Sie sind auf Seite 1von 8

www.afm-journal.

de

FULL PAPER
Spontaneous Outcropping of Self-Assembled Insulating
Nanodots in Solution-Derived Metallic Ferromagnetic
La0.7Sr0.3MnO3 Films
By César Moreno, Patricia Abellán, Awatef Hassini, Antoine Ruyter, Angel
Pérez del Pino, Felip Sandiumenge, Marie-Jose Casanove, José Santiso,
Teresa Puig, and Xavier Obradors*

generate dramatically new functionalities.


A new mechanism is proposed for the generation of self-assembled nanodots Nanomagnetism has become a very rich
arena where new phenomena are being
at the surface of a film based on spontaneous outcropping of the secondary
discovered with many potential applica-
phase of a nanocomposite epitaxial film. Epitaxial self-assembled Sr–La oxide tions in magnetic recording, spintronics,
insulating nanodots are formed through this mechanism at the surface of an sensors, biomedicine, etc.[1,2] Also, it has
epitaxial metallic ferromagnetic La0.7Sr0.3MnO3 (LSMO) film grown on been shown that vortex pinning centers can
SrTiO3 from chemical solutions. TEM analysis reveals that, underneath the be generated in superconducting films
La–Sr oxide (LSO) nanodots, the film switches from the compressive out-of- grown on top of surface nanotemplates,[3]
similarly to nanocomposite films.[4] The two
plane stress component to a tensile one. It is shown that the size and
general approaches to nanostructure for-
concentration of the nanodots can be tuned by means of growth kinetics and mation in surfaces, lithography and self-
through modification of the La excess in the precursor chemical solution. The assembling, have been already found to be
driving force for the nanodot formation can be attributed to a cooperative extremely successful in metallic and semi-
effect involving the minimization of the elastic strain energy and a conducting nanostructure generation.[2,5–7]
Complex oxides, instead, have only recently
thermodynamic instability of the LSMO phase against the formation of a
become an object of intense study. The
Ruddelsden–Popper phase Sr3Mn4O7 embedded in the film, and LSO surface formation of self-assembled oxide nano-
nanodots. The mechanism can be described as a generalization of the structures by vapor deposition techniques
classical Stranski–Krastanov growth mode involving phase separation. LSO have been found to follow either the
islands induce an isotropic strain to the LSMO film underneath the island Volmer–Weber or Stranski–Krastanov
which decreases the magnetoelastic contribution to the magnetic anisotropy. mechanisms, similarly to semiconductor
systems,[7] while chemical solution deposi-
tion (CSD) has been only recently started to
be investigated as a powerful bottom-up nanofabrication meth-
1. Introduction odology.[8,9] Three-dimensional self-organized nanocomposite
films have also very recently appeared as a very rich new arena
Nanoscale manipulation of functional materials is generating a
for the generation of new physical phenomena, such as multi-
very broad interest in many fields such as electronics, optics,
ferroic behavior,[10] or to tune their physical properties, such as
magnetism, ferroelectricity, or superconductivity owing to the
magnetoresistance in a percolative transport system.[11]
capability of modifying the fundamental properties or even to
Among the complex oxides, rare-earth manganites of the type
Ln1xAxMnO3 (Ln is a rare earth, A is a alkaline earth) have become
[*] Prof. X. Obradors, C. Moreno, P. Abellán, Dr. A. P. del Pino, an extremely rich family of compounds displaying a wide variety of
Dr. F. Sandiumenge, Dr. J. Santiso, Dr. T. Puig behavior, such as ferromagnetism, metallic conductivity, anti-
Institut de Ciència de Materials de Barcelona, CSIC
Campus Universitat Autònoma de Barcelona
ferromagnetism, colossal magnetoresistance, charge order, phase
Bellaterra, Catalonia 08193 (Spain) separation, insulating states, etc.[12–14] Additionally, a strong link
E-mail: xavier.obradors@icmab.es among these states can be induced, either by modified composi-
Dr. A. Hassini, Dr. A. Ruyter tion, oxygen stoichiometry or strain state, or by several types of
LEMA-Universitè de Tours, CNRS-CEA UMR 6157 external parameters, such as temperature, pressure, magnetic, and
Parc de Grandmont, 37000 Tours (France) electric fields, X-ray irradiation, etc. The metallic ferromagnetic
Dr. M.-J. Casanove phase La1xSrxMnO3 (LSMO) with x  0.3–0.4 has been found
CNRS, CEMES, Universite de Toulouse
UPS, 29 rue J. Marvig, 31055 Toulouse (France)
to be one of the most robust manganites and hence it is often
selected as a model system to investigate nanoscale phenomena. In
DOI: 10.1002/adfm.200900095 the present work, a new strategy of general validity is proposed to

Adv. Funct. Mater. 2009, 19, 2139–2146 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2139
www.afm-journal.de

generate self-assembled nanostructures from a nanocomposite


FULL PAPER

film which generalize the classical Stranski–Krastanov growth


mechanism. We propose that the same elastic strain energy
behaving as driving force for the formation of nanodots over a
wetting layer controls the topological arrangement of a phase
segregated system consisting of a wetting layer with nanodots on
its surface in the nanocomposite system. The crystalline phase
having the higher interfacial misfit will be that outcropping as self-
assembled nanodots at the surface of the strained film.

