10
research papers J. Appl. Cryst. (2014). 47 doi:10.1107/S1600576714021049 1 of 10 Journal of Applied Crystallography ISSN 1600-5767 Received 28 July 2014 Accepted 22 September 2014 # 2014 International Union of Crystallography Self-organization of Fe clusters on mesoporous TiO 2 templates Patrick Ziegler, a Neelima Paul, b Peter Mu ¨ller-Buschbaum, c Birgit Wiedemann, a Wolfgang Kreuzpaintner, a Jaru Jutimoosik, d Rattikorn Yimnirun, d Annette Setzer, e Pablo Esquinazi, e Peter Bo ¨ni a and Amitesh Paul a * a Technische Universita ¨t Mu ¨ nchen, Physik Department E21, Lehrstuhl fu ¨ r Neutronenstreuung, James-Franck-Strasse 1, D-85748 Garching bei Mu ¨ nchen, Germany, b Forschungs-Neutronenquelle Heinz Maier-Leibnitz, Lichtenbergstrasse 1, 85748 Garching bei Mu ¨ nchen, Germany, c Technische Universita ¨ t Mu ¨ nchen, Physik Department E13, Lehrstuhl fu ¨ r Funktionelle Materialien, James-Franck- Strasse 1, D-85748 Garching bei Mu ¨ nchen, Germany, d School of Physics, Institute of Science and NANOTEC-SUT Center of Excellence on Advanced Functional Nanomaterials, Suranaree University of Technology, Nakhon Ratchasima 30000, Thailand, and e Division of Superconductivity and Magnetism, University of Leipzig, D-04103 Leipzig, Germany. Correspondence e-mail: [email protected] Fe layers with thicknesses between 5 and 100 nm were sputtered on mesoporous nanostructured anatase TiO 2 templates. The morphology of these hybrid films was probed with grazing-incidence small-angle X-ray scattering and X-ray reflectivity, complemented with magnetic measurements. Three different stages of growth were found, which are characterized by different correlation lengths for each stage. The magnetic behavior correlates with the different growth regimes. At very small thicknesses the TiO 2 template is coated and a porous Fe film results, with in-plane and out-of-plane magnetization components. With increasing thickness, agglomeration of Fe occurs and the magnetization gradually turns mostly in plane. At large thicknesses, the iron grows independently of the template and the magnetization is predominantly in plane with a bulk-like characteristic. 1. Introduction In the pursuit of spintronic materials, suitable for devices that are compatible with silicon technology, self-organized struc- tures are often preferred. Self-organization (regular distribu- tion) is a low-cost alternative way to fabricate materials structured down to a few nanometres. It can result in inter- acting magnetic memory dots which are suitable for spin injection in semiconductors. Recently, Matthieu et al. (2006) found a high-ordering-temperature (T C ) ferromagnetic phase in self-organized Mn-rich nanocolumns that are surrounded by an Mn-poor matrix. It is obvious that understanding the properties of metal– oxide interfaces and their formation processes is of great importance to optimize the characteristics of spintronic devices. Depending on the preparation conditions and the reactivity of the involved species, single metal atoms can diffuse into the magnetic oxide and act as a dopant. However, so far little work has been done on the growth and structure development of metal layers on self-structured semi- conducting oxides. The metal–semiconductor morphology depends on the chemical interaction, the growth behavior and the interface properties, which determine the diffusivity. A promising way to achieve room-temperature ferro- magnetism in ferromagnetic semiconductors was proposed by Dietl et al. (2000) in the form of the so-called diluted magnetic semiconductors (DMSs). Since then, there have been numerous reports of ferromagnetic semiconductors, involving a variety of semiconducting materials such as ZnO, SnO 2 , Ge, GaN or GaAs doped with transition metals like Fe, Co, Mn, Ni etc. (Kobayashi et al., 2005; Fitzgerald et al., 2006; Dumm et al., 2000; Filipe & Schuhl, 1997). However, there are several problems for these DMS systems as they suffer from secondary phases (inhomogeneities), leading to magnetic nanoclustering inside the semiconductor. Recent research revealed that certain defects, like vacancies and/or atoms, help to stabilize a magnetically ordered phase in several oxides (Esquinazi et al., 2013). In particular, TiO 2 has often been used as wide-gap dilute magnetic oxide. Even though the rutile phase of TiO 2 is thermodynamically stable, it has a high mobility of n-type charge carriers and a large thermopower. In the present investigation we use nanostructured porous (pore size about 20–50 nm) TiO 2 network films as templates for the self-assembly (random distribution) of Fe. The film morphology is probed with grazing-incidence small-angle X-ray scattering (GISAXS) (Naudon et al., 1998; Mu ¨ ller- Buschbaum, 2003; Perlich et al., 2009). In principle, the TiO 2 network is expected to restrict the mobility, forming a dot-like mesh of Fe atoms. However, one should note that we are not

Self-organization of Fe clusters on mesoporous TiO 2 templates

  • Upload
    amitesh

  • View
    213

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Self-organization of Fe clusters on mesoporous TiO               2               templates

research papers

J. Appl. Cryst. (2014). 47 doi:10.1107/S1600576714021049 1 of 10

Journal of

AppliedCrystallography

ISSN 1600-5767

Received 28 July 2014

Accepted 22 September 2014

# 2014 International Union of Crystallography

Self-organization of Fe clusters on mesoporous TiO2

templates

Patrick Ziegler,a Neelima Paul,b Peter Muller-Buschbaum,c Birgit Wiedemann,a

Wolfgang Kreuzpaintner,a Jaru Jutimoosik,d Rattikorn Yimnirun,d Annette Setzer,e

Pablo Esquinazi,e Peter Bonia and Amitesh Paula*

aTechnische Universitat Munchen, Physik Department E21, Lehrstuhl fur Neutronenstreuung,

James-Franck-Strasse 1, D-85748 Garching bei Munchen, Germany, bForschungs-Neutronenquelle

