Você está na página 1de 10

Nano Today (2009) 4, 438447

available at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/nanotoday

REVIEW

Nanoparticle magnetism
Georgia C. Papaefthymiou a,b,
a
b

Department of Physics, Villanova University, Villanova, PA 19073, United States


Institute of Materials Science, NCSR Demokritos, 15310 Athens, Greece

Received 8 July 2009; received in revised form 11 August 2009; accepted 13 August 2009
Available online 11 September 2009

KEYWORDS
Superparamagnetism;
Finite-size effects;
Surface effects;
Intrinsic spin
structure;
Exchange-bias;
Monte Carlo
simulations

Summary The current status of fundamental physics that govern nanoparticle magnetism
is reviewed. Emphasis is placed on studies of the particles intrinsic spin structure as inuenced by surface and nite-size effects. Theories of dynamic spin uctuation and spin reversal
processes for uniaxial, isolated magnetic nanoparticles are presented, as applied primarily to
the interpretation of magnetization and Mssbauer spectroscopic measurements. Monte Carlo
simulation studies that corroborate experimental ndings and advance elucidation of core vs.
surface contributions to magnetic behavior are also presented. In addition, applications to
nano and bio technology and future research directions in core/shell, matrix embedded and
interacting magnetic nanoparticles are also addressed.
2009 Published by Elsevier Ltd.

Introduction
Materials engineering utilizes an interdisciplinary, sciencebased approach to the production of nanophase materials
with novel microstructure, in order to tap the physical
resources of the quantum/classical boundary. The present
explosion in nanoscale research derives primarily from
recent advances in synthetic approaches, biomimetic processes and the development of enabling technologies that
allow the manipulation, stabilization and characterization
of matter at the atomic level. Among the fundamental scientic disciplines that have contributed to extraordinary
recent advances in nanoscience and nanotechnology, magnetism holds a prominent place. Finite-size effects in the

Correspondence address: Department of Physics, Villanova University, Mendel Hall, 800 Lancaster Ave., Villanova, PA 19073, United
States. Tel.: +1 610 519 4883.
E-mail address: gcp@villanova.edu.

1748-0132/$ see front matter 2009 Published by Elsevier Ltd.


doi:10.1016/j.nantod.2009.08.006

magnetic properties of matter have unraveled interesting


new magnetic phenomena in nano-materials, not manifested in the bulk [1,2]. This has led to new physics and new
technological applications in diverse areas of science and
technology ranging from ultra-high density magnetic recording [35], improved nanocomposite permanent magnet
materials [6], giant-magneto-resistance phenomena [7,8],
spintronics [9,10], biogenic nano-magnets [11,12] to fast
expanding bio-medical applications [1316]. The discovery
of giant magneto-resistance earned the 2007 Nobel Prize
in Physics due to its important device applications in new
read-head technologies [17]. Thus, magnetism, arguably the
oldest scientic discipline to have been continuously investigated since the mining of magnetite (loadstone) in ancient
Greece in the 6th century BC, still offers fertile ground for
scientic innovation today, within the realm of nanomagnetism.
In this article the fundamental physics that govern
nanoparticle magnetism is reviewed as garnered from experimental and theoretical studies of the magnetic properties

Nanoparticle magnetism

439

of nanoparticles pertained to their intrinsic spin structure


and spin dynamical phenomena unique to the nanostructured state. In addition, applications to nano and bio
technology and future research directions in core/shell,
matrix embedded and interacting magnetic nanoparticles
are briey addressed.

From bulk to nano


The magnetic properties of matter are classied according
to its response to an external magnetic eld [18,19]. Upon
the application of a eld, diamagnetic and paramagnetic
materials develop negative and positive magnetization,
respectively, which readily disappears upon removal of the
eld. To the contrary, magnetic materials develop positive magnetization which persists upon removal of the
eld, giving rise to hysteresis loops that determine the
technological properties of the material [19]. Magnetic
order in materials is a collective phenomenon anchored in
inter-atomic quantum-mechanical exchange [2022], which
couples neighboring atomic moments. The Pauli exclusion principle leads to the Heisenberg model of magnetic
exchange [23], where inter-atomic exchange interactions
can be described phenomenologically according to the
Hamiltonian:
Hex = 2

Jij Si Sj

(1)

i<j

rst introduced by Dirac [19]. Here Si and Sj are atomic spins
at different neighboring sites. The value of the exchange
coefcient or exchange constant, Jij , is strongly related to
the overlap of the two atoms and thus the inter-nuclear
distance. Positive values of Jij favor ferromagnetism, while
negative values lead to anti-ferromagnetism. In ferromagnetic materials neighboring moments are parallel while in
antiferromagnetic materials are anti-parallel. The exchange
interaction between spins in Eq. (1) is purely isotropic; there
is no built-in preferred direction. However, within a crystal
lattice symmetry is reduced; not all directions are equivalent. The direction of spin orientation is found to be along
some preferred crystallographic axis, which necessitates an
additional anisotropic term to the Hamiltonian, Eq. (2):
Hex = 2


i<j

Jij Si Sj Kmc

(Szi )2

(2)

Here, the second term stands for the simplest example


of uniaxial anisotropy, not corresponding to any particular crystal symmetry. Nevertheless, it is crystal anisotropy
that provides this anisotropic term, whereby the spinorbit
interaction couples the spin to the preferred crystallographic axis assumed along the z-direction [24]. The
magnetocrystalline anisotropy, Kmc , determines the direction and stability of the magnetization within the material.
The distance over which the spins retain their collinearity
gives the size of a magnetic-domain [18,19]. Within a singlemagnetic-domain, typically 100 nm, the magnetization is
uniform, that is, all spins are collinear. However, in the bulk
material single-magnetic-domains are oriented attempting
to minimize magnetostatic energy, due to the presence of
long-range magnetostatic interactions that compete with
the short-range exchange interactions in minimizing the

overall magnetic energy of the system [25]. Thus, in the bulk


the material consists of non-aligned magnetic-domain structures, separated by domain walls [18,19], which produce
magnetic ux closures, rendering the bulk material macroscopically non-magnetic. The application of even small
external magnetic elds, however, induces magnetic-wall
movement and nite magnetization. Only in high applied
elds, where magnetic saturation is reached, are all spins in
the material collinear.
Magnetic particles in the nanometer-size range are necessarily single-magnetic-domain (SMD) structures and thus
magnetically saturated in the absence of an external magnetic eld. It has been estimated that the critical size
below which a spherical particle exists as a single-magneticdomain is given by [19]:
RSMD

