Sei sulla pagina 1di 9

Magnetism of 3d transition metal

nanoparticles on surfaces probed with synchrotron


radiation from ensembles towards individual objects
Joachim Bansmann
*
,1
, Armin Kleibert
2,4
, Mathias Getzlaff
3
, Arantxa Fraile Rodrguez
4
, Frithjof Nolting
4
,
Christine Boeglin
5
, and Karl-Heinz Meiwes-Broer
2
1
Institut fur Oberachenchemie und Katalyse, Universitat Ulm, Albert-Einstein-Allee 47, 89081 Ulm, Germany
2
Institut fur Physik, Universitat Rostock, Universitatsplatz 3, 18051 Rostock, Germany
3
Institut fur Angewandte Physik, Universitat Dusseldorf, Universitatsstr. 1, 40225 Dusseldorf, Germany
4
Swiss Light Source, Paul Scherrer Institut, 5232 Villigen, Switzerland
5
Institut de Physique et de Chimie de Strasbourg, Universite de Strasbourg, CNRS UMR 7504, 23 rue de Loess, 67034 Strasbourg,
France
Received 5 November 2009, revised 7 December 2009, accepted 11 December 2009
Published online 18 January 2010
PACS 73.22.f, 75.30.Gw, 75.75.a, 75.25.z, 78.70.Dm, 81.07.b
*
Corresponding author: e-mail joachim.bansmann@uni-ulm.de, Phone: 49 731 50 25469, Fax: 49 731 50 25452
Mass-ltered Fe, Co and FeCo nanoparticles in the size regime
from 7 to 25 nm have been deposited under soft-landing
conditions onto ferromagnetic lms, non-magnetic surfaces as
well as embedded into Al matrices. In situ X-ray magnetic
circular dichroism (XMCD) measurements reveal a ferromag-
netic behaviour of FeCo nanoparticles (size: 10 nm) on Si-
substrates at room temperature whereas the respective Co
nanoparticles are superparamagnetic. Besides measurements
on ensembles of nanoparticles, we have also carried out in situ
measurements on individual Fe nanoparticles using X-ray
photoemission electron microscopy at the Fe L
3,2
edges. Fe
nanoparticles on Co lms show a magnetic contrast depending
on the direction of the underlying poly-crystalline Co domains.
This technique also allows to record XMCD spectra on
individual nanoparticles.
Spectroscopy and spectro-microscopy on magnetic particles.
2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
1 Introduction In the last decade, ferromagnetic
nanoparticles have attracted much interest due to their
possible applications in data storage media, chemistry,
biotechnology and medicine (see, e.g. Refs. [13]).
Nanoparticles with a homogeneous size distribution have
the advantage of similar magnetic properties and are thus
required for most technical applications. Besides chemistry-
based preparations routes which are more appropriate for
large-area technical applications (e.g. Refs. [46]), physics-
based methods are often used for more fundamental research
and characterisation of their magnetic properties. Here, we
report on investigations on the magnetic properties (such as
the magnetic anisotropy energy) of well-dened mass-
ltered 3d transition metal (TM) nanoparticles in a size
regime from 7 to about 25 nm prepared by deposition of
preformed particles on surfaces. X-ray magnetic circular
dichroism (XMCD) is a well-established technique for
analysing the magnetism, particularly the magnetic spin and
orbital moments using tuneable, circularly polarised syn-
chrotron radiation at third generation synchrotron facilities.
The method is capable of investigating magnetic nanopar-
ticles and clusters down to a size of only a fewatoms [7] with
Phys. Status Solidi B 247, No. 5, 11521160 (2010) / DOI 10.1002/pssb.200945516
p s s
basic solid state physics
b
s
t
a
t
u
s
s
o
l
i
d
i
www.pss-b.com
p
h
y
s
i
c
a
2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
a high accuracy (for an overview, see e.g. Refs. [811]). As a
matter of fact, most of the XMCD measurements average
over a certain area on the sample, usually the area that is
illuminated by the incoming radiation. In third generation
storage rings equipped with undulator beamlines these areas
are about 100 mm100 mm or larger. Thus, hundred
thousands of individual nanoparticles are excited which
might differ in their orientation with respect to the surface,
their magnetic anisotropy axes and their shapes. However,
even at microfocus beamlines with spots down to only a few
nanometer, one averages over a large number of particles. As
a consequence of these variations, the magnetic properties of
individual nanoparticles might differ signicantly from the
average behaviour of all particles being detected in XMCD
spectroscopy. Thus, imaging the magnetic properties of
individual particles and, moreover, to combine this infor-
mation with structural information on the particle being
investigated is a highly interesting goal for a detailed
understanding of magnetic phenomena of nanoscaled
objects. It has already been demonstrated that certain
magnetic properties of individual nanoparticles can be
investigated; one possibility has been shown by Jamet
et al. [12] who used a micro-SQUID set-up [13] to measure
the magnetic anisotropy energy of a single Co nano-
particle with about 1000 atoms. Further attempts have been
made by Lorentz-microscopy [14, 15], revealing magnetic
domains in Fe nanoparticles as small as 25 nm or switching
events in 10 nm sized FeCo nanoparticles. Furthermore,
chiral TEM [16] may allow in the near future to analyse
the magnetic spin and orbital moments of magnetic
nanoparticles by combining high-resolution transmission
electron microscopy (HR-TEM) with a magnetic circular
dichroism technique.
In our approach, we use photoemission electron