Heinz Maier-Leibnitz, Lichtenbergstrasse 1, 85748 Garching bei Munchen, Germany, cTechnische

Universitat Munchen, Physik Department E13, Lehrstuhl fur Funktionelle Materialien, James-Franck-

Strasse 1, D-85748 Garching bei Munchen, Germany, dSchool of Physics, Institute of Science and

NANOTEC-SUT Center of Excellence on Advanced Functional Nanomaterials, Suranaree University

of Technology, Nakhon Ratchasima 30000, Thailand, and eDivision of Superconductivity and

Magnetism, University of Leipzig, D-04103 Leipzig, Germany. Correspondence e-mail:

[email protected]

Fe layers with thicknesses between 5 and 100 nm were sputtered on mesoporous

nanostructured anatase TiO2 templates. The morphology of these hybrid films

was probed with grazing-incidence small-angle X-ray scattering and X-ray

reflectivity, complemented with magnetic measurements. Three different stages

of growth were found, which are characterized by different correlation lengths

for each stage. The magnetic behavior correlates with the different growth

regimes. At very small thicknesses the TiO2 template is coated and a porous Fe

film results, with in-plane and out-of-plane magnetization components. With

increasing thickness, agglomeration of Fe occurs and the magnetization

gradually turns mostly in plane. At large thicknesses, the iron grows

independently of the template and the magnetization is predominantly in plane

with a bulk-like characteristic.

1. IntroductionIn the pursuit of spintronic materials, suitable for devices that

are compatible with silicon technology, self-organized struc-

tures are often preferred. Self-organization (regular distribu-

tion) is a low-cost alternative way to fabricate materials

structured down to a few nanometres. It can result in inter-

acting magnetic memory dots which are suitable for spin

injection in semiconductors. Recently, Matthieu et al. (2006)

found a high-ordering-temperature (TC) ferromagnetic phase

in self-organized Mn-rich nanocolumns that are surrounded

by an Mn-poor matrix.

It is obvious that understanding the properties of metal–

oxide interfaces and their formation processes is of great

importance to optimize the characteristics of spintronic

devices. Depending on the preparation conditions and the

reactivity of the involved species, single metal atoms can

diffuse into the magnetic oxide and act as a dopant. However,

so far little work has been done on the growth and structure

development of metal layers on self-structured semi-

conducting oxides. The metal–semiconductor morphology

depends on the chemical interaction, the growth behavior and

the interface properties, which determine the diffusivity.

A promising way to achieve room-temperature ferro-

magnetism in ferromagnetic semiconductors was proposed by

Dietl et al. (2000) in the form of the so-called diluted magnetic

semiconductors (DMSs). Since then, there have been

numerous reports of ferromagnetic semiconductors, involving

a variety of semiconducting materials such as ZnO, SnO2, Ge,

GaN or GaAs doped with transition metals like Fe, Co, Mn, Ni

etc. (Kobayashi et al., 2005; Fitzgerald et al., 2006; Dumm et al.,

2000; Filipe & Schuhl, 1997). However, there are several

problems for these DMS systems as they suffer from

secondary phases (inhomogeneities), leading to magnetic

nanoclustering inside the semiconductor. Recent research

revealed that certain defects, like vacancies and/or atoms, help

to stabilize a magnetically ordered phase in several oxides

(Esquinazi et al., 2013). In particular, TiO2 has often been used

as wide-gap dilute magnetic oxide. Even though the rutile

phase of TiO2 is thermodynamically stable, it has a high

mobility of n-type charge carriers and a large thermopower.

In the present investigation we use nanostructured porous

(pore size about 20–50 nm) TiO2 network films as templates

for the self-assembly (random distribution) of Fe. The film

morphology is probed with grazing-incidence small-angle

X-ray scattering (GISAXS) (Naudon et al., 1998; Muller-

Buschbaum, 2003; Perlich et al., 2009). In principle, the TiO2

network is expected to restrict the mobility, forming a dot-like

mesh of Fe atoms. However, one should note that we are not

Page 2: Self-organization of Fe clusters on mesoporous TiO               2               templates

concentrating on achieving a high-TC dilute magnetic oxide in

this article; rather, we focus on the aspect of self-organization

of a magnetic metal on a structured semiconducting oxide film.

As a consequence, in the present case, the Fe thicknesses are

chosen to be so large that they cannot be considered as a

dilute species in the semiconductor. We focus on three

different stages of growth of the nanoclusters. Interestingly, we

find a direct correlation of the growth morphology and

structure with the magnetic response for these stages.

2. Sample preparation

The nanostructured templates used for depositing Fe films

were anatase TiO2 mesh-like structures. These TiO2 templates

were prepared on Si substrates using the block copolymer

assisted sol–gel method, which resulted in foam-like porous

titania network films. Details of this preparation method are

described by Cheng & Gutmann (2006) and Niedermeier et al.

(2012). Fe layers with five different thicknesses were grown on

top of the nanostructured TiO2 surface by magnetron sput-

tering. The corresponding samples are labeled as Fed, with d =

0, 5, 10, 20, 50 and 100 nm. We note that Fe layers with

thicknesses of �1–3 nm were also deposited on similar titania

templates (but grown with a different block copolymer base).

We exclude them in the present work since they do not exhibit

any magnetic response for the as-deposited specimens, as this

usually becomes apparent after annealing. During deposition,

the Ar pressure in the magnetron sputtering chamber was 3 �

10�3 mbar (1 mbar = 100 Pa). The process was started at a

base pressure of 5� 10�7 mbar. The Fe was deposited at room

temperature and at a rate of 0.07 nm s�1.

3. Experimental methods

The nanostructure at the sample surface has been probed with

scanning electron microscopy (SEM). Atomic force micro-

scopy (AFM) to identify the size and the structure and

magnetic force microscopy (not shown) to identify the

potential magnetic domains were also performed. The struc-

tures of the TiO2 template and of the Fe films on top of these

templates have been investigated with conventional X-ray

diffraction (XRD) methods. Thin-film characterization was

based on measurements using GISAXS, diffuse X-ray scat-

tering (XDS) and X-ray reflectivity (XRR).

