• Nem Talált Eredményt

Asymmetric alloy formation at the Fe-on-Ti and Ti-on-Fe interfaces

N/A
N/A
Protected

Academic year: 2022

Ossza meg "Asymmetric alloy formation at the Fe-on-Ti and Ti-on-Fe interfaces"

Copied!
10
0
0

Teljes szövegt

(1)

Journal of Physics: Condensed Matter

PAPER

Asymmetric alloy formation at the Fe-on-Ti and Ti-on-Fe interfaces

To cite this article: J Balogh et al 2018 J. Phys.: Condens. Matter 30 455001

View the article online for updates and enhancements.

This content was downloaded from IP address 148.6.26.237 on 22/10/2018 at 18:41

(2)

1. Introduction

The special magnetic properties of surfaces and interfaces induce a variety of new magnetic phenomena [1] in nanoscale magnetic-nonmagnetic bilayers or multilayer heterostructures.

Understanding such phenomena and their technological appli- cation alike requires the knowledge of atomic structure and element distribution across the interfaces. Most nano-devices are hetero-structures of different layers of various elements and alloys. Concentration distribution of the magnetic ele- ments along the growth direction generally implies a distribu- tion of the local magnetic moments and may affect the overall magnetization directly, as well as through the variation of the

anisotropies and interactions between the magnetic moments.

The knowledge of the relevant interface asymmetries may help in designing the proper deposition sequence of the layers in order to achieve the desired magnetic behaviour.

Fe–Ti multilayers have been extensively studied in the past and, mainly as a consequence of interface formation, the magn etic properties depend on the thickness of both the magnetic Fe and the nonmagnetic Ti layers [2–4] in the few nm thickness range. Sequential deposition of Fe and Ti by sputtering [2–4] results in chemical mixing at the interface of body centered cubic (bcc) Fe and hexagonal close packed (hcp) Ti layers. The interface regions may contain crystal- line bcc-Fe100−xTix (for x < 10), Fe2Ti, FeTi, hcp-Ti100−yFey

Journal of Physics: Condensed Matter

Asymmetric alloy formation at the Fe-on-Ti and Ti-on-Fe interfaces

J Balogh1, P Süle2, L Bujdosó1, Z E Horváth2, D Kaptás1 , A Kovács3, D G Merkel1 , A Nakanishi4, Sz Sajti1 and L Bottyán1

1 Wigner Research Centre for Physics, Hungarian Academy of Sciences, H-1525 Budapest 114, PO Box 49, Hungary

2 Centre for Energy Research, Hungarian Academy of Sciences, H-1525 Budapest 114, PO Box 49, Hungary

3 Ernst Ruska-Centre for Microscopy and Spectroscopy with Electrons, Forschungszentrum Jülich 52425, Germany

4 Department of Physics, Shiga University of Medical Science, Shiga 520-2192, Japan E-mail: balogh.judit@wigner.mta.hu

Received 26 January 2018, revised 6 September 2018 Accepted for publication 28 September 2018 Published 22 October 2018

Abstract

The Fe-on-Ti and Ti-on-Fe interfaces were studied experimentally by Mössbauer spectroscopy (MS), transmission electron microscopy (TEM) and x-ray reflectometry (XRR) on Ti/Fe/Ti trilayers grown on Si(1 1 1) substrates by vacuum evaporation. The nanoscale structure and composition were explored in cross sections using TEM, the layer structure and the interface widths by specular x-ray reflectometry. MS was applied to identify the interface alloy phases and to determine the pure and alloyed Fe layer fractions. The experimental results were compared with molecular dynamics (MD) simulations of layer growth on Fe or Ti underlayers of different orientations. The concentration distributions provided by MD simulations show an asymmetry at the interfaces in the layer growth direction. The transition is atomically sharp at the Ti-on-Fe interface for the (0 0 1) and (1 1 0) crystallographic orientations of the Fe underlayer, while it spreads over a few atomic layers for Fe(1 1 1) underlayer and for all studied Ti underlayer orientations at the Fe-on-Ti interface. MS and XRR data on Ti/Fe/Ti trilayers confirm the asymmetry between the bottom and top Fe interface, but the inferred interface widths considerable exceed those deduced from the MD simulations.

Keywords: interfaces, Mössbauer spectroscopy, x-ray reflectivity, molecular dynamics (Some figures may appear in colour only in the online journal)

J Balogh et al

Printed in the UK 455001

JCOMEL

© 2018 IOP Publishing Ltd 30

J. Phys.: Condens. Matter

CM

10.1088/1361-648X/aae508

Paper

45

Journal of Physics: Condensed Matter IOP

2018

1361-648X

https://doi.org/10.1088/1361-648X/aae508 J. Phys.: Condens. Matter 30 (2018) 455001 (9pp)

(3)

J Balogh et al

2

(for y < 1) alloys. Moreover the presence of amorphous com- pounds similar to those obtained by co-sputtering were also reported [5, 6]. The phase formation under ion-irradiation [7]

or heat treatment [8, 9] of multilayers has also been inves- tigated, but the Ti-on-Fe and the Fe-on-Ti interfaces were assumed equivalent in all these multilayer studies. Trilayers or single interfaces have rarely been investigated [10, 11].

An elaborate high-resolution transmission electron micros- copy (TEM) study of multilayer samples [12] claimed non- equivalent Fe-on-Ti and Ti-on-Fe interfaces, but later TEM investigations [3, 9, 13, 14] gave no evidence for any interface asymmetry and the results on the phase formation were also controversial. Namely, references [9, 14] find no evidence of crystalline or amorphous alloys at the interfaces, while reference [12] identifies amorphous-like interface regions for small repetition periods (7.2 nm) and crystalline regions with chemical gradient along the layer-growth direction for large repetition periods (40 nm). The presence of amorphous and crystalline interface alloys is also observed by [3, 13].

Note, that bcc-Ti is only reported by [12] (for small multi- layer periods). Beyond the exper imental studies molecular dynamics (MD) simulations were performed for Ti deposited onto Fe(0 0 1) substrates [15]. The results indicated an almost atomically sharp Ti-on-Fe interface under room temperature (RT) growth. These controversial results necessitate a more elaborate experimental and theoretical work in order to reveal the interface properties of polycrystalline Fe–Ti multilayers.