microscopy (PEEM) in combination with circularly
polarised synchrotron radiation in order to analyse the
magnetic properties of individual Fe nanoparticles on a
polycrystalline Co lm by resonantly exciting photo-
electrons at the Fe L
3,2
edges. The Co lms allow to
magnetically saturate the Fe nanoparticles [17] without the
need of an external magnetic eld which would interfere
with PEEM.
2 Experimental The experiments have been carried
out at the electron storage rings BESSYin Berlin (Germany)
and the Swiss Light Source SLS (Paul Scherrer Institut) in
Villigen (Switzerland). A UHV compatible and transport-
able cluster source, which is described in the following
section, serves for in situ experiments.
2.1 Cluster source The arc cluster ion source (ACIS)
[18, 19] has been designed to produce mass-ltered metal
nanoparticles in a size regime from 4 to 25 nm. The typical
size distribution has a width of 15% of the mean particle
diameter. The cluster source itself, as shown in Fig. 1,
consists of three parts: the hollowcathode, where the clusters
are created by an arc erosion in an inert seeding gas, a
pressure reducing part and a mass-ltering unit with an
electrostatic quadrupole deector. In the rst part, the
material for the cluster production is eroded from the metal
cathode (in our case Fe, Co or FeCo of high purity), the
condensation of the particles starts in the seeding gas (argon)
at pressures around 2040 mbar and is nally nished in the
expansion channel.
The particles, more than half of them being in a charged
state, leave the aggregation part of the cluster source after a
supersonic expansion. The rare gas required for the seeding
process is pumped off in the next unit consisting of two
skimmers, and the remaining molecular beam enters the
mass-ltering unit. The electrostatic quadrupole selects only
charged particles which are deected and may pass the exit
slit (at the ux monitor) depending on their kinetic energy.
Since all particles have nearly identical velocities after the
supersonic expansion givenby the velocity of the seeding gas
(in our case of pure Ar around 500 ms
1
), the kinetic energy
only depends on the mass of the particle when they are single
charged (which is usually the case). Thus, the quadrupole
deector acts as a mass selector with a moderate resolution,
but a quite high transmission efciency. The continuous
working mode (no pulsed units such as lasers or pulse valves)
moreover contributes to a high cluster deposition rate and
enables the application of static electric elds for the mass
selection. The kinetic energies of the particles before
deposition are in the order of 0.1 eV per atom and thus
below the limit where fragmentation might occur [20].
Phys. Status Solidi B 247, No. 5 (2010) 1153
Original
Paper
Figure 1 (online colour at: www.
pss-b.com) Schematic drawing of the
arc cluster ionsource withits functional
units.
www.pss-b.com 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
2.2 Experimental set-up at electron storage
rings and XMCD technique At BESSY, the cluster
source was attached to several beamlines (e.g. UE46-PGM1,
UE56/1-PGM and PM3) providing circularly polarised
radiation (6501000 eV). Investigations regarding high
magnetic elds (up to 5 T) with a superconducting 7 T
magnet (IPCMS Strasbourg, France) were carried out at the
beamline UE46-PGM1. The PEEM measurements on the
magnetic behaviour of individual nanoparticles were
performed in ultra-high vacuum at the surface/interface
microscopy (SIM) beamline [21] of the Swiss Light Source
using an Elmitec LEEM III setup equipped with an energy
analyser. The PEEM provides a magnied image of the
emitted secondary photoelectrons, thus generating a spatial
map of the absorption with a probing depth of a few
nanometers [22]. These measurements were performed at
room temperature (RT) without external magnetic elds.
The magnetic behaviour of the 3d (TM) nanoparticles
deposited on surfaces has been analysed using XMCD in
photoabsorption (for an overviewand applications, see Refs.
[8, 23, 24]) which requires tuneable soft X-rays with circular
polarisation at the L
3,2
absorption edges of the ferromagnetic
3d materials Fe, Co and Ni. The XMCD sum rules [2426]
allow to determine the magnetic orbital m
orb
and spin
moments m
spin
. The XMCD signal is recorded by measuring
the photoabsorption, in spectroscopy via the total electron
yield, in microscopy via the partial electron yield. Technical
details regarding the data analysis are described in earlier
publications of our group [19, 27, 28].
3 Results and discussion We focus on the magnetic
properties of 3d metal nanoparticles (Co, FeCo and Fe) on
magnetic and non-magnetic surfaces as well as particles
embedded in non-magnetic matrices. The correlation of the
shape and structure of deposited, mass-ltered 3d TM
nanoparticles with their magnetic properties investigated by
XMCD has already been discussed in earlier publications
(see, e.g. Refs. [19, 29, 30]).
3.1 Magnetic spin and orbital moments of Co
nanoparticles on surfaces For lms [31] and nanopar-
ticles [32], it has been pointed out, that both, the spin and
especially the orbital moments, are usually signicantly
underestimated when the XMCD sum rules are applied
to spectra obtained by TEY detection. In a preceding
publication [30], we have in detail analysed the TEY
detection in XMCD on supported nanoparticles. First of all,
due to the short mean free path of the excited electrons in Fe
(1.7 A