2. Results and Discussion


Thin films of La1xSrxMnO3 (LSMO, x ¼ 0.3) with self-assembled
La–Sr oxide (LSO) nanoislands grown on top of them have been Figure 1. Reciprocal space map of an LSMO film grown at 1000 8C during
prepared by CSD process. The precursors were metallic 12 h determined with X-ray diffraction around the (103) reflection. The
propionates synthesized from their corresponding acetates, position of the Bragg peaks of the STO substrate and the LSMO film are
[15,16]
following a well-established route. Essentially, the chemical indicated.
solution was obtained by dissolving the metallic propionates
in propionic acid solvent, either in a nominally stoichiometric
ratio (x ¼ 0.3) or with an excess of La (up to 20 mol %). All fully strained state is not usually observed in films grown by
the films reported here were prepared using a solution CSD.[21–23]
concentration 0.3 M respect to the manganese content leading The surface morphology of the CSD-LSMO films was found to
to a final viscosity of 2.5 mPa  s which remains stable over long strongly evolve with thermal annealing conditions. While at 900 8C
periods of time (10 days). The solution metal content was a granular structure (epitaxial grains with a grain size of a few tens
checked by induced coupled plasma (ICP) measurements and they of nm) is observed in the atomic force microscope (AFM) images,
were found to correspond to the nominal one within 5% at. at 1000 8C atomically flat surfaces are developed over more than
Heteroepitaxial growth of LSMO films was confirmed by XRD 90% of the surface area.[17] Strikingly, however, annealing the films
analysis, indicating a cube-on-cube orientation relationship at 1000 8C for extended periods (several hours) leads to the
(100)LSMO//(100)STO[17] (pseudo-cubic indexation for LSMO spontaneous formation of self-assembled nanodots at the film
will be used throughout for simplicity). Figure 1 presents a typical surface. Figure 2a–c displays three typical AFM images, with the
reciprocal space map scanned around the (103) reflections of corresponding profiles, obtained on films with different island
STO and LSMO. The coordinate system is oriented with the sizes developed on atomically flat LSMO films. As it is seen,
horizontal axis, Qx//[100], parallel to the substrate interface polyhedral nanodots with heights up to 50 nm and a fairly narrow
and the vertical axis, Qy//[010], perpendicular to the substrate. In size distribution outcrop on the atomically flat LSMO surface. The
Figure 1, rlu means ‘‘reciprocal lattice units’’ and 1 rlu ¼ 2/l with annealing conditions (temperature and time), together with
l ¼ 1.54184 Å. Figure 1 clearly reveals that the in-plane the metal concentration in La–Mn ratio, allow a fine tuning of the
component Qx of (103)LSMO matches
perfectly with the STO lattice para-
meter, indicating that the film is fully
strained (a// ¼ 3.904(8) Å, very close to
the STO lattice parameter 3.905 Å[18]),
while a corresponding com-pressive
strain occurs perpendicular to the
substrate with a? ¼ 3.847(6) Å. The
diffraction averaged in-plane and out-
of-plane strains of the LSMO films
are therefore tensile, e// ¼ (a//  abulk)/
abulk ¼ þ0.83%, and compressive,
e? ¼ (a?  abulk)/abulk ¼ 0.67%, res-
pectively, where abulk ¼ 3.873 Å.[19]
The fully strained structure was found
to exist independently of the thermal
treatment and is consistent with the fact
that critical film thickness is below of Figure 2. Topographic AFM images (5  5 mm2), binarized area and profile across several nanodots of
the critical value for misfit dislocation LSMO films with different concentrations and size of LSO nanodots. a,b) Stoichiometric LSMO films
formation measured for PLD-LSMO to grown at 1000 8C on an STO substrate during 8 and 12 h, respectively; c) LSMO film grown at 1000 8C
be >100 nm.[20] We also note that this during 12 h on STO with 20 mol % excess of La propionates precursor content in the starting solution.

2140 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2009, 19, 2139–2146
www.afm-journal.de

