Consequences of Magnetic
Consequences of Magnetic
1.1 Introduction
Magnetic granular materials are composites in which a matrix is filled with magnetic
particles that have sizes in the nanometer range. Examples of such materials are
those prepared by sputtering, where two or more elements that do not easily form
compounds are deposited onto a substrate; an incomplete multilayer deposition; a
heterogeneous alloy; nanoparticles incorporated into a matrix by physical methods;
or even agglomerations of nanoparticles covered with another substance. Figure 1.1
shows two such granular materials. For several decades nanostructures have been
produced by physical and chemical methods. Enhanced techniques for the manip-
ulation of matter at the nanoscale have enabled the preparation of better controlled
nanostructures. Such magnetic granular materials have interesting physical and
chemical properties that are also of technological interest. Some of those physical
properties are related to magnetism, like superparamagnetism, or, in connection with
magnetotransport, giant magnetoresistance (GMR), tunneling magnetoresistance
(TMR), and the giant Hall effect (GHE). Some of these are currently being exploited
in computer hard disks, perpendicular recording of magnetic bits, and in biomedical
applications such as drug delivery and magnetohyperthermia, among others. The
read heads of modern hard disks, which use the GMR effect, represent the first
Fig. 1.1 Above: A granular material can be formed by the dispersion of nanoparticles in a matrix
(left) or by the clustering of grafted nanoparticles (right). Below: (left) cobalt nanoparticles in a
silica matrix; (right) magnetite nanoparticles grafted with silica (Image courtesy Dr. D. Muraca)
1.1.1 Superparamagnetism
.AK/1=2
rc 9 ; (1.1)
0 Ms2
4 L.M. Socolovsky and O. Moscoso Londoño
1.1.2 Anisotropies
In real systems, nanoparticles are not very far from each other, so in principle, the
magnetism of one nanoparticle should affect others. Principally, two long-range
interactions, known as dipolar and RKKY, must be considered in granular materials.
A dipolar interaction is described by
ED
0
!
m ! 3
m ! !
m r C r ;
! !
m (1.4)
1 2 1 2
4r3 r2
0 m2
Ed : (1.5)
4 r3
In this formulation, m and r are assumed to be the averages of the magnetic
moment and the nanoparticle separation, respectively [14].
In a granular material, another important type of long-range interaction is
the well-known RKKY interaction. Such interactions, formulated by Ruderman,
Kittel, Kasuya, and Yoshida in the mid-1950s, appears in materials with itinerant
electrons. In the particular case of nanoparticles supported in a metallic matrix, the
RKKY interaction does not involve a direct coupling between magnetic nanoparticle
moments but is mediated by itinerant electrons, which are subject to a localized
magnetic moment that polarizes them [8], that is, the mediated electrons are affected
by a spin-dependent local potential. This interaction is present in granular materials
where the matrix is metallic. RKKY interactions are oscillatory and fall with the
1 Consequences of Magnetic Interaction Phenomena in Granular Systems 7
cubic power of distance. At large distances the term that regulates the strength of
this interaction, the exchange parameter J, is
cos .2kF r C '/
J D J0 ; (1.6)
.2kF r/3
!
! !
HDM D dij Si Sj : (1.7)
When nanoparticles are very close to each other, or even in contact, a dipolar
interaction is not the predominant one. Direct interaction between magnetic ions
located on different nanoparticles can be common, so a new magnetic state emerges.
This state is usually called superspin glass (SSG) [5]. This name might no seem
appropriate at first sight because the origin of the new state does not lie in the
competing RKKY interaction, as is the case for spin glasses [17]. When percolation
is established, then we have a state called superferromagnetism (SFM) [5, 18]. In
the authors’ opinion, more appropriate names could be used, such as interacting
superparamagnetism (I-SPM) instead of SSG or percolated superparamagnetism
(P-SPM) instead of SFM. A qualitative magnetic phase diagram is shown in Fig. 1.3.
We can think of these magnetic behaviors in granular systems as balls of the same
size that fill a large glass box (Fig. 1.4). Let us imagine black balls as magnetic
nanoparticles and the white matter that surrounds them as a nonmagnetic matrix.
We begin by considering the box filled exclusively by white matter: our system
is nonmagnetic. If we randomly add more balls, which may be far apart, we will
have a canonical superparamagnetic system. With the addition of more balls, the
system more closely resembles a real superparamagnet. We can easily see that as
our magnetic nanoparticle concentration increases (with the addition of more balls),
some of the balls will come into close contact with each other, sometimes with more
than two balls, and we see SSG or I-SPM states. If we continue to fill the receptacle,
we will see more and more chains of balls, until we see a pathway spanning the
whole box. This point is called the percolation limit, and its magnetic manifestation
is the SFM or P-SPM state. When this percolated state is reached, there will still be
isolated balls, so we must expect a complex magnetic behavior that will show both
SPM and ferromagnetic features. By adding more balls, we finally obtain a fully
ferromagnetic state.
Percolation is a geometrical concept [19]. The situation presented here just shows
contacting particles. Because we are talking about magnetism, which is force that
8 L.M. Socolovsky and O. Moscoso Londoño
Fig. 1.3 Changing nanoparticle concentration, in three phases. Less-concentrated granular mate-
rials behave as predicted by theory. The intermediate range (I-SPM or SSG) can be described,
at lower concentrations, by Dormann–Bessais–Fiorani, Mørup, or T* models [4]. At higher
concentrations, only a T* model can describe magnetic behavior. Close to the percolation limit
all models fail. At that percolation limit, a long-range magnetic order, based on direct interchange
between atoms, starts to become established until a full ferromagnetic-like (e.g., FM or AFM)
order sets in
Fig. 1.4 Black balls represent magnetic nanoparticles. As the receptacle is filled with magnetic
nanoparticles, the system evolves from a noninteracting system to a fully percolated one. Pictures
(a–c) represents granular materials above percolation threshold, and (d–f) represent percolated
systems
1 Consequences of Magnetic Interaction Phenomena in Granular Systems 9
Note that in the preceding function, MZFC (T) is dependent on the magnetization
saturation MS , applied magnetic field H, effective anisotropy Kef , the measurement
time m , and the attempt time 0 , which is assumed to be on the order of
1 Consequences of Magnetic Interaction Phenomena in Granular Systems 11
109 –1010 s, as mentioned earlier. Also, such a dependence was weighted using
a blocking temperature distribution, which of course is directly related to the size
distribution.