) and Co (2.2 A

), the XMCD signal mainly stems


from the outer layers of the nanoparticles whereas the inner
part does not contribute to the magnetic signal. In order to
correct for these underestimated magnetic moments (m
TEY
),
we have developed an algorithm for spherically shaped Fe
nanoparticles that enables us to recalculate the magnetic spin
and orbital moments. It has been shown that correction
factors can be obtained by numerical simulations [30]. The
correction factors m
TEY
/m
m
(the superscript m denotes
moments calculated from reference data, for details see
Ref. [30]) for the orbital and spin moments of Co
nanoparticles are shown in Fig. 2 as a function of the particle
diameter. For the spin moment, the corrections factors are
close to 1 and thus, the deviations that have to be considered,
are small. The correction factors for the orbital moments are
higher and reach a value of m
TEY
orb
/m
m
orb
0.75 for Co
nanoparticles with a size close to 10 nm. Compared to Fe,
the correction factors are generally smaller due to the lower
absorption coefcient of Co.
As an overview, the corrected values for the ratio of the
magnetic orbital and spin moments m
orb
/m
spin
as well as the
spin moments m
spin
of 7.6 nm Co nanoparticles deposited
onto Ni(111) and Fe(110) lms on W(110) and Au(111) are
given in Table 1. First of all, the experimentally determined
values differ signicantly depending on the surfaces; the spin
moments by about 10% (between 1.46 and 1.66 m
B
) and
the orbital moments even more (cf. Table 1). It should be
mentioned that the measurements are performed under
slightly different conditions. On ferromagnetic Ni(111) and
1154 J. Bansmann et al.: Magnetism of 3d TM nanoparticles on surfaces
p
h
y
s
i
c
a
s s p
s
t
a
t
u
s
s
o
l
i
d
i
b
Figure 2 (online colour at: www.pss-b.com) Correction factors
for the experimentally determined magnetic spin m
TEY
spin
(solid
black dots) and orbital moments m
TEY
orb
(open red dots) of spherical
Co nanoparticles with diameter D on surfaces. The dependence
originates from the limited escape depth of electrons in the TEY
mode. Lines serve as a guide for the eyes.
Table 1 Ratio of the magnetic orbital and spin moments m
orb
/
m
spin
as well as the spin moments m
spin
of Co nanoparticles (size:
7.6 nm) on different supports (room temperature) after correcting
for saturation effects in the TEY detection mode.
m
orb
/m
spin
m
spin
/(m
B
/atom)
Co/Ni(111) [27, 19] 0.090 0.02 1.54 0.06
Co/Fe(110) [27] 0.132 0.008 1.77 0.03
Co/Au(111) (08) [28] 0.29 0.03 1.46 0.01
Co/Au(111) (408) [28] 0.162 0.03 1.66 0.015
bulk hcp Co [24] 0.099 1.55
bulk fcc Co [33] 0.078 1.72
2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.pss-b.com
Fe(110) lms on W(110), the Co nanoparticles were
magnetised by direct exchange interaction with the under-
lying lm (without externally applied magnetic elds) along
the easy magnetisation direction of these lms (i.e. for
Ni(111)/W(110) along the W[001] direction, for Fe(110)/
W(110) along the W[110] direction of the underlying
tungsten single crystal, for details regarding these lms,
see Ref. [27]. Moreover, the saturation magnetisation of Fe
and Ni lms (thickness: 13 ML) are quite different
(corresponding bulk values Fe: 2.0 T, Ni: 0.65 T) and
thus, the Co nanoparticles might not be completely
magnetised due to the much lower exchange interaction of
Ni. Additionally, the different symmetries of the lms
(quasi-hexagonal bcc(110) and fcc(111) symmetry) and the
different in-plane magnetisation directions of the lms might
also affect the magnetic properties of the deposited Co
nanoparticles. In the case of non-magnetic substrates, the Co
nanoparticles were magnetised with an external magnetic
eld (either by a superconducting magnet at B1.5 T, as in
the case of Au(111), or via a small electromagnet with
magnetic elds below 100 mT at the position of the sample,
cf. Fig. 3). As already discussed in detail in Ref. [28], the
magnetic orbital moments of Co nanoparticles on Au(111)
strongly change with the temperature and the angle of
incidence, however, the ratio of orbital to spin moment is
always strongly enhanced with respect to the corresponding
bulk value. The smaller magnetic spin moment in normal
incidence and magnetisation perpendicular to the surface
(08) may be (partly) caused by the magnetic dipole term m
T
[3436] which automatically enters the corresponding
equation for m
spin
when using XMCD. In the case of Co
nanoparticles and Co lms on Au(111), the magnetic dipole
term does not vanish and has a signicant inuence when
determining the spin moment [37], it especially lowers the
value in normal incidence and a magnetisation perpendicular
to the surface. For uncapped Co nanoparticles on Si, we were
not able to determine the spin moment due to the limited
external magnetic eld (cf. Fig. 3).
In conclusion, we would like to point out, that the
magnetic orbital and spin moments of deposited Co
nanoparticles strongly depend on the surface where depo-
sition took place. The orbital moment in Co nanoparticles
with a size of 9.6 nm is generally larger than the
corresponding Co bulk value, not least due to the enhanced
magnetic orbital moments at ferromagnetic surfaces due to
the reduced symmetry.
3.2 Magnetic behaviour of Co and FeCo alloy
particles on silicon surfaces with a native oxide In
the preceding section, we have reported on the orbital and
spin moments of Co nanoparticles on (mainly ferromagnetic)
surfaces and their deviation from the respective bulk values.
In the rst part of this subsection, we will compare the
magnetic anisotropy energy of Co and FeCo alloy nano-
particles, in the second part, the magnetic behaviour of Co
nanoparticles on surfaces and in matrices. Fe and Co
nanoparticles with a diameter of about 10 nm or less are
expected to show superparamagnetism at RT, the respective
blocking temperature T
b
is about 17 K (bcc Fe) and 160 K
(fcc Co), when assuming bulk-like magnetic anisotropy
values. The blocking temperature can be estimated from
NeelBrown model, cf. Eq. (1) (see, e.g. Wernsdorfer et al.
[38]), which relates the uni-axial magnetic anisotropy energy
E
MAE
K
u
V (V volume, K
u
uni-axial magnetic anisotropy
constant) of a particle to the thermal energy E
th
k
B
T (T
temperature, k
B
Boltzmann constant):
1
t
0
f
0
exp
K
u
V
k
B
T