6 AK
=
0 M2s

(3)

Here A, the exchange stiffness, is a constant characteristic of the material related to its critical temperature for
magnetic ordering [19], K is the magnetic anisotropy of the
particle, Ms the saturation magnetization and 0 the permeability of free space. The response of nanoparticles to
external magnetic elds reects characteristically different spin dynamical processes compared to those of their
bulk counterparts. Namely, they reect spin rotation within
a single-domain rather than magnetic-domain-wall movement.
Technological applications of SMD-particles depend critically on our fundamental understanding of spin rotation
dynamics and their dependence on the particles intrinsic
spin structure. The simplest model proposed by Stoner and
Wohlfarth [26] in their pioneering work on moment reversal
in SMD-particles, assumes coherent spin rotation whereby
all spins within the single-domain particle are collinear and
rotate in unison. It predicts the eld strength necessary
to reverse the spin orientation direction or coercive eld,
Hc . The model assumes uniform magnetization throughout
the particle, which remains so throughout the rotation process, an assumption that allows for exact solutions, but
not generally observed experimentally. Additional models,
allowing for non-uniform, incoherent spin reversal mechanisms have also been developed [2730]. Generally the
energies required to reverse the spin orientation within a
SMD-particle are larger than those needed to induce domainwall movement, yielding coercivities larger than those of the
bulk. Fig. 1 shows the dependence of coercivity as a function
of particle size for Co, Fe and iron oxide particles reproduced from the classic paper by Luborsky [31]. The increased
coercivity is the result of the transition from multi-domain
to single-domain behavior. The maximum of coercivity is
controlled by the anisotropy energy which opposes rotation
away from the easy axis of magnetization. Various sources
of anisotropy have been identied; crystal anisotropy, strain
anisotropy, exchange anisotropy and shape anisotropy [31].
The high coercivity of magnetic nanoparticles in the range of
10100 nm diameters is what the design of nanocrystalline
permanent magnets and of media for device applications is
based on. For smaller particle sizes the coercivity decreases
rapidly to zero due to thermal effects that result in rapid
spin reversals, a regime known as superparamagnetism.

440

G.C. Papaefthymiou
sals between the two minima can be achieved by thermal
excitations over the energy barrier, when kT > KV, where k
is Boltzmanns constant and T is the temperature. According
to the NelBrown theory [32,33] the spin relaxation time,
 s , follows the Arrhenius Eq. (5):

s = 0 exp

Figure 1 Experimental relation between coercivity and diameter for particles deriving their coercive force principally from
crystal anisotropy energy [Reproduced by permission from F.E.
Lubrosky, J. Appl. Phys. 32 (1961) S171, Ref. [31]].

The superparamagnetic properties of magnetic nanoparticles determine many of their bio-medical applications.

Spin reversal in uniaxial magnetic


nanoparticles
Since the pioneering work of Nel [32], Brown [33] and
Aharoni [34] on uniaxial magnetic nanoparticles, superparamagnetic relaxation processes have been extensively
studied. Generally, a particle with uniaxial anisotropy is
modeled by an ellipsoid of revolution, or prolate spheroid,
with the easy axis of magnetization along the major axis, as
shown in Fig. 2(a). The magnetic anisotropy energy:
Eani () = KV cos2

(4)

where is the angle between the direction of magnetiza and the easy axis, V is the volume of the particle
tion M
and K is the uniaxial magnetic anisotropy constant. Fig. 2(b)
gives a graphical representation of the magnetic anisotropy
energy as a function of the angle, . The potential energy
landscape has two minima of equal depth, for = 0 and ,
separated by an energy barrier U = KV. In the absence of an
external eld the particle moment has equal probability to
lie along either direction of the easy axis. Moment rever-

KV
kT

(5)

where  0 is a characteristic attempt time for spin reversal,


a constant characteristic of the material and of the order of
109 to 1012 [35]. The application of an external magnetic
eld at a random angle  relative to the anisotropy axis intro and away
duces competition for moment alignment along H
from the easy axis. Fig. 2(b) depicts how the energy land is applied along the
scape is modied for the case where H
anisotropy axis, by lifting the energy degeneracy of the wells
and decreasing the energy barrier. According to the Stoner
and Wohlfarth theory [26], in the absence of temperature
assisted spin reversals, uniformly magnetized single-domain
particles with uniaxial anisotropy undergoing coherent spin
rotation would exhibit a maximum coercivity Hc = 2K/Ms at
zero absolute temperature. Non-temperature induced spin
reversals have also been observed to occur through macroscopic quantum-mechanical tunneling of the magnetization
vector between the two wells [3638].

Intrinsic spin structure and dynamic spin


relaxation
In real systems the dynamics of spin reversals are sensitive to the internal spin structure of the particles. Mounting
experimental evidence in studies of well characterized,
monodispersed magnetic nanoparticles point to an intrinsic spin structure of greater complexity compared to the
simple collinear model of StonerWohlfarth [26]. The complexity arises from the abrupt interruption of the crystaland spin-lattice structure at the particles surface. It is generally recognized that novel properties at the nanoscale
arise from the large number of atoms that lie on the surface, along grain boundaries or particle/support interfaces.
Lattice distortions at the surface trap atoms in thermodynamically non-equilibrium states, which are not generally
encountered in the bulk, as depicted in Fig. 3 [39]. Thus, the

Figure 2 (a) Schematic of a prolate spheroid depicting a nanoparticle with uniaxial magnetic anisotropy in the presence of
 at an angle  relative to the direction of the anisotropy axis. Angles , give the orientation the
an external magnetic eld H
 relative to the anisotropy axis and the magnetic eld, respectively. (b) Magnetic orientational
magnetization of the particle, M,
potential energy as a function of angle in the absence of an applied eld, solid line (), and in the presence of an applied eld
along the anisotropy axis, broken line ( ). The minima occur at = 0 and .