All real systems contain relatively larger and smaller particles, regardless of the
method used to fabricate them. This distribution of sizes usually follows a lognormal
or Gaussian function. This means that the exponential term in Eq. 1.3 is different,
not only because of the volume but because of the anisotropy. Most nanoparticle
systems have a lognormal size distribution. Because they are nonsymmetric, the
mean value is different from the median and mode values. Knobel et al. present an
interesting discussion on this point [4].
On the other hand, a large number of magnetic nanostructured systems have
been reported as superparamagnetic and studied in the framework of this theory.
Nevertheless, a lack of agreement between experimental data and standard super-
paramagnetic theory, as well as the obtained spurious parameters, is commonly
observed. Anisotropy constant K is generally calculated using the approximate
equation
which derives from Eq. 1.3, in which are applied logarithms evaluated at the
blocking temperature TB . It is considered that a characteristic measurement time
for DC magnetometry lies around 100 s [4]. The first challenge, which involves
the use of Eq. 1.9, is to correctly determine the blocking temperature, for which
there is no established protocol, meaning some research groups use the maximum
of the ZFC curve, whereas others, knowing that the sample has distributed properties
(e.g., distribution of sizes, blocking temperatures, anisotropies), first determine the
blocking temperature distribution, generally through
d
f .TB / D .MZFC MFC / ; (1.10)
dT
from which the mean blocking temperature hTB i can be extracted [26]. According
to the Ref. [27], the derivative method is the best for obtaining the blocking
temperature distribution [25].
We depict this process to illustrate the following remarks. Figure 1.6 shows
six different types of ZFC curves commonly seen in real granular materials.
Each curve corresponds to the physical situation described in Fig. 1.4. Here,
a, b, and c correspond to nonpercolated systems. The first one, a, is typical of a
noninteracting system. Characteristic features we can observe in such curves include
(1) a concavity at low temperatures, (2) a rounded peak, and, at T Tpeak , (3)
a decrease in the magnetization with temperature, approximately following a 1/T
dependence.
For other interacting systems, the peak shifts to higher temperatures and the
high-temperature side of the curve departs from a Curie-type law at temperatures
close to that of the maximum, as seen in b. The third one, c, corresponds to a
12 L.M. Socolovsky and O. Moscoso Londoño
Fig. 1.6 The maximum of a ZFC is called the peak temperature, Tpeak . The irreversibility
temperature, Tirr , is also marked. (a) Magnetite nanoparticles dispersed in paraffin at very low
concentration. (b) Magnetite nanoparticles grafted with SiO2 ; dipolar interaction plays a role. (c)
Magnetite nanoparticles grafted with acetic acid, compressed in pellets. (d) Fe–Cu alloy, with
iron concentration slightly above the percolation limit. (e) Fe–Cu alloy, percolated; its blocking
temperature is above 300 K. FC was not recorded. (f) Fully percolated Fe–Cu alloy. A TB above
550 K is evident. The FC curve seems to show that a structural change could happen at those
high temperatures. Each panel can be viewed as corresponding to the situations in Fig. 1.4.
Measurements were done in a field of 20 Oe
N2 H M 2 VH H
M .H; T/ D D s DC ; (1.12)
3kB T 3kB T T
14 L.M. Socolovsky and O. Moscoso Londoño
M C
DC D D : (1.13)
H T
We can use Eqs. 1.12 and 1.13 to calculate the mean magnetic moment of each
particle, if we have their volume, by plotting the magnetization vs. the inverse
temperature. Alternatively, we can plot 1/M or 1/ vs. T to obtain a more viewable
graph (Fig. 1.7). In this case, the slope of the high-temperature straight part is
1 3kB
D 2 : (1.14)
C Ms V
M C
DC D D ; (1.15)
H T
Fig. 1.7 Plot of inverse susceptibility vs. temperature of iron nanoparticles dispersed in paraffin;
the high-temperature part can be fitted to a Curie-type law
1 Consequences of Magnetic Interaction Phenomena in Granular Systems 15
1.2.1.3 Susceptometry
0 D cos ';
(1.17)
00 D sin ':
1
A more advanced version of this technique is used in modern SQUID devices.
16 L.M. Socolovsky and O. Moscoso Londoño
Fig. 1.8 In-phase 0 (dots) and out-of-phase 00 (triangles) AC susceptibilities of a lightly
interacting system composed of packed silica-covered Fe3 O4 nanoparticles (upper left), a strongly
interacting system composed of Fe14 Cu86 (upper right), an almost percolating interacting system
composed of Fe17 Cu83 (lower left), and a percolated system composed of Fe30 Au70 (lower right).
High-frequency measurements are indicated by open symbols, low-frequency by black symbols
shapes of the curves can also be used (Fig. 1.8). The 00 curve shows an interesting
feature. Because this phase is representative of the energy dissipation in the system,
it gives an additional view on it. The maximum in 0 or 00 can be used as the
blocking temperature, but the blocking temperature is best represented by the peak
of 00 [33]. This peak of the 00 curve usually occurs at lower temperatures than the
peak of the 0 measurement. In a normal susceptometer, 00 is small compared with
0 . In adapted SQUID devices, this 00 signal is bigger. Based on the relationships
of both susceptibilities Cole–Cole plots, information about the frequency attempt 0
and energy barrier distribution can be obtained. A useful article on the underlying
physics is that of García Palacios [34]. A measurement of 00 vs. temperature in
strongly interacting systems, even percolated ones, displays a peak that more or less
still corresponds to the effect of individual nanoparticles.
