You are on page 1of 8

Eur. Phys. J.

B 64, 211218 (2008)


DOI: 10.1140/epjb/e2008-00294-6

THE EUROPEAN
PHYSICAL JOURNAL B

Eective anisotropy eld variation of magnetite nanoparticles


with size reduction
J.M. Vargas1,a , E. Lima Jr1 , R.D. Zysler1 , J.G.S. Duque2 , E. De Biasi2 , and M. Knobel2
1
2

Centro At
omico Bariloche and Instituto Balseiro, 8400 S. C. de Bariloche, RN, Argentina
Instituto de Fsica Gleb Wataghin, Universidade Estadual de Campinas, Campinas (SP) 13081-970, Brazil
Received 28 February 2008 / Received in nal form 9 June 2008
c EDP Sciences, Societ`
Published online 25 July 2008 
a Italiana di Fisica, Springer-Verlag 2008
Abstract. Size eect on the internal magnetic structure has been investigated on weakly interacting magnetite (Fe3 O4 ) nanoparticles by ferromagnetic resonance experiments at 9.5 GHz as a function of temperature (4300 K). A set of three samples with mean particle size of 2.5 nm, 5.0 nm and 13.0 nm, respectively,
were prepared by chemical route with narrow size distribution ( < 0.27). To minimize the dipolar interaction, the particles were dispersed in a liquid and a solid polymer matrix at 0.6% in mass. By freezing the
liquid suspension with an applied external eld, a textured was obtained. Thus, both random and textured
suspensions were studied and compared. The ferromagnetic resonance experiments in zero-eld-cooled and
eld-cooled conditions were carried out to study the size eect on the eective anisotropy eld. The dc
magnetization measurements clearly show that the internal magnetic structure was strongly aected by
the particle size.
PACS. 75.50.Tt Fine-particle systems; nanocrystalline materials 75.30.Gw Magnetic anisotropy
75.60.Ch Domain walls and domain structure 76.50.+g Ferromagnetic, antiferromagnetic, and ferrimagnetic resonances; spin-wave resonance

1 Introduction

picture of the particle as a perfectly ordered single-domain


nanoparticle [68].

Although many nanostructured systems (i.e., systems


with particle sizes of the order of nanometers) have been
reported as superparamagnetic, in general the experimental data show deviations from the standard superparamagnetic theory. These dierences are usually ascribed
to the intrinsic particle size distribution, anisotropies of
dierent origins, and interactions among magnetic particles. An ideal superparamagnetic system is composed
by magnetic single-domain nanoparticles whose atomic
magnetic moments rigidly align through exchange interaction. All the individual magnetic moments add to form a
magnetic supermoment, usually of the order of thousands
Bohr magnetons (B ). The description of superparamagnetism, following the Langevin formalism, considers a negligible anisotropy for each particle and neglects the eect
of interparticle interactions [13]. More recently, Kodama
et al. [4,5] suggested that the modication of the structural
and electronic properties near to the particle surface would
give rise to site specic surface anisotropy and weakened
exchange coupling, which would lead to surface spin misalignment with respect to the ordered core spins and frustration. Such disorder would propagate from the surface
to the particle core, actually making no longer valid the

Among many publications on magnetic nanoparticles,


there is a considerable number of studies performed by
means of ferromagnetic resonance (FMR) due to the fact
that it yields relevant information regarding the dynamics of the system, and it allows one to discriminate the
dierent internal magnetic contributions. For example
Koksharov et al. [9], studied ferrite nanoparticles embedded in polyethylene matrix by ferromagnetic resonance
technique, where the coexistence of both ferromagnetic
and antiferromagnetic phases were clearly determined.
Berger et al. [10], studied the eect of annealing on the
nucleation and growth of nanoparticles in borate glasses
doped with low concentration of iron oxide, where a
new resonance component was observed when the annealing temperature increased, ascribed to superparamagnetic
nanoparticles of a crystalline iron-containing compound.
From the viewpoint of nite size eect, Gazeau et al. [11],
studied the internal elds in colloidal maghemite nanoparticles with particle size between 510 nm, where two
contributions to the internal elds were determined, one
anisotropic and other isotropic. In the eort to better understand the evolution of the magnetic properties from the
atomic level to a bulk solid, De Biasi et al. [12,13] considered the internal magnetic structure for isolated particles.
In their model the site-to-site spin was followed by Monte