Nanoparticle magnetism

441

Figure 3 Schematic depiction of disorder encountered at the


surfaces of grains and grain boundaries of nanostructured materials [Reprinted by permission, H. Gleiter, Acta. Mater. 48 (2000)
1, Ref. [39]].

unusually large surface to volume ratio in nanosized materials provides for the stabilization of articial structures with
novel physical properties. With diminishing particle size the
surface to volume ratio increases, rendering strongly sizedependent behavior.
One property that shows extreme sensitivity to surface
structure is the particles magnetic anisotropy, which governs spin relaxation time and coercivity values. In the bulk,
crystal eld splitting induced magnetocrystalline anisotropy
is the primary source of anisotropy. In small particles,
additional contributions due to surface effects and surface
strain become dominant. In magnetic nanoparticles values
of magnetic anisotropy have been observed up to two orders
of magnitude larger compared to the magnetocrystalline
anisotropy of the bulk [40,41]. For spherical particles of
diameter D the empirical formula of Eq. (6) [40,4245] has
satisfactorily interpreted many experimental results:
K = Kc +

6Ks
D

Figure 4 Total effective anisotropy energy constant for


metallic iron particles as a function of reciprocal particle diameter [Reprinted by permission from F. Bdker, S. Mrup and S.
Lideroth, Phys. Rev. Lett. 72 (1994) 282, Ref. [43]].

magnetic saturation even in large magnetic elds [4649],


and hysteresis loop shifts in antiferromagnetic nanoparticles due to exchange-bias between the core and surface
magnetization [50].
Mssbauer spectroscopic studies have provided extensive
experimental data on various nanoparticle systems pertaining to their magnetic anisotropies and intrinsic spin structure
[2]. The Mssbauer effect detects the nuclear Zeeman splitting caused by the internal magnetic eld at the site of the
iron nucleus. An example to demonstrate the methodology
is shown in Fig. 6(a) for the biomineral core of ferritin, the
iron storage protein where antiferromagnetic ferrihydrite
naturally sequesters within the interior cavity of a protein
nanotemplate of 7 nm interior and 12 nm exterior diameters
[51,52]. The thick protein coat insures magnetic isolation
of the particles. At 4.2 K two magnetic sites are resolved
associated with core and surface atoms, respectively, with
different internal magnetic hyperne elds of Hc 495 kOe
and Hs 450 kOe, as measured by the overall splitting of
the external lines of each magnetic subspectrum (Fig. 6(a)).

(6)

Here Kc is the magnetocrystalline anisotropy of the core


of the particle characteristic of the material, and Ks
is the surface anisotropy characteristic of the particle,
both anisotropies treated as uniaxial. The above equation
assumes a simple additive effect of the surface ignoring any
cross-linking terms and predicts effective total anisotropies
that scale as 1/D, as reported in Fig. 4 [43] for small metal
iron particles in the range of 27 nm diameter.
Magnetic particles are, thus, modeled as two-phase systems composed of a highly crystalline core surrounded by a
disordered surface layer. Vacancies, broken bonds and lattice strain at the surface produce not only atomic disorder
but also spin frustration which destabilizes the collinear
spin arrangement of the StonerWohlfarth model at the surface, producing various canted spin structures (Fig. 5). This
complex intrinsic spin structure is supported by both theoretical and experimental studies in small particle systems,
as discussed below. Consequences are reduced Ms , lack of

Figure 5 (a) Schematic representation of a spherical particle


with uniform magnetization according to the Stoner and Wohlfarth model: all spins are collinear and rotate in unison. (b)
Schematic representation of two-phased particle with different
core and surface magnetization. Surface spins are canted relative to the cores magnetization and are subject to low energy
excitations (see text).

442

G.C. Papaefthymiou
H/H0 , is shown in Fig. 8. Here, H = Hhf (T) and H0 = Hhf (4.2 K),
assumed to be the saturation hyperne eld as T 0. For
the interior iron sites of the core of the particle, the temperature dependence of the reduced hyperne magnetic elds
is consistent with the CME model of Mrup and Topse [55],
described by Eq. (7):
H
kT
=1
H0
2KV

(7)

A maximum of 15% diminution in hyperne eld is expected


through this process, before the spectrum collapses to a
paramagnetic doublet due to superparamagnetism. The precipitous collapse of the reduced hyperne eld at surface
sites observed in Fig. 8 indicates the presence of additional
surface spin excitation modes, consistent with theories of
multi-spin nano-magnet systems, where surface anisotropies
introduce spin canting and a greater complexity in the
potential energy landscape at the surface [56].
The temperature at which the spectra change from
primarily magnetic (six-line absorption spectra) to paramagnetic (two-line absorption spectra) determines the blocking
temperature, TB , of the ensemble of particles in the sample. For the sample of Fig. 7, TB is estimated to be about
40 K, where the absorption area of the spectrum is com-

Figure 6 (a) Mssbauer spectra of iron ferrihydrite nanoparticles grown within horse spleen apo-ferritin nanotemplates
measured at 4.2 K. Two magnetic subsites are resolved associated with core and surface iron atoms. (b) Core/surface model
of the iron biomineral core of ferritin [Reprinted by permission from R.A. Brooks, J. Vymazal, R.B. Goldfarb, J.W. Bulte, P.
Aisen, Magn. Res. Med. 40 (1998) 227, Ref. [53]].