We performed GISAXS measurements in noncoplanar

scattering geometry to map the lateral correlations using a

Ganesha 300 XL SAXS–WAXS system (SAXSLABApS,

Copenhagen, Denmark). The X-ray radiation was produced at

50 kV/0.6 mA from a Cu anode with a wavelength of � =

0.154 nm. The horizontal and vertical beam size was 0.4 and

0.3 mm with a divergence of 1 and 0.1 mrad, respectively. The

in-plane � were in the range of �1.15�. The chosen incident

angle of the X-rays of �i ¼ 0:5� was near the critical angle �c

for total reflection of our samples. A Pilatus 300K solid-state

two-dimensional photon-counting detector with a resolution

of 172 mm at a detector-to-sample distance of around 1.058 m

was used to record the intensity at room temperature. The

XDS and XRR measurements were performed with a D5000

instrument operating at 40 kV/40 mA from a Cu anode with a

wavelength of � = 0.154 nm.

X-ray absorption near-edge structure (XANES) spectra

were acquired at Beamline 5 (SUT-NANOTEC-SLRI) (with

electron energy of 1.2 GeV, bending magnet, beam current 80–

150 mA, 1.1–1.7 � 1011 photons s�1) at the Synchrotron Light

Research Institute (SLRI), Nakhon Ratchasima, Thailand.

The Fe K-edge spectra were measured in the fluorescent mode

with a four-element Si drift detector. A double-crystal

Ge(220) monochromator was used to scan the energy of the

synchrotron X-rays with the range of 7100–7200 eV and an

energy step of 0.2 eV for Fe K-edge XANES spectra. The

XANES region corresponds to the excitation of core electrons

to unoccupied bound states or to low-lying continuum states.

Thus, XANES provides information about the chemical

composition of the deposited Fe with various thicknesses.

Different scattering geometries at grazing incidence,

measuring specular reflection, scattering in the plane of inci-

dence and scattering perpendicular to the plane of incidence,

were employed. This way one can estimate the range of

accessible correlation lengths from the momentum transfers

along the three different axes owing to the scattering

geometry for small angles.

The total scattering vector is the difference of the wave-

vector of the incident beam and the scattered beam:

Q ¼ ki � kf; ð1Þ

Qx

Qy

Qz

0@

1A ¼ k

cos �f cos�� cos�i

cos �i sin�sin �i þ sin �f

0@

1A; ð2Þ

where �i is the incidence angle of the X-rays and �f is the angle

of the scattered wavevector k (k = |k| = 2�/�) in the xz plane.

Here � is the wavelength and � is the angle in the xy plane,

which is relevant to determine lateral correlation lengths. A

schematic of the GISAXS geometry is shown in Fig. 1.

XRR measurements only provide information on the

thickness and roughness of the layers, since the signal is

research papers

2 of 10 Patrick Ziegler et al. � Self-organization of Fe clusters J. Appl. Cryst. (2014). 47

Figure 1Sketch of the typical scattering geometry for GISAXS measurementsfrom a sample. The sample is represented here by an atomic forcemicroscopy image of Fe100 in the QxQy plane and the correspondingdetector image in the QzQy plane.

Page 3: Self-organization of Fe clusters on mesoporous TiO               2               templates

independent of Qy and Qx because �i ¼ �f and � ¼ 0 in this

geometry. The corrections in these measurements are

obtained by longitudinal off-specular measurements (XDS) at

an offset angle (� ¼ 0:15�) of the specular direction. The

lateral components of the scattering vector Q are measured

using off-specular scans and GISAXS. The transverse scans

along Qx are measured as a function of Qz. Information about

horizontal structures was extracted from horizontal line cuts

from the two-dimensional GISAXS data, which were taken at

the position of the Yoneda peak (Yoneda, 1963), the maxima

of the Fresnel transmission coefficients.

Conventional in-plane and out-of-plane magnetization

loops were measured at various temperatures and fields using

a superconducting quantum interference device (SQUID)

from Quantum Design. The temperature dependence of the

magnetization was measured in a field of 10 mTafter zero-field

cooling (ZFC) and after field cooling (FC) in a field of 1 T.

4. Results and discussion

4.1. Scanning probe microscopy by SEM and AFM

The nanostructure of the substrate (Fe0) is shown in the

SEM image in Fig. 2(a). The surface indicates a very rough

texture, which is expected because the preparation process

leads to a random network structure (Kaune et al., 2009, 2010).

Fig. 2(b) shows a SEM image of Fe100 with identical magnifi-

cation. The surface looks smoother, indicating that the

template is poorly replicated in the case of the thick Fe layer.

Although the topography gets blurred, the distance between

the two brighter or sharper regions (�25 nm) remains of the

same order of magnitude.

The AFM image of Fe0, the pure titania surface, is shown in

Fig. 3(a), and as an example that of Fe100, the thickest Fe film

on the template, is shown in Fig. 3(b). The height of the

structure is roughly estimated to be h0 ¼ 33 (2) nm. Note that

h0 is fairly constant for all six samples, which means that there

are no filling effects caused by sputtering and the change in the

SEM images is due to a variation in their jaggedness only. A

Fourier transformation of the images, which is shown in

Figs. 3(c) and 3(d), indicates that the mesh-like titania struc-

ture does not have any well pronounced distances, as they

would be visible in the reciprocal space representation.

Instead, the titania templates are fairly isotropic. This isotropy

is maintained if Fe is deposited on top of the titania templates,

in particular for Fe100.

4.2. XRD, XRR and XDS measurements and analysis

Fig. 4 shows an example of the XRD data of Fe20, as a

representative deposited Fe film on the titania nanostructure

template. Note that all six samples are grown on identical

substrates using identical methods and values for deposition

and can be assumed as identical except for the layer thickness.