The atomistic simulation of the interface formation during the atomic deposition process is a feasible theoretical approach up to about 100 nm2 surfaces. The most frequent characteriza- tion methods of multilayer interfaces, viz. grazing incidence x-ray or neutron reflectivity, Rutherford backscattering or Auger electron spectroscopy, use samples of macroscopic size and they are unable to distinguish between the topolog- ical roughness/waviness of the interface and chemical mixing of the layers. The various TEM methods may give atomic scale information about both the topological features and the chemical mixing, but they sample only a nanoscale section of a macroscopic sample and sample preparation might acciden- tally change the structure. Consequently, a quantitative com- parison of the experimental results and the MD simulation is not unambiguous. Mössbauer spectroscopy (MS), being sen- sitive to the atomic scale neighbourhood of a specific isotope, can give unique information on the extent of the chemical mixing and on the nature of the compound phases formed at the interface. However, additional information on the morph- ology of the sample is inevitable for a sound identification of the different Mössbauer spectral components. For example,

57Fe probe nuclei surrounded exclusively by Fe atoms in the first and second coordination shell in a body centered cubic structure experience a hyperfine field similar to that in pure bcc-Fe, no matter if they sit at the interface of an iron layer, inside a small iron cluster or in a random dilute alloy.

In this work we study the alloy formation at the Ti-on-Fe and Fe-on-Ti interfaces in Ti/Fe/Ti trilayers by applying Mössbauer spectroscopy complemented by cross-sectional TEM and x-ray reflectometry (XRR). To attain the best Mössbauer signal to noise ratio conversion electron Mössbauer

spectroscopy (CEMS) was applied. Samples were prepared using iron metal highly enriched in the 57Fe Mössbauer reso- nant isotope. In order to determine the phase fractions within the top (Ti-on-Fe) and bottom (Fe-on-Ti) Fe interface Ti/57Fe/

Ti and Ag/57Fe/Ti sample pairs were compared exploiting the non-mixing property [16, 17] of Fe and Ag. Analysing the spectra of Ag/57Fe/Ti samples provides information on the Fe-on-Ti interface, since the Ag/Fe interface is indeed chemically sharp and the Ag layer causes only a small and well documented change [18] in the hyperfine parameters of

57Fe within the atomic layers nearest and next-nearest to the Fe–Ag interface. Additional spectral components appearing in the Mössbauer spectra of Ti/57Fe/Ti samples reveal the spe- cific properties of the Ti-on-Fe interface. The experimental results are compared with MD simulations of the layer growth performed for all basic crystal orientations of the respective underlayer, Ti or Fe. The samples were prepared by evapora- tion in high vacuum, which, to our knowledge, has not yet been used in the interface studies of this system.

2. Experimental methods

The metallic layers were prepared by vacuum evaporation onto Si(1 1 1) substrates in an evaporation chamber which per- forms 10−7 Pa base vacuum and is equipped with two substrate holders. Ti and Ag were e-beam evaporated, while the iron (95 per cent enriched in the 57Fe isotope) was evaporated by resistive heating from a tungsten crucible onto RT substrates.

The evaporation rates were about 0.1 nm s−1. Three sample pairs with 1.5, 2.5 and 3.5 nm 57Fe layer thickness were pre- pared, as given in table 1.

The equal thickness of the 57Fe layers within one sample pair was ensured by the simultaneous deposition of 57Fe from a crucible placed at equal distances from two identical Si(1 1 1) substrates of 2″ diameter. Similarly, the respective Ti layers of the sample pairs were deposited simultaneously, but the distance to the substrates was somewhat different, which may result in some (below 1 nm) difference in the Ti layer thicknesses within the pair. The (Ti) capping layers in (Ti)/Ag/

Fe/Ti/Si samples are meant to prevent oxidation of the 57Fe layer. The thin Ag layers on Fe are insufficient for the purpose, since they were found to get easily fractured [16].

Structural characterization was performed on sample Ti/3.5 Fe/Ti by TEM and XRR. Specimens for cross-sectional TEM studies were prepared in a dual-beam scanning electron microscopy and Ga+ focused ion beam (FIB) system (FEI Helios 440), which allowed to keep the sample temperature

Table 1. Layer sequence of the studied samples.

Sample

identifier Sample structure

Ti/1.5Fe/Ti 20 nm Ti/1.5 nm 57Fe/10 nm Ti/Si(1 1 1)

Ag/1.5Fe/Ti 10 nm Ti/10 nm Ag/1.5 nm 57Fe/10 nm Ti/Si(1 1 1) Ti/2.5Fe/Ti 20 nm Ti/2.5 nm 57Fe/10 nm Ti/Si(1 1 1)

Ag/2.5Fe/Ti 10 nm Ti/10 nm Ag/2.5 nm 57Fe/10 nm Ti/Si(1 1 1) Ti/3.5Fe/Ti 20 nm Ti/3.5 nm 57Fe/20 nm Ti/Si(1 1 1)

Ag/3.5Fe/Ti 10 nm Ti/10 nm Ag/3.5 nm 57Fe/20 nm Ti/Si(1 1 1)

J. Phys.: Condens. Matter 30 (2018) 455001

(4)

unaltered during the preparation. Pure carbon was depos- ited on the surface as a protective layer. The surface damage induced by the high-energy Ga beam was reduced by low- energy Ar+ ion beam milling using Fischione NanoMill 1040 system at energies below 1 keV. The structure and chemical composition of the trilayer was studied using a conventional TEM (FEI Tecnai G2) and a probe-aberration-corrected scan- ning TEM (FEI Titan G2 80-200) equipped with in-column energy-dispersive x-ray spectroscopy (EDXS) detectors [19].

The high-resolution images and elemental maps were pro- cessed using the FEI Velox software.

XRR measurements were performed by a Bruker AXS D8 Discover diffractometer equipped with a Göbel-mirror and a scintillation detector using Cu Kα radiation. The XRR probes the refractive index depth profile in flat samples perpend- icular to the surface. Interpretation of the reflectivity involves modelling the depth profile by an approximate multilayer structure of constant scattering length densities of the layers and interface regions (or the respective indices of refraction determined by the composition and density). Specular reflec- tivity does not distinguish the interface roughness and the lat- erally homogeneous intermixing e.g. due to inter-diffusion.