EELS analyses were performed across the nanodot–film and

FULL PAPER
film–substrate interfaces. Results are shown in Figure 3b. The set
of spectra were taken every 3.5 nm using an electron probe with a
diameter of 0.5 nm. The positions of the starting and final analyses
are indicated by 1 (substrate) and 2 (nanodot), as indicated in the
inset and the plot. As expected, the O–K edge peak at 532 eV is
observed through the whole profile. It is also clearly observed
that the La M5,4 edge persists across the film–nanodot interface
while the Mn L3,2 line is confined within the LSMO layer. This
indicates that the nanodots have a significant content in La and lack
Mn. EELS did not yield reliable information about the occurrence
of Sr because of the unfavorable energies associated to its edges.[24]
Accordingly, the occurrence of Sr was checked by energy dispersive
X-ray spectroscopy (EDS). Results yielded significant amounts of
Sr (20–30 wt %).
In addition to the outcropped LSO islands, TEM observations
revealed the occurrence of rounded inclusions, about 10–20 nm in
diameter, embedded within the LSMO film. Figure 4a is a low
magnification cross-section image including an LSO island and
such an inclusion, labeled Ruddelsden–Popper (R–P). In order to
gather high spatial resolution chemical information from such
Figure 3. TEM-EELS analysis of an LSMO film grown on an STO substrate
inclusions, energy filtered imaging was carried out using a Cs-
during 12 h at 1000 8C. a) Low magnification cross-sectional TEM
image showing a general view of the system where several interfacial corrected electron microscope operated at 200 kV, enabling a
nanodots are observed; b) EELS line scan across the interfaces nanodot– spatial resolution of 1 nm. To that purpose, an inclusion in contact
film and film–substrate, as shown in the inset. The scan was performed with the roots of an LSO island was selected enabling direct
from 1 to 2. comparison between both phases. Figure 4b and c presents the
conventional elastic-electron image corresponding to the analyzed
area, and the La-M5,4-edge distribution map, respectively. Clearly,
size and concentration of the polyhedral nanodots. For instance, the island appears La-rich while the inclusion exhibits a contrast
the equivalent diameter (D), mean height (h), concentration (N) of even darker than the surrounding LSMO matrix, as expected from
the nanodots, and total volume of outcropped nanodots per stoichiometric balance considerations.
unit surface (V) seen in Figure 2a–c were estimated to evolve In the analyzed LSMO [010] zone axis images, such inclusions
as follows: a) D  60  20 nm, h  4  3 nm, N ¼ 14 dots mm2, typically exhibit either cubic-like or tetragonal-like structural
V  0.45  103 nm3 mm2; b) D  336  56 nm, h  40  10 nm, projections. As an example, Figure 4d and e shows a high
N ¼ 2 dots mm2, V  45.2  103 nm3 mm2; c) D  270  60 nm, resolution TEM image of an [100] oriented inclusion and its
h  50  20 nm, N ¼ 8 dots mm2, V  114.5  103 nm3 mm2. corresponding FFT spectrum, respectively. Figure 4f is a [001]
The increase of the nanodot volume with annealing time suggests zone axis FFT spectrum corresponding to a different inclusion.
that the outcropping process is kinetically limited while the This indicates that these inclusions hold an epitaxial relationship
reduction of the nanodot concentration N with the annealing with the surrounding LSMO matrix. Considering that the observed
time indicates that coarsening occurs simultaneously and so two types of projection correspond to two major zone axis, namely
surface diffusion is important. On the other hand, the strong [001] and [100] or [010], the lattice parameters derived from the FFT
increase of nanoisland volume in samples with La excess Figure 2c spectra yield lattice parameters 3.8 and 19.7 Å, which
clearly demonstrates that the nanodot size and concentration can interestingly fit with the n ¼ 2 R–P structure[25] (S.G. I4/mmm,
be tuned through the initial composition and annealing process. A a  b  ap, c  20 Å).[26] Thus, the epitaxial relation is given by
high resolution TEM cross-section investigation was therefore (001)[100]R–P//(001)[100]LSMO. This structure is formed by the
carried out to investigate the structural response of the LSMO film insertion of a rock-salt-like unit, SrO, every n (¼2) perovskite-type
to the phase segregation process as well as its consequences on the blocks, SrMnO3. The Sr cations can in fact be replaced by La,
physical properties. resulting in an increase of the lattice parameters.[26] As a general
A typical low magnification image of a 12 h annealed behavior, [001] images exhibit rather perfect crystallinity,
stoichiometric film, viewed along [010] is shown in Figure 3a. while [100] or [010] ones feature strong disorder, both in the
The LSMO film thickness was found to change smoothly from one stacking sequence and in the lateral continuity of atomic layers. In
observed area to another with values comprised between 20 and fact, stacking disorder is a common feature in these phases.[27]
26 nm. The image clearly shows flat islands with lateral facets Thus, it is likely that both, structural mismatch and cationic
and sizes in the range 20–55 nm in height and 90–280 nm in balance are further accommodated through structural and
length. Nanodots facets are found to form an apparent angle of chemical disorder (mostly in the form of stacking faults and
about 45 8 with the substrate surface. It is furthermore observed lateral discontinuity of atomic layers) in addition to elastic strains
that the nanodots penetrate into the LSMO film, thus thinning the within the inclusions.
LSMO film below the nanodots which in some cases reduce Figure 5a and c compares high resolution cross-sectional TEM
considerably its thickness. images corresponding to an island partly embedded in the film,

Adv. Funct. Mater. 2009, 19, 2139–2146 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2141
www.afm-journal.de

and (200)LSMO spots and those labeled 1 and


FULL PAPER

2 for the island, respectively.