The data confirm the presence of Fe on a template which has

an anatase structure – a polymorph of TiO2 (Cheng &

Gutmann, 2006). The average grain size of the polycrystalline

Fe20 sample is calculated from the FWHM of the peaks in the

diffraction pattern as �1.98 nm for anatase TiO2 and 3.33 nm

([110]) and 25 nm ([100]) for Fe crystallites using the Scherrer

research papers

J. Appl. Cryst. (2014). 47 Patrick Ziegler et al. � Self-organization of Fe clusters 3 of 10

Figure 2SEM images of (a) the pure anatase surface of TiO2 in Fe0 and (b) thesurface of Fe in Fe100 with the thickest Fe layer (100 nm). The thick Felayer leads to a blurring of the surface image. The insets show a zoomedportion of the images.

Figure 3AFM images of (a) the pure anatase surface of TiO2 in Fe0 and (b) thesurface of Fe in Fe100 with the thickest Fe layer (100 nm). CorrespondingFourier transformations of the AFM data, showing ring-shaped intensity,for (c) Fe0 and (d) Fe100, which have been analyzed to identify somesignature of a regular structure from that of a random structure.

Figure 4Representative XRD measurement of sample Fe20 with marked Braggpeaks of the Si substrate, of Fe 110 and of anatase TiO2 101. The jumps inthe intensity around the Si 400 peaks are due to misalignment of the beamwith respect to the substrate plane.

Page 4: Self-organization of Fe clusters on mesoporous TiO               2               templates

formula. Grains can often be agglomerates of many small

crystallites that are seen by XRD.

The XRR and the XDS data (scans along Qz) are shown in

Fig. 5(a) and the true specular data (off-specular corrected)

along with their fits in Fig. 5(b). Subtracting the longitudinal

offset data from the specular data gives cleaner specular data

(true specular) devoid of the off-specular contribution that

may have been included owing to possible relaxed resolution

along Qx. For most practical purposes it does not make any

significant difference in the fit parameters, particularly for

bilayer-type systems (unlike the case of a periodic multilayer

system) with small roughness, since the off-specular intensities

remain much lower than the specular part. The reflectivity

data were analyzed by means of a Parratt fitting routine used

to calculate the optical reflectivity following an iterative model

within the dynamic scattering theory (Parratt, 1954; Daillant &

Gibaud, 2009). The film was modeled as consisting of layers of

specific thickness, roughness and scattering length density

(SLD). An intermediate layer is not required, indicating that

there is little interdiffusion. The roughness is very high owing

to the nanostructure of the sample and physical information

about the roughness is irrelevant. The thickness of the anatase

titania layer [assumed from the sputtering rate to be da =

100 (10) nm] and SLD (estimated to be around 3.68 �

10�3 nm�2) were found to be very similar for all samples, as

shown in Fig. 6(a). A variation of thickness (which matches

quite well with the nominal ones) and SLD for the Fe layer

and the top oxide layer was obtained. The Kiessig oscillations

are very poor owing to the high roughness (�3.0 nm), which

can give some error in the estimation.

The information about the SLD (re�) or the electron

density (�) that can be extracted from the position of the

critical angle at the edge of total reflection (�2c = �2re�/�) can

also be used to determine the porosity � using the following

formula: � ¼ 1� �simul=�theory. Here re is the electron radius

and � is given by the atomic density (N) � the atomic number

(Z). The term SLD is generally used in the case of neutron

reflectivity but can be used for X-rays as well owing to its

expressional similarity. The porosity of Fe0 could be deter-

mined as � ¼ 0:56 for the TiO2 template. Compared with the

highest reported porosities of titania network-type nano-

structured films, which were prepared via a block copolymer

assisted sol–gel synthesis, this value is very moderate (Kaune

et al., 2009, 2010). For samples with Fe, the scattering length

density (�simul) is a mixture of the SLDs of Fe and its oxides

(FeO, Fe2O3, Fe3O4) along with Ti and its oxide (TiO2). It is

therefore difficult to differentiate the porosity of Ti from that

of Fe, as the corresponding critical angles are juxtaposed. The

variation of SLDFe has been plotted in Fig. 6(b), showing its

evolution as it approaches the bulk value of Fe (�5.8 �

10�3 nm�2).

The Qz versus Qx map for the example of sample Fe20 is

shown in Fig. 7. No Bragg sheet intensities along the hori-

zontal axis are seen, indicating no vertical correlation from

correlated interface roughness of the structure (Holy &

Baumbach, 1994; Daillant & Belorgey, 1992). A simulation of

the data was calculated within the distorted-wave Born

research papers

4 of 10 Patrick Ziegler et al. � Self-organization of Fe clusters J. Appl. Cryst. (2014). 47

Figure 6(a) SLD profiles of the samples. (b) SLD profile of Fe layer versus dFe.

Figure 5(a) XRR and XDS data along with (b) the off-specular corrected (truespecular) data (symbols) and fits (solid cyan lines) calculated with theParratt algorithm to determine the SLD and the roughness of eachsample. The different samples are indicated as explained in the text. Thecurves are shifted along the intensity axis for clarity of the presentation.The XRR signals from Fe50 and Fe100 are particularly overshadowed bythe comparable off-specular signal from the underlying nanostructureeven at low angles, signifying their low reflectivity.

Page 5: Self-organization of Fe clusters on mesoporous TiO               2               templates

approximation (DWBA) according to the model of Ming et al.

(1993), yielding a lateral correlation length �k ’ 30 nm, as

shown in Fig. 8. This model describes an intermediate case

between vertically uncorrelated and vertically correlated

layers. It assumes that the vertical correlations do not depend

on the lateral size of the roughness.