The collected reflectivity data were evaluated by using the FitSuite code, a free multipurpose software [20]. FitSuite uses a transfer matrix method described in [21] for calcul- ation of x-ray reflectograms. It treats the interface roughness as described in [22, 23] assuming a Gaussian height distri- bution and a corresponding error function-type interface depth profile characterized by the standard deviation of the Gaussian, the (rms) roughness, σ. The Gaussian’s full width at half maximum FWHM = 2·√

2·ln2·σ ≈ 2.355 σ covers the central region of the error function profile between 4.8%

and 95.2%. Independent of the shape, we will characterize the interface width by W5-95, the distance between the 5% and 95% quanti les of the profile.

The CEMS measurements were carried out by a conven- tional constant acceleration Mössbauer spectrometer. For the detection of the conversion electrons a low background gas filled proportional counter was used with pure H2 gas at low temperatures [24], and 96%He–4%CH4 gas mixture at RT. The spectra were measured by a 50 mCi 57Co(Rh) single line Mössbauer source. The hyperfine field (HF) dis- tributions were evaluated according to the Hesse–Rübartsch method [25], by fitting the amplitudes of a number of sex- tets with HFs increasing with equal step values. The isomer shift (IS) values are given relative to that of α-Fe at room temperature.

3. Atomistic simulation methods

Classical molecular dynamics calculations were used in par- allel environment using GPU nodes reaching a significant speedup, as implemented in the LAMMPS code (Large-scale Atomic/Molecular Massively Parallel Simulator) [26] The applied MD scheme is as follows:

–determination of the decomposed spatial regions within the periodic simulation box to be used for the message inter- face passing (MPI) parallel calculations on supercomputers

–the code first builds a neighbour list within the predefined cut-off distance (3–4 Angstrom) and stores this list throughout the full simulation time

–initial guess: the atoms’ initial positions and initial veloc- ities according to the initial temperature with Boltzmann distribution

–analytical calculation of the forces within the cut-off dis- tance and moving the atoms according to the forces acting on them–dynamical time stepping via numerically adjustable time step which roughly inversely depends on the forces acting on the atoms (Close to the equilibrium state the time stepping becomes faster.)

–integrating the equations of motion using Verlet algorithm to find the new atomic positions and velocities

–introducing temperature effects via the NPT (Nose– Hoover prestostat and thermostat) algorithm

–the above scheme is continued until the predefined time limit is reached.

The isobaric-isothermal (NPT ensemble) simulations were carried out at 300 K, allowing for the variation of the supercell size and shape, which ensured further flexibility in absorbing pressure waves and/or in relaxing the occasional strain in the system. Periodic boundary conditions were ensured along the lateral directions (x,y) and vacuum regions were inserted above and below the slab of the film/substrate system. The size of the vacuum region was chosen a few nanometers in order to keep sufficient space between the evaporated particles and the forming surface of the deposit. The temperature of the substrate was kept at 300 K exploiting the variable time step algorithm. The atomic and nanoscale structures were dis- played using the OVITO code [27]. The specific (fix deposit) algorithm of the LAMMPS was applied for the simulation of the thin film growth. The flux of the deposited atoms was kept at 2.25 · 1025 atoms s−1 cm−2 (one particle per 10 k simulation steps). In total, at least 15–20 monolayers (ML) of adatoms were deposited on the substrate in which each ML corre- sponded to nearly 200 adatoms for the Fe and Ti layers on the 4 × 4 nm2 substrate surface. The EAM potentials were used for both Fe and Ti as generated by the EAM database tool [28]

provided by the LAMMPS code. The FeTi cross-interaction was taken into account by the Johnson mixing rule [29] which provides reasonable alloy properties. Further details of the simulated thin film growth can be found in [17].

4. Experimental results

4.1. Electron microscopy

The structure of the deposited layers of the Ti/3.5 Fe/Ti was studied in cross-section using various TEM methods, as shown in figure 1. Figure 1(a) displays a bright-field (BF) scanning TEM (STEM) image of the Ti and Fe layers deposited on Si substrate. The Fe layer thickness on the STEM images was measured to be 4.2 nm, about 20% above the nominal thick- ness. The ~1 nm thick native SiO2 layer is also visible on the surface of the Si substrate. Both Ti layers show polycrystalline structure with grain sizes ranging from 5 to 15 nm. The light

(5)

J Balogh et al

4

contrast of ~3.5–4 nm thickness on the top of the upper Ti layer is due to surface oxidation. Figure 1(b) shows the colour coded chemical composition map of the Ti and Fe layers extracted from EDXS measurement. The standard background removal and the Cliff–Lorimer method were applied to calcu- late the atomic fraction of the elements. Figure 1(c) shows the corresponding variation across the Fe interfaces measured in the map of figure 1(b). The Fe concentration peaks at approxi- mately 0.94 in the middle of the Fe layer. EDXS does not reveal differences between the Fe-on-Ti and the Ti-on-Fe transition zones, both appear around 3 nm thick. Figure 1(d) displays the high-resolution BF STEM image. In figure 1(e) the digital diffractogram of the high-resolution image shows the Ti and Fe lattice arrangements relative to each other. The analysis of the pattern suggests [0 1 0] orientation of the anal- ysed Ti grain and [1 0 0] orientation of the Fe grain.

4.2. X-ray reflectivity

Results of the XRR measurements performed on the Ti/3.5 Fe/Ti sample using the Cu Kα line (0.154 nm) are shown in figure 2 and table 2. The layer thicknesses were found to be

21.5 nm, 3.63 nm, 20.4 and 4 nm for the bottom Ti, Fe, top Ti, and TiOx layers, respectively. The native SiO2 layer of 1 nm thickness on the Si substrate and the surface TiOx layer were included in the layer model in accordance with the TEM results. Significant intermixing occurs only at the bottom of the Fe layer, i.e. at the Fe-on-Ti interface, as the roughness of the bottom Ti layer is 0.86 nm while that of the Fe layer is found quite sharp, 0.23 nm.