In order to investigate the interaction of
mismatch strains originated at the island
interface with those originated at the substrate,
SAD patterns containing the LSMO–STO
interface beneath and far from the islands
were compared. The SAD pattern shown in
Figure 5b corresponds to the island free region.
The solid square indicates spots due to the
STO substrate while the dashed square
shows the spots coming from the LSMO film.
The pattern is consistent with the rhombohe-
dral structure of LSMO (S.G. R-3c) and
indicates a cube-on-cube epitaxial relationship.
Spots contributed by both, STO and LSMO,
appear more intense and are elongated
along [001], perpendicular to the interface.
Careful inspection of higher order spots
reveals splitting in that direction, as demon-
strated in the magnified view of the (103)
STO and LSMO reflections. On the other hand,
within experimental resolution, the pattern
indicates complete matching between film
and substrate parallel to the interface, i.e.,
a//(LSMO)  a//(STO) ¼ 3.905 Å. Thus, LSMO
film regions not affected by the interaction with
nanodots appear under an in-plane tensile
stress and the corresponding compression
perpendicular to the substrate, in agreement
with the averaged view provided by X-ray
diffraction (Fig. 1). The strain values deter-
mined for the LSMO regions not affected by the
Figure 4. TEM cross-section analysis of inclusions embedded within the film. a) Low magnifi- interaction with the nanodots are e// ¼ þ0.83%
cation image showing the three-phases building the epitaxial system: An (La,Sr)-oxide island, an and e? ¼ 1.03%. These values yield a Poisson
embedded R–P type inclusion, and the LSMO film; b,c) correspond to the elastic electron and La- ratio n ¼ e /(e  2e ) of 0.38, close to
? ? //
M5,4 edge filtered images, respectively, of an area including an (La,Sr)-oxide island and a
that determined by Maurice et al.[20] for a
neighboring R–P type inclusion; d) high resolution image of an R–P inclusion viewed along
its [100] zone axis; e) FFT spectrum corresponding to (d); f) [001] zone axis FFT spectrum
L2/3Sr1/3MnO3 film, n ¼ 0.34, thus signaling
corresponding to an R–P type inclusion different from (d) (not shown). that the film keeps its structural integrity even
after the complex segregation process.
Figure 5e corresponds to the SAD pattern
with an island free film region. Strain contrasts can be observed obtained across the LSMO/STO interface in the region beneath an
associated to the root of the island, while the free-film exhibits a LSO nanodot. Again, the diffraction spots due to the STO and
high structural perfection. The most significant lattice spacings LSMO structures are connected by solid and dashed squares,
obtained from the high resolution image of the island are: 8.37 and respectively. In this region, the film also corresponds to a
5.41 Å, which, to the authors knowledge, do not correspond neither rhombohedral LSMO phase. However, careful inspection of the
to the sesquioxide Sr0.2La1.8O2.97,[28] nor to any other Sr–La–O pattern did not reveal any signature of spot splitting or even
reported phase, namely: Sr3La4O9,[29] La4SrO7,[30] La2Sr2O5,[31] and elongation in any direction, i.e., diffraction spots coming from the
La2SrOx.[32] A structural determination of these nanodots falls two different lattices are apparently fully coincident. Thus, within
beyond the scope of the present study. However, careful analyses of experimental accuracy, comparison between the SAD patterns
the buried roots of the island reveal that they sustain an epitaxial obtained from regions far from and below the nanodots, do not
relationship both perpendicularly and parallel the film. This is reveal any apparent difference concerning the in-plane lattice
indicated in Figure 5c showing a SAD pattern taken across the parameters. However, the c-axis parameter of the LSMO film
island-film interface, where the coincident lattice is marked by a becomes tensile stretched below the LSO nanodots by þ0.83%,
dashed line. Matching distances are found to be d// ¼ 4.055 Å and resulting in an expanded LSMO unit cell volume DV/V  þ2.5%,
d? ¼ 4.048 Å, parallel and perpendicular to the interface, respec- as compared to the bulk LSMO unit cell.
tively, yielding misfit strains e// ¼ 4.7% and e? ¼ 4.5%. Spot Results reported above indicate that a complex chemical
splitting between both diffraction patterns are better observed for segregation phenomenon lies at the root of the new mechanism of
higher order reflections, as for instance between the (002)LSMO surface nanodot formation. The observation of a coarsening effect

2142 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2009, 19, 2139–2146
www.afm-journal.de

formed with the same epitaxial order described

FULL PAPER
above, and with respective surface occupancies
(1  x)/x, the total elastic strain energy per unit
area Eelnc could be written as:

Eelnc ¼ ð1  xÞ Eel ðLSMOÞ


þ xEel ðLSOÞ þ Eint (1)