1 Consequences of Magnetic Interaction Phenomena in Granular Systems 17
Using Eq. 1.3 we show how energy barriers affect relaxation times. This equation is
known as the Néel–Arrhenius (N–A) law [35]. The relaxation time for a system of
noninteracting superparamagnetic particles follows that law, which we repeat here
to compare with the Vogel–Fulcher (V–F) law in this section:
Kef V
D 0 exp : (1.18)
kB T
Despite the fact that the V–F model does not take into account the surface effects
or polydisperse properties present in all real nanoparticle systems, the use of this
model provides a more physically relevant result. Table 1.1 presents the values of 0
obtained for some systems composed of magnetic nanoparticles by means of N–A
and V–F laws.
As can be seen in Table 1.1 and in most cases where nanoparticle interactions
are relevant, the response time takes nonphysical values when the N–A law is
used. Despite the fact that these values are physically incorrect, it is interesting
to observe that in interacting or weakly interacting nanoparticle systems, there is
a clear tendency to underestimate these parameters. However, such values can be
corrected by adopting the Vogel–Fulcher expression from which values ranging
between 109 and 1011 s are obtained.
18 L.M. Socolovsky and O. Moscoso Londoño
Table 1.1 Values of 0 obtained by applying both Néel–Arrhenius and Vogel–Fulcher models
on granular materials composed of monodomain magnetic nanoparticles
Sample Mean NP size (nm) 0 using N–A law (s) 0 using V–F law (s) Reference
”-F2 O3 4.9 – 2.5 1011 [36]
Fe3 O4 3.8 4.9 1036 4.7 1010 [37]
4.5 1.6 1016 3.2 1010
Fe3 O4 5.5 1032 1013 [38]
”-F2 O3 (powder) 4 3 1018 5.2 107 [39]
”-F2 O3 (compacted) 1.6 1042 9.7 108
Ni (1.9 wt%)-SiO2 4.3 3.4 1010 1.7 109 [40]
Ni (2.7 wt%)-SiO2 5 2.9 1011 2.1 1010
Ni (4 wt%)-SiO2 4.9 4.2 1011 1.4 1010
Ni (7.9 wt%)-SiO2 5.3 1.3 1012 1.2 1010
Ni (12.8 wt%)-SiO2 5.5 5.8 1015 6.3 1010
Fe3 O4 4.9 – 6 106 [41]
CoO-Pt core-shell 4 6 1019 2.4 1011 [42]
Fe3 O4 7.5 9.5 1010 9.8 1010 [43]
Au-Fe3 O4 dimer 11 1.8 1016 1.4 109
Au-Fe3 O4 core-shell 15 5.9 1014 8.9 109
CoFe2 O4 4.5 6.8 1027 – [44]
6.3 1.5 1032 –
M D M0 C 1 H C 2 H 2 C 3 H 3 : (1.20)
In this kind of study, higher-order terms of susceptibility are analyzed. The peak
in 3 corresponds to a blocking temperature or spin-glass transition temperature
[12, 32].
Another alternative that allows for a different way to study magnetization in a
sample is provided by Mössbauer spectroscopy, which is based on the resonant
absorption on certain isotopes. The effect is quite remarkable on iron, so for many
magnetic materials it is a very useful technique. One feature that a Mössbauer
experiment can measure is the hyperfine magnetic field, Bhf , which is a measure
of the internal fields sensed at the nuclear level. Bhf is seen as a sextet of peaks in
an absorption vs. energy graph. It has been observed in granular systems that such
a sextet collapses to a singlet or doublet when the temperature is raised. Because
the dominant frequency corresponds to a Larmor precession, which in this case
corresponds to that of the nuclear magnetic moment with the surrounding magnetic
field, we can make a distinct measure of a sample’s magnetization. By measuring
1 Consequences of Magnetic Interaction Phenomena in Granular Systems 19
There are commonly two schools of thought when it comes to analyzing magnetic
interactions in systems of nanoparticles. One of those schools is based on the
Dormann–Bessais–Fiorani [54] model, which uses energy barriers to describe the
effect on a nanoparticle of a full ensemble; the other school uses the Mørup
20 L.M. Socolovsky and O. Moscoso Londoño
Fig. 1.9 M vs. H curves measured at 5 and 350 K of a sample composed of iron nanoparticles in
a SiO2 matrix. Both measurements (which are S-shaped) were taken at temperatures far from the
blocking temperature (TB 35 K). The 5 K curve (up black triangle), which is in the blocked
state, shows an important coercive field, while at 350 K (down red triangle), which is in the
superparamagnetic state, it shows little coercivity. Saturation magnetization is lower in the high-
temperature measurement, as it should be
model [55], in which a mean field approximation is proposed. Both models are
based on approximations commonly used in physics and are discussed in detail
in Ref. [56]. The first one seems to represent experimental results better than the
second. But in more interacting systems, a third approach, developed by Allia et al.,
the so-called T* model, has shown better depictions of real systems [23, 57].