4.3. Mössbauer spectroscopy

The CEMS spectra of the three as-received sample pairs measured at RT and 15 K are shown in figures 3(a) and (b), respectively. All spectra (except for sample Ti/1.5Fe/Ti at RT) were fitted by a set of sextets with hyperfine fields allowed in the 5–40 T range and a broadened doublet with a single quadrupole splitting (QS). The evaluated normalized HF dis- tributions are shown as insets to each spectrum in figure 3. In the spectrum fitting the line-width of the sextet components of the HF distribution was fixed to 0.24 mm s−1, the HF step value was an iteration parameter in the 0.7–0.8 T range. The isomer shifts were assumed proportional to the HFs, the lower the HF the larger negative value to have. The intensities of the second and fifth lines relative to that of line three and four of the sextets were close to 4 in each case, indicating an in-plane orientation of the Fe magnetic moments.

The most important parameters of the spectra are summa- rized in table 3. The magnetic component is divided into two subgroups for each spectrum. The one indicated as bcc-Fe, is a sum of the sextets in the 0.5 T vicinity of 33 and 33.8 T, the literature HF values of bcc-Fe at RT and 15 K, respec- tively. This component can be attributed to Fe atoms without Ti or Ag as first neighbours. The component stemming from the sextet intensities below 32.5 and 33.3 T at RT and 15 K, respectively is labelled as magnetic alloy and belongs to Fe atoms with Ti (or Ag) neighbours in the first and/or second

Figure 1. Structure of the Ti/3.5 Fe/Ti sample studied in cross- section using various TEM methods; bright-field STEM image of the Ti and Fe layers deposited on Si substrate (a), colour coded chemical composition map of the Ti and Fe layers extracted from EDXS recordings (b), atomic fraction of the elements across the interfaces calculated from the composition map (c), high resolution bright-field STEM image (d), digital diffractogram of the area marked in the high resolution image (e).

Figure 2. Measured and fitted x-ray reflectivity curve for the Ti/3.5 nm Fe/Ti sample. The inset shows the concentration profiles of the different materials corresponding to the fitted parameters.

Red (dark) and green (light) thick lines indicate Ti and Fe. The components appearing around 0 and 50 nm depth and above indicated by thin black linesbelong to TiOx, SiO2, and Si, respectively. The asymmetry of the FeTi interface is salient.

J. Phys.: Condens. Matter 30 (2018) 455001

(6)

neighbour shells [31]. The average HF of this component is largely different for samples Ti/1.5Fe/Ti and Ag/1.5Fe/Ti, but remains constant within the ~0.5 T experimental error for Fe thicknesses of 2.5 and 3.5 nm. The spectral fraction of the different components (bcc-Fe, paramagnetic alloy, magnetic alloy) was transformed into Fe equivalent layer thicknesses

by multiplying it with the nominal thickness of the Fe layer.

(The Mössbauer-Lamb factors were assumed equal for all components.) The Fe equivalent layer thicknesses will be labelled as dbcc, dpara, and dmag for the bcc-Fe, for the par- amagnetic alloy and for the magnetic alloy components, respectively.

Table 2. The parameters obtained from the fit of x-ray reflectivity measurements. The roughness values refer to the top interface of the respective layer. The complex refractive indices (n=1δ+iβ), see [30], were kept constant in the fit. The densities of the TiOx and SiO2 were assumed to be 3.78 and 2.2 g·cm−3, respectively. These densities correspond to TiO2 in the anatase and SiO2 in the amorphous phase, respectively. The errors indicated are statistical errors. Based on various model simulations we estimate the overall uncertainty of the layer thickness and the interface roughness below ±5% and ±20%, respectively.

TiOx Top Ti 57Fe Bottom Ti SiO2

δ · 106 11.62 13.52 22.45 13.52 7.12

β · 106 0.58 1.13 2.90 1.13 0.09

Thickness (nm) 4.0 ± 0.02 20.4 ± 0.02 3.63 ± 0.02 21.5 ± 0.04 1 ± 0.03

Roughness (nm) 0.87 ± 0.02 1.36 ± 0.01 0.23 ± 0.05 0.86 ± 0.03 0.53 ± 0.01

Figure 3. Mössbauer spectra of the as received samples measured at RT (a) and at 15 K (b). The sub-spectra belonging to the doublet component and to the HF distributions (shown in the insets) are indicated by blue (light) and red (dark) lines, respectively. In the insets the horizontal (HF) scales are the same for all HF distributions and are labelled in the topmost insets in tesla units. The vertical intensity scales of the insets, however, are normalized to unit area and vary from graph to graph one tick belonging to p(B) = 0.1 T–1 in each inset.

Table 3. Mössbauer parameters of the spectra measured at RT and 15 K; isomer shift (IS), quadrupole splitting (QS) and line width (W) of the doublet, average hyperfine field of the magnetic alloy component (HFmag) and the iron content of the different components expressed in Fe equivalent layer thicknesses, as calculated from the spectral fractions and the nominal thickness of the deposited Fe layer. Typical errors are given in the second row.

IS (mm s−1) QS (mm s−1) W (mm s−1) HFmag (T) dbcc (nm) dpara (nm) dmag (nm)

Typical error ±0.02 ±0.02 ±0.05 ±0.5 ±0.05 ±0.05 ±0.05

Temperature RT/15 K RT/15 K RT/15 K RT/15 K RT/15 K RT/15 K RT/15 K

Ti/1.5 Fe/Ti −0.17/− 0.10 0.29/0.52 0.40/0.62 0/16.4 0/0 1.50/0.66 0.00/0.84

Ag/1.5 Fe/Ti −0.18/− 0.10 0.38/0.51 0.61/0.57 25.0/25.2 0.04/0.25 0.46/0.24 1.00/1.01 Ti/2.5 Fe/Ti −0.21/− 0.07 0.47/0.50 0.56/0.69 26.3/27.5 0.61/0.66 0.33/0.39 1.56/1.45 Ag/2.5 Fe/Ti −0.22/− 0.09 0.42/0.46 0.51/0.43 27.1/27.0 0.54/0.91 0.21/0.19 1.75/1.40 Ti/3.5 Fe/Ti −0.21/− 0.08 0.45/0.50 0.65/0.56 27.1/27.5 1.53/1.46 0.35/0.33 1.62/1.71 Ag/3.5 Fe/Ti −0.19/− 0.11 0.46/0.54 0.47/0.61 27.0/28.3 1.73/1.80 0.21/0.26 1.56/1.44

(7)

J Balogh et al

6

5. Atomistic simulations

The classical MD simulations of the thin film growth of Fe–Ti bilayers were carried out with different orientations of either the Fe or the Ti underlayer. The cross-sectional view of the grown interfaces is shown in figures 4(a) and (b) and the cal- culated concentration depth profiles for different substrate ori- entations are displayed in figure 5. The Fe-on-Ti interfaces are slightly intermixed and the interface broadening depends on the orientation of the substrate, although the variation is small.