where Eel(LSO) is the corresponding elastic


energy of an epitaxial LSO film growing on
STO, which is higher than Eel(LSMO) due to
the enhanced misfit strain and Eint is the elastic
strain energy associated to the LSMO/LSO
interfaces which would depend on the geome-
trical arrangement of the two phases within the
nanocomposite film. In the present case, due
to the high lattice misfit of LSO with, both the
STO substrate and the LSMO phase, it is
anticipated that Eelnc will be high. Previous
investigations of nanocomposite epitaxial
films, such as LCMO/MgO or LSMO/
ZnO,[34,11] showed that a columnar phase-
separated structure could be formed with
stress accommodation through a structural
phase transition of LCMO. The main differ-
ence in the present system is that both phases
of the nanocomposite are elastically strained
Figure 5. a) Cross-sectional HRTEM image of the LSMO film and STO/LSMO interface in an while in LCMO/MgO nanocomposites, one of
area far away from the LSO nanodots influence; b) electron diffraction pattern of a selected area the phases (MgO) was not strained because the
displayed in (a) covering STO substrate and LSMO film; c) HRTEM image of the LSMO film under substrate was also MgO. Our observation of an
a nanodot showing the STO/LSMO and LSMO/island interfaces; d) SAD pattern taken across the outcropping phenomenon should be driven
island–LSMO interface illustrating the matching distances between both phases. 1 and 2 mark then by a decrease in the total elastic energy of
spots originated by the island; e) electron diffraction pattern taken at an area covering LSMO film the nanocomposite film in a way which can be
under a nanodot and the STO substrate. considered somehow similar to that described
in the Stranski–Krastanov growth mechanism
enhanced by temperature and a higher La concentration suggests for films with a single composition. In the present case, a
that the LSO nanodots arise as a consequence of a kinetically generalized theoretical analysis would require consideration of
limited surface outcropping process, probably cooperatively driven the complex equilibrium phenomena revealed by the observed
by the thermodynamic instability leading the simultaneous chemical segregation.
formation of the outcropped LSO islands and the Sr3Mn2O7 In the Stranski–Krastanov nanodot formation mechanism,[6,35]
embedded inclusions, and a reduction of the total elastic energy of a strained wetting layer persists while coherently strained
the resulting three phase nanostructure. The observed topological nanodots are formed at the surface driven by an elastic relaxation
distribution of segregated phases is governed by the energy cost term (negative energy term), even if an extra surface is formed
associated with the creation of new interface and surface area. In (generally contributing a positive energy term). The more complex
this sense, it is worth noting that the n ¼ 2 R–P structure consists of problem of morphological evolution of nanocomposite films still
the alternation of two perovskite layers with a single rock-salt-type lacks of theoretical analysis and so there is no prediction of the
layer.[33] Therefore, there exists a favorable structural compatibility novel outstanding behavior reported here for the LSMO/LSO
between this structure and the LSMO and STO ones, that would system: the spontaneous formation of nanodots at the surface with
explain why this phase remains typically embedded in the film and a composition La–Sr–O naturally selected on the basis of
occasionally in contact with the STO surface (Fig. 4a), while the minimization of the elastic energy of the nanocomposite film.
LSO phase tends to minimize its interface area with the film. Thus, The remaining wetting layer keeps the composition having the
to a first approximation the contribution of the R–P inclusions to lowest elastic energy with the substrate, i.e., LSMO. A thermo-
the elastic energy of the coherently strained LSMO film can dynamic analysis of the new self-assembled nanostructure could
be neglected. Thus, the elastic energy per unit surface can be be made where the increase of the energy of the self-assembled
approximated, as usual, by Eel(LSMO)a e2t, where e is the lattice island/LSMO nanodot–film Esa, as compared to an equivalent
misfit and t the film thickness. This energy is the driving force for LSMO/LSO nanocomposite film (Eq. 1), can be then written as:
defining a critical thickness for misfit dislocation formation. If a
coherently strained LSMO/LSO nanocomposite film would be E sa ¼ x½Eel ðLSMOÞ  Eel ðLSOÞ  Eint þ Esnd þ Eelnd (2)

Adv. Funct. Mater. 2009, 19, 2139–2146 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2143
www.afm-journal.de