This model, sometimes called the T-star model or interacting superparamagnetic
model (ISP), was proposed as a phenomenological approach that treats nanoparticle
magnetic moments as interacting through dipolar-type long-range fields, and the
overall effect can be modeled by a fictitious temperature, T*, which should be added
to the real temperature in the denominator of the Langevin function argument. The
new temperature, called the apparent temperature, has the effect of slowing the
approach to saturation and can be written TA D T C T*, where T* is related to
the dipolar interaction energy through "D D kB T*. Then the modified Langevin
function has the form
Z 1
CO H
M .H; T/ D nCO CO L f .CO / dCO ; (1.21)
0 kB .T C T /
1 Consequences of Magnetic Interaction Phenomena in Granular Systems 21
where the suffix CO refers to corrected values. To solve the modified Langevin
equation, it is necessary first to determine T*, which can be accomplished by starting
with the low-field inverse susceptibility, given by
T
D 3kB NCO C ˇ; (1.22)
.ıM s /2
where AP is the magnetic moment magnitude obtained from the standard Langevin
equation, called apparent because its values are screened by the nanoparticle
interaction effects. The susceptibility is obtained from the low-field region of M
vs. H curves, and is estimated from the lognormal distribution obtained from the
parameters and hAP i. By plotting / vs. T/(ıMs )2 and following the relationship
between apparent and real values shown in Ref. [23], one obtains ˇ and, therefore,
T* [23].
But even this model fails when systems are close to the percolation limit [57, 58].
As discussed earlier, in almost and fully percolating systems, magnetic behavior
is complex. Particularly since the publication of the aforementioned works of
Dormann–Bessais–Fiorani and Mørup, many experiments have been conducted and
theoretical developments made on mildly interacting systems, but few advances
have been made in theoretical models that can accurately describe systems that are
close to percolation or fully percolating.
Fig. 1.10 Zero field cooling and FC curves obtained for both PVA loaded with (left) 7 nm and
(right) 10 nm magnetite nanoparticles. The effect of interactions can be seen as the nanoparticle
concentration increases. Powder samples are composed of single nanoparticles in close contact,
with only the coating of citric acid separating them
Now, the question arises as to what the correct anisotropy value is, and what is its
real dependence with magnetic strenght. We have magnetite nanoparticles in which
the effective anisotropy’s largest contribution comes from the magnetocrystalline
anisotropy (due to the spin–orbit interaction). In addition, recall that in the presented
systems the magnetic nanoparticles are not percolated. We believe that the value
that is closest to that of a nanoparticle alone should come from a system in which
the magnetite nanoparticles are in the lower interacting state. This means that
interactions modify the anisotropy constant, as was observed by Ferrari et al. [62].
24 L.M. Socolovsky and O. Moscoso Londoño
Fig. 1.11 Blocking temperature distribution obtained for both PVA loaded with (left) 7 nm and
(right) 10 nm magnetite nanoparticles. Symbols: obtained from derivative method. Straight line:
lognormal fit from where the mean blocking temperature TB and the lognormal standard
deviation were determined
Note that when anisotropy constants are studied or compared, careful attention
should be paid specifically to the dilution conditions, besides the size and grafting
of the nanoparticle.
Table 1.2 Anisotropy values obtained for both systems using hTB i through the derivative method,
KhTBi (column 4), maximum of ZFC curve, KT-ZFC (last column)
NP mean size wt. % of NP hTB i (K) KhTBi (104 J/m3 ) TZFC (K) KT-ZFC (104 J/m3 )
7 nm (system A) 0:5 47 9:03 80 15.3
3 48 9:22 92 17.6
15 49 9:41 115 22.1
30 135 26:0 148 28.4
10 nm (system B) 0:5 49 3:22 125 8.24
3 66 4:35 150 9.88
15 71 4:68 236 15.5
30 86 5:66 240 15.8
(less than 1 at.%) some features appear that resemble those of superparamagnetic
systems, like a cusp in a magnetization or susceptibility vs. temperature measure-
ment, or irreversibilities in a ZFC-FC curve. The origin of that magnetic order
is frustration because of the competing interactions among magnetic ions, like
RKKY and others mentioned earlier, which makes it different from the thermally
activated process that is SPM [8, 16, 63]. When the concentration of magnetic atoms
is increased, some ions will be close enough to establish direct interactions like
direct interchange, super interchange, or Dzyalonskii–Moriya interchange. Those
interactions are stronger than RKKY, so the magnetic landscape becomes more
complicated. If we look at the conformation of the sample, at a concentration of
atomic ions of a few atomic percent, we see that those that are in close contact can
be regarded as nanoparticles, despite their origin. Those regions in close contact
have interactions like direct interchange or super interchange, as in a regular
nanoparticle. This is called cluster glass (CG) [64, 65]. For face-centered-cubic (fcc)
arrangements, percolation takes place at concentrations of approximately 17 at.%
for ferromagnetic ordering and 45 at.% for antiferromagnetic ordering. Chains of
interconnected magnetic ions are distinguishable, which produces regions of fer-
romagnetic (or antiferromagnetic) order. This state is called quasi-ferromagnetism
(QFM). They used to be called mictomagnetic (mixed magnetism) because it was
believed that they were mixtures of antiferro and ferromagnetic interactions [64,
65]. In addition, the term reentrant magnetism is also used. Earlier, discussions took
place about the physical situation for this state, but now the state is interpreted as a
transverse component freezing of the magnetization vector [66].
Because in some systems it can be difficult to adequately characterize an
unknown material and determine whether it is a spin-glass or superparamagnetic
system, we can use a criterion that takes into account blocking (or glassy) tempera-
tures and frequencies, obtained in susceptibility vs. temperature measurements [16].
Such a criterion is known as the Mydosh parameter ˚ [16] and can be calculated
from the maximum of the imaginary part of AC susceptibility (TM ), according to
the equation
26 L.M. Socolovsky and O. Moscoso Londoño
Fig. 1.12 Magnetic phases when magnetic atom concentration is increased. At very low concen-
trations of less than 1 at.%, we have canonical spin glasses. Increasing the concentration produces
a phase where spin-glass structures coexist with short-range magnetic order (SRO), such as direct
interchanges between neighboring magnetic atoms. Above 10 at.%, bigger SRO regions appear,
forming clusters that can be viewed as nanoparticles. Crossing the percolation limit, we encounter
a long-range magnetic order (LRO) that causes ferromagnetic features in the magnetic behavior
TM
ˆD ; (1.24)
TM log .!/
where TM is the frequency shift for the temperature of that maximum, and ! is
the frequency of the measurement [16]. Values of ˚ < 0.018 are found for canonical
spin glasses. Higher values, on the order of 101 , are typical of superparamagnetic
systems. In almost and fully percolated systems, ˚ values are small, independently
of whether the system is a spin glass or SPM in origin (Fig. 1.12).