W5-95, the distance between the 5 and 95 at% intersection points of the interface profile is around 1 nm for Fe/Ti(1 1 0 1) and 0.8 nm for Fe/Ti(0 0 0 1). In the Ti-on-Fe case no atomic mixing can be seen for the Fe(0 0 1) substrate orientation, but the interface is slightly wavy, as shown in figure 4(b), which results in a W5-95 = 0.3 nm width of the concentration depth profile. For the Fe(1 1 0) orientation a modest intermixing also takes place and the intermixed region is slightly wider, W5-95 = 0.5 nm. The largest intermixing takes place in case of the Fe(1 1 1) substrate orientation, but W5-95 = 0.7 nm still remains below the Fe-on-Ti values. In summary, the simula- tions give evidence of a somewhat orientation dependent inter- mixing which is asymmetric with respect to the interchange of the constituents of the growing film and the underlayer.

6. Discussion

The most direct experimental information on the range of the chemical mixing is obtained from Mössbauer spectroscopy therefore we discuss those results first. The room temper- ature spectra shown in figure 3(a) are in line with previous Fe/

Ti multilayer studies [2, 4]. The observed room temperature IS and QS values of the Ti/1.5Fe/Ti sample, −0.18 mm s−1 and 0.38 mm s−1, respectively, suggest that below 2 nm thick- ness of the Fe layer, all Fe atoms are alloyed with Ti in an amorphous structure. Although the paramagnetic component may contain minor contributions from crystalline Fe2Ti, FeTi and FeTi2 [5, 6], it will be referred as amorphous component.

Further on the low temperature CEMS spectra in figure 3(b) reveal that the amorphous component is not homogeneous.

Indeed, the Ti/1.5Fe/Ti spectrum reveals 40% paramagnetic

Fe atoms at 15 K, while the remaining Fe exhibits a broad HF distribution in the 5 to 30 T range. The well resolved magnetic component that can be observed in the spectrum of Ag/1.5 Fe/Ti at RT, is not present in the spectrum of Ti/1.5 Fe/Ti, which demonstrates that the Fe-on-Ti and the Ti-on-Fe inter- faces overlap in Ti/1.5 Fe/Ti. The single Fe-on-Ti interface of the Ag/1.5 Fe/Ti sample, in addition to amorphous alloys, certainly contains crystalline bcc-Fe100−xTix alloy phase, since it is the only magnetic Fe–Ti phase at RT. All other crystalline or amorphous alloys are known to exhibit Curie temperatures below RT in the entire concentration range.

With increasing the thickness of deposited Fe the formation of a bcc-Fe layer and two separated interfaces, Fe-on-Ti and Ti-on-Fe, are expected. The formation of two well separated interfaces is signalled by constant values of the HF param- eters and of the interface width as a function of the depos- ited Fe thickness. In samples Ti/2.5 Fe/Ti and Ti/3.5 Fe/Ti the interface components give an Fe equivalent layer thickness of dpara + dmag ≈ 2 nm, in accordance with former publications [2, 4] and with our above conclusion on the overlapping inter- faces below 2 nm Fe thickness. In these two samples the HF distributions (insets in figure 3) and the HFmag values in table 3

Figure 4. Typical cross-sectional view with 1 nm slab thickness of Fe/Ti(0 0 0 1) (a) and Ti/Fe(0 0 1) (b) interfaces as obtained from MD simulations. Blue (dark) and red (light) spheres are Fe and Ti atoms, respectively.

Figure 5. Concentration profiles across the Fe-on-Ti and Ti-on- Fe interfaces as calculated from MD simulations for different orientations of the substrate layer. The substrate orientations which belong to the different curves and the thickness (W5-95) of the respective interface alloy are indicated in the inset.

J. Phys.: Condens. Matter 30 (2018) 455001

(8)

agree with each other within the experimental error and differ from those in Ti/1.5 Fe/Ti. The bcc-Fe component in Ti/2.5 Fe/Ti and Ti/3.5 Fe/Ti forms a layer of dbcc = 0.64 nm and 1.50 nm, respectively. Although the bcc-Fe component might as well be assigned to Fe atoms with no Ti neighbours in a dilute alloy, but the formation of a continuous bcc-Fe layer in Ti/3.5 Fe/Ti seems to be justified by the following observa- tions: (i) for a 1 nm increase of the nominal Fe thickness dbcc

was found to increase almost the same amount, 0.86 nm, (ii) the interface properties (dpara, dmag, QS, HFmag) do not depend on the deposited Fe thickness for the Ti/2.5 Fe/Ti and Ti/3.5 Fe/Ti samples and finally (iii) TEM and XRR (see figures 1 and 2) are indicative of a continuous Fe layer separating the Fe-on-Ti and Ti-on-Fe interfaces in the Ti/3.5 Fe/Ti sample.

The minor difference between the spectra of Ti/3.5 Fe/Ti and Ag/3.5 Fe/Ti samples (see figure 3) implies that the Ti-on-Fe interface is chemically rather sharp and the interface alloys are formed predominantly at the Fe-on-Ti interface.

The interface asymmetry will be analysed below by com- paring the Ti/3.5 Fe/Ti and Ag/3.5 Fe/Ti samples. Since the amorphous alloy and the crystalline bcc alloy phases can be clearly distinguished in the RT Mössbauer spectra, the inter- face widths will be calculated from the RT spectral fractions.