e-mail: vargasjm@ifi.unicamp.br

212

The European Physical Journal B

Carlo simulations considering dierent spin-spin coupling


for the core, shell and between them. Thus, the observed
ferromagnetic resonance quantities (line shape, line width
and resonant eld) were interpreted in terms of the particle parameters (magnetic moment, anisotropy) [1315].
Undoubtedly, a major challenge to better understand
the internal magnetic structure in nanosystems and test
the existing models is to produce good quality samples. In
fact, the limited knowledge of nanoparticle shape, composition gradient, size distribution, or even the existance of
touching nanoparticles can hinder an accurate picture of
the whole system. We have contributed to overcome this
problem by producing nanosized magnetite nanoparticles
of approximately cubic-faceted shape and narrow size distribution by chemical synthesis method. The presence of
organic insulating capping on the surface of the magnetic
particles prevents the formation of agglomerates or chains.
Furthermore, the particles have been dispersed in polymer
and solvent in order to minimize the inter-particle dipolar interactions, allowing us to systematically study the
nite size eect on the internal magnetic structure of such
systems.

The morphological characterization was performed by


transmission electron microscopy (TEM), in a Philips CM
200 UT-TEM (200 kV). Samples for TEM analysis were
prepared by drying a high-concentrated solution of the asmade nanoparticles, on amorphous carbon cooper grids.
The crystalline characterization was performed by X-ray
diraction (XRD), in a Philips W1700 diractometer under Cu-K radiation source. The ferromagnetic resonance
(FMR) measurements were performed using a Bruker ESP
300 spectrometer operating at 9.5 GHz (X-band), modulation eld frequency at 100 kHz and amplitude 10 Oe.
A commercial gas ow cryostat was used which allowed
to achieve temperatures in the range of 1.4300 K. The
cavity itself was kept at room temperature and its quality factor was not changed upon cooling. The dc magnetization curves as functions of temperature M (T ) (4 K
< T < 300 K, under zero-eld-cooling and eld-cooling
conditions, H = 20 Oe) and applied eld M (H) (H 50
kOe) were measured in a MPMS XL7 SQUID magnetometer (Quantum Design).

3 Results and discussions


2 Experimental procedure
The magnetite nanoparticles were prepared by chemical synthesis as described in reference [16]. Briey, in
a reaction vessel a metallic precursor, iron (III) acetylacetonate, in presence of 1,2-hexadecanediol and oleic
acid-oleylamine surfactant were vigorously stirred in
phenyl-ether solvent under inert gas ux. The reaction
was carried out at high temperature (270 C) and the
mean size of the particles was tuned by the molar ratio
between metallic precursor and surfactant (rp:s ), without size-selection process. This method leads to colloidal
magnetite nanoparticles with characteristics close to ideal
nanoparticle systems, i.e. particles with excellent morphological (size and shape) and crystalline homogeneity.
The determination of the spinel oxide phase is a dicult
task due to the oxygen stoichiometry variations and the
small size of the particles. It is well established that small
changes in the oxygen stoichiometry results in variations
in the Verwey transition temperature [17]. Therefore, in
our nanoparticles, we expect a smoothing in the Verwey
transition. In this aspect, Mossbauer Spectroscopy (MS)
is a suitable technique since it probes the local environmental and the oxidation state of the Fe atoms. MS analysis in similar systems showed that the samples usually
present both Fe+2 characteristic of magnetite and a lack
of the expected stoichiometry of the particles [18]. Actually, as we are interesting in the magnetic behaviour
of the spinel structure our results are independent of
the oxygen stoichiometry. The as-made nanoparticles were
coated by surfactant molecules to avoid the agglomeration
and the direct exchange-like interactions among particles.
We have obtained three samples with dierent mean grain
sizes, L, when rp:s was xed at the values: 1:20 (2.5 nm),
1:10 (5 nm) and 1:2 (13 nm), respectively. The samples
were labeled as SI, SII and SIII, respectively.