These systems have been studied systematically since the 1960s. Interestingly,
one of the most studied systems is Au: Fe, both of whose elements are metallic.
Since they do not form alloys, they are prone to segregation, so a tendency to form
aggregates of pure iron is expected. The enthalpy of formation is positive [67]. At
room temperature and concentrations above 1 at % it is not difficult to imagine that
nanoparticles of iron can form from the migration of isolated atoms. Experiments
on such a system started as early as the 1930s, but the term spin glass was coined
only in the early 1970s. The subject attracted the attention of theoreticians who
developed models to describe these systems [12].
An interesting experiment on a special type of nanoparticle was performed
by Nunes et al. [68]. They prepared by chemical synthesis nickel nanoparticles
with an almost amorphous internal atomic structure. Those nanoparticles have an
approximate diameter of 12 nm, but inside, the atoms are highly disordered, forming
small clusters of just a few atoms. This leads to a magnetic behavior that is closer
to a cluster-glass system, despite the fact that it is formed by very well-defined
nanoparticles. In this experiment, ZFC-FC measurements showed a maximum close
1 Consequences of Magnetic Interaction Phenomena in Granular Systems 27
to 20 K that did not change if higher fields were used, at least up to 1000 Oe, a
characteristic of spin glasses. AC susceptibility measurements showed, by applying
the Mydosh criterion, that the system could be treated as a spin-glass one ( D 0.01),
so it must be regarded as a cluster glass. The same researchers also investigated
smaller Ni nanoparticles, approximately 5 nm, prepared using a similar synthesis
route [69]. They found that SPM order no longer existed to give rise to a mixture
of paramagnetic and ferromagnetic orders, again despite the existence of very
well-defined nanoparticles. An important increase in surface anisotropy was also
registered.
ŒR .Hmax / R.0/
GMR D ; (1.25)
R.0/
where R(Hmax ) and R(0) are the electrical resistances measured at the maximum
magnetic field Hmax and at the zero field.
Alternatively, we can use R(Hc ) instead of R(0) if the coercive field is well
marked or if it is large, as is the case for multilayered materials. The minus sign
means only that the change is negative, and sometimes it is calculated without it, as
shown in Fig. 1.13.
Giant magnetoresistance is produced by differential spin-dependent scattering
due to the relative orientation of moments of magnetic entities, as in sandwiched
28 L.M. Socolovsky and O. Moscoso Londoño
GMR was observed and interpreted first in multiple Fe-Cr layers in two
contemporary experiments done almost simultaneously by the groups of Albert Fert
and Peter Grünberg [73, 74]. Fert had already studied theoretically and developed
a two-subband model for the resistivity in ferromagnetic metals [75]. Experimental
realization of those sandwiches with thicknesses of just a few nanometers was only
possible when MBE machines were available, in the late 1980s. GMR in granular
materials was first reported by Chien in 1992 [76], but it had been previously
observed by other researchers, as noted even in that article. Systems that are
regarded as canonical spin glasses, like Au-Fe, show negative magnetoresistance.
Nigam and Majumdar, in an article published in 1983, analyzed magnetoresistance
measured in several samples of canonical spin glasses, in concentrations of magnetic
solute that reached approximately 10 at.%, which cannot be considered a diluted
magnetic sample (see our earlier discussion on the magnetism of spin glasses) [77].
Some results on magnetoresistance on diluted alloys that form spin glasses were
collected in Refs. [78, 79], and in Ref. [80]. The first report on a so-called negative
magnetoresistance on a spin glass–like system, at least from what we investigated,
was reported by Nakhimovich in 1941, in a Soviet physics journal [81].
In granular systems, GMR depends on the relative orientation of nanoparti-
cles’ supermoments (Figs. 1.13 and 1.14). When a magnetic field is applied,
supermoments align. The greater the field strength, the greater the alignment of
supermoments and, hence, the lower the electrical resistance [72, 82]. An interesting
theory of GMR in granular materials was presented in an article by Ferrari et al.,
where spin-dependent scattering was taken into account within nanoparticles and
their boundaries (Fig. 1.15) [83].
At low nanoparticle concentrations, GMR scales with the square of magneti-
zation. In concentrated systems, as an effect of increased magnetic interaction,
this relationship no longer exists. In such systems it is seen, in low fields, that
magnetoresistance vs. magnetization parabola flattens [84]. This is clearly seen in
the article of Liu et al., where 8–9 nm magnetite nanoparticles were compressed
into pellets, to show the GMR. This effect does not scale with M2 , as is clearly seen
in Fig. 1.15. It is also seen in Refs. [84–85].
The effect is greater when NP concentration is close to the percolation limit.
When this limit is crossed, mixed effects take place. A lower GMR effect can
be recorded, owing to the remaining nanoparticles that are still relatively free to
orientate with the field and the chains of nanoparticles that have ferromagnetic
interactions among them [86].
Fig. 1.14 Alignment of a nanoparticle’s supermoment with an external magnetic field causes
electrical resistance to drop
ŒR .Hs / R .Hc /
TMR D ; (1.26)
R .Hc /
where R(Hs ) and R(Hc ) are respectively the resistances measured at the magnetic
saturation Hs and coercive field Hc . Alternatively, we can use R(0) instead of R(Hc )
1 Consequences of Magnetic Interaction Phenomena in Granular Systems 31
0
Fe10Cu90
MR [%]
-1 0 1
H [T]
3
0
Fe19Cu81
2
MR [%]
MR [%]
1 0
-1 0 1 0 10 20 30 40 50
H [T] Fe concentration [at %]
if the coercive field is not well defined or is small, and both cases are common in
granular materials at room temperature or with a wide size distribution, as is the
case with GMR presented earlier.