The properties of the Fe-on-Ti interface are directly mani- fested in the Ag/3.5 Fe/Ti spectra. The paramagnetic and the magnetic alloy component contains dpara (Fe-on-Ti) = 0.21 nm and dmag (Fe-on-Ti) = 1.56 nm Fe, respectively, however, this latter value in Ag/3.5 Fe/Ti contains two monolayers (~0.3 nm) Fe contribution of the Ag-on-Fe interface, there- fore in further calcul ations we use a corrected value, dcmag (Fe-on-Ti) = 1.26 nm. To unravel details of the Ti-on-Fe interface, the Ti/3.5 Fe/Ti and Ag/3.5 Fe/Ti spectra will be compared. In Ti/3.5 Fe/Ti dpara and dmag are slightly larger than in Ag/3.5 Fe/Ti at the expense of dbcc, while the hyper- fine parameters (IS, QS and HFmag) are practically the same in the two samples. This indicates the presence of similar phases in different thicknesses in the corresponding interface regions. The small difference in the intensity of the paramagn- etic comp onent is clearly seen in the spectra of figure 3. For Ti/3.5 Fe/Ti dpara is larger than for Ag/3.5 Fe/Ti and this way dpara(Ti-on-Fe) = 0.14 nm. The slight difference in the HF dis- tributions, with a 0.06 nm larger dmag for Ti/3.5 Fe/Ti than for Ag/3.5 Fe/Ti, results in dmagc (Ti-on-Fe) = 0.36 nm.

In the following, we attempt to estimate the relative phase volumes/thicknesses (Dpara, Dmag) from the Fe equivalent layer thicknesses (dpara, dmagc ) by considering the known concentra- tion limits of the amorphous and crystalline alloys identified by MS. In lack of experimental data on the density of the dif- ferent compounds, the alloyed thicknesses will be calculated using the bulk densities and the molar masses of the elements (ρFe, ρTi, MFe, and MTi, resp.), and the Fe concentration of the respective alloy (cFe) as follows:

D=d Å

1+ρFeMTi(1−cFe) ρTiMFecFe

ã .

(1) Determination of the alloy concentrations or the concentra- tion distributions is impossible from the present Mössbauer spectra both for the amorphous and for the crystalline alloy.

Trials to fit the HF distributions of the crystalline alloy as a random solution by binomial distribution remained unsuc- cessful even for the separated RT spectral fractions. The small variation of IS and QS with the concentration of the amorphous alloy also make the estimation futile, therefore we use average concentrations by considering the known concentration limits of the amorphous and crystalline alloys. Consequently, in our modelling the possible concentration ranges 20 to 80at%

Ti for the amorphous [5] and 0 to 20at% Ti for the crystal- line [32] components will be replaced by their average, 10 and 50at% Ti. Applying equation  (1) to dpara and dmagc pro- vides Dpara(Ti-on-Fe) = 0.35 nm, Dpara(Fe-on-Ti) = 0.52 nm, Dmag(Ti-on-Fe) = 0.42 nm, Dmag(Fe-on-Ti) = 1.47 nm thick- ness for the amorphous (Fe50Ti50) and crystalline (Fe90Ti10) alloy phases, respectively. The total Fe-on-Ti and Ti-on-Fe interface widths are 1.99 nm and 0.77 nm, respectively. The asymmetry of the interface width is beyond dispute, although the concentrations and the distribution of the interface alloys are somewhat uncertain. Allowing for relative deviations of the average concentrations as large as 50% (viz. 5 or 15 at% and 25 or 75 at% Ti for the magnetic and the paramagn- etic alloy, respectively) results in 2.74 nm and 1.23 nm total Fe-on-Ti and Ti-on-Fe interface widths for the high Ti content and 1.67 nm and 0.60 nm for the low Ti content case.

From the above discussion of the Mössbauer results we may conclude that the Fe-on-Ti and Ti-on-Fe interfaces show a large difference in the total width, as displayed in figure 6, and in the relative amount of the Ti-rich amorphous and the Fe-rich magnetic alloys of the interface. The Fe-on-Ti inter- face appears almost three times thicker than the Ti-on-Fe interface. The Fe-rich bcc alloy dominates over the Ti-rich amorphous alloy in the Fe-on-Ti interface in contrast to the thinner Ti-on-Fe interface in which the two are of similar thickness (~0.4 nm). Such thin layers are often discontinuous.

The XRR measurements (see figure 2 and table 2) also indicate an asymmetry of the Fe-on-Ti and Ti-on-Fe inter- faces beyond the estimated uncertainty of the roughness. The interface width is defined as the full width at half maximum 2·√

2·ln2·σ ≈ 2.355σ of the corresponding normal distri- bution, where σ is the interface (rms) roughness. The Gaussian distribution comprises 90% of its integral intensity (from 5 to 95%) within the full width at half maximum, therefore σ is a special case of a W5-95 interface width for an error-function like interface profile. Roughness values of 0.86 nm and 0.23 nm result in W5-95 = 2.0 and 0.54 nm for (the bottom) Ti and the Fe layers, respectively, to be compared with histogramic width of 1.99 and 0.77 nm estimated from the Mössbauer data.

The thickness of the Ti, TiOx and SiO2 deduced from the XRR measurements are consistent with those appearing in the TEM picture in figure 1(a). The derived layer structure is displayed in figure 6 along with the MS and EDXS data. XRR indi- cates a continuous Fe layer in Ti/3.5 Fe/Ti. The scaling of the experimental interface width may be model dependent, but an asymmetry between the (thinner) Ti-on-Fe and the (thicker) Fe-on-Ti interface is undoubtedly revealed by both the XRR and MS data.

The TEM measurements sample a much smaller area than XRR or MS, but reveal some further details of the structure.

(9)

J Balogh et al

8

The 4.0 nm–4.2 nm Fe layer thickness, estimated from the dark contrast region in figure 1(a), appears larger than the XRR value of 3.6 nm, and exceed the nominal thickness of 3.5 nm.

Figures 1(a) and (d) clearly reveal a surface roughness of the Fe layer on both sides. From the contrast of the Fe layer the roughness can be estimated to be around 1–2 nm. Somewhat surprisingly, no alloyed bcc-Fe fraction can be identified and amorphous or disordered regions cannot be seen at the Fe–Ti interfaces, although the EDXS results indicate mixing between the layers. The Fe concentration peaks only at 95 at% and the concentration change is not abrupt but extends over 2–3 nm in the elemental concentration depth profile of figure 1(c) derived from the EDXS measurement. Moreover, EDXS does not show noticeable asymmetry between the Fe-on-Ti and Ti-on-Fe interfaces, which contradicts to the MS and XRR results, as shown if figure 6. To understand the mes- sages of figure 1, one has to consider the specimen geometry used in the TEM experiments. The specimen was approxi- mately 40 nm thick (in the view direction) and the measure- ment shows the projected signal in the maps. The elemental distribution is a result of averaging the signal across the spec- imen thickness parallel with the substrate surface. Moreover, the grain size of Fe (and of Ti) is smaller than the thickness of the specimen, the signal at a given point stems from sev- eral grains. The contributions of the grain boundaries and of the interface roughness along the electron beam direction are averaged, resulting in a broadened elemental distribution. The small crystallite size might explain an Fe concentration profile peaking at 95 at%, since the Ti deposition on Fe may results in Ti diffusion along the grain boundaries, which appears as the presence of Ti in the middle of the Fe layer, as shown in figure 1(c). Diffusion along the grain boundaries might also explain that the experimentally observed interface thickness is about one order of magnitude larger than the one obtained by MD simulations assuming perfect single crystal substrate layers. One may conclude that the EDXS measurement across small-sized crystalline layers using TEM may not give clear view on the distribution of the elements. Here we note, that

the high-resolution BF STEM image of figure 1(d) also allow alloying up to a few atomic percent at the interfaces, since it may be characteristic only to the change of the structure.