Figure 1 shows the TEM images (Figs. 1a1c) and diraction patterns (Figs. 1d1f) for the samples SI, SII and SIII,
respectively. As the mean size increases, the crystalline degree rises from amorphous-like to well-crystalline pattern,
with narrower peaks owing to the increment of sizes [19].
From several TEM images, the particle size histogram has
been built by counting more than 300 particles for each
sample. The cubic-faceted shape particles display a rather
narrow Lognormal size distribution, with 0.27. For the
well-crystalline samples SII and SIII, the volume average
crystalline domain size (Lv ) was calculated by the Scherrer
formula [20] applied to the principal (113) peak (see patterns in Figs. 1e1f). Table 1 summarizes the crystalline
domain size and their comparison with the TEM particle
size characterization. Since the diraction analysis gives
a volume average value for the mean crystalline domain
size, the volume-average particle sizes (LT W ) were estimated from the TEM histograms [21]. Therefore, an excellent agreement was obtained between both crystalline
and morphological mean particle size for the samples SII
and SIII. For the dc magnetic measurements, the particles
were dispersed in a polymer (PEI) at 0.6%wt., to minimize
the dipolar interaction. Roughly, the calculated mean distance between particles (), correspond to /L > 8 [22].
Figure 2 shows the temperature variation of the magnetization measurement for the samples SI, SII and SIII under zero-eld-cooling (ZFC) and eld-cooling (FC) conditions, for applied elds of H = 20 Oe. The irreversibility
temperature, Tirr , is dened as the threshold temperature
above which FC and ZFC curves coincide. Therefore, as
the mean size increases, the Tirr rises from 4 K to nearly
room temperature (see values in Tab. 2). Interestingly,
the shape of the ZFC-FC magnetization curves is similar
to the one expected for ideal weakly interacting and randomly oriented particle systems [23]. On the other hand,
the FC magnetization curve of sample SIII suggests the

J.M. Vargas et al.: Eective anisotropy eld variation of magnetite nanoparticles with size reduction

213

Fig. 1. TEM images for the samples SI (a), SII (b) and SIII (c); and the diraction patterns SI (d), SII (e) and SIII (f).
Table 1. Mean size values obtained from TEM and XRD characterization.
sample
SI
SII
SIII

L (nm)
2.5
5.0
13.0

L3 (nm3 )
15.6
125.0
2197.0

Lv (nm)

5.5
12.7

LT W (nm)
2.8
5.5
14.4

appearance of dipolar interactions [24]. The main blocking


temperature, TB , is dened as the maximum value in the
energy barrier distribution occurs. For weakly-interacting
nanoparticles, according to the Neel-Brown model for a
single superparamagnetic particle, the blocking temperature distribution function (f (T )) can be obtained from
the ZFC and FC magnetization curves as, d[MZF C (T )
MF C (T )]/dT T f (T ) [25]. Then, the K value of each
sample was estimated by using the TB value and the Arrhenius law. As is commented below, although for samples
SI and SII the Arrhenius law is not completely satised
due to the complex internal magnetic structure, it was
possible to estimate a rough value for K for each sample. Figure 3 shows the magnetization loops for the three
samples at room temperature and T = 4 K. At room temperature (Figs. 3a3c) the three samples are in the thermal equilibrium i.e., loops without hysteretic or memory
eects, showing reversible magnetization. Moreover, from
the slope of the magnetization curves it was possible to observe that the curve is more pronounced as the particle size
increases, in good agreement with the Langevin formalism
[13]. At T = 4 K (Figs. 3d3f), as the particle size increases the hysteretic loops show striking variation: from
wasp-waisted-like shape for the sample SI to square-like
shape for the sample SIII. Comparing the three samples,
the maximum value of HC was achieved for the sample
SII, probably due to the strong interaction between the
core and shell magnetic regions. Moreover, in agreement
with the theoretical predictions, the sample SI shows the
biggest Hirr 10 kOe (where Hirr correspond to the higheld value where both ascending and descending magnetization curves splitting), associated to the strong surface

Fig. 2. Temperature variation of the dc magnetization measured under ZFC and FC conditions for the samples SI (a),
SII (b) and SIII (c), with H = 20 Oe. The open circle correspond to the ferromagnetic resonance susceptibility (IF M R ),
measured in the ZFC condition.
Table 2. Values and parameters from the dc magnetic characterization.
ZFC-FC
hysteresis loops
sample Tirr (K) TB (K) K (erg/cm3 ) HC (Oe) Hirr (Oe)
SI
SII
SIII

6
50
272

<2
19
86

<5.0 105
5.8 105
1.5 105

23
310
274

10 000
5200
2400

anisotropy contribution, which leads to magnetic frustration in surface at low temperature [15]. Table 2 summarize
the dc magnetic results.