The iron site with larger magnetic eld exhibits sharper


absorption lines indicating a highly crystalline core, while
the smaller hyperne eld site exhibits broadened absorption lines indicating a more disordered or amorphous state
at the surface. A two-phase model of the ferritin biomineral
core depicted in Fig. 6(b) was originally proposed by Brooks
et al. [53] on the basis of relaxometry measurements.
Fig. 7 shows the full spectral prole of reconstituted
horse spleen ferritin in the temperature range from 4.2
to 80 K [54]. As temperature increases above 4.2 K, the
spectra broaden due to the onset of dynamic spin relaxation processes prior to full magnetization reversals at
thermal energies above the superparamagnetic energy barrier KV (Fig. 2). Simultaneously, the magnitudes of the
observed hyperne elds decrease, prior to their collapse
due to superparamagnetism at high temperatures and the
appearance of a quadrupole doublet. The reduction in the
magnitude of the magnetic hyperne eld is due to the
precession of the particles magnetization vector about
its anisotropy axis, at temperatures insufcient to induce
spin reversals, as proposed by Mrup and Topse in their
collective magnetic excitation (CME) model [55], which
successfully describes spin dynamics below the blocking
temperature, where the magnetization vector is trapped
or blocked within a single potential well. The temperature dependence of the reduced hyperne magnetic elds,

Figure 7 Mssbauer spectra of lyophilized, in vitro reconstituted horse spleen ferritin at various temperatures. Solid lines
(black) through the experimental points are least square ts,
to a superposition of iron subsites: purple, interior core sites;
green, surface sites; red, superparamagnetic sites [from Ref.
[54]].

Nanoparticle magnetism

443

Figure 8 Temperature dependence of the reduced hyperne


magnetic elds at interior, core (purple) and surface (green)
iron sites. The steeper reduction of surface elds is indicative
of a more complex potential energy landscape at the surface
[from Ref. [54]].

prised by 50% magnetic and 50% paramagnetic components.


Theoretically the blocking temperature is obtained from Eq.
(8) which follows from the Arrhenius Eq. (5):
TB =

KV/k
ln(m /0 )

(8)

Here  m is the characteristic measuring time of the experimental technique,  m 108 s for Mssbauer spectroscopy.
Thus, TB is not an intrinsic property of the nanoparticles
but rather a technique dependent parameter that records
spin dynamical behavior relative to the time window of the
experimental measurement.
Spin dynamics in small magnetic particles have also been
extensively studied through magnetization measurements
employing superconducting quantum interference device
(SQUID) magnetometers. Zero eld-cooled (ZFC) and eldcooled (FC) magnetization curves are routinely taken in
order to obtain the blocking temperature of the sample
and thus characterize the magnetic anisotropy of the particles. An example is shown in Fig. 9, again for the case
of the ferrihydrite core of reconstituted horse spleen ferritin [54]. The sample is initially cooled in zero eld. Under
this condition the two minima in the potential landscape
of Fig. 2(b) are equally populated and the sample exhibits
zero magnetization. The application of an external eld,
 introduces competition for spin reorientation away from
H,
the anisotropy axis and into the direction of the magnetic
eld (Fig. 2(a)). Small applied elds (500 Oe) are generally insufcient to induce spin reorientation away from the
anisotropy axis at low temperature. With increasing temperature, however, thermal activation assists in reorienting
the moment in the direction of the magnetic eld, inducing a magnetization. The recorded magnetization passes
through a maximum before it is reduced again at higher
temperatures due to thermally induced fast reversals of the
particles magnetization vector. The maximum in the ZFC
magnetization curve marks the blocking temperature of the
sample for magnetization measurements, which have a characteristic measuring time of 10100 s. Determination of TB
with two techniques of characteristic measuring times in

different time windows allows determination of both parameters [46], K and  0 , which govern the superparamagnetic
relaxation properties of the particles.
The type of SQUID experiments just described, while
consistent with the NelBrown spin reversal theory, lack
the ability to rigorously test spin reversal mechanisms, as
they are performed on samples containing a collection of
particles; details in spin reversal processes are masked
by particle-size distributions and the possible presence of
structural defects. Sophisticated recent experiments using
a Nb micro-bridge-dc-SQUID, fabricated by e-beam lithography [57], have tested both the static and dynamic properties
of magnetization on single ferromagnetic microstructures.
Experiments on individual Ni, Co and Fe nanowires have
observed non-uniform spin rotation, nucleation and propagation of domain walls, while in nanoparticles smaller than
20 nm coherent spin reversal of the Stoner and Wohlfarth
type has been veried, together with thermally assisted spin
reversal mechanisms consistent with the Arrhenius Eq. (5) of
the NelBrown theory [57].
Mssbauer spectroscopy has also uniquely contributed to
the conrmation of the existence of canted spin structures
at the surface of nanoparticles, by recoding of their Mssbauer spectra in the presence of external magnetic elds
[5863]. Due to selection rules imposed by the principle
of conservation of angular momentum upon absorption of
the -ray photon by the 57 Fe nucleus, the intensity of the
m = 0 (second and fth) absorption lines should be reduced
to zero for iron atoms with internal elds perpendicular to
the direction of propagation of the -ray [64]. Fig. 10 [58]
demonstrates the principle for spectra of ne particles of
-Fe2 O3 in the presence and absence of an applied eld
in the direction of propagation of the -ray. Separation of
the antiferromagnetic sublattices of the spinel structure of
maghemite is observed upon application of the eld and
diminution of the m = 0 absorption lines. Spin canting at
the surface has also been reported for Co covered -Fe2 O3
[65]. While the magnetic moment of the core of the particle
orients in the direction of the magnetic eld, the majority of surface spins are canted perpendicular to the cores
magnetic moment due to surface and nite-size effects.

Figure 9 ZFC and FC magnetization of ferrihydrite core of


reconstituted horse spleen ferritin recorded in an applied eld
of 500 Oe [from Ref. [54]].

444

G.C. Papaefthymiou
with recent theoretical calculations that take into account
surface anisotropy in spherical many-spin magnetic particles [56,81,82], which indicate that the surface introduces
a cubic anisotropy component to the energy landscape.
Such surface spin excitations leading to a sharp decline of
hyperne elds at the surface have also been observed in
antiferromagnetic nanoparticles [83]. Faster spin uctuations at the surface compared to the core have also recently
been reported for 7 nm diameter -Fe2 O3 nanoparticles by
Desautels et al. [84,85]. Other computer simulation studies of surface and nite-size effects in ferromagnetic and
antiferromagnetic nanoparticles based on Monte Carlo techniques have also been published [86,87].

Figure 10 Mssbauer spectra of ultra-ne -Fe2 O3 particles


at 5 K: (a) without and (b) with an applied eld of 50 kOe parallel
to the -ray propagation direction. Separation of the antiferromagnetic sublattices of the spinel structure is observed upon
application of the eld and diminution of the m = 0 absorption
lines [Reprinted by permission from J.M. Coey, Phys. Rev. Lett.
27 (1971) 1140, Ref. [58]].

