26001−26011.
(24) Makovec, D.; Kodre, A.; Arčon, I.; Drofenik, M. J. Nanopart. Res.
2009, 11, 1145−1148.
(25) Kamiyama, T.; Haneda, K.; Sato, T.; Ikeda, S.; Asano, H. Solid
State Commun. 1992, 81, 563−566.
(26) Goya, G. F.; Rechenberg, H. R.; Chen, M.; Yelon, W. B. J. Appl.
Phys. 2000, 87, 8005−8007.
(27) Hamdeh, H. H.; Ho, J. C.; Oliver, S. A.; Willey, R. J.; Oliveri, G.;
Busca, G. J. Appl. Phys. 1997, 81, 1851−1857.
(28) Gomes, J. A.; Sousa, M. H.; Tourinho, F. A.; Mestnik-Filho, J.;
Itri, R.; Depeyrot, J. J. Magn. Magn. Mater. 2005, 289, 184−187.
(29) Oliver, S. A.; Hamdeh, H. H.; Ho, J. C. Phys. Rev. B 1999, 60,
3400−3405.
(30) Li, F. S.; Wang, L.; Wang, J. B.; Zhou, Q. G.; Zhou, X. Z.;
Kunkel, H. P.; Williams, G. J. Magn. Magn. Mater. 2004, 268, 332−340.
(31) Shim, J. H.; Lee, S.; Park, J. H.; Han, S. J.; Jeong, Y. H.; Cho, Y.
W. Phys. Rev. B 2006, 73, 064404.1−064404.4.
(32) Jeyadevan, B.; Tohji, K.; Nakatsuka, K. J. Appl. Phys. 1994, 76,
6325−6328.
(33) Oliver, S. A.; Harris, V. G.; Hamdeh, H. H.; Ho, J. C. Appl. Phys.
Lett. 2000, 76, 2761−2763.
(34) Ammar, S.; Jouini, N.; Fiévet, F.; Stephan, O.; Marhic, C.;
Richard, M.; Villain, F.; Cartier dit Moulin, C.; Brice, S.; Sainctavit, P.
J. Non-Cryst. Solids 2004, 658, 345−346.
(35) Williamson, G. K.; Hall, W. H. Acta Metall. 1951, 1, 22−31.
(36) Giacovazzo, C.; Monaco, H. L.; Viterbo, D.; Scordari, F.; Gilli,
G.; Zanotti, G.; Catti, M. In Fundamentals of Crystallography;
Giacovazzo, C., Ed.; Oxford University Press: New York, 1995; p 109.
(37) Larson, C.; Von Dreele, R. B. General Structure Analysis
System, Los Alamos National Laboratory (ftp://ftp.lanl.gov/public/
gsas), 2001.
(38) Newville, M. J. Synchrotron Radiat. 2001, 8, 322−324.
(39) Ankudinov, A. L.; Ravel, B.; Rehr, J. J.; Conradson, S. D. Phys.
Rev. B 1998, 58, 7565−7576.
(40) See for instance the FEFFIT Web site: cars1.chigago.
edu:80∼newville.feffit/.
(41) Aquino, R.; Depeyrot, J.; Sousa, M. H.; Tourinho, F. A.; Dubois,
E.; Perzynski, R. Phys. Rev. B 2005, 72, 184435.1−184435.10.
(42) Kameli, P.; Salamati, H.; Aezami, A. J. Appl. Phys . 2006, 100,
053914.1−053914.20.
(43) Bardhan, A.; Ghosh, C. K.; Mitra, M. K.; Das, G. C.; Mukherjee,
S.; Chattopadhyay, K. K. Solid State Sci. 2010, 12, 839−844.
(44) Zhang, F.; Jin, Q.; Chan, S. W. J. Appl. Phys. 2004, 95, 4319−
4326.
(45) Liu, B. H.; Ding, J.; Dong, Z. L.; Boothroyd, C. B.; Yin, J. H.; Yi,
J. B. Phys. Rev. B 2006, 74, 184427.1−184427.10.
(46) Gomes, J. A.; Azevedo, G. M.; Depeyrot, J.; Mestnik-Filho, J.; da
Silva, G. J.; Tourinho, F. A.; Perzynski, R. J. Magn. Magn. Mater. 2011,
323, 1203−1206.
(47) Revel, R.; Bazin, D.; Elkaim, E.; Kihn, Y.; Dexpert, H. J. Phys.
Chem. B 2000, 104, 9828−9835.
(48) Guaiata, F. J.; Beltrán, H.; Cordoncillo, E.; Carda, J. B.;
Escribano, P. J. Eur. Ceram. Soc. 1999, 19, 363−370.
(49) Šepelák, V.; Steinike, U.; Uecker, D. C.; Wißmann, S.; Becker, K.
D. J. Solid State Chem. 1998, 135, 52−58.
(50) Druska, P.; Steinike, U.; Šepelák, V. J. Solid State Chem. 1999,
146, 13−21.
(51) Vaishnava, P. P.; Senaratne, U.; Buc, E. C.; Naik, R.; Naik, V.
M.; Tsoi, G. M.; Wenger, L. E. Phys. Rev. B 2007, 76, 024413.1−
024413.10.
(52) Lucas, I. T.; Durand-Vidal, S.; Dubois, E.; Chevalet, J.; Turq, P.
J. Phys. Chem. C 2007, 111, 18568−18576.
(53) Pascal, P. Les Ferrites, Nouveau Traité de Chimie Minérale, Tome
XVII; Masson et Cia: Paris, 1967; p 665.
(54) Silaban, A.; Harrison, D. P.; Berggren, M. H.; Jha, M. C. Chem.
Eng. Commun. 1991, 107, 55−71.

24291 dx.doi.org/10.1021/jp3055069 | J. Phys. Chem. C 2012, 116, 24281−24291

You might also like