: (1)
In this Arrhenius law, the frequency f
0
is the attempt
frequency (for reversing the magnetisation) with a typical
value of f
0
10
9
s
1
and t
0
the relaxation time. A system is
often considered as blocked when t
0
is larger than about
100 s; details are given in, e.g. Refs. [2, 3840]. Equation (1)
can also be used to estimate the magnetic anisotropy required
of a nanoparticle of a given size to be ferromagnetic at RT.
Hcp Co nanoparticles with a diameter 9.6 nm should thus
have a magnetic anisotropy constant of K13 meVper atom
and would be ferromagnetic at RT. In case of a cubic
anisotropy with a constant K(as in fcc Co) one can replace K
u
in Eq. (1) by (K/4) [39]. This leads to remarkably different
blocking temperatures (or the required magnetic anisotropy
constants for observing ferromagnetic behaviour at RT)
between fcc (cubic anisotropy) and hcp (uni-axial aniso-
tropy) Co nanoparticles of the same size. In case of a cubic
anisotropy, at least 51 meV per atom is required to show
ferromagnetism at RT, which is clearly above the bulk value
for fcc Co. For an overview, the respective magnetic
anisotropy values for bulk materials are given in Table 2.
In the following we will focus on the magnetic behaviour
of Co and FeCo nanoparticles of identical size deposited on
silicon wafers with native oxide surfaces. The black dots in
Fig. 3 represent the relative magnetisation of Co nanopar-
ticles with a size of 9.6 nm, the particles show a
Phys. Status Solidi B 247, No. 5 (2010) 1155
Original
Paper
25 0 50 75 100
-1,0
-0,5
0,0
0,5
1,0
applied magnetic field mT /
-50 -25 25 0 50
-1,0
-0,5
0,0
0,5
1,0
M
/
M
S
applied magnetic field mT /
n
o
r
m
a
l
i
s
e
d
X
M
C
D
s
i
g
n
a
l
(
a
r
b
.
u
n
i
t
s
)
Figure 3 Normalised XMCD signals from a hysteresis curve of
9.6 nm Co nanoparticles deposited onto a Si wafer. The solid line
represents a Langevin t for superparamagnetic Co nanoparticles
(m
Co
1:5 m
B
) of this size at 300 K. The inset shows a hysteresis
curve recorded from9.6 nmFeCo nanoparticle also deposited ontoa
Si wafer [46].
www.pss-b.com 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
superparamagnetic behaviour at RT. For comparison, a
Langevin-type magnetisation curve is plotted in the same
viewgraph (solid line), the curve is calculated for Co
nanoparticles of this size (50,000 atoms per particles, spin
moment m
Co
1:5 m
B
per atom) at RT. The fact that
Co particles of this size are superparamagnetic at RT can
be ascribed to their fcc structure [19, 45]. As discussed
above, the cubic symmetry results in much lower blocking
temperatures when compared to hcp particles. Interestingly,
FeCo (again 9.6 nm) nanoparticles deposited on a silicon
wafer showed a ferromagnetic hysteresis curve (cf. inset in
Fig. 3) with a magnetic remanence M
r
of about M
r
=M
s
0:2
(M
s
saturation magnetisation) and a coercive eld H
c
of
about 8 mT.
The presence of a nite coercive eld and a higher
remanent magnetisations in FeCo alloy nanoparticles
compared to Co nanoparticles of the same size (and
deposited under the same conditions onto identical sub-
strates) is unexpected. FeCo alloys are usually soft magnetic
materials with very high saturation magnetisation elds and
very low coercive elds [47]. Thus, the respective magnetic
anisotropy energies should be very small, at least much
smaller than those of Co. Recent calculations [48], however,
showed that tetragonally distorted FeCo alloys (lms and
particles) might exhibit (with a Co concentration between 50
and 70%) both, high magnetic saturation elds (around
2.2 T) and very high magnetic anisotropies (up to a factor of
20 larger than those of Co, cf. Table 2). Experiments by
Andersson et al. [44] and Winkelmann et al. [49] have proven
the general prediction of strongly enhanced magnetic
anisotropies in FeCo alloys, e.g. Andersson et al. observed
magnetic anisotropies of 150 meV per atom and saturation
elds of 2.2 T at RT. For comparison, the bulk magnetic
anisotropy constants of Fe and Co are also given in Table 2.
Our FeCo alloy particles consist of about 50% Fe and
48% Co, the remaining parts are mainly vanadium (1.8%)
and nickel (0.3%), for details see also Refs. [17, 46]. The
small contributions fromthe 3d metal (Ni and V) could not be
detected in XAS and are not expected to signicantly
inuence the magnetic anisotropy of the respective nano-
particles. Based on the calculations from Burkert et al. [48],
extremely high values for the magnetic anisotropy are not
expected, since the Co content is only slightly smaller than
the Fe content. Nevertheless, strongly enhanced magnetic
anisotropy values are also predicted for materials consisting
of equal amount of Fe and Co, provided the lattice is
tetragonally distorted with a c/a axes ratio of more than 1.05.
However, our data clearly indicate that FeCo nanoparticles
deposited onto a Si-wafer show strongly enhanced magnetic
anisotropies compared to bulk FeCo materials. For ferro-
magnetic single domain particles without interaction, the
respective coercive eld H
c
is related to the magnetic
anisotropy constant K, the saturation magnetisation M
s
and
the thermal energy as given in Eq. (2) [39]:
H
c