the temperature dependent magnetoresis-


FULL PAPER

tance. The temperature dependence of the


magnetization was measured under 5 mT in all
the samples and we found that a sharp
ferromagnetic to paramagnetic transition
occurs with a Curie temperature Tc  365 K
while high field magnetization measurements
at 10 K indicated a saturation magnetization mo
Ms  590 mT, a value very close to that of bulk
LSMO,[13,14,37] see Figure 6a. The temperature
dependence of the resistivity measured at zero
magnetic field and under a magnetic field of 5 T
Figure 6. a) Temperature dependence of the magnetization measured at 5 mT for a typical LSMO
film with (La,Sr)Ox nanodots and electrical resistivity measured at zero magnetic field and under a
is also displayed in Figure 6a. The observed
magnetic field of 5 T; b) isothermal magnetic hysteresis loops of an LSMO film measured at 300 K resistivity [r(300 K) ¼ 8.42  104 V cm] and
with the magnetic field oriented parallel (//) and perpendicular (?) to the film substrate. maximum magnetoresistance (DR/R  11%)
are very typical of those observed in high
quality single crystalline thin films.[13,14,37]
nd These results mean that the CSD grown LSMO films exhibit
Where Es is a positive term including the extra surface and
interface energies associated to the nanodots formation, Eelnd is intrinsic macroscopic magnetic parameters (Ms and Tc) and
the elastic relaxation energy of the LSO nanodots and the other transport properties very close to those of bulk values and that
terms are defined in (Eq. 1). For a parallepipedic nanodot with the bare macroscopic behavior of the films is not modified by the
dimensions a  b  h, Esnd can be written as:[6,36] additional strain induced by the interfacial LSO nanodots.
Isothermal magnetization loops revealed the macroscopic
  anisotropy of the films. Figure 6b displays typical loops measured
h
Esnd ¼ g 1  1 ab  G1 ða þ bÞh (3) at 300 K under in-plane and out-of-plane configurations. The
t
magnetization curves reveal that the macroscopic behavior
corresponds to an in-plane easy axis orientation with an anisotropy
where g 1 ¼ us  ui  ut and G1 ¼ utcosu þ uscotu  uicotu  field mo HK  500 mT, as deduced from the saturation field on the
2uecscu. The u terms correspond to surface energies (per unit hard axis magnetization hysteresis loop. This planar easy magnetic
area) of substrate, of the island’s top and edge side facets and axis is that expected for in-plane tensile LSMO/STO films[38,39]
the dot–substrate interface, respectively, h is the dot height, where both the shape anisotropy energy (Es ¼ mo Ms2/2) and the
u is the angle of the lateral surfaces with the substrate, and t is the magnetoelastic coupling contribute to the in-plane easy axis
thickness of the equivalent flat film. magnetic anisotropy. Magnetoelastic energy can also contribute to
The third term Eelnd in (Eq. 2) is the elastic relaxation energy of the magnetic anisotropy. This term can be written for strained thin
the epitaxial nanodots associated to a non-uniform strain films as Eme ¼ 3/2l(s?cos2u þ s//cos2u) where u is the angle
distribution within the island and it can be written for 3D islands between the strain and magnetization, l is the magnetostriction
as:[36] coefficient (l > 0), and si is the stresses along the two mutually
perpendicular directions (si ¼ 3eY/2 with Y the Young’s modulus,
    Y ¼ 5  1011 N m2).[40,41]
a b
Eelnd ¼ Lh2 b ln þ ln (4) The observed anisotropy field (mo HK  500 mT) is very close to
Qh Qh
that estimated from the shape anisotropy (mo Ms  465 mT),
therefore, we must conclude that the overall in-plane macroscopic
L is defined as L ¼ f 2(1  n)/pm, where m is the shear modulus, magnetic anisotropy of the nanostructured LSMO films is
n is Poisson’s ratio, f ¼ (sa þ sb)/2 is the elastic force monopole preserved when interfacial Sr–La oxide nanodots are created
along the island edge induced by the lattice mismatch, si are the and it is dominated by shape anisotropy. This is consistent with the
diagonal island stress tensor components and Q ¼ eb3/2, where b estimated ratio of Eme/Es  0.1, taking into account that at 300 K,
is a numerical factor.[36] The spontaneous occurrence of out- l  104.[41] Actually, below the self-assembled nanodots the
cropping means that the self-assembled phase-separated state has LSMO film has a null magnetoelastic coupling term because
a lower energy Esa < Eelnc. In Equation (2), the only term that it is s? ¼ s// and hence Eme ¼ 3/2l(s?cos2u þ s//sin2u) ¼ 3/
usually positive is Esnd, while the other three are negative, and 2ls(cos2u þ sin2u) ¼ 3/2lsi which means that this term does
hence in the present LSMO/LSO nanocomposite films they are not contribute below the islands to the magnetic anisotropy (there
dominant. We can then conclude that the outcropping process is is no dependence on the orientation of the magnetization).
enhanced when the in-plane stresses between the island and the Nanoscale electrical analysis was performed by CSAFM at room
wetting layer is large. temperature. By applying a voltage bias between the LSMO surface
We turn now to the analysis of the consequences of the new self- and the conducting tip, a current flow is generated. This current is
assembled nanostructures on the local properties of the LSMO used to create a spatially resolved current (and hence, conductance)
films. We first measured, however, the macroscopic magnetic and image simultaneously to the topography one. The measurements
transport properties of the films through magnetic hysteresis loop were recorded at room temperature, either in a dry nitrogen gas
measurements, including the anisotropic magnetic behavior, and environment or under vacuum conditions (109 mbar) to avoid

2144 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2009, 19, 2139–2146
www.afm-journal.de

It is proposed that the driving force for the

FULL PAPER
generation of the surface nanodots is a
partial release of the elastic energy when
the nanodots are formed at the surface.
This is the same driving force of island
formation that is in the Stranski–Krastanov
mechanism and hence the new mechan-
ism reported here can be considered as a
generalized case, though the problem is
considerably more complex because it also
involves compositional separation of two
phases. The spontaneous outcropping of
self-assembled nanodots on top of an
epitaxial wetting layer with a different
composition appear then as a new oppor-
tunity to modulate at the nanoscale the
physical properties of the underlying
wetting film or to generate nanotemplates
in a film surface.