In Jullière’s model, TMR is related to the polarization, P, of spins of the tunneling
electrons by
P2
TMR D : (1.27)
1CP2
Similarly to what happened with GMR, TMR was measured first in multilayered
systems, followed by granular materials. Impressive values have been recorded,
sometimes greater than their multilayered counterparts [91]. Some groups have
measured large TMRs at room temperature in granular samples of Fe3 O4 in
polymeric matrices [92, 93].
Magnetoresistance has been found in nanoparticles grafted with oleic acid, a
system that can also be considered granular. Those experiments were based on
the fact that grafting prevents direct contact between nanoparticles and retains a
distance of around 2–4 nm. By thermal annealing or chemical treatment, the distance
between nanoparticles can be reduced to less than 1 nm [94–97]. Unfortunately, in
those functionalized nanoparticles grafting seems to be different if nanoparticles are
close enough, so in concentrated samples nanoparticles have a tendency to form
chains, as was observed experimentally [98].
32 L.M. Socolovsky and O. Moscoso Londoño
H D R0 B C REHE 0 M; (1.28)
where R0 is the ordinary Hall coefficient, and REHE is the extraordinary one [100].
The first term is the usual effect that appears as a consequence of the Lorentz force
that is exerted by the external magnetic field on charge carriers. The second term is
the extraordinary part, which is due to the scattering of the polarized conduction d
electrons and in granular systems can be very large (Fig. 1.16).
The sources of this effect are not clear. The effect is associated with spin-
dependent scattering at nanoparticle surfaces. This breaks the spatial symmetry on
the trajectories of charge carriers and produces two effects, skew scattering and a
side jump [101]. Skew scattering is a constant deviation of the trajectories of charge
carriers [102]. It is usually described by
Fig. 1.16 A typical Hall resistivity curve. The low-field part is almost straight, in which the
contribution of the first term in Eq. 1.28, which is the ordinary part, contributes slightly to the Hall
resistivity. Thus, the slope of this part is almost the extraordinary contribution, and REHE can be
extracted from this value. The high-field part, where magnetization is reversible, is almost straight,
so a second slope can be calculated. From that, the ordinary Hall resistivity can be calculated
1 Consequences of Magnetic Interaction Phenomena in Granular Systems 33
Frequently the second term is dropped, so it is accepted that REHE has a linear
dependence on . At low temperatures or in metals, where resistivity is small, skew
scattering is usually considered the main mechanism for this extraordinary Hall
resistivity [101].
The second mechanism is the side jump, a lateral quantum displacement of
charges, which results in a quadratic dependence on of REHE [101, 103]. This
mechanism is usually considered in concentrated granular materials, and it is
important at high temperatures.
Thus, it is considered that a measurement of the extraordinary Hall resistivity
should scale with ordinary resistance to a power of 1 or 2, depending on the
mechanism. A shorter expression is used:
1.4 Conclusions
Granular materials show features that are interesting for both academic studies
and technological applications. A basic set of characterizations can be made using
equipment available in several laboratories. Zero field cooling and field cooling
magnetization vs. temperature measurements provide important information, in
addition to unambiguously showing the existence of nanoparticles and an approx-
34 L.M. Socolovsky and O. Moscoso Londoño
References
1. Fert A (2008) The present and the future of spintronics. Thin Solids Films 517(1):2–5
2. Martin DH (1967) Magnetism in solids. Iliffe Books Ltd, Londres
3. O’Handley RC (2000) Modern magnetic materials: principles and application. Wiley, New
York
4. Knobel M, Nunes WC, Socolovsky LM, De Biasi E, Vargas JM, Denardin JC (2008)
Superparamagnetism and other magnetic features in granular materials: a review on ideal
and real systems. J Nanosci Nanotech 8:2836
5. Bedanta S, Petracic O, Kleemann W (2015) Supermagnetism. In: Buschow KHJ (ed)
Handbook of magnetic materials. Elsevier B.V., Amsterdam
6. Guimarães AP (2009) Principles of nanomagnetism. Springer, Berlin
7. Blundell S (2001) Magnetism in Condensed Matter. Oxford University Press, London
8. Hurd CM (1982) Varieties of magnetic order in solids. Contemp Phys 23(5):469
9. Aharoni A (1996) Introduction to the theory of ferromagnetism. Clarendon Press, Oxford
10. Bedanta S, Kleemann W (2009) Supermagnetism. J Phys D: Appl Phys 42:013001
11. Battle X, Labarta A (2002) Finite-size effects in fine particles: magnetic and transport
properties. J Phys D: Appl Phys 35:R15–R42
12. Majetich SA, Sachan M (2006) Magnetostatic interactions in magnetic nanoparticle assem-