2K
M
s
1
k
B
T
KV

: (2)
Based on this equation, we obtain a magnetic anisotropy
energy of about K 190 200 kJ m
3
for FeCo nanopar-
ticle of 10 nm in diameter, considering T 300 K, and a
coercive eld of H
c
8 mT. The fact, that these FeCo
nanoparticles have a non-vanishing remanence at RTresults,
most likely, from the presence of a distorted lattice (such as
the tetragonal distortion for FeCo predicted in Ref. [48] and
observed in Ref. [44]) and a reduced symmetry which leads
to uni-axial anisotropy K
u
and higher blocking temperatures
T
b
. Our experimentally observed value of H
c
8 mTts well
to corresponding data from Peng et al. [50] for Fe
70
Co
30
nanoparticles of the same size at RT. Based on their
temperature-dependent data, Peng and co-workers deter-
mined a magnetic anisotropy energy of about 40 kJ m
3
, a
value that coincides with the Fe bulk value and the respective
value of FeCo alloys with small contents of Co (cf. Ref. [51]).
However, contrary to our approach with well-separated
nanoparticles, the authors worked with cluster-assembled
lms where magnetic dipole and exchange interactions play
a dominant role. Hence, the results obtained from this study
are only partly comparable to our experimental data.
3.3 Co nanoparticles on Au(111) on surfaces
and embedded into an Al matrix Magnetic particles
on surfaces and embedded in matrices might show different
magnetic properties due to their different symmetry and
interfaces which will be important in any application. As an
example, we present experimental data on Co nanoparticles
(size: 7.6 nm) embedded in an Al matrix and discuss the
results with respect to data on Au(111) [28]. In Fig. 4, the
relative magnetisation M/M
s
(given by the normalised
XMCD signal at the L
3
edge) is displayed for T 200 K.
Analogously to the results obtained earlier on uncapped Co
nanoparticles on Au(111) (cf. red triangles in Fig. 4), an
external magnetic eld of 1 T is required to saturate the
magnetic spin moments. A Langevin-type magnetisation
curve [52] (black solid curve), plotted for 7.6 nmCo particles
with a magnetic moment of 1.5 m
B
(T200 K), agrees fairly
well with the experimental data points (black dots) for Co
nanoparticles in an Al matrix. Additional experimental data
points for embedded Co nanoparticles (not displayed in this
graph) recorded at 300 K for magnetic elds of 1 and 5 T
show nearly identical results.
In the following, we will focus on the magnetic
anisotropies in the orbital moments of Co nanoparticles
1156 J. Bansmann et al.: Magnetism of 3d TM nanoparticles on surfaces
p
h
y
s
i
c
a
s s p
s
t
a
t
u
s
s
o
l
i
d
i
b
Table 2 Magnetic anisotropy constants K for bulk-like Fe and Co
[4143] as well as for tetragonally distorted FeCo lms [44] with a
c/a axes ratio of 1.18.
KkJ m
3
K (meV/atom)
bcc Fe (bulk) 42 3.3
hcp Co (bulk) 430 43
fcc Co (bulk) 270 27
tetragonally distorted FeCo lms 2900 150
2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.pss-b.com
using appropriate external magnetic elds for magnetising
our deposited particles. On Au(111), we showed that Co
nanoparticles exhibit a temperature-dependent anisotropy of
the orbital moment with high anisotropy values at 300 K and
a nearly vanishing anisotropy at 140 K (see inset in the left
part of Fig. 5 [28]). In contrast to these ndings, our
experimental data (cf. Fig. 5 for 200 K(left) and 300 K(right
part)) display, within the uncertainty, more or less identical
XMCDcurves (normalised to 1 at the Co L
2
edge) for normal
incidence (red triangles) and 608 off-normal incidence (black
dots), both at 200 and 300 K. Anisotropies in the magnetic
orbital moments would lead to different L
3
intensities in such
XMCD spectra; however, the spectra do not exhibit any
signicant hint for the presence of pronounced temperature
or angular magnetic anisotropies in embedded Co nanopar-
ticles. This nding is in agreement with results from Fe
nanoparticles in Al matrices [53].
3.4 Spectromicroscopy of Fe nanoparticles on
ferromagnetic surfaces X-ray magnetic circular dichro-
ism spectroscopy averages the magnetic properties of the
system under research over the area illuminated by the
incoming photon beam. In the case of non-interaction
particles, we assume that the ensemble of Fe or Co
nanoparticles deposited on ferromagnetic lms consists of
nearly identical particles and experiences similar inter-
actions with the underlying lm. Amore detailed analysis of
individual particles usually requires a lateral resolution
higher than the size of the particles. Spin resolved STM has
nicely achieved these boundary conditions for truly 2D
objects (see, e.g. Refs. [54, 55]), however, the method is up to
now not suited for visualising magnetic properties of nearly
spherical 3D nanoparticles with dimension exceeding 4 nm.
The combination of photoelectron microscopy and XMCD
analysis based on circularly polarised radiation offers the
possibility to get access to laterally resolved magnetic
properties of objects in the nanometer regime. In the
following, we will demonstrate the capabilities of spectro-
microscopy on Fe nanoparticles deposited from the ACIS
with sizes down to 5 nm onto a 13-nm-thick polycrystalline
Co lmwith in-plane magnetic anisotropy. In order to enable
us to study size-dependent properties, a wide range of
particle sizes (from 5 to 25 nm) has been deposited on the
same sample. Since the diameter and height of the Fe
particles cannot directly be quantied from the PEEM
images (the size of our Fe nanoparticles is below the spatial
resolution of the PEEM), the heights of the particle were
determined ex situ by atomic force microscopy (AFM) (see,
e.g. Fig. 6c). Using lithographic markers on the substrates it
is possible to identify the same particles with PEEM and
AFM.
The upper part of Fig. 6 displays individual Fe
nanoparticles (bright spots) which are visible via a chemical
contrast in X-ray PEEM. Although the size of the Fe
nanoparticles is smaller than the spatial resolution of the
PEEM, bright spots appear at the positions of widely spaced
Fe nanoparticles due to the resonant excitation of photo-
electrons at the Fe L
3,2
edges. Such images are obtained by
recording separately images before (703 eV) and on the Fe L
3
absorption edge (708.1 eV) edge and by dividing the
intensities in these images. Differences in the brightness of
the spots representing Fe nanoparticles are often related to
different particles sizes, and a small number of these spots
(10%) refer to nanoparticles having double mass (and thus,
also a double charge) which may occur during cluster growth
or due to statistics by deposition. The fraction of single
charged nanoparticles was (in the case of FeCo) determined
to about 90%, cf. [56]. An important requirement for imaging
of individual nanoparticles is that the average nearest
Phys. Status Solidi B 247, No. 5 (2010) 1157
Original
Paper
2 1 0 5
0,4
0,6
0,8
1,0
externally applied magnetic field T /
n
o
r
m
a
l
i
s
e
d
X
M
C
D
s
i
g
n
a
l
(
a
r
b
.
u
n
i
t
s
)
Figure 4 (online colour at: www.pss-b.com) Normalised magnet-
isation (M/M
s
) of Co nanoparticles in an Al matrix (black dots)
recorded at 200 K together with the respective magnetisation of
uncappedConanoparticlesof thesamesizeonAu(111) [28] at 300 K
(red triangles). The lines refer to the respective Langevin-curves for
superparamagnetic Co nanoparticles of 7.6 nm (27,000 atoms) at
200 K (solid line) and 300 K (red dashed line).
770 780 790 800
-3
-2
-1
0
1
780 790 800
300K 200K
binding energy eV /
0
60
binding energy eV /
0
60
0 100 200 300
0,0
0,2
0,4
0
40
bulk
m
o
r
b
(