4.3. XANES

Fig. 9 shows a comparison of the measured Fe K-edge

XANES spectra from the samples (solid symbols) and the

reference spectra from each of the possible constituents that

can produce the absorption edge, for example, Fe, FeO, Fe3O4

and Fe2O3. By considering Fe, FeO, Fe3O4 and Fe2O3 as the

parent components, the XANES spectra of the samples were

fitted (lines) with a superposition of XANES profiles of the

parent components using the linear combination analysis

(LCA) method. The fitting was performed using the package

ATHENA (Ravel & Newville, 2005) with the LCA tool. The

fits are shown in the figure together with the measured

XANES spectra. In this way we estimate the weighted

proportions of the constituents in the samples. They are

composed of phase-separated regions that differ negligibly in

the proportion of their respective constituents (Fe, FeO, Fe3O4

and Fe2O3) from sample to sample. The results indicate that

there are negligible traces of FeO in the samples. The main

constituents are Fe with proportions of Fe2O3 and Fe3O4. The

results can be seen in Table 1. The results show an obvious

trend of decreasing percentage of oxidation within the Fe

layer with increasing thickness of the layer.

4.4. GISAXS measurements and analysis

4.4.1. GISAXS measurements. Fig. 10 shows the two-

dimensional GISAXS data as reciprocal space maps Qz versus

Qy for all samples. The GISAXS measurements were taken at

�i = 0.5�. The reason for choosing an angle above the total

reflection edge was to allow the beam to penetrate the entire

film. One can expect to see the specular peak (S) and the

material-specific Yoneda peak (Y) well separated. Moreover,

other scattering features arising from possible correlations of

interfaces can be detected (Salditt et al., 1995; Muller-Busch-

baum & Stamm, 1998).

For an Fe layer thickness of 5–50 nm, the GISAXS patterns

show intensity modulations caused by resonant diffuse scat-

tering (due to scattering from the total thickness) (Lee et al.,

2008). For a further increase in thickness (�50 nm), no

research papers

J. Appl. Cryst. (2014). 47 Patrick Ziegler et al. � Self-organization of Fe clusters 5 of 10

Table 1Overview over the percentage of iron oxide that has been determined byfitting the XANES measurements.

Sample Fe3O4 ratio (%) Fe2O3 ratio (%)

Fe5 43.4 35.3Fe10 39.9 13.1Fe20 21.1 2.5Fe50 14.1 0Fe100 8.1 0

Figure 8XDS measurement (solid symbols) on sample Fe20 for Qz = 0.8 nm�1,shown together with a simulation in the framework of the DWBA todetermine the lateral correlation length along Qx. The positions of theYoneda wings (Y) are indicated with arrows. The mismatch in the widthof the specular peak is due to multiple surfaces of the nanostructures andthe resolution limit of the detector.

Figure 9XANES measurements (symbols) of all samples and corresponding fitsbased on different proportions of Fe oxides. The curves are shifted alongthe y axis for clarity of the presentation.

Figure 7Example of a two-dimensional Qz–Qx map of sample Fe20 measured inX-ray diffuse scattering geometry. The light-blue lines indicate thepositions of the Yoneda wings and the yellow line is the Qz position at0.8 nm�1, where a line cut has been extracted.

Page 6: Self-organization of Fe clusters on mesoporous TiO               2               templates

intensity modulations are visible as they are no longer resol-

vable with the used setup.

The Yoneda peaks are well resolved in all cases as �i = 0.5�

has been chosen. The vertical position of the Yoneda peak

(Y = �f + �c) moves towards higher angles with increasing Fe

thickness, indicating the increased loading with Fe. If the

samples had had well defined lateral structures, well defined

side maxima of the Yoneda peak would be visible in the

horizontal direction of the two-dimensional GISAXS data.

Since there are no such side peaks visible, we can exclude a

high degree of ordering of the nanostructures (e.g. no well

defined nearest neighbor order). Instead of such peaks, the

intensity that we see is broadly spread out in the horizontal

direction in all cases. This is due to the presence of well

resolved lengths scales for structures with a lower degree of

ordering.

4.4.2. GISAXS analysis alongQz. Vertical line cuts from the

two-dimensional GISAXS data (intensity profiles along Qz at

Qy = 0) are shown in Fig. 11 for all six samples. The main

features are the sharp specular peaks and broad Yoneda peaks

for the respective material constituents. With an increase in

the Fe film thickness, the critical angle is gradually shifted till it

reaches �c ’ 0:4�. Fig. 11 shows the evolution of �c with the Fe

layer thickness. Note that the critical angle obtained from the

GISAXS data, for example in the case of Fe20, is very similar

to that obtained from the XRR data (Qz ’ 0.64 nm�1). The

position of �c saturates at an Fe thickness of around 30 nm,

which approximately matches the height (h0) of the template.

By analyzing the difference in the Yoneda position between

Fe0 and Fe100 one can see that there is a superposition of two

distinct Yoneda peaks. One Yoneda peak arises from the

anatase titania network film (marked with an arrow pointing

downwards) and another one from the Fe deposited onto this

film (marked with arrows pointing upwards). From the

Yoneda peak positions one can determine the porosity of the

Fe films on titania. Since the variation of the position of the

Yoneda peak shifts with Fe thickness, we restrict the calcula-

tion to Fe0 only.

The porosity of Fe0 from the position of the Yoneda peak in

the GISAXS pattern gives � ¼ 0:61, which is similar to that

determined by XRR.

4.4.3. GISAXS analysis along Qy. Horizontal line cuts from

the two-dimensional GISAXS data (intensity profiles along

Qy) were taken at the Yoneda peak position of Fe for each

sample. These line cuts are shown in Fig. 12 on a log–log scale,

together with a fit to determine the lateral structure lengths.

The Yoneda wings arise because of the increase in scattering

caused by the maxima in the Frensnel transmission function

and are thus a signature of the overall roughness. The diffuse

scattering along the peak position is due to the regularity of

the nanostructures. The horizontal line cuts do not show well

pronounced maxima but have a broad distribution. Thus the

samples exhibit a distribution of small-scale lateral structures.