The apparently larger thickness of the Fe layer as compared the nominal value or to the MS and XRR values, also support this notion. One can conclude that while TEM provides very important information on the structure of the layers, but the cross-sectional analysis of the current polycrystalline samples is not suitable to unambiguously reveal the structure and the chemical composition of the interface.

It is worth putting the present results into context with the adhesion and surface energies of the layers. According to a rule of thumb, the smaller the surface energy of the underlayer as compared to that of the over-layer, the larger the tendency for clustering and consequent interface roughening [33]. Since the layer growth occurs under highly non-equilibrium condi- tions the kinetics of the process and the adhesive energies of the layers gain importance [34, 35]. The lower the adhesive energy of the underlayer, the more often atomic exchanges between the ad-atoms and the surface atoms of the underlayer take place. The experimental values for the surface energies [36–38] range from 1.99 to 2.10 (J m−2) for Ti and from 2.42 to 2.48 (J m−2) for Fe, which predict a slight asymmetry of the interfaces, the Ti-on-Fe being sharper than the Fe-on-Ti.

Our MD simulation results on interface mixing are in good agreement with the above expectations. It is worth a mention that the MD simulations provided a much larger asymmetry for the Fe–Al system [17] in accordance with the surface energy values, being smaller for Al (1.14–1.16 (J m−2)) than for Ti. The cohesive energy is also smaller for Al than for Ti [39], thereby making Al-for-Fe exchanges easier at the Al sur- face than Ti-for-Fe exchanges at the Ti surface. The exper- imental values obtained by Mössbauer spectroscopy and by x-ray reflectometry also show the asymmetry of the Fe-on-Ti and Ti-on-Fe interfaces, but the estimated interface width is larger than those derived from MD simulations. The larger intermixing is most probably due to the large amount of grain boundaries seen in the TEM pictures. An enhanced diffusivity along the grain boundaries certainly increases the width of the intermixed region and it may play a role in the formation of amorphous regions.

Solid state amorphization at thin film interfaces may occur when the diffusion coefficients of the elements of the layers in touch are largely different [40]. Iron (similarly to Co and Ni) is well known to exhibit fast diffusion in hcp-Ti, the Fe impu- rity diffusivities being more than 104 times larger than the Ti self-diffusivity [41], while the diffusivity of Ti in bcc-Fe is about the same magnitude as in Ti [42]. In view of the diffu- sivities, the amorphous alloy formation is expected predomi- nantly at the Fe-on-Ti rather than at the Ti-on-Fe interface, but fast diffusion along the grain boundaries or defect sites may alter the picture and explain why the estimated average thick- ness of the amorphous layers at the Fe-on-Ti and the Ti-on-Fe interfaces are so little different, Dpara = 0.52 nm and 0.35 nm, respectively. The absence of amorphous or disordered regions in the atomic layer structure visible in figure 1(b) supports the notion that the component identified as amorphous by Mössbauer spectroscopy does not form a continuous layer.

Figure 6. Layer and interface profiles of Fe in Ti/3.5Fe/Ti as inferred from MS, XRR and EDXS data. The common depth zero is chosen arbitrarily for all methods as the abscissa of the half maximum Fe concentration point at the Ti-on-Fe interface.

J. Phys.: Condens. Matter 30 (2018) 455001

(10)

7. Conclusion

Conversion electron Mössbauer spectroscopy and x-ray reflectometry measurements on Ti/Fe/Ti trilayers grown on Si(1 1 1) substrates by vacuum evaporation are indicative of asymmetric Fe interfaces, with the Fe-on-Ti interface being a factor of 2–3 times wider than the Ti-on-Fe interface. The alloys formed at the interfaces are identified by Mössbauer spectroscopy; Fe-rich bcc alloy as the main component and Ti-rich amorphous alloy as the minor component. The Fe-rich bcc alloy dominates over the Ti-rich amorphous alloy in the Fe-on-Ti interface in contrast to the thinner Ti-on-Fe inter- face in which the two are of similar thickness Atomic reso- lution TEM shows domains with rather sharp lattice spacing transition at the interfaces, supporting the idea that the amor- phous alloy does not form a continuous layer at the interface.

Molecular dynamics simulations of the layer growth also show asymmetric concentration distribution along the layer growth direction; it is atomically sharp at the Ti-on-Fe interface for the (0 0 1) and (1 1 0) crystallographic orientations of the Fe layer, while it varies over a few atomic layers for Fe(1 1 1) substrate and at the Fe-on-Ti interface for all basic crystal- lographic orientations of Ti. The larger experimental interface width as compared to the simulated values is probably due to the polycrystalline nature of the layers, the small size of the Fe crystallites and the grain boundary diffusion during deposi- tion. Experiments on single crystal or perfect epitaxial layers could explore the role of crystal orientations, grain boundaries and interface roughness and would allow for a more direct comparison of the experiments and the MD calculations.

Acknowledgment

The authors acknowledge financial support of the Hungarian Scientific Research Fund (OTKA) Grant K112811.

ORCID iDs

D Kaptás https://orcid.org/0000-0002-2084-8928 D G Merkel https://orcid.org/0000-0001-9644-2521 Sz Sajti https://orcid.org/0000-0002-8748-8242

References

[1] Hellman F et al 2017 Rev. Mod. Phys. 89 25006

[2] Rodmacq B, Hillairet J, Laugier J and Chamberod A 1990 J.