B
/
a
t
o
m
)
temperature (K)
n
o
r
m
a
l
i
s
e
d
X
M
C
D
s
i
g
n
a
l
(
a
r
b
.
u
n
i
t
s
)
Figure 5 (online colour at: www.pss-b.com) Normalised XMCD
spectra (08: normal incidence, and 608 off-normal) recorded from
Co nanoparticles embedded in an Al matrix at 200 K (left)
and 300 K (right). The inset (left panel) refers to the temperature-
dependent anisotropy measured for uncapped Co nanoparticles on
Au(111) [28].
www.pss-b.com 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
distance between them is signicantly larger than the spatial
resolution of the PEEM (about 50 nm). Additionally,
the secondary photoelectron background signal from the
underlying silicon wafer at the Fe L
3,2
absorption edges has
to be small compared to the signal fromthe Fe nanoparticles.
Due to the element-specic excitation process, the method
allows to separate the magnetic signal of both, the Fe
nanoparticles and the Co lm. The XMCD contrast depends
on the relative orientation of the incoming circularly
polarised radiation with respect to the local magnetisation
state of the sample. In the case of the XMCDimage shown in
Fig. 6b, the contrast is obtained from the asymmetry of two
images with right- and left-handed circular polarisation, i.e.
Is

Is

=Is

Is

. The XMCD image


(Fig. 6b) recorded at the Fe L
3
edge (708.1 eV) has been
collected at the same area as the image with the pure
chemical contrast (Fig. 6a), typical acquisition times of the
Fe XMCDimages being 5 min. Magnetic images at the Fe L
2
edge (721.2 eV) were also recorded (not shown here)
verifying that the observed intensity contrast is truly of
magnetic origin. Additionally to the PEEMimages, we show
an AFM image (Fig. 6c) of a single Fe particle (height:
23 nm) marked by the box in the upper panels of Fig. 6. More
detailed results of these investigations will be published
elsewhere [57]. Using PEEM, it is also possible to measure
XMCD spectra (in remanence) of individual Fe nanoparti-
cles by collecting stacks of XMCDimages around the Fe L
3,2
edges for both photon helicities. The possibility of recording
such photoabsorption spectra has already been demonstrated
in a recent publication of ex situ prepared Co nanoparticles in
an Al matrix on Si-wafers with markers [58]. Here, we show
results on XMCD spectra recorded from Fe nanoparticles
deposited in situ onto a Co lm on a Si wafer (cf. Fig. 7). We
observe a well-pronounced contrast at the Fe L
3
edge, but
only a marginal contrast at the Fe L
2
edge. For a magnetic
effect, the contrast at the Fe L
3
and the L
2
edges should be of
opposite sign (e.g. cf. XMCD sum rules, [24]). In the case of
Fe, the effect at the L
2
edge should be smaller (a factor of 2 or
3) than at L
3
edge (the detailed ratio determines the magnetic
orbital and spin moments). Although the contrast at the Fe L
2
edge is quite small and nearly vanishing, clear evidence is
provided that XMCD spectroscopy on individual three-
dimensional nanoparticles down to 10 nm can be achieved.
We like to mention that the Fe nanoparticles in this sequence
of data are partly oxidised (cf. double peak structure at the
Fe L
3
edge) due to a very long illumination time with
synchrotron radiation. The additional shoulder observed at a
binding energy of 710 eV is only visible in the case of partly
oxidised Fe, it is also responsible for the reduced magnetic
contrast at the L
2
edge. We usually observed such a
behaviour after more than 12 h, but only on a small area
around the spot on the incoming radition. When moving the
sample laterally, we easily nd a non-oxidised area for
further investigations.
1158 J. Bansmann et al.: Magnetism of 3d TM nanoparticles on surfaces
p
h
y
s
i
c
a
s s p
s
t
a
t
u
s
s
o
l
i
d
i
b
Figure 6 (online colour at: www.pss-b.com) (a and b) PEEM
images with elemental (a) and magnetic contrast (b) obtained from
Fe nanoparticles deposited onto a polycrystalline Co layer; (c) AFM
imageof thesingleFenanoparticlemarkedintheupper image, height
23 nm. Black and white contrast refers to different magnetisation
directions of the individual nanoparticles, the redarrowindicates the
direction of the incoming radiation.
Figure 7 (online colour at: www.pss-b.com) XMCD spectra
recorded with left (s

