You are on page 1of 9

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Research Article

Cite This: ACS Appl. Mater. Interfaces 2019, 11, 39315−39323 www.acsami.org

Interfacial Silicide Formation and Stress Evolution during Sputter


Deposition of Ultrathin Pd Layers on a‑Si
Bärbel Krause,*,† Gregory Abadias,∥ Clarisse Furgeaud,∥ Anny Michel,∥ Andrea Resta,⊥
Alessandro Coati,⊥ Yves Garreau,⊥,# Alina Vlad,⊥ Dirk Hauschild,†,‡ and Tilo Baumbach†,§

Institut für Photonenforschung und Synchrotronstrahlung (IPS), Karlsruher Institut für Technologie, Eggenstein-Leoplodshafen
76344, Germany

Institut für Technische Chemie und Polymerchemie (ITCP), and §Laboratorium für Applikationen der Synchrotronstrahlung
(LAS), Karlsruher Institut für Technologie, Karlsruhe 76131, Germany

See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Institut PPrime, UPR 3346, Université de Poitiers-CNRS-ENSMA, Chasseneuil-Futuroscope 86960, France



Synchrotron SOLEIL, Gif sur Yvette 91192, France
#
Laboratoire Matériaux et Phénomènes Quantiques, Université de Paris, Paris 75013, France
Downloaded via 27.53.253.135 on September 29, 2022 at 06:16:58 (UTC).

ABSTRACT: Synchrotron experiments combining real-time stress, X-ray diffraction,


and X-ray reflectivity measurements, complemented by in situ electron diffraction and
photon electron spectroscopy measurements, revealed a detailed picture of the
interfacial silicide formation during deposition of ultrathin Pd layers on amorphous
silicon. Initially, an amorphous Pd2Si interlayer is formed. At a critical thickness of 2.3
nm, this layer crystallizes and the resulting volume reduction leads to a tensile stress
buildup. The [111] textured Pd2Si layer continues to grow up to a thickness of ≈3.7 nm
and is subsequently covered by a Pd layer with [111] texture. The tensile stress relaxes
already during Pd2Si growth. A comparison between the texture formation on SiOx and
a-Si shows that the silicide layer serves as a template for the Pd layer, resulting in a surprisingly narrow texture of only 3° after
800 s Pd deposition. The texture formation of Pd and Pd2Si can be explained by the low lattice mismatch between Pd(111) and
Pd2Si(111). The combined experimental results indicate a similar interface formation mechanism for Pd on a-Si and c-Si,
whereas the resulting silicide texture depends on the Si surface. A new strain relaxation mechanism via grain boundary diffusion
is proposed, taking into account the influence of the thickness-dependent crystallization on the material transport through the
silicide layer. In combination with the small lattice mismatch, the grain boundary diffusion facilitates the growth of Pd clusters,
explaining thus the well-defined thickness of the interfacial silicide layer, which limits the miniaturization of self-organized
silicide layers for microelectronic devices.
KEYWORDS: in situ, real-time, X-ray reflectivity, X-ray diffraction, stress, silicide, sputter deposition, XPS

1. INTRODUCTION Due to their low crystallization temperature, Ni- and Pd-based


Silicide layers have two important functions in silicon-based silicides are of great interest for microelectronic applications.1,3,7
microelectronic devices:1−3 They lower the thermal budget for The model system Pd/Si forms only one silicide phase (Pd2Si)
the production of polysilicon, and they are employed as ohmic over a wide temperature range, reducing thus the number of
contact and local interconnect. The layers are typically produced competing structure formation mechanisms.8,9 Bulk Pd has an
in a two-step process. A metal layer is deposited on silicon and fcc unit cell with a = 3.891 Å,10 whereas Pd2Si is hexagonal with a
subsequently annealed. However, already during deposition, a = 6.496 Å and c = 3.433 Å.11 Our study focuses on the interfacial
nanometer-scale silicide interlayer might develop at the silicide formation and its influence on the structure during
interface.3 Such a layer can serve as a seed for the annealing- subsequent metal growth. This includes the stress buildup and
induced silicide growth, influencing thus its microstructure and relaxation during deposition, which is a key factor for all
resulting electronic properties. Furthermore, the naturally industrial applications.
forming interlayer represents a lower limit for the accessible The silicide formation during thin film deposition is difficult
silicide thickness. to access. Depending on the envisioned application, the metal
Numerous studies focus on the silicide formation during post- films are thermally evaporated or deposited by magnetron
growth annealing including, e.g., refs 1, 4−6. With the ongoing sputtering on amorphous (a-Si) or crystalline (c-Si) silicon
miniaturization of microelectronic devices, however, the inter- surfaces. Only few authors report on the deposition of Pd on a-Si
face formation during deposition becomes increasingly relevant.
In the following, we will present a detailed in situ and real-time Received: July 5, 2019
study of the chemical interface reaction and structure formation Accepted: September 24, 2019
during sputter deposition of palladium on amorphous silicon. Published: September 24, 2019