214

The European Physical Journal B

Fig. 3. Magnetization loops at room temperature for the samples SI (a), SII (b) and SIII (c); and T = 4 K for the samples SI
(d), SII (e) and SIII (f).

Fig. 4. Temperature variation of the magnetic resonance spectra at (i) 300 K, (ii) 250 K, (iii) 110 K, (iv) 50 K, (v) 4 K in the
ZFC condition for the samples SI (a), SII (b) and SIII (c). The dotted lines correspond to the computer simulated spectra.

In the FMR experiments the particles were dispersed


in toluene (with freezing temperature 180 K), where the
volume ratio between the as made colloidal solution and
solvent was 1:20. Approximately, for this dilution, the
mean distance between nanoparticles was /L > 8 [22].
Figure 4 shows the temperature variation of the FMR
spectra in the ZFC condition for the samples SI, SII and
SIII. From the absorption-derivative spectra, two characteristic parameters were dened: the resonant eld H0 as
the point where the spectrum equals to zero; and the peakto-peak line width Hpp . Therefore, the following features
were observed in the FMR-ZFC experiments:
The spectrum of the sample SI (Fig. 4a) shows a narrow symmetric line, with a Lorentzian-like shape, cen-

tered at H0 = 3379 Oe and Hpp 110 Oe. As


the temperature decreases the H0 does not change,
although a small continuous broadened eect is observed. The integrated FMR absorption intensity,
IF MR , was calculated through the double integration
of the absorption-derivative spectra at each temperature. Previous normalization by the dc magnetic susceptibility at room temperature, the IF MR does not
show appreciable temperature variation. This feature
could be explained by a strong surface anisotropy contribution, with spin-glass-like characteristics.
The spectrum of the sample SII (Fig. 4b) consists in an
asymmetric line that can be described as a sum of two
lines: a broad line with the Hpp 380 Oe, associated
with the particle-core contribution, and a narrow one

J.M. Vargas et al.: Eective anisotropy eld variation of magnetite nanoparticles with size reduction

215

Fig. 5. The H and H resonance elds obtained from the simulated ZFC curves at temperatures 4 K < T < 110 K for the
samples SI (a), SII (b) and SIII (c). These values are compared with the respective ones measured in the FMR-FC spectra. The
temperature variation of the center eld, H0 (T ), is plotted for comparison (values obtained from the FMR-ZFC spectra).

with Hpp 50 Oe, associated with the particle-shell


contribution [9,10]. Cooling the sample, the narrow
line decreases steeply, and the broad line shifts towards
lower elds and its width increases. The temperature
variation of IF MR for the broad line was calculated
(previously discounting the narrow line, i.e., the shell
magnetic contribution) and it is plotted in Figure 2b.
At T > 150 K, a good agreement is observed between
both dc and IF MR susceptibilities. Interestingly, this
agreement was predicted for ideal paramagnetic systems, where IF MR is proportional to the paramagnetic susceptibility [26], i.e., the temperature variation
of the broad line follows the superparamagnetic-core
characteristics. At lower temperatures IF MR (T ) increases monotonically, although it is not enough to
follow the dc susceptibility.
The spectrum of sample SIII (Fig. 4c) shows a broad
and symmetric line, Hpp 420 Oe, without the narrow line component. As the temperature decreases, the
line shifts toward lower elds and its width increases.
Moreover, IF MR (T ) slows down as the temperature
increases (see Fig. 2c), as it would be expected in the
nanoparticle FMR model [13].
Therefore, as the particle size increases, the sharp line feature in the FMR-ZFC spectra was gradually replaced by
the broad one, and the two line-component coexist at intermediate sizes. Actually, similar trends were observed in
computer-generated spectra by Berger et al. [30], for ideal
nanoparticles with size and shape distribution. The FMRZFC spectra were simulated at temperatures lower than
110 K, where the ferrouid was frozen, with randomly oriented easy magnetization axis. The theoretical absorption
expression for a randomly oriented anisotropy axis system
can be written as [28]:

A(H)

H
H

K(H  )F (H  )dH  ,

(1)

where F (H  ) = 1/((H H  )2 + H02 ) is the Lorentzian


function and K = (H2 + H 2 )/H2 is the weight probably factor. The H , H and Hpp are the parameters
of the tting. In particular, in the linear model [13], the
correlation between both resonant elds and the eective
anisotropy eld HA can be written as:
3
HA = H H .
(2)
2
Figure 5 shows the temperature variation of the calculated
parameters H and H . Whereas the mean particle size
increases, the dierence between both resonance elds increases. Interestingly, the calculated parameter Hpp for
the three samples not show an appreciable temperature
variation between 4300 K, where for the sample SI the
Hpp 300 Oe, and for the samples SII and SIII the
Hpp 800 Oe.
The FMR spectra measured in the FC condition
(FMR-FC), i.e., after the ferrouid was frozen in an applied eld (H = 10 kOe) and the easy magnetization
axes became more aligned along the external eld direction, provide fundamental information about the dierent anisotropy contributions. From the FMR-FC spectra,
the measured parallel and perpendicular resonance elds
were compared with the calculated ones parameters (see
Fig. 5). Whereas for the samples SI and SIII an excellent agreement is observed, the disagreement observed for
the sample SII at low temperatures can be associated to a
core-shell magnetic structure (60/40%), which in this case
has not been considered. From the angular variation of the
FMR-FC spectra it is evident that as the mean particle
size increases, the amplitude of the angular variation is
more accentuated (Fig. 6). Moreover, the angular variation is well tted with the cos2 dependence, expected for
uniaxial anisotropy systems [11].
As is mentioned above, the shape spectrum in the
FMR-FC condition is very sensitive to the coexistence of
the dierent anisotropy contributions. Figure 7 shows the
temperature variation (4 K < T < 110 K) of the FMR

216

The European Physical Journal B

Fig. 6. The angular variation of the center eld H0 at 4 K, 50 K and 110 K for the samples SI (a), SII (b) and SIII (c), after
freezing the uid with an applied eld (H = 10 kOe).

Fig. 7. Temperature variation of the magnetic resonance spectra measured in the FC condition at (i) 4 K, (ii) 10 K, (iii) 20 K,
(iv) 30 K, (v) 50 K, (vi) 110 K, for the sample SI. The spectra
do not show angular variation.

spectra measured in the FC condition for the sample SI.


At low temperatures (T < 27 K), a very broad line is
observed. For temperatures higher than 27 K, the broad
line is steeply narrowed, centered at H0 = 3379 Oe. Also,
the IF MR calculated from the FMR-FC spectra reaches
a maximum at Tf z = 27 K, where Tf z is dened as the
freezing temperature of surface contribution. Remarkably,
the dc magnetization vs temperature measured after cooling with H = 50 kOe (M50 kOe (T ) (gure not shown) is
correlated to magnetic variation observed in the FMR-FC
spectrum at low-temperatures. The M50 kOe (T ) curve
rises at T < 27 K, owing to the formation of a clusterglass-like phase at low temperatures [32]. Concomitantly,
the shape of the FMR-FC spectra becomes more asymmetric as the mean particle size increases (see Figs. 7
and 8), due to the relative increase of the crystalline
anisotropy contribution and the vanishing of the surface