4. Experimental
LSMO films with superficial Sr–La oxide nano-
dots were deposited on (100) SrTiO3 (STO)
substrates by spin coating at 6000 rpm for 2 min
Figure 7. a) AFM topography image of an LSMO thin film grown on an STO substrate at 1000 8C and annealed at temperatures in the range 900–
during 12 h with the nonmagnetic-insulating (La,Sr)Ox nanodots imaged on the surface; b) CSAFM 1000 8C under flowing oxygen gas for different
image corresponding to the same surface displayed in (a); c) profile following the lines indicated in (a) annealing times up to 12 h [17]. Final film
and (b) of the topography and the current measured with CSAFM. thickness was estimated to be in the range 20–
26 nm by AFM and cross-sectional TEM. X-Ray
humidity-induced electrochemical reactions. Figure 7a shows a diffraction reciprocal space maps were obtained using a Bruker D8
typical topography image with its corresponding current map Advance difractometer equipped with a four-angle goniometer and a
general angle diffraction detection system (GADDS) 2D detector (Cu Ka
(Fig. 6b) obtained with a bias of 2.7 V scanning the biased tip from
1 radiation, no monochromator). The films surface morphology was
left to right at a speed of 0.5 mm s and with 50 nN load. The investigated with a PicoSPM AFM from Molecular Imaging, using
current measured on the LSMO film reach 5 nA, the maximum intermittent contact mode at room temperature which allowed to evidence
current measurable by the detector. different evolution stages of the nanostructures. Statistical analyses of the
Moreover, the current recorded on the LSO islands drastically AFM images were carried out by the software Mountains (Digital Surf).
drops to zero, which accounts for their insulating nature. In Additionally, current sensing AFM (CSAFM) measurements were
performed to analyze the metallic or insulating character of the
conclusion, these results confirm that while the LSMO film
nanostructured films with an Omicron Instruments UHV AFM/STM
preserves its conducting character the behavior of the LSO system by standard contact mode with a silicon AFM tip coated with a
nanodots is fully insulating. boron-doped diamond conducting film (CDT-FMR, Nanosensors). TEM
investigations were carried out with a 200 kV Jeol JEM-2010F field emission
gun microscope equipped with an electron energy loss spectrometer
(EELS) with a nominal energy resolution of 0.8 eV, and a Jeol JEM-2011
3. Conclusions microscope equipped with an EDS. Energy filtered images were obtained
using a 200 kV Cs corrected F20-SACTEM Tecnai microscope fitted with a
In conclusion, we have presented a new mechanism for the GIF-Tridiem energy filter. Cross-sectional samples for TEM examination
creation of self-assembled epitaxial nanostructures in a film were prepared by the conventional cutting, gluing and grinding procedure,
surface. When an epitaxial nanocomposite film is formed, based followed by a final Ar milling step down to perforation. Magnetic
on the phase separation of two crystallographic dissimilar measurements have been performed using a superconducting quantum
structures, diffusion towards the surface of the film of the interference device (SQUID) magnetometer provided with a 7T magnet to
check the properties of the LSMO layers. Electrical resistivity and
minoritary phase may occur spontaneously leading to the
magnetoresistance were obtained using a physical properties measure-
formation of self-assembled islands. The two crystalline phases ment system (PPMS, Quantum Design) equipped with a 9T magnet with a
investigated here have been LSMO and an unknown insulating standard four-point probe method.
LSO phase which appear by TEM to display a coherently strained
structure, even though the strain state and the lattice symmetry of
the LSMO film underneath the islands have been modified. We
have shown that the concentration and size of the insulating Acknowledgements
nanoislands can be tuned through modification of the initial We acknowledge the financial support from Spanish MEC (NANOARTIS,
composition of the chemical solution and growth kinetics control. MAT2005-02047; NANOFUNCIONA, NAN2004-09133-CO3-01; NANO-

Adv. Funct. Mater. 2009, 19, 2139–2146 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2145
www.afm-journal.de

SELECT, CSD2007-00041 FPU, AP2005-4669 FPU), Generalitat de [18] J. Brous, I. Fankuchen, E. Banks, Acta Crystallogr. 1953, 6, 67.
FULL PAPER