The presence of spin canting explains the observed lack of


magnetic saturation in small particles of iron oxides, oxyhydroxides and ferrites even in the presence of large external
magnetic elds up to 5 T [6668]. With decreasing size,
spin disorder may extend throughout the particle volume
[65,6872], rather than be conned only within the surface layer. Spin canting in antiferromagnetic small particles
has also been conrmed by polarized [73] and inelastic [74]
neutron scattering, as well as, by ferromagnetic resonance
[75].
Iglesias and Labarta [76] have used Monte Carlo calculations to simulate core and surface contributions to the
magnetization of small -Fe2 O3 nanoparticles in the size
range of 2.58.3 nm diameter. Their results indicate that the
surface contribution to the magnetization increases from
48% for the 8.3 nm diameter particles to 95% for the 2.5 nm
diameter particles. Hysteresis loop simulations in this subsingle-magnetic-domain range shown in Fig. 11 [76] indicate
different loop shapes for surface vs. core magnetizations.
The core shows square hysteresis loops with saturated magnetization, akin to coherent spin reversal; while the surface
exhibits elongated, non-saturated, rounded hysteresis loops
indicative of a multi-step spin reversal mechanism. These
results indicate that the presence of disorder and spin frustration at the surface allows for the easier reversal of
surface spins compared to the core, in agreement with the
enhanced thermal excitations of surface spins observed by
Mssbauer spectroscopy in Fig. 8. Other studies indicate that
the breakage of super-exchange bonds results in the creation of a surface shell within which spin disorder leads to
a spin-glass-like phase at the surface with closely spaced
equilibrium states [77,78]. This surface shell may penetrate
quite far into the interior of the particle and may be uncoupled from the core, behaving as a quasi-independent layer
[79,80]. This introduces greater complexity in the potential
energy landscape at the surface, compared to the simple
double-welled potential of Fig. 2(b), which in turn supports
low-energy spin wave excitations. This is also consistent

Applications: current trends and future


directions
We have reviewed in some detail current fundamental
concepts in nanoparticle magnetism related to spin reversal processes, nite-size and surface effects in isolated
magnetic nanoparticles. Most applications of magnetic
nanoparticles require, however, that particles be coated
for stabilization, biocompatibility and functionalization, be
supported on substrates or embedded within a matrix for
device applications. Tailored particle/support interfacial
interactions can be protably utilized to further enhance
surface effects in matrix embedded nanoparticles or
core/shell nanoarchitectures, by engineering additional
stress or strain at the interface. Depending on the nature of
the shell or matrix the magnetic properties of the particles
can be modied to meet specic requirements. A multitude
of synthetic techniques have been devised in bottom-up
and top-down approaches, including biomimetic synthesis
utilizing ferritin protein templates, for the production of
nanostructured, arrayed or nanocomposite systems for
a plethora of practical applications in diverse areas of
nanotechnology.
Advanced magnetic recording media, for example,
require the production of well ordered two-dimensional
nanoparticle arrays which have traditionally been produced
by nanolithography, molecular beam epitaxy or chemical
vapor deposition techniques and most recently by laser
interference lithography and focused ion-beam milling [1].
The continued demand for scaling to smaller bit and grain
sizes to achieve higher magnetic recording densities is limited by the thermal instability inherent in the NelBrown
spin switching model of Eq. (5), governed by the superparamagnetic energy barrier as compared to thermal energy or
the stability ratio KV/kT. It is predicted that a stability
ratio of 60 will be required for projected 40 Gbits/in2 conditions and 10-year storage time, putting stringent demands
on the need to design high anisotropy nanoparticles [88].
In most conducting media the conduction electron
mean free-path is in the range from a few to several
nanometers rendering magnetic nanoparticles useful as
magneto-resistive sensors and in spin electronics. Giant
magneto-resistance (GMR) [7,8], which occurs in magnetically inhomogeneous media, was rst observed in magnetic
multilayers (Fe/Cr, Co/Cu, Fe/Cu) but has also been
observed in granular metals composed of ferromagnetic
metaliron nanoparticles embedded in non-magnetic con-

Nanoparticle magnetism

445

Figure 11 Monte Carlo simulations for surface (continuous line) and core (dashed line) contributions to the hysteresis loops
of -Fe2 O3 nanoparticles of diameters D = 2.5 nm, T = 10 K (a); D = 2.5 nm, T = 20 K (b); D = 5 nm, T = 10 K (c); D = 5 nm, T = 20 K (d)
[Reprinted from O. Iglesias, A. Labarta, Phys. Rev. B 63 (2001) 184416, Ref. [76]].

ducting Cu matrix [89]. Spin dependent scattering at the


Cu/nanoparticle interface is responsible for the effect. The
magnitude of the GMR effect was found to depend on the orientation of the anisotropy axis, the density and size of the
magnetic nanoparticles. In spintronics, the focus is on the
generation and control of carrier-spin polarization in order
to incorporate spins into existing semiconductor technology
and create new functionalities not achievable by conventional electronics [9,10].
The fast growing area of biomedical applications has produced novel functionalized nanoparticle systems for in vitro
and in vivo applications. These include magnetic separation
of cells, proteins and DNA fragments, bio-imaging, magnetic resonance imaging (MRI) enhancement, targeted drug
delivery and hyperthermia cancer therapy [1316]. In MRI
enhancement, superparamagnetic iron oxide particles are
introduced in order to enhance the relaxation process of
the proton nuclear spins, while in hyperthermia studies targeted cancerous cells are locally heated by radio-frequency
electromagnetic energy absorbed by the magnetic nanoparticles and subsequently dissipated in the surrounding tissue.
One of the challenges in vivo applications still to be overcome is the opsonization or agglomeration of the magnetic
nanoparticles in the ferrouid to form larger clusters. Large
agglomerates become recognizable by the bodys defense
system and are quickly removed from the blood stream.
Understanding particle agglomeration in ferrouids is, thus,
a very active area of research [90].
Many applications of magnetic nanoparticles require platforms in which the particles are in close proximity. In such