Because the polycrystalline material probed here is on top

of a mesoporous material, one can choose the model of

supported islands. In the first approximation, the form factor

of the host material is well represented by a cylindrical model

research papers

6 of 10 Patrick Ziegler et al. � Self-organization of Fe clusters J. Appl. Cryst. (2014). 47

Figure 11Vertical line cuts from the two-dimensional GISAXS data, which havebeen extracted at the position Qy = 0. The arrows mark the Yoneda peaksof the anatase (down arrow) and the Fe structure (up arrows). Theintensity is normalized with respect to the Fe0 data. The line cuts areplotted with an offset for clarity. In the inset we plot the variation of �c,depicting its dependence on the Fe layer thickness. The dash–dot linerepresents the �c value of the anatase phase of TiO2.

Figure 10Two-dimensional GISAXS data measured at an incident angle of �i =0.5� for all samples. All images show the same color scale and an equalarea of the detector. The specular peak (S) and the Yoneda (Y) peak areindicated by arrows in each case.

Page 7: Self-organization of Fe clusters on mesoporous TiO               2               templates

and the structure factor by a Percus–Yevick three-dimensional

type with a local monodisperse approximation of size distri-

bution consisting of two coexistent dominant length scales. For

the Fe deposits on top, although the island sizes are rather

irregular, the assumption of cylindrical particles fits to the data

very well. The peak shape of the Qy profile is a convolution of

the form factor and structure factor involved. The form factor

was considered to be cylindrical, and the cluster size (2R) and

the intercluster distance (�) were taken as the fitting para-

meters. A Gaussian distribution of R was allowed in the fits. A

schematic of the model geometry is shown in Fig. 13.

The fits to the data were done using the effective surface

approximation (ESA) of the DWBA (Muller-Buschbaum,

2009) and are shown in Fig. 12. We have considered an

assembly of two isolated cylinders with radii R1 and R2. The

details of the model used here can be found elsewhere (Sarkar

et al., 2014). The parameters that have been extracted from the

fit are shown in Table 2. For the anatase phase we assume a

cylinder of radius R1 with a correlation length �a (or �R1) in

our model. For Fe, we consider another cylinder of radius R2

with a correlation length �Fe (or �R2). Initially we found �a =

160 (12) nm, while �Fe shows some variation with the thickness

of the Fe film.

The correlation of Fe can be divided into three regimes.

Regime I is for thickness up to 5 nm, regime II is between 10

and 20 nm, and region III is beyond 20 nm. In regime I, Fe

takes a form similar to the underlying anatase template. Note

that the Fe with a cylinder of radius R2 is different from the

underlying cylinder with radius R1. It has a smaller correlation

length (�R2 <�R1) than the template. This is a stage of surface

bonding of Fe to the template. The atoms do not agglomerate

with other atoms because of their low mobility, as the template

radius R1 increases marginally after Fe is deposited, probably

owing to some imbalance in the deposits.

For the next stage, in regime II, we find increasing surface

coverage with the onset of agglomeration. Thus, a drastic

decrease in �R2 (from 80 to 30 nm) occurs as the atoms

agglomerate. However, this does not affect the correlation

length of the underlying template. This trend continues until

Fe20.

In regime III, a slight increase in R2 along with the lateral

correlation length (�R2) is observed for samples beyond Fe50.

In this regime the structures become independent of the

template, as the Fe layer matches the average height of the

underlying TiO2 template, h0, which is �30 nm. With

increasing Fe adatoms the layer grows in height with packed

grains. A schematic of the growth process at different stages is

shown in Fig. 14.

4.5. Magnetization measurements

4.5.1. Magnetization versus field. Fig. 15 shows the SQUID

hysteresis loops at T = 5 K and T = 300 K for Fe5–Fe100. The

diamagnetic signal from Fe0 is subtracted from the hysteresis

loops. The applied field is parallel to the film plane. The

research papers

J. Appl. Cryst. (2014). 47 Patrick Ziegler et al. � Self-organization of Fe clusters 7 of 10

Figure 14Schematic of the Fe layer growth process. Three different stages ofgrowth are identified and indicated in the figure. The thickness dFe wasdetermined from XRR, while the average height h0 is estimated from theatomic force microscopy images.

Figure 12Horizontal line cuts from the two-dimensional GISAXS data for allsamples are plotted in a log–log presentation. The fits to the data (cyanlines) are also shown.

Table 2Overview over the lateral correlation lengths that has been determinedby fitting the vertical line cuts of the two-dimensional GISAXS data.

Sample�R1 (�a) (nm)� 12

R1 (nm)� 2.5

�R2 (�Fe) (nm)� 3.0

R2 (nm)� 0.7

Fe0 160 19 NA NAFe5 150 21 80 6.4Fe10 150 14 30 5.7Fe20 150 14 29 5.7Fe50 200 20 40 7.5Fe100 200 20 42 8.3

Figure 13Sketch of the cylindrical model used for analyzing the GISAXSmeasurements with cluster sizes (2R1, 2R2) and intercluster distances(�R1, �R2).

Page 8: Self-organization of Fe clusters on mesoporous TiO               2               templates

hysteresis loops of Fe50 (green curve) and Fe100 (orange curve)

show a classical ferromagnetic behavior of bulk iron. Note that

the magnetic moment for bulk Fe is 2.2 B per atom (Billas et

al., 1993). The slight increase in moment of Fe20 (2.4 B per

atom) seen in the figure for Fe20 (magenta curve) could

originate from the comparable sizes of the grain and dFe

(Tiago et al., 2006). This can be compared with the reduced

moment and higher saturation fields for Fe5 (red curve) and

Fe10 (blue curve).

The lowering of magnetic moment with thickness could be

caused by a demagnetization factor due to lowering of the

thickness and/or loss of neighbors. One may note that the

saturation field increases with lowering of thickness. It varies

from around 60 mT (Fe100) to 300 mT (Fe5) when measured at

5 K in the film plane. This gives an indication that the easy

direction of magnetization is gradually turning out of plane.

We plot in Fig. 16 the out-of-plane magnetization for the Fe5

sample at 5 and 300 K. A saturation moment, comparable to

the in-plane moment, is clearly evident in the out-of-plane

measurement as well. This indicates the demagnetization

effect in the film plane with the lowering of thickness.