Phys.: Condens. Matter 2 95

[3] Fnidiki A, Juraszek J, Teillet J, Duc N H, Danh T M, Kaabouchi M and Sella C 1998 J. Appl. Phys. 84 3311 [4] Kopcewicz M, Stobiecki T, Czapkiewicz M and Grabias A

1997 J. Phys.: Condens. Matter 9 103

[5] Nakamura Y, Sumiyama K and Ezawa H 1986 Hyperfine Interact. 27 361

[6] Chien C L and Liou S H 1985 Phys. Rev. B 31 8238 [7] Kopcewicz M, Jagielski J and Stobiecki T 2005 J. Phys.:

Condens. Matter 17 2149

[8] Wang W-H, Bai H Y and Wang W K 1999 J. Appl. Phys.

86 4262

[9] Wu Z L, Peng T X, Cao B S and Lei M K 2009 Thin Solid Films 517 6553

[10] Rayput P, Gupta A and Avasthi D K 2008 Nucl. Instrum.

Methods Phys. Res. B 266 1680

[11] Carbucicchio M and Palombarini G 1993 Appl. Surf. Sci.

65/66 331

[12] Bourret A and Rouviere J L 1990 Phil. Mag. 62 415 [13] Wang W and Wen L 2001 J. Mater. Sci. Technol. 17 521 [14] Lei M K, Wu Z L, Chen T and Cao B S 2006 Thin Solid Films

500 174

[15] Choi H, Yoon G and Chung Y 2011 Japan. J. Appl. Phys.

50 01BE07

[16] Balogh J, Bujdosó L, Kaptás D, Dézsi I and Nakanishi A 2012 Phys. Rev. B 85 195429

[17] Süle P, Kaptás D, Bujdosó L, Horváth Z E, Nakanishi A and Balogh J 2015 J. Appl. Phys. 118 135305

[18] Lugert G and Bayreuther G 1988 Phys. Rev. B 38 11068 [19] Ernst Ruska-Centre for Microscopy and Spectroscopy with

Electrons (ER-C), Forschungszentrum Jülich and RWTH Aachen 2016 FEI Titan G2 80-300 CREWLEY J. Large Scale Res. Facil. 2 A43

[20] Sajti S, Deák L and Bottyán L FitSuite 1.0.5.rc11 (arXiv:0907.2805v1 [cond-mat.other])

[21] Deák L, Bottyán L, Nagy D L and Spiering H 2001 Physica B 297 113

[22] Röhlsberger R 1999 Hyperfine Interact. 123/124 301 [23] Deák L 1999 PhD Thesis Roland Eötvös University

Budapest

[24] Fukumura K, Nakanishi A and Kobayashi T 1994 Nucl.

Instrum. Methods Phys. Res. B 86 386 [25] Hesse J and Rübartsch A 1974 J. Phys. E 7 526 [26] Plimpton S J 1995 J. Comput. Phys. 117 1

[27] Stukowski A 2010 Modelling Simul. Mater. Sci. Eng.

18 015012

[28] Zhou X W, Johnson R A and Wadley H N G 2004 Phys. Rev.

B 69 144113

Zhou X W et al 2001 Acta Mater. 49 4005

Wadley H N G, Zhou X, Johnson R A and Neurock M 2001 Prog. Mater. Sci. 46 329

[29] Johnson R A 1989 Phys. Rev. B 39 12554

[30] Henke B L, Gullikson E M and Davis J C 1993 At. Data Nucl.

Data Tables 54 181342

[31] Vincze I and Campbell I A 1973 J. Phys. F 3 647

[32] Gunnlaugsson H P, Zenkevich A and Weyer G 2005 J. Alloys Compd. 398 33

[33] Himpsel F J, Ortega J E, Mankey G J and Willis R F 1998 Adv.

Phys. 47 511

[34] Chang T-M and Carter E A 1995 J. Phys. Chem. 99 7637 [35] Buchanan J D R, Hase T P A, Tanner B K, Chen P J, Gan L,

Powell C J and Egelhoff W F Jr 2002 Phys. Rev. B 66 104427

[36] Vitos L, Ruban A V, Skriver H L and Kollár J 1998 Surf. Sci.

411 186

[37] Tyson W R and Miller W A 1977 Surf. Sci. 62 267 [38] de Boer F R, Boom R, Mattens W C M, Miedema A R and

Niessen A K 1988 Cohesion in Metals (Amsterdam: North- Holland) p 121, 219

[39] Kittel C 2005 Introduction to Solid State Physics 8th edn (Hoboken, NJ: Wiley) p 50

[40] Schwarz R B and Johnson W L 1983 Phys. Rev. Lett. 51 415 [41] Perez R A, Nakajima H and Dyment F 2003 Mater. Trans.

[42] Liu Y, Zhang L, Du Y and Liang D 2009 CALPHAD, Comput. 44 2 Coupling Phase Diagr. Thermochem. 33 732

Hivatkozások

KAPCSOLÓDÓ DOKUMENTUMOK

Two routes of formaldehyde formation were proposed: (i) on reduced sur- faces the reduction of formic acid to formaldehyde caused the oxidation of surface Ti cations and (ii) on

This trend means that Fe atoms in off- stoichiometric Fe-Al alloys with excess Fe atoms sitting on Al sites have magnetic moments and the number of magnetic Fe atoms can be

The layer thickness range of the trilayers (8 nm for Ag and Cr and 25 nm for Fe) was small enough that the interface magnetoresistance was not shunted by the resistance of the

On the other hand, a recent 57 Fe M¨ossbauer spectroscopy (MS) study of a low B content nanocrystalline Fe–Zr–B–Cu alloy did not find [5] any appreciable smearing out of the

M € ossbauer parameters of the spectra measured at RT and 20 K; Isomer shift (IS), quadrupole splitting (QS) and linewidth (W) of the doublet, average hyperfine field of the bcc

Below a critical Fe-layer thickness, depending on t Ag and n, the iron atoms do not form a continuous layer but form clusters with average size decreasing with decreasing t Fe and

Analysis of the results obtained by MS for sequence permutated sample pairs had shown [2] that the composition of the amorphous Fe/B and B/Fe interfaces are different, chemical

With the assumption that all components form a continuous layer, the widths of the paramagnetic and the mag- netic layers within the interfaces can be estimated from the d para and