platforms the particles can interact strongly magnetically


via two distinct mechanisms: magnetic exchange across particle boundaries in contact, and dipoledipole interactions
[9195]. Then, the magnetic properties of the nanoparticles are strongly inuenced by the presence of interactions
between particles, whereby the particle moments no longer
switch independently [96,97]. Current and future efforts in
nanoparticle magnetism research will concentrate in further elucidating the magnetic properties of ensembles of
interacting magnetic nanoparticles. Dense magnetic particle systems exhibit collective spin dynamics, leading to
spin-glass ordering and a very complex potential energy
landscape of metastable spin states, due to the presence
of competing ferromagnetic and antiferromagnetic interactions known as frustration. Below their spin-glass-freezing
temperature, they exhibit very slow spin relaxation dynamics to thermal equilibrium that depend on temperature and
the length of time they have spent below their glass-freezing
temperature, a dependence known as aging. A great deal
of theoretical effort is currently been expended in understanding non-equilibrium scaling phenomena in such systems
[98102].
New synthetic approaches have produced highly monodispersed core/shell nanoarchitectures with the ability to ne
tune the thickness of the shell, leading to excellent experimental systems for the investigation of dipolar interparticle
magnetic interactions, as a function of interparticle distance
[103,104]. These interactions affect spin relaxation rates
and lead to spin-glass-like formation in the case of strong
interaction energies compared to kT [105].

446
Exchange-bias between ferromagnetic and antiferromagnetic lattices appears to induce drastic changes in the
magnetic properties of ferromagnetic nanoparticles coated
with an antiferromagnetic shell, which promises to overcome limitations due to thermal uctuations of small
magnetic particles. For example Skumryev et al. [106]
report that Co nanoparticles embedded into a paramagnetic
Al2 O3 matrix exhibit a blocking temperature of 10 K, while
when surrounded by an antiferromagnetic CoO matrix, they
show a blocking temperature of 290 K, a 30-fold increase.
Interfacial exchange interactions between the ferromagnetic Co particles with the antiferromagnetic CoO lattice
are believed to be responsible for the observed moment
stabilization, an interpretation that is supported by the
observation of a strong exchange-bias effect manifested in
shifted hysteresis loops. It is expected that future directions in nanoparticle magnetism [107] will seek to utilize
such interactions in order to overcome thermal uctuations
in the quest to beat the superparamagnetic limit in the continued miniaturization of magnetic information storage and
the design of high density magnetic recording media.

Acknowledgements
The author thanks the National Science Foundation for support at Villanova University under contract DMR-0604049 and
the European Commission for support at NCSR Demokritos
under the Marie Curie Program.

References
[1] R. Skomski, J. Phys.: Condens. Matter 15 (2003) R841.
[2] J.L. Dormann, D. Fiopranni, E. Tronc, Adv. Chem. Phys. XCVIII
(1997) 283.
[3] M.L. Plumer, J.v. Ek, D. Weller (Eds.), The Physics of UltrahighDensity Magnetic Recording Series: Springer Series in Surface
Sciences, vol. 41, 2001.
[4] T. Hayashi, S. Hirono, M. Tomita, S. Umemura, Nature 381
(1996) 772.
[5] R.L. Costock, Introduction to Magnetism and Magnetic Recording, Wiley, New York, 1999.
[6] R. Skomski, J.M.D. Coey, Permanent Magnetism, Institute of
Physics, Bristol, 1999.
[7] M.N. Baibich, J.M. Broto, A. Fert, F. Nguyen Van Dau, F.
Petroff, P. Eitenne, et al., Phys. Rev. Lett. 21 (1988) 2472.
[8] G. Binasch, P. Grnberg, F. Saurenbach, W. Zinn, Phys. Rev. B
39 (1989) 4828.
uti
[9] I. Z
c, J. Fabia, S. Das Sarma, Rev. Mod. Phys. 76 (2004) 323.
[10] S.A. Wolf, D.D. Awschalom, R.A. Buhrman, J.M. Daughton, S.
von Molnr, M.L. Roukes, et al., Science 294 (2001) 1488.
[11] R.P. Blakemore, Annu. Rev. Microbiol. 36 (1982) 217.
[12] D.A. Bazylinski, R.B. Frankel, H.W. Jannasch, Nature 334
(1988) 518.
[13] W.C.W. Chan, Bio-applications of Nanoparticles, Springer,
2007.
[14] C.C. Berry, A.S.G. Curtis, J. Phys. D: Appl. Phys. 36 (2003)
R198.
[15] Y.-W. Jun, J.-W. Seo, J. Cheon, Accounts Chem. Res. 41 (2008)
179.
[16] J.R. McCathy, R. Weissleder, Adv. Drug Deliv. Rev. 60 (2008)
1241.
[17] K. Nagasaka, J. Magn. Magn. Mater. 321 (2009) 508.
[18] A.H. Morrish, The Physical Principles of Magnetism, Wiley,
New York, 1965 (Reprinted IEEE Press, New York, 2001).