anisotropy contribution. Generally speaking, for an ideal


crystal with cubic symmetry two maxima are expected
in the angular variation, fact that does not occur in the
studied samples [11]. Therefore, the shape anisotropy is
the most important factor to the eective anisotropy in
the present case. To check whether the asymmetric line
shape of the FMR-FC experiments is related to the coexistence of the shape and crystalline magnetic contributions,
we simulated the spectra with both contributions, i.e., the
shape uniaxial anisotropy and the magnetocrystalline with
cubic-distortion symmetry [34]. The inset of Figure 8b,
shows the good agreement between the experimental and
simulated shape spectra. The eective anisotropy eld
HA (T ) was calculated from equation (2) and it is shown
in Figure 9 for the three studied samples. In particular, for
the smaller particle size (2.5 nm), without angular variation, HA H H = 0. In this case, the dierence observed in the FMR experiments between the ZFC and FC
conditions evidences the shell magnetic contribution with
strong surface anisotropy eld. Thus, at low temperatures,
in the FMR-ZFC experiments the disordered spin conguration leads to a small amounts of accessible states for the
resonance condition, with narrow anisotropy eld distribution. On the other hand, in the FMR-FC experiments,
all spins are quasi-aligned and in the resonance condition,
with a subsequent broad anisotropy eld distribution. In
the case of intermediary (5.0 nm) and bigger (13.0 nm)
particles, HA = 800 Oe at T = 4 K, and both follow similar trends. However, for the sample SII (particles with
intermediate sizes) the variation is more accentuated due
to the surface contribution. Similar values were reported
by Barker et al. [33], for the case of uniaxial anisotropy.

4 Conclusions
In summary, the present study shows that the FMR technique in association with the dc magnetic measurements
yields an eective tool for exploring the shape, crystalline and surface magnetic contributions of the magnetic
nanoparticles. In particular, such combination sheds some
light into the eect of size on the magnetic structure of the

J.M. Vargas et al.: Eective anisotropy eld variation of magnetite nanoparticles with size reduction

217

Fig. 8. The parallel and perpendicular ferromagnetic resonance lines at 4 K, 50 K and 110 K for the samples SII (a) and SIII
(b). The continuous and dotted lines correspond to the parallel and perpendicular measurements, respectively. The inset in (b)
correspond to the computer simulated parallel and perpendicular curves.

cle nearly corresponds to a ferromagnetic single-domain


state. In this case the eective HA = 800 Oe at 4 K and
slowly down as the temperature increase. In the sample
with intermediate grain sizes (SII), the coexistence of both
surface and volume contribution was clearly observed and
the HA slowly down as the temperature increase (where
HA = 800 Oe at 4 K). However, compared with the sample SIII, the temperature variation is more accentuated
due to the surface contribution.
Fig. 9. The temperature variation of the eective anisotropy
eld, HA , for the samples SI (a), SII (b) and SIII (c).

nanoparticle. In the present study, a set of three samples


of magnetite nanoparticles were systematically prepared
and investigated. The mean particle size was tuned in the
synthesis process and samples with mean sizes of 2.5 nm,
5.0 nm and 13.0 nm were obtained with corresponding surface to volume ratios of 90%, 40% and 10%, respectively.
The degree of spin uctuations was expected to decrease with increasing the particle size [11,12,32]. This
eect was conrmed by FMR and dc magnetic measurements in the present study. For the sample with smaller
nanoparticles (SI) the magnetic properties are absolutely
dominated by surface magnetism, where the surface spin
is strongly disordered. It was shown that the surface spin
uctuation decreases as the temperature is diminished,
leading to a frozen disordered state of the surface spins at
low temperatures, which the eective HA = 0. In the other
extreme, for the sample with bigger particles (SIII), the
magnetic properties are completely dominated by the volume contribution, following the Stoner-Wohlfarth model,
where the internal magnetic structure of the nanoparti-

References
1. E.C. Stoner, E.P. Wohlfarth, Philos. Trans. R. Soc.
London Ser. A 240, 599 (1948)
2. L. Neel, Ann. Geophys. 5, 99 (1949)
3. X. Battle, A. Labarta, J. Phys. D: Appl. Phys. 35, R15
(2002)
4. R.H. Kodama, A.E. Berkovitz, E.J. Mc Ni, S. Foner Jr,
Phys. Ref. Lett. 77, 394 (1996)
5. R.H. Kodama, S.A. Makhlouf, A.E. Berkovitz, Phys. Rev.
Lett. 79, 1393 (1997)
6. R.H. Kodama, J. Magn. Magn. Mater. 200, 359 (1999)
7. O. Iglesias, A. Labarta, Phys. Rev. B 63, 184416 (2001)
8. D. Fiorani, A.M. Testa, F. Lucari, F. DOrazio, H.
Romero, Phys. B 320, 122 (2002)
9. Y.A. Koksharov, D.A. Pankratov, S.P. Gubin, I.D.
Kosobudsky, M. Beltran, Y. Khodorkovsky, A.M. Tishin,
J. Appl. Phys. 89, 2293 (2001)
10. R. Berger, J. Kliava, J. Bissey, V. Baeto, J. Appl. Phys.
87, 7389 (2000)
11. F. Gazeau, J.C. Bacri, F. Gendron, R. Perzynski, Y.L.
Raikher, V.I. Stepanov, E. Dubois, J. Magn. Magn. Mater.
186, 175 (1998)
12. E. De Biasi, C.A. Ramos, R.D. Zysler, H. Romero, Phys.
Rev. B 65, 144416 (2002)