) and right (s

) circularly polarised soft X-


rays from a single 12 nm Fe nanoparticle (indicated by a yellow
squareintheinset). Theredarrow(intheinset) indicatesthedirection
of the incoming photons.
2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.pss-b.com
4 Conclusion We have investigated the magnetic
properties of deposited, mass-ltered Co nanoparticles in a
size regime between 7.6 and 25 nm on surfaces, thin
ferromagnetic lms and non-magnetic substrates, as well
as embedded into an Al matrix. In the rst part, we have
presented the correction factor for magnetic orbital and spin
moments obtained in the TEY mode and applied them to
results from earlier studies. The results clearly show that the
magnetic orbital moments of Co nanoparticles on surfaces
are enhanced with respect to the corresponding bulk values
as expected from the reduced symmetry at the surface of the
nanoparticles. Moreover, the values differ signicantly
depending on the type of surface. The mist between the
lattice parameters of the Co nanoparticles and the underlying
single-crystalline surfaces might give rise to elastic tetra-
gonal distortion in the Co lattice parameters (i.e. a magneto-
elastic contribution to the complex magnetic anisotropy
energy of the particle), and thus to modied magnetic orbital
moments. Compared to Co on silicon substrates with native
oxide surface, FeCo nanoparticles with a size of 9.6 nmshow
an unusal high magnetic anisotropy energy which is most
likely related to a small tetragonal distortion of the lattice.
Such phenomena have earlier been observed in FeCo lms
on Pd and Pt surfaces resulting in strongly enhanced
magnetic anisotropy energies [44, 49]. The effect is
remarkable since FeCo alloys are usually soft magnetic
materials with very low anisotropy energies. Based on the
ferromagnetic behaviour at RT we estimate the uni-axial
magnetic anisotropy of these FeCo nanoparticle to be about
K 190 200 kJ m
3
. Fcc Co particles of the same size
with even higher magnetic anisotropy energy, however, are
superparamagnetic at RT due to their cubic symmetry.
Additionally, we have compared the magnetic anisotropies
in the orbital moments of Co nanoparticles on Au(111)
surfaces with those embedded into Al matrices. The
pronounced anisotropy in the orbital moments as observed
for Co nanoparticles on Au(111) vanishes when the particles
are embedded in Al. A similar effect has recently been
observed for Fe nanoparticles on W(110) and embedded into
Al [53]. Finally, we have demonstrated that we are able to (i)
chemically identify and (ii) magnetically image individual
Fe nanoparticles (down to less than 10 nm) on Co lms.
Additionally to the images at the L
3
edge, we also recorded
XMCD spectra from individual nanoparticles. This method
opens the possibility to investigate the magnetic properties of
individual nanoparticles and not only on an ensemble of
usually not identical nanoparticles.
Acknowledgements We like to thank our co-workers
J. Passig (Universitat Rostock, D) and L. J. Heyderman,
A. Weber, R. Schelldorfer (Paul Scherrer Institut, Switzerland).
Furthermore, we are indebted to D. Schmitz and P. Imperia (Hahn-
Meitner-Institut Berlin) for their help during the experiments at the
HMI beamline UE-46PGM. We gratefully acknowledge nancial
support by the Deutsche Forschungsgemeinschaft (DFG) within the
priority program 1153 Clusters in Contact with Surfaces via DFG
BA 1612/3-3, DFG KL 2188/1-3 and DFG GE 1026/4-3.
References
[1] D. L. Huber, Small 1, 482 (2005).
[2] K.-H. Meiwes-Broer, Clusters on Surfaces (Springer-Verlag,
Berlin, 2000).
[3] A. Hutten, D. Sudfeld, I. Ennen, G. Reiss, K. Wojczykowski,
and P. Jutzi, J. Magn. Magn. Mater. 293, 93 (2005).
[4] S. Sun, C. B. Murray, D. Weller, L. Folks, and A. Moser,
Science 287, 1989 (2000).
[5] G. S. Chaubey, C. Barcena, N. Poudyal, C. Rong, J. Gao,
S. Sun, and J. P. Liu, J. Am. Chem. Soc. 129, 7214 (2009).
[6] A. Ethirajan, U. Wiedwald, H. G. Boyen, B. Kern, L. Han, A.
Klimmer, F. Weigl, G. Kastle, P. Ziemann, K. Fauth, C. Jun,
R. J. Behm, P. Oelhafen, P. Walther, J. Biskupek, and U.
Kaiser, Adv. Mater. 19, 406 (2007).
[7] P. Gambardella, S. Rusponi, M. Veronese, S. S. Dhesi, C.
Grazioli, A. Dallmeyer, I. Cabria, R. Zeller, P. H. Dederichs,
K. Kern, C. Carbone, and H. Brune, Science 300, 1130
(2003).
[8] J. B. Kortright, D. D. Awschalon, J. Stohr, S. D. Bader, Y. U.
Idzerda, S. S. P. Parkin, I. K. Schuller, and H. C. Siegmann, J.
Magn. Magn. Mater. 207, 7 (1999).
[9] J. Bansmann, S. H. Baker, C. Binns, J. A. Blackman, J. P.
Bucher, J. Dorantes-Davila, V. Dupuis, L. Favre, D.
Kechrakos, A. Kleibert, K. H. Meiwes-Broer, G. M. Pastor,
A. Perez, O. Toulemonde, K. N. Trohidou, J. Tuaillon, and Y.
Xie, Surf. Sci. Rep. 56, 189 (2005).
[10] H. A. Durr, T. Eimuller, H. J. Elmers, S. Eisebitt, M. Farle, W.
Kuch, F. Matthes, M. Martins, H. C. Mertins, P. M. Oppeneer,
L. Plucinski, C. M. Schneider, H. Wende, W. Wurth, and H.
Zabel, IEEE Trans. Magn. 45, 15 (2001).
[11] R. H. Kodama, J. Magn. Magn. Mater. 200, 359 (1999).
[12] M. Jamet, W. Wernsdorfer, C. Thirion, D. Mailly, V. Dupuis,
P. Melinon, and A. Perez, Phys. Rev. Lett. 86, 4676 (2001).
[13] W. Wernsdorfer, D. Mailly, and A. Benoit, J. Appl. Phys. 87,