4.5.2. Magnetization versus temperature. The coercivity of

the samples exhibits a typical ferromagnetic behavior and is

shown in Fig. 17 for Fe5 and Fe100 as representatives of two

stages of growth. The decrease in coercivity with decreasing T

for Fe5 is possibly due to the out-of-plane orientation of

magnetization which freezes at low temperatures. A similar

kind of enhancement and drop of coercivity with temperature

has been reported previously (Xu, 2007). The origin for the

anomalous temperature dependence of the coercivity is not

well understood, but some of the hints provided in that work

may also apply in our case, namely the surface anisotropy and

stress distribution, especially at the near-surface region. The

coercivity for Fe100, on the other hand, remains monotonic.

The temperature dependence of the magnetization (FC and

ZFC) is shown in Fig. 18 for Fe5, Fe20 and Fe50. The curves

beyond Fe50 are very similar. They also show a ferromagnetic

behavior. The change in the profiles could possibly be due to

the magnetization component turning out of plane for lower

Fe thickness. This is evident as we follow the ZFC curves,

where the magnetic moment at low temperature (below

150 K) is seen to increase gradually with Fe thickness. The

grains (kinetically arrested) are not rotating similarly with

temperature for different thicknesses. Furcation of ZFC and

research papers

8 of 10 Patrick Ziegler et al. � Self-organization of Fe clusters J. Appl. Cryst. (2014). 47

Figure 18Plots of FC and ZFC out-of-plane magnetic moment with temperature forsamples (a) Fe5, (b) Fe20 and (c) Fe100 as representatives of the threeregimes of growth. The applied in-plane field was 10 mT.

Figure 16Out-of-plane magnetization measurements of the Fe5 sample at 5 and300 K.

Figure 15Magnetization measurements of all samples with a magnetic layer for (a)5 K and (b) 300 K. The field sweeps have been done up to 1 T, but thefield range has been zoomed for clarity around the coercive fields. Adashed line indicating the bulk Fe magnetic moment is shown forcomparison.

Figure 17Temperature dependence of the coercivity of the in-plane magneticmoment for samples Fe5 and Fe100 as representatives of the extremestages of growth.

Page 9: Self-organization of Fe clusters on mesoporous TiO               2               templates

FC shows some irreversibility but no signature of super-

paramagnetism (Hansen & Mørup, 1998). The ZFC curves

indicate a broad distribution of grain sizes which can be

inferred from the absence of a peak. These grains respond at

different temperatures, and there is no definite blocking

permitted. The in-plane moments at various temperatures

show an increase with the increase in thickness for the FC

curves.

In regime I, we find a significant out-of-plane component of

magnetization, even though the thickness of the Fe layer in Fe5

and Fe10 is large enough to turn the moments in plane. This

may be due to the high percentage of oxidation, which may

reduce the effective magnetic thickness drastically. We can see

a clustering type of effect as we follow the ZFC curve for Fe5.

However, this does not lead to superparamagnetism. It shows

thermal stability beyond 200 K. Most likely, the deposition

process of Fe on the mesoporous template leads to statistically

distributed magnetic anisotropy axes. The competition

between the anisotropy energy and the exchange interaction

can often lead to noncollinear spin structures (Fraile Rodrı-

guez et al., 2010). As a result, the magnetic moments of the Fe

islands have become canted by a certain angle.

In regime II, for Fe20, we find a transition of the out-of-

plane moments to in plane. This is expected as the agglom-

erated species start to form a complete layer. Note that the

oxidation of the Fe layer is limited to a few monolayers only.

The superparamagnetic behavior of the clusters is also absent

in this regime. In this regime, we can expect the competition

between the anisotropy energy and exchange interaction to

reduce a little as the exchange energy gains its strength from

the number of Fe agglomerates. This obviously increases the

magnetic moment to its maximum value. The net magnetiza-

tion of a cluster ensemble can be stabilized by interparticle

and particle–template interactions.

Beyond this regime (regime III), for Fe50, the magnetization

remains stable at low temperatures but shows thermal

instability beyond 150 K, typical of bulk-like characteristics.

The fact that we observe bulk-like moments in Fe nano-

particles reflects the fact that the exchange interaction clearly

dominates over the magnetic anisotropy energies in this

regime.

5. Conclusions

In conclusion, we report on the growth of Fe on a mesoporous

structure of titania anatase. We investigated the lateral and

longitudinal correlation of the Fe layers as a function of

thickness using XRR and GISAXS. We find three different

regimes of lateral correlation, which can be correlated with the

magnetic properties of the layer. According to our results,

regime I has a range of layer thickness up to 5 nm. The Fe

diffuses into the TiO2 anatase structure (leading to porous Fe).

However, the correlation length (80 nm) remains smaller than

the correlation length of the template (150 nm), indicating a

different mode of growth. The magnetization has in-plane and

out-of-plane components responding differently with a

variation of the temperature. Regime II includes layer thick-

nesses around 10–20 nm. In this regime we have a change over

from wetting to agglomeration on the semiconductor

template. This reduces the correlation length drastically to

around 30 nm. The magnetization gradually turns mostly in

plane. The third regime is above 20 nm. In this regime Fe

grows with a correlation length of approximately 40 nm. The

magnetization is predominantly in plane and the behavior and

moment value have a bulk-like characteristic.

The experimental results demonstrate the impact of the

deposition kinetics on the physical properties of supported

islands in the nanometric regime. The complex relation of

lateral structure, particle orientation and morphology raises

the question of exploring the same with single-particle sensi-

tivity. On the basis of these findings, we are confident that our

results will contribute to a further understanding of self-

assembly of Fe within nanostructured dilute magnetic oxides.

We would like to thank A. Bauer for his initial trial

magnetization measurements using the physical property

measurement system. This work was supported by the

Deutsche Forschungsgemeinschaft via the Transregional

Collaborative Research Center TRR 80 and partially

supported by the collaborative project SFB 762 ‘Functionality

Oxide Interfaces’. We thank Ezzeldin Metwalli for help with

setting up the GISAXS instrument. Martin Niedermeier is

acknowledged for his help in preparation of the TiO2

template. PMB acknowledges Nanosystems Initiative Munich

for funding.