Catalunya (Catalan Pla de Recerca SGR-0029 and CeRMAE), CSIC (PIF- [19] A. Hammouche, E. Siebert, A. Hammou, Mater. Res. Bull. 1989, 24,
CANNAMUS) and EU (HIPERCHEM, NMP4-CT2005-516858). The Cs 367.
corrected F20-FEI electron microscope was used through the European [20] J. L. Maurice, F. Pailloux, A. Barthelemy, O. Durand, D. Imhoff, R. Lyonnet,
project ESTEEM (contract no. 026019). Valuable discussions with I.C.
A. Rocher, J. P. Contour, Philos. Mag. 2003, 83, 3201.
Infante are gratefully acknowledged. We also acknowledge electron
[21] A. Cavallaro, F. Sandiumenge, J. Gazquez, T. Puig, X. Obradors, J. Arbiol,
microscopy facilities from Servei de Microscòpia de la UAB and Servei
H. C. Freyhardt, Adv. Funct. Mater. 2006, 16, 1363.
de Microscòpia de la UB.
[22] P. A. Langjahr, F. F. Lange, T. Wagner, M. Ruhle, Acta Mater. 1998, 46, 773.
Received: January 19, 2009 [23] A. Pomar, M. Coll, A. Cavallaro, J. Gazquez, J. C. Gonzalez, N. Mestres, F.
Published online: June 2, 2009 Sandiumenge, T. Puig, X. Obradors, J. Mater. Res. 2006, 21, 1106.
[24] T. Walther, Ultramicroscopy 2003, 96, 401.
[1] C. A. Ross, Annu. Rev. Mater. Res. 2001, 31, 203. [25] S. N. Ruddlesden, P. Popper, Acta Crystallogr. 1958, 11, 54.
[2] H. Zheng, J. Wang, S. E. Lofland, Z. Ma, L. Mohaddes-Ardabili, T. Zhao, L. [26] R. Seshadri, C. Martin, M. Hervieu, B. Raveau, C. N. R. Rao, Chem. Mater.
Salamanca-Riba, S. R. Shinde, S. B. Ogale, F. Bai, D. Viehland, Y. Jia, D. G. 1997, 9, 270.
Schlom, M. Wuttig, A. Roytburd, R. Ramesh, Science 2004, 303, 661. [27] R. Seshadri, M. Hervieu, C. Martin, A. Maignan, B. Domenges, B. Raveau,
[3] S. R. Foltyn, L. Civale, J. L. Manus-Driscoll, Q. X. Jia, B. Maiorov, H. Wang, A. N. Fitch, Chem. Mater. 1997, 9, 1778.
M. Maley, Nat. Mater. 2007, 6, 631. [28] M. Foex, Bull. Soc. Chim. France 1961, 109.
[4] J. Gutierrez, A. Llordes, J. Gazquez, M. Gibert, N. Roma, S. Ricart, A. [29] A. R. Schulze, H. Mullerbuschbaum, Z. Anorg. Allg. Chem. 1980, 471,
Pomar, F. Sandiumenge, N. Mestres, T. Puig, X. Obradors, Nat. Mater. 59.
2007, 6, 367. [30] International Centre for Diffraction Data No. 00–022–1430, 2008.
[5] J. V. Barth, G. Costantini, K. Kern, Nature 2005, 437, 671. [31] International Centre for Diffraction Data No. 00–022–1431, 2008.
[6] V. A. Shchukin, D. Bimberg, Rev. Mod. Phys. 1999, 71, 1125. [32] International Centre for Diffraction Data No. 00–042–0343, 2008.
[7] C. Teichert, Phys. Rep. 2002, 365, 335. [33] G. van Tendeloo, O. I. Lebedev, M. Hervieu, B. Raveau, Rep. Prog. Phys.
[8] M. Gibert, T. Puig, X. Obradors, A. Benedetti, F. Sandiumenge, R. Hühne, 2004, 67, 1315.
Adv. Mater. 2007, 19, 3937. [34] B. S. Kang, H. Wang, J. L. Manus-Driscoll, Y. Li, Q. X. Jia, I. Mihut, J. B.
[9] M. Gibert, T. Puig, X. Obradors, Surf. Sci. 2007, 601, 2680. Betts, Appl. Phys. Lett. 2006, 88, 192514.
[10] H. M. Zheng, F. Straub, Q. Zhan, P. L. Yang, W. K. Hsieh, F. Zavaliche, Y. H. [35] J. Tersoff, F. K. Legoues, Phys. Rev. Lett. 1994, 72, 3570.
Chu, U. Dahmen, R. Ramesh, Adv. Mater. 2006, 18, 2747. [36] J. C. Nie, H. Yamasaki, Y. Mawatari, Phys. Rev. B 2007, 70, 195421.
[11] V. Moshnyaga, B. Damaschke, O. Shapoval, A. Belenchuk, J. Faupel, O. I. [37] A. Urushibara, Y. Moritomo, T. Arima, A. Asamitsu, G. Kido, Y. Tokura,
Lebedev, J. Verbeeck, G. van Tendeloo, M. Mucksch, V. Tsurkan, R. Tidecks, Phys. Rev. B 1995, 51, 14103.
K. Samwer, Nat. Mater. 2003, 2, 247. [38] A. M. Haghiri-Gosnet, J. Wolfman, B. Mercey, C. Simon, P. Lecoeur, M.
[12] J. M. D. Coey, M. Viret, S. von Molnar, Adv. Phys. 1999, 48, 167. Korzenski, M. Hervieu, R. Desfeux, G. Baldinozzi, J. Appl. Phys. 2000, 88,
[13] M. B. Salamon, M. Jaime, Rev. Mod. Phys. 2001, 73, 583. 4257.
[14] Y. Tokura, Rep. Prog. Phys. 2006, 69, 797. [39] M. Ziese, H. C. Semmelhack, P. Busch, J. Magn. Magn. Mater. 2002, 246,
[15] U. Hasenkox, C. Mitze, R. Waser, J. Am. Ceram. Soc. 1997, 80, 2709. 327.
[16] U. Hasenkox, C. Mitze, R. Waser, R. R. Arons, J. Pommer, G. Guntherodt, J. [40] B. D. Cullity, Introduction to Magnetic Materials, Addison-Wesley, Reading,
Electroceram. 1999, 3, 255. MA 1972.
[17] A. Hassini, A. Pomar, J. Gutierrez, M. Coll, N. Roma, C. Moreno, A. Ruyter, [41] K. Steenbeck, T. Habisreuther, C. Dubourdieu, J. P. Senateur, Appl. Phys.
T. Puig, X. Obradors, Supercond. Sci. Technol. 2007, 20, S230. Lett. 2002, 80, 3361.

2146 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2009, 19, 2139–2146

Das könnte Ihnen auch gefallen