blies: energy, time and length scales. J Phys D: Appl Phys 39:R407–R422
13. Zener C (1954) Classical theory of the temperature dependence of magnetic anisotropy
energy. Phys Rev 96(5):1335
14. Torres TE, Lima E Jr, Mayoral A, Ibarra A, Marquina C, Ibarra MR, Goya GF (2015) Validity
of the Néel-Arrhenius model for highly anisotropic Cox Fe3x O4 nanoparticles. J Appl Phys
118:183902
15. Skomski R (1999) RKKY interactions between nanomagnets of arbitrary shape. Europhys
Lett 48(4):455–460
16. Mydosh JA (1993) Spin Glasses. Taylor & Francis, Washington, DC
1 Consequences of Magnetic Interaction Phenomena in Granular Systems 35
17. Skomski R (2001) Are there superspin glasses? J Appl Phys 109:07E149
18. Mørup S (1994) Superferromagnetic nanostructures. Hyp Int 9:171–185
19. Essam JW (1980) Percolation theory. Reports on Progress in Physics 43(7):833–912
20. Bakuzis AP, Morais PC (2004) Dilution effect upon the superferromagnetic ordering of two-
dimensional magnetic nanodots. Phys. Stat. Sol. (C) 1(12):3332–3335
21. Yang HT, Liu HL, Song NN, Du HF, Zhang XQ, Cheng ZH, Shen J, Li LF (2011)
Determination of the critical interspacing for the noninteracting magnetic nanoparticle
system. Appl Phys Lett 98:153112
22. Vargas JM, Nunes WC, Socolovsky LM, Knobel M, Zanchet D (2005) Phys Rev B 72:184428
23. Allia P, Coisson M, Tiberto P, Vinai F, Knobel M, Novak MA, Nunes WC (2001) Granular
Cu-Co alloys as interacting superparamagnets. Phys Rev B 64:144420
24. Skumryev V, Stoyanov S, Zhang Y, Hadjipanayis G, Givord D, Nogués J (2003) Beating the
superparamagnetic limit with exchange bias. Nature 423:850–853
25. Gamino M, Michea S, Denardin JC, Schelp LF, Dorneles LS. Exchange-biased extraordinary
Hall effect in SiO2 /Co/CoO multilayers (unpublished results)
26. Luis F, Petroff F, Torres JM, García LM, Bartolomé J, Carrey J, Vaurès A (2002) Magnetic
relaxation of interacting Co clusters: crossover from two- to three-dimensional lattices. Phys
Rev Lett 88:217205
27. Bruvera IJ, Zélis PM, Calatayud MP, Goya GF, Sánchez FH (2016) Determination of the
blocking temperature of magnetic nanoparticles: the good, the bad, and the ugly. J Appl Phys
118(18):184304
28. Socolovsky LM, Denardin JC, Brandl AL, Knobel M (2003) Magnetotransport, magnetic, and
structural properties of TM–SiO2 (TMDFe, Co, Ni) granular alloys. Mat Char 50:117–121
29. Tournus F, Bonet E (2011) J Mag Magn Mat. 323:1109
30. Tournus F, Tamion A (2011) J Mag Magn Mat 323:1118
31. Knobel M, Nunes WC, Winnischofer H, Rocha TCR, Socolovsky LM, Mayorga CL,
Zanchet D (2007) Effects of magnetic interparticle coupling on the blocking temperature of
ferromagnetic nanoparticle arrays. M. Journal of Non-Crystalline Solids 353:743–747
32. Socolovsky LM, Sánchez FH, Shingu PH (2002) Physica B: Phys Cond Matter 3201-4:149–
152
33. Fernández van Raap MB, Sánchez FH, Rodríguez Torres CE, Casas LI, Roig A, Molins E
(2005) J Phys: Cond Matter 17:6519
34. García-Palacios JL (2000) On the statics and dynamics of magneto-anisotropic nanoparticles.
Adv Chem Phys 112:1–210
35. Shtrikmann S, Wohlfarth EP (1981) The theory of the Vogel-Fulcher law of spin glasses. Phys
Lett A 85:467
36. Bittova B, Poltierova-Vejpravova J, Roca AG, Morales MP, Tyrpekl V (2010) Relaxation
phenomena in ensembles of CoFe2 O4 nanoparticles. J Phys: Conference Series 200:072012
37. Fang M, Strom V, Olsson RT, Belova L, Rao KV (2012) Particle size and magnetic properties
dependence on growth temperature for rapid mixed co-precipitated magnetite nanoparticles.
Nanotechnology 23:145601
38. Barbeta VB, Jardim RF, Kiyohara PK, Effenberger FB, Rossi LM (2010) Magnetic properties
of Fe3 O4 nanoparticles coated with oleic and dodecanoic acids. J Appl Phys 107:073913
39. Nadeem K, Krenn H, Traussnig T, Wurschum R, Szabo DV, Letofsky-Papst I (2011) Effect of
dipolar and exchange interactions on magnetic blocking of maghemite nanoparticles. J Mag
Magn Mat 323:1998–2004
40. Masunaga SH, Jardim RF, Fichtner PFP, Rivas J (2009) Role of dipolar interactions in a
system of Ni nanoparticles studied by magnetic susceptibility measurements. Phys Rev B
80:184428
41. Parekh K, Upadhyay RV, Aswal VK (2009) Monodispersed superparamagnetic Fe3 O4
nanoparticles: synthesis and characterization. J Nanosci Nanotech 9:2104–2110
42. Zelenaková A, Zelenak V, Michalík S, Kovác J, Meisel MW (2014) Structural and magnetic
properties of CoO-Pt core-shell nanoparticles. Phys Rev B 89:104417
36 L.M. Socolovsky and O. Moscoso Londoño
68. Nunes WC, De Biasi E, Meneses CT, Knobel M, Winnischofer H, Rocha TCR, Zanchet D
(2008) Appl Phys Lett 92(18):183113
69. De Biasi E, León-Vanegas A, Nunes WC, Sharma SK, Haddad P, Rocha TCR, Santos Duque
JG, Zanchet D, Knobel M (2008) Eur Phys J B 66(4):503–508
70. Levy PM (1994) Solid State Physics 47:367
71. Chien CL, Xiao JQ, Jiang JS (1993) Giant negative magnetoresistance in granular ferromag-