References

Billas, I. M. L., Becker, J. A., Chatelain, A. & de Heer, W. A. (1993).Phys. Rev. Lett. 71, 4067–4070.

Cheng, Y.-J. & Gutmann, J. S. (2006). J. Am. Chem. Soc. 128, 4658–4674.

Daillant, J. & Belorgey, O. (1992). J. Chem. Phys. 97, 5824–5836.Daillant, J. & Gibaud, A. (2009). Editors. X-ray and Neutron

Reflectivity: Principles and Applications, Lecture Notes in Physics,Vol. 770. Berlin, Heidelberg: Springer.

Dietl, T., Andrearczyk, T., Lipinska, A., Kiecana, M., Tay, M. & Wu,Y. (2000). Science, 11, 1019–1022.

Dumm, M., Zolfl, M., Moosbuhler, R., Brockmann, M., Schmidt, T. &Bayreuther, T. G. (2000). J. Appl. Phys. 87, 5457–5459.

Esquinazi, P., Hergert, W., Spemann, D., Setzer, A. & Ernst, A.(2013). IEEE Trans. Magn. 49, 4668–4674.

Filipe, A. & Schuhl, A. (1997). J. Appl. Phys. 81, 4359–4361.Fitzgerald, C. B., Venkatesan, M., Dorneles, L. S., Gunning, R.,

Stamenov, P., Coey, J. M. D., Stampe, P. A., Kennedy, R. J., Moreira,E. C. & Sias, U. S. (2006). Phys. Rev. B, 74, 115307.

Fraile Rodrıguez, A., Kleibert, A., Bansmann, J., Voitkans, A.,Heyderman, L. & Nolting, F. (2010). Phys. Rev. Lett. 104,127201.

Hansen, M. F. & Mørup, S. (1998). J. Magn. Magn. Mater. 184, 262–274.

Holy, V. & Baumbach, T. (1994). Phys. Rev. B, 49, 10668–10676.Kaune, G., Haese-Seiler, M., Kampmann, R., Moulin, J.-F., Zhong, Q.

& Muller-Buschbaum, P. M. (2010). J. Polym. Sci. B, 48, 1628–1635.

Kaune, G., Memesa, M., Meier, R., Ruderer, M. A., Diethert, A.,Roth, S. V., D’Acunzi, M., Gutmann, J. S. & Muller-Buschbaum,P. M. (2009). ACS Appl. Mater. Interfaces, 1, 2862–2869.

research papers

J. Appl. Cryst. (2014). 47 Patrick Ziegler et al. � Self-organization of Fe clusters 9 of 10

Page 10: Self-organization of Fe clusters on mesoporous TiO               2               templates

Kobayashi, M., Ishida, Y., Hwang, J. l., Mizokawa, T., Fujimori, A.,Mamiya, K., Okamoto, J., Takeda, Y., Okane, T., Saitoh, Y.,Muramatsu, Y., Tanaka, A., Saeki, H., Tabata, H. & Kawai, T.(2005). Phys. Rev. B, 72, 201201.

Lee, B., Lo, C.-T., Thiyagarajan, P., Lee, D. R., Niu, Z. & Wang, Q.(2008). J. Appl. Cryst. 41, 134–142.

Matthieu, J., Andre, B., Devillers, T., Poydenot, V., Dujardin, R.,Bayle-Guillemaud, P., Rothman, J., Bellet-Amalric, E., Marty, A.,Cibert, J., Mattana, R. & Tatarenko, S. (2006). Nat. Mater. 5, 653–659.

Ming, Z. H., Krol, A., Soo, Y. L., Kao, Y. H., Park, J. S. & Wang, K. L.(1993). Phys. Rev. B, 47, 16373–16381.

Muller-Buschbaum, P. M. (2003). Anal. Bioanal. Chem. 376, 3–10.Muller-Buschbaum, P. M. (2009). Applications of Synchrotron Light

to Noncrystalline Diffraction in Materials and Life Sciences, LectureNotes in Physics, Vol. 776, edited by T. A. Ezquerra, M. Garcia-Gutierrez, A. Nogales & M. Gomez, pp. 61–90. Berlin: Springer.

Muller-Buschbaum, P. M. & Stamm, M. (1998). Macromolecules, 31,3686–3692.

Naudon, A., Babonneau, D., Petroff, F. & Vaures, A. (1998). ThinSolid Films, 319, 81–83.

Niedermeier, M. A., Magerl, D., Zhong, Q., Nathan, A., Korstgens, V.,Perlich, J., Roth, S. V. & Muller-Buschbaum, P. M. (2012).Nanotechnology, 23, 145602.

Parratt, L. G. (1954). Phys. Rev. 95, 359–369.Perlich, J., Kaune, G., Memesa, M., Gutmann, J. S., Muller-

Buschbaum, P. M. (2009). Philos. Trans. R. Soc. London Ser. A,367, 1783–1798.

Ravel, B. & Newville, M. (2005). J. Synchrotron Rad. 12, 537–541.Salditt, T., Rhan, H., Metzger, T. H., Peisl, J., Schuster, R. & Kotthaus,

J. P. (1995). Z. Phys. B, 96, 227.Sarkar, K., Schaffer, C. J., Gonzalez, D. M., Naumann, A., Perlich, J.

& Muller-Buschbaum, P. M. (2014). J. Mater. Chem. A, 2, 6945–6951.

Tiago, M. L., Zhou, Y., Alemany, M. M. G., Saad, Y. & Chelikowsky,J. R. (2006). Phys. Rev. Lett. 97, 147–201.

Xu, X.-J., Ye, Q.-L. & Ye, G.-X. (2007). Phys. Lett. A, 361, 429–433Yoneda, Y. (1963). Phys. Rev. 131, 2010–2013.

research papers

10 of 10 Patrick Ziegler et al. � Self-organization of Fe clusters J. Appl. Cryst. (2014). 47