218

The European Physical Journal B

13. E. De Biasi, C.A. Ramos, R.D. Zysler, J. Magn. Magn.


Mater. 262, 235 (2003)
14. E. De Biasi, C.A. Ramos, R.D. Zysler, H. Romero, Phys.
B 354, 286 (2004)
15. E. De Biasi, C.A. Ramos, R.D. Zysler, D. Fiorani, Phys.
B 372, 245 (2006)
16. J.M. Vargas, R.D. Zysler, Nanotech. 16, 1474 (2005)
17. R. Arag
on, D.J. Buttrey, J.P. Shepherd, J.M. Honig, Phys.
Rev. B 31, 430 (1985)
18. A.D. Arelaro, E. Lima, L.M. Rossi, P.K. Kiyohara, H.R.
Rechenberg, J. Magn. Magn. Mater. 320, e335 (2008)
19. D. Zanchet, B.D. Hall, D. Ugarte, Characterization of
nanophase materials, edited by Z.L. Wang (VCH-Wiley,
Weinheim, 1999)
20. R.M. Cornell, U. Schwertmann, The Iron Oxides:
Structure, Properties, Reactions, Occurrence and uses
(VCH, New York, 1996)
21. J.M. Vargas, L.M. Socolovsky, M. Knobel, D. Zanchet,
R.D. Zysler, J. Appl. Phys. 101, 023903 (2007)
22. It is worth mentioning that in polymer and solvent, the
degree of the particle dispertion was cheked previously
by Small Angle X-ray Scattering (SAXS). For this sample preparation the SAXS data conrms the homogeneity
in the particle-dispersion, without agglomerates or particle
chains.
23. J.M. Vargas, W.C. Nunes, L.M. Socolovsky, M. Knobel,
D. Zanchet, Phys. Rev. B 72, 184428 (2005)

24. M. Hanson, C. Johansson, M.S. Pedersen, S. Mrup, J.


Phys. Cond. Matter 7, 9269 (1995)
25. J.M. Vargas, J. G
omez, R.D. Zysler, A. Butera, Nanotech.
18, 115714 (2007)
26. G.E. Pake, Paramagnetic Resonance: An Introductory
Monograph (W.A. Benjamin Inc., New York, 1962)
27. R.S. de Biasi, T.C. Devezas, J. Appl. Phys. 49, 2466 (1978)
28. C.P. Poole Jr, Electron Spin Resonance: A comprehensive
Treatise on Experimental Techniques (Dover Publications
Inc. New York, 1983)
29. The surfactants and metallic precursor do not show magnetic resonance signal. We also have performed magnetic
resonance measurements on the surfactants after the same
treatment as them have in the synthesis route and we have
registred a very small signal (comparable with the noise
amplitude) probably due to free radicals or broken bonds
created in the process
30. J. Kliava, R. Berger, J. Magn. Magn. Mater. 205, 328
(1999)
31. R. Berger, J. Bissey, J. Kliava, H. Daubric, C. Estournes,
J. Magn. Magn. Mater 234, 535 (2001)
32. R.H. Kodama, A.E. Berkowitz, Phys. Rev. B 59, 6321
(1999)
33. A.J. Barker, B. Cage, S. Russek, C.R. Stold, J. Appl. Phys.
98, 063528 (2005)
34. C. Ramos, E. De Biasi, R. Zysler, E. Vassallo, M. V
azquez,
J. Magn. Magn. Mater. 316, E6 (2007)

You might also like