5094 (2000).
[14] T. Tanji, M. Maeda, N. Ishigure, and N. Aoyama, Phys. Rev.
Lett. 84, 1038 (1999).
[15] S. A. Majetich and Y. Jin, Science 284, 470 (1999).
[16] P. Schattschneider, S. Rubino, C. Hebert, J. Rusz, J. Kunes, P.
Novak, E. Carlino, M. Fabrizioli, G. Panaccione, and G.
Rossi, Nature 441, 486 (2006).
[17] J. Bansmann and A. Kleibert, Appl. Phys. A 80, 957 (2005).
[18] R. P. Methling, V. Senz, E. D. Klinkenberg, T. Diederich, J.
Tiggesbaumker, G. Holzhuter, J. Bansmann, and K. H.
Meiwes-Broer, Eur. Phys. J. D 16, 173 (2001).
[19] A. Kleibert, J. Passig, K. H. Meiwes-Broer, M. Getzlaff, and
J. Bansmann, J. Appl. Phys. 101, 114318 (2007).
[20] H. Haberland, Z. Insepov, and M. Moseler, Phys. Rev. B 51,
11061 (1995).
[21] C. Quitmann, U. Flechsig, L. Patthey, T. Schmidt, G. Ingold,
M. Howell, M. Janousch, and R. Abela, Surf. Sci. 480, 173
(2001).
[22] J. Stohr, H. A. Padmore, S. Anders, T. Stammler, and M. R.
Scheinfein, Surf. Rev. Lett. 5, 12973 (1998).
[23] J. Stohr, J. Electron Spectrosc. Relat. Phenom. 75, 253
(1995).
[24] C. T. Chen, Y. U. Idzerda, H. J. Lin, N. V. Smith, G. Meigs, E.
Chaban, G. H. Ho, E. Pellegrin, and F. Sette, Phys. Rev. Lett.
75, 152 (1995).
[25] P. Carra, B. T. Thole, M. Altarelli, and X. Wang, Phys. Rev.
Lett. 70, 694 (1993).
Phys. Status Solidi B 247, No. 5 (2010) 1159
Original
Paper
www.pss-b.com 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
[26] B. T. Thole, P. Carra, F. Sette, and G. van der Laan, Phys.
Rev. Lett. 68, 1943 (1992).
[27] J. Bansmann, M. Getzlaff, A. Kleibert, F. Bulut, R. K.
Gebhardt, and K. H. Meiwes-Broer, Appl. Phys. A 82, 73
(2006).
[28] J. Bansmann, A. Kleibert, F. Bulut, M. Getzlaff, P. Imperia,
C. Boeglin, and K. H. Meiwes-Broer, Eur. Phys. J. D 45, 521
(2007).
[29] A. Kleibert, F. Bulut, R. K. Gebhardt, W. Rosellen, D.
Sudfeld, J. Passig, J. Bansmann, K. H. Meiwes-Broer, and
M. Getzlaff, J. Phys.: Condens. Matter 20, 445005 (2008).
[30] A. Kleibert, K. H. Meiwes-Broer, and J. Bansmann, Phys.
Rev. B 79, 125423 (2009).
[31] R. Nakajima, J. Stohr, and Y. U. Idzerda, Phys. Rev. B 59,
6421 (1999).
[32] K. Fauth, Appl. Phys. Lett. 85, 3271 (2004).
[33] M. Tischer, O. Hjortstam, D. Arvanitis, J. H. Dunn, F. May,
K. Baberschke, J. Trygg, J. M. Wills, B. Johansson, and O.
Eriksson, Phys. Rev. Lett. 75, 1602 (1995).
[34] J. Stohr and H. Konig, Phys. Rev. Lett. 75, 3748 (1995).
[35] R. Wu and A. J. Freeman, Phys. Rev. Lett. 73, 1994
(1994).
[36] C. Ederer, M. Komelj, M. Fahnle, and G. Schutz, Phys. Rev.
B 66, 094413 (2002).
[37] D. Weller, J. Stohr, R. Nakajima, A. Carl, M. G. Samant, C.
Chappert, R. Megy, P. Beauvillain, P. Veillet, and G. A. Held,
Phys. Rev. Lett. 75, 3752 (1995).
[38] W. Wernsdorfer, E. Bonet Orozco, K. Hasselbach, A. Benoit,
B. Barbara, N. Demoncy, A. Loiseau, H. Pascard, and D.
Mailly, Phys. Rev. Lett. 78, 1791 (1997).
[39] A. Aharoni, Introduction to the Theory of Ferromagnetism
(Oxford University Press, Oxford, 2000).
[40] D. L. Beke, S. Szabo, and M. Kis-Varga, in: Advances
in Condensed Matter and Materials Research, edited
by F. Gerard (Nova Science Publishers, New York, 2004).
[41] T. Sugimoto, Monodispersed Particles (Elsevier Science B.V,
Amsterdam, 2001).
[42] M. Jamet, W. Wernsdorfer, C. Thirion, V. Dupuis, P.
Melinon, A. Perez, and D. Mailly, Phys. Rev. B 69,
024401 (2004).
[43] M. Getzlaff, Fundamentals of Magnetism (Springer-Verlag,
Berlin, 2007).
[44] G. Andersson, T. Burkert, P. Warnicke, M. Bjorck, B. Sanyal,
C. Chacon, C. Zlotea, L. Nordstrom, P. Nordblad, and O.
Eriksson, Phys. Rev. Lett. 96, 037205 (2006).
[45] O. Kitakami, H. Sato, Y. Shimada, F. Sato, and M. Tanaka,
Phys. Rev. B 56, 13849 (1997).
[46] M. Getzlaff, J. Bansmann, F. Bulut, R. K. Gebhardt, A.
Kleibert, and K. H. Meiwes-Broer, Appl. Phys. A 82, 95
(2006).
[47] D. Wu, Q. Zhang, J. P. Liu, D. Yuan, and R. Wu, Appl. Phys.
Lett. 92, 052503 (2008).
[48] T. Burkert, L. Nordstrom, O. Eriksson, and O. Heinonen,
Phys. Rev. Lett. 93, 027203 (2004).
[49] A. Winkelmann, M. Przybylski, F. Luo, Y. Shi, and J.
Barthel, Phys. Rev. Lett. 96, 257205 (2006).
[50] D. L. Peng, Y. Chena, H. Shea, R. Katoh, and K. Sumiyama,
J. Alloys Compd. 469, 276 (2009).
[51] R. C. Hall, J. Appl. Phys. 31, 1037 (1960).
[52] C. P. Bean and J. D. Livingston, J. Appl. Phys. 30, 120S
(1959).
[53] A. Kleibert, F. Bulut, W. Rosellen, K. H. Meiwes-Broer, J.
Bansmann, and M. Getzlaff, J. Phys.: Conf. Ser., in press
(2010).
[54] M. Bode, Rep. Prog. Phys. 66, 523 (2003).
[55] A. Kubetzka, P. Ferriani, M. Bode, S. Heinze, G. Bihlmayer,
K. von Bergmann, O. Pietzsch, S. Blugel, and R.
Wiesendanger, Phys. Rev. Lett. 94, 087204 (2005).
[56] M. Getzlaff, A. Kleibert, R. P. Methling, J. Bansmann, and K.
H. Meiwes-Broer, Surf. Sci. 566568, 332 (2004).
[57] A. Fraile Rodr guez, A. Kleibert, J. Bansmann, A. Voitkans,
L. J. Heyderman, and F. Nolting, submitted (2009).
[58] A. Fraile Rodr guez, F. Nolting, J. Bansmann, A. Kleibert,
and L. J. Heyderman, J. Magn. Magn. Mater. 316, 426 (2007).
1160 J. Bansmann et al.: Magnetism of 3d TM nanoparticles on surfaces
p
h
y
s
i
c
a
s s p
s
t
a
t
u
s
s
o
l
i
d
i
b
2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.pss-b.com

Potrebbero piacerti anche