G.C. Papaefthymiou
[19] R.C. OHandley, Modern Magnetic Materials, Principles and
Applications, John Wiley and Sons, New York, 2000.
[20] W. Heisenberg, Z. Phys. 49 (1928) 619.
[21] F. Bloch, Z. Phys. 57 (1929) 545.
[22] J.C. Slater, Phys. Rev. 49 (1936) 537.
[23] G. Burns, Solid State Physics, Academic Press, New York, 1985.
[24] P. Bruno, Phys. Rev. B 39 (1989) 865.
[25] C. Kittel, Rev. Mod. Phys. 21 (1949) 541.
[26] E.C. Stoner, E.P. Wohlfarth, Trans. R. Soc. Lond. A 240 (1948)
599.
[27] E.H. Frei, S. Shtrikman, D. Treves, Phys. Rev. 106 (1957) 446.
[28] I.S. Jacobs, C.P. Bean, Phys. Rev. 100 (1955) 1060.
[29] H.B. Braun, Phys. Rev. Lett. 71 (1993) 3557.
[30] U. Nowak, R.W. Chantrell, E.C. Kennedy, Phys. Rev. Lett. 84
(2000) 163.
[31] F.E. Luborsky, J. Appl. Phys. 32 (1961) 171S (Supplement to
volume).
[32] L. Nel, Ann. Geophys. 5 (1949) 99;
L. Nel, J. Phys. Radium 20 (1959) 215.
[33] W.F. Brown Jr., Phys. Rev. 130 (1963) 1677 (Reprinted in J.
Appl. Phys. 39 (1968) 993).
[34] A. Aharoni, Phys. Rev. A 135 (1964) 447.
[35] D.P.E. Dickson, N.M.K. Reid, C. Hunt, H.D. Williams, M. ElHilo, K. OGrady, J. Magn. Magn. Mater. 125 (1993) 345.
[36] D.D. Awschalom, J.F. Smyth, G. Grinstein, D.P. DiVincenzo, D.
Loss, Phys. Rev. Lett. 68 (1992) 3092.
[37] S. Gider, D.D. Awschalom, T. Douglas, S. Mann, M. Chaparala,
Science 268 (1995) 77.
[38] E.M. Chudnovsky, L. Gunther, Phys. Rev. Lett. 60 (1998) 661.
[39] H. Gleiter, Acta. Mater. 48 (2000) 1.
[40] M. Respaud, J.M. Broto, H. Rakoto, A.R. Fert, L. Thomas, B.
Barbara, et al., Phys. Rev. B 57 (1998) 2925.
[41] P. Gambardella, S. Rusponi, M. Veronese, S.S. Dhesi, C. Grazioli, A. Dallmeyer, et al., Science 300 (2003) 1130.
[42] N. Prez, P. Guardia, A.G. Roca, M.P. Morales, C.J. Serna, O.
Iglesias, et al., Nanotechnology 19 (2008) 475704.
[43] F. Bdker, S. Mrup, S. Lideroth, Phys. Rev. Lett. 72 (1994)
282.
[44] F. Gazeau, J.C. Bacri, F. Gendron, R. Perzynski, Y.L. Stepanov,
E. Dubois, J. Magn. Magn. Mater. 186 (1998) 175.
[45] I. Hrianca, C. Caizer, Z. Schlett, J. Appl. Phys. 92 (2002) 2125.
[46] L. Zhang, G.C. Papaefthymiou, J.Y. Ying, J. Appl. Phys. 81
(1997) 6892.
[47] B.H. Sohn, R.E. Cohen, G.C. Papaefthymiou, J. Magn. Magn.
Mater. 182 (1997) 216.
[48] L. Zhang, G.C. Papaefthymiou, J.Y. Ying, J. Phys. Chem. B 105
(2001) 7414.
[49] S.R. Ahmed, S.B. Ogal, G.C. Papaefthymiou, R. Ramesh, P.
Konas, Appl. Phys. Lett. 80 (2002) 1616.
[50] T.-J. Park, G.C. Papaefthymiou, A.J. Viescas, A.R. Moodenbaugh, S.S. Wong, Nano Lett. 7 (2007) 766.
[51] N.D. Chasteen, P.M. Harrison, J. Struct. Biol. 126 (1999) 182.
[52] S. Mann, Biomineralization: Principles and Concepts in
Bioinorganic Materials Chemistry, Oxford University Press,
New York, 2001.
[53] R.A. Brooks, J. Vymazal, R.B. Goldfarb, J.W. Bulte, P. Aisen,
Magn. Res. Med. 40 (1998) 227.
[54] G.C. Papaefthymiou, A.J. Viescas, E. Devlin, A. Simopoulos,
Mater. Res. Symp. Proc. 1056E (2007) HH03HH27.
[55] S. Mrup, H. Topse, Appl. Phys. 11 (1976) 63.
[56] H. Kachkachi, E. Bonet, Phys. Rev. B 73 (2006) 224402.
[57] W. Wernsdorfer, E. Bonet Orozco, B. Barbara, K. Hasselbach, A. Benoit, D. Mailly, et al., J. Appl. Phys. 81 (1997)
5543.
[58] J.M. Coey, Phys. Rev. Lett. 27 (1971) 1140.
[59] F.T. Parker, A.E. Berkowitz, Phys. Rev. B 44 (1991) 7437.
[60] A. Millan, A. Urtizberea, N.J.O. Silva, F. Palacio, V.S. Amaral,
E. Snoeck, V. Serin, J. Magn. Magn. Mater. 312 (2007) L5.