netic systems. J Appl Phys 73(10):5309
72. Camblong HE, Levy PM, Zhang S (1995) Phys Rev B 51:2216052
73. Baibich MN, Broto JM, Fert A, Van Dau FN, Petroff F, Etienne P, Creuzet G, Friederich A,
Chazelas J (1988) Giant magnetoresistance of (001) Fe/(001) Cr magnetic superlattices. Phys
Rev Lett 61:2472
74. Binasch G, Grünberg P, Saurenbach F, Zinn W (1989) Enhanced magnetoresistance in layered
magnetic structures with antiferromagnetic interlayer exchange. Phys Rev B 39(7):4828(R)
75. Fert A (1969) Two-current conduction in ferromagnetic metals and spin wave-electron
collisions. J Phys C 2(2):1784
76. Xiao JQ, Jiang JS, Chien CL (1992) Giant magnetoresistance in nonmultilayer magnetic
systems. Phys Rev Lett 68(25):3749
77. Nigam AK, Majumdar AK (1979) Anomalous magnetoresistance in AuFe alloys. J App Phys
50:1712
78. Gerritsen AK (1953) The magnetoresistances of alloys of a noble metal and a transition metal
at low temperatures. Physica 19(1–12):61–73
79. Gerritsen AK (1957) Resistance and magneto-resistance of dilute alloys of gold with iron at
low temperatures. Physica 23(6–10):1087–1099
80. Rohrer H (1968) Magnetoresistance of dilute alloys. Phys Rev B 174(2):583
81. Nakhimovich NM (1941) Dependence of the electrical resistance of gold and several of its
alloys on a magnetic field al low temperatures. J Phys V 2–3:141
82. Maekawa S, Inoue J, Itoh C (1996) J Appl Phys 79:8
83. Ferrari EF, da Silva FCS, Knobel M (1999) Phys Rev B 59(13):8412
84. Allia P, Knobel M, Tiberto P, Vinai F (1995) Magnetic properties and giant magnetoresistance
of melt-spun granular Cu100x Cox alloys. Phys Rev B 52:15398–11541
85. Liu K, Zhao L, Klavins P, Osterloh FE, Hiramatsu H (2003) Extrinsic magnetoresistance in
magnetite nanoparticles. J App Phys 93:7951
86. Kechkrakos D, Trohidou KN (2002) Physica B 318:360
87. Julliere M (1975) Phys Lett A 54(3):225
88. Abeles R, Sheng P, Coutts MD, Arie Y(1975) Structural and electrical properties of granular
metal films. Adv Phys 24–3:407–461
89. Kechkrakos D, Trohidou KN (2005) Phys. Rev. B 71:054416
90. Inoue J, Maekawa S (1996) Theory of tunneling magnetoresistance in granular magnetic
films. Phys. Rev. B 53(18):11927–11929
91. Denardin JC, Knobel M, Dorneles LS, Schelp LF (2004) Structural and magnetotransport
properties of discontinuous Co/SiO2 multilayers. Mat Sci Eng B 112(2–3):120–122
92. Varfolomeev AV, Zavyalov AS, Volkov AV, Volkov IA, Moskvina MA, Polyakov SN,
Malashko AP, Bayburtskiy FS, Baldokhin YV, Stepanov GV (2006) Preparation and inves-
tigation of magnetic composite materials properties on the basis of magnetite and polyvinyl
alcohol. Oxidation Comm 29(3):693–697
93. Wang W, Yu M, Chen Y, Tang J (2006) Large room-temperature spin-dependent tunneling
magnetoresistance in a Fe3 O4 -polymer composite system. Phys Rev B 73:134412
94. Black CT, Murray CB, Sandstrom RL, Sun S (2000) Spin-dependent tunneling in self-
assembled cobalt-nanocrystal superlattices. Science 290:1131–1134
95. Taub N, Tsukernik A, Markovich G (2009) J Mag Magn Mater 321:1933
96. Poddar P, Fried T, Markovich G (2002) First-order metal-insulator transition and spin-
polarized tunneling in Fe3 O4 nanocrystals. Phys Rev B 65:172405
97. Markovich G (2013) Magneto-transport and magnetization dynamics in magnetic nanoparti-
cle assemblies. Mat Today 38:939
38 L.M. Socolovsky and O. Moscoso Londoño
98. Bonini M, Fratini E, Baglioni P (2007) SAXS study of chain-like structures formed by
magnetic nanoparticles. Mat Sci Eng C 27:1377–1381
99. Pakhomov AB, Yan X, Zhao B (1995) Appl Phys Lett 67:3497
100. Hurd M (1972) The Hall effect in metals and alloys. Plenum Press, New York
101. Gerber A, Milner A, Finkler A, Karpovski M, Goldsmith L, Tuaillon-Combes J, Boisron O,
Mélinon P, Perez A (2004) Correlation between the extraordinary Hall effect and resistivity.
Phys Rev B 69:224403
102. Karplus R, Luttinger M (1954) Hall effect in ferromagnetics. Phys Rev 95:1154
103. Berger L (1970) Side-jump mechanism for the Hall effect on ferromagnets. Phys Rev
B2:4559
104. Jing XN, Wang N, Pakhomov AB, Fung KK, Yan X (1996) Effect of annealing on the giant
Hall effect. Phys Rev B 53:14032
105. Socolovsky LM, Oliveira CLP, Denardin JC, Knobel M, Torriani IL (2005) Nanostructure of
granular Co-SiO2 thin films modified by thermal treatment and its relationship with giant Hall
effect. Phys Rev B 72:184423
106. Denardin JC, Pakhomov AB, Brandl AL, Socolovsky LM, Knobel M, Zhang XX (2003) Appl
Phys Lett 82(5):763–765
107. Socolovsky LM, Oliveira CLP, Denardin JC, Knobel M, Torriani IL (2006) Nanostructure
and giant Hall effect in TMx (SiO2 )1-x (TM D Co, Fe, Ni) granular system. J Appl Phys
99:08C511
108. Socolovsky LM, Sánchez FH, Shingu PH (2001) Giant magnetoresistive properties of
Fex Au100–x alloys produced by mechanical alloying. J Mag Magn Mat 226–230:736