Nanoparticle magnetism
[61] O. Helgason, H.K. Rasmussen, S. Mrup, J. Magn. Magn. Mater.
302 (2006) 413.
[62] G.J. Long, F. Grandjean (Eds.), Mssbauer Spectroscopy
Applied to Magnetism and Materials Science, Plenum Press,
New York, 1993.
[63] A.H. Morrish, K. Haneda, J. Appl. Phys. 52 (1981) 2496.
[64] N.N. Greenwood, T.C. Gibb, Mssbauer Spectroscopy, Chapman Hall, London, 1971.
[65] Q.A. Pankhurst, R.J. Pollard, Phys. Rev. Lett. 67 (1991) 248.
[66] A.H. Morrish, K. Haneda, J. Magn. Magn. Mater. 35 (1983) 105.
[67] J.M. Coey, Can. J. Phys. 65 (1987) 1233.
[68] E. De Grave, R.E. Vandenberghe, L.H. Bowe, in: J. Stanek,
A.T. Pedziwiatr (Eds.), Condensed Matter Studies by Nuclear
Methods, World Scientic, Singapore, 1990.
[69] F. Parker, M.W. Foster, D.T. Margulies, A.E. Berkowitz, Phys.
Rev. B 47 (1993) 7885.
[70] S. Linderoth, P.V. Hendriksen, F. Bodker, S. Wells, K. Davis,
S.W. Charles, S. Mrup, J. Appl. Phys. 75 (1994) 6583.
[71] R.H. Kodama, S.A. Makhlouf, Berkowitz, Phys. Rev. Lett. 79
(1997) 5461.
[72] M.P. Morales, C.J. Serna, F. Bdker, S. Mrup, J. Phys. Condens. Matter 9 (1997) 5461.
[73] D. Lin, A.C. Nunes, C.F. Majkrzak, A.E. Berkowitz, J. Magn.
Magn. Mater. 145 (1995) 343.
[74] F. Gazeau, E. Dubois, M. Hennion, R. Perzynski, Y.L. Raikher,
J. Magn. Magn. Mater. 40 (1997) 575.
[75] F. Gazeau, J.C. Bacri, F. Gendron, R. Perzynski, Y.L. Raikher,
V.I. Stepanov, et al., J. Magn. Magn. Mater. 186 (1998) 175.
[76] O. Iglesias, A. Labarta, Phys. Rev. B 63 (2001) 184416.
[77] P.V. Hendriksen, S. Linderoth, P.-A. Lindgrd, Phys. Rev. B 48
(1993) 7259.
[78] R.H. Kodama, A.E. Berkowitz, E.J. McNiff Jr., S. Foner, Phys.
Rev. Lett. 77 (1996) 394.
[79] B. Martinez, X. Obradors, L. Barcells, A. Rouanet, C. Monty,
Phys. Rev. Lett. 80 (1997) 181.
[80] G.H. Kachkachi, D.A. Garanin, Phys. A: Stat. Mech. Appl. 300
(2001) 487.
[81] H. Kachkachi, D.A. Garanin, in: D. Fiorani (Ed.), Surface
Effects in Magnetic Nanoparticles, Book Series in Nanostructure Science and Technology, Springer, US, 2005, p. 75;
H. Kachkachi, D.A. Garanin, arXiv:cond-mat/0310694, v129,
October 2003, Fig. 14.
[82] R. Yanes, O. Chubykalo-Fesenko, H. Kachkachi, D.A. Garanin,
R. Evans, R.W. Chantrell, Phys. Rev. B 76 (2007) 064416.
[83] F. Bou-Abdallah, E. Carney, N.D. Chasteen, P. Arosio, A.J.
Viescas, G.C. Papaefthymiou, Biophys. Chem. 139 (2007) 114.
[84] R.D. Desautels, E. Skoropala, J. van Lierop, J. Appl. Phys. 10
(2008) 07D512.
[85] T.N. Shendruk, R.D. Desautels, B.W. Southern, J. van Lierop,
Nanotechnology 18 (2007) 455704.
[86] K.N. Trohidou, J.A. Blackman, Phys. Rev. B 41 (1990) 9345.
[87] X. Zianni, K.N. Trohidou, J. Appl. Phys. 85 (1999) 1050.
[88] D. Weller, A. Moser, IEEE Trans. Magn. 35 (1999) 4423.
[89] J.Q. Xiao, J.S. Jiang, C.L. Chien, Phys. Rev. Lett. 68 (1992).

447
[90] T. Kruse, H.-G. Krauthuser, A. Spanoudaki, R. Pelster, Phys.
Rev. B 67 (2003) 094206.
[91] S. Mrup, C.A. Oxborrow, P.V. Hendriksen, M.S. Pedersen, M.
Hanson, C. Johansson, J. Magn. Magn. Mater. 140144 (1995)
409.
[92] S. Mrup, D.E. Madsen, C. Frandsen, C.R.H. Bahl, M.F. Hansen,
J. Phys.: Condens. Matter 19 (2007) 213202.
[93] S. Mrup, F. Bodker, P.V. Hendriksen, S. Linderoth, Phys. Rev.
B 52 (1995) 287.
[94] S. Mrup, Europhys. Lett. 28 (1994) 671.
[95] L. Rebbouh, R.P. Hermann, F. Grandjean, T. Hyeon, K. An, A.
Amato, G.J. Long, Phys. Rev. B 76 (2007) 174422.
[96] X. Chen, S. Sahoo, W. Kleemann, S. Cardoso, P.P. Freitas, Phys.
Rev. B 70 (2004) 172411.
[97] P. Allia, M. Coisson, P. Tiberto, F. Vinai, M. Knobel, M.A. Novak,
et al., Phys. Rev. B 64 (2001) 144420.
[98] T. Jonsson, P. Nordblad, P. Svedlindh, Phys. Rev. B 57 (1998)
497.
[99] H. Eissfeller, W. Kinzel, J. Phys. A: Math. Gen. 25 (1992) 1473.
[100] D.S. Fisher, H. Sompolinsky, Phys. Rev. Lett. 54 (1985) 1063.
[101] S.A. Majetich, M. Sachan, J. Phys. D: Appl. Phys. 39 (2006)
R407.
[102] V. Dupuis, F. Bert, J.-P. Bouchaud, J. Hammann, F. Ladieu, D.
Parker, et al., Pramana 64 (2005) 1109.
[103] T. Hyeon, S.S. Lee, J. Park, Y. Chung, H.B. Na, J. Am. Chem.
Soc. 123 (2001) 12798.
[104] D.K. Yi, S.S. Lee, G.C. Papaefthymiou, J.Y. Ying, Chem. Mater.
18 (2006) 614.
[105] G.C. Papaefthymiou, E. Devlin, A. Simopoulos, D.K. Yi, S.S.
Lee, J.Y. Ying, et al., Phys. Rev. B 80 (2009) 024406.
[106] V. Skumryev, S. Stoyanov, Y. Zhang, G. Hadjipanayis, D.
Givord, J. Nogus, Nature 423 (2003) 850.
[107] O. Iglesias, A. Labarta, K. Battle, J. Nanosci. Nanotechnol. 8
(2008) 2761.
G.C. Papaefthymiou is a professor of physics
at Villanova University in Villanova, PA, USA
and a visiting research professor at NCSR
Demokritos in Athens Greece, where she
has recently held the Marie Curie Chair of
Excellence within the Institute of Materials
Science. Her research interests are in condensed matter physics encompassing areas
in: (1) cluster science and the transition
from molecular to bulk behavior with increasing cluster size; (2) nanoscale magnetism,
fundamental studies and applications to nanotechnology and nanomedicine; (3) multiferroic materials; (4) iron containing proteins.
She uses Mssbauer spectroscopy to probe the electronic and magnetic structure of molecular and solid-state materials (amorphous or
crystalline) in combination with SQUID (superconducting quantum
interference device) magnetometry. She is the author or co-author
of over 150 publications that include peer reviewed articles, book
chapters, conference proceedings and opinion pieces.

Você também pode gostar