© 2019 American Chemical Society 39315 DOI: 10.1021/acsami.9b11492


ACS Appl. Mater. Interfaces 2019, 11, 39315−39323
ACS Applied Materials & Interfaces Research Article

Figure 1. Measurement geometry (left column) and simultaneous real-time measurements (right column) during deposition of Pd on a-Si: (a) Stress
measurements, (b) diffraction data, and (c) X-ray reflectivity measurements (black) and fit curve (red). The deposition periods of a-Si and Pd as well as
the observed Bragg peaks of Pd and Pd2Si (yellow bars) are indicated. tc indicates the onset of crystallization. The scale of F/w was optimized to the Pd
signal. The inset of (a) shows the signal during Si deposition, and the red circle marks a weak but reproducible stress change at t = 180 s.

at room temperature.9,12−15 It was found that the chemical Detailed information about the electronic and chemical
reaction between Pd and c-Si occurs instantly during enrivonment of Pd close to the interface was revealed by XPS
deposition16 and is already activated at 150 K.16,17 During and Auger electron spectroscopy (AES).
early growth stages, several reports propose the formation of The combined experimental in situ and real-time approach
amorphous silicide on c-Si.18−21 For thicker Pd layers, epitaxial provides an encompassing picture of the structure formation
Pd2Si(001) was found on Si(111).4,20,22,23 Layers with gyroaxial processes during deposition. In the following, this will be
texture were observed on Si(100).6 demonstrated for the complex Pd/a-Si interface, which exhibits
The reported silicide thickness varies significantly. The main a competition between high interface reactivity and crystal
reason is that the Pd2Si formation can also be induced by ion growth.
beam irradiation24 as used, e.g., for X-ray photoelectron
spectroscopy (XPS) depth profiles and transmission electron 2. EXPERIMENTAL SECTION
microscopy (TEM) sample preparation. However, for both All samples were prepared in a modular and portable sputter deposition
evaporation and sputter deposition of Pd, the lowest reported chamber designed for in situ X-ray experiments.29 For the real-time
values are in the range of 3−5 nm.16,18,25,26 growth studies, the chamber with a base pressure of 10−6 Pa was
To cover different aspects of the Pd/a-Si interface formation, installed on the multi-environment diffractometer (MED) of the SIXS
the here presented study combines real-time growth experi- beamline (synchrotron SOLEIL, France).27 For the lab-based growth
ments, performed at the beamline SIXS of the synchrotron studies, the chamber was docked to a UHV cluster with a base pressure
SOLEIL,27 with lab-based in situ electron spectroscopy and 10−8 Pa.
The grounded Si(100) substrates were covered by native oxide. Pd
reflection high-energy electron diffraction (RHEED) measure- films with up to 40 nm thickness were sputter-deposited on a-Si buffer
ments. During the real-time measurements, we monitored layers with thicknesses of 4.5 and 9 nm. Within the experimental error,
simultaneously the X-ray diffraction (XRD) signal, the reflected the results were independent of the selected buffer layer thickness.
X-ray beam, and the evolution of the sample curvature, which Therefore, in the following, we will focus on the results for 9 nm a-Si.
reflects the stress buildup during deposition.28 As we will show, For comparison, several Pd reference layers were deposited without a
the stress signal is highly sensitive to microstructural changes buffer layer. During preparation, no additional sample heating was
within the deposited film and is thus ideally suited as low-cost applied. Due to the requirements of the stress measurements, the real-
time experiments were performed on unclamped Si wafers with 100 μm
growth monitor. However, its interpretation in terms of
thickness. All other samples were deposited on clamped Si wafers with 1
microstructure is often ambiguous. This problem is overcome mm thickness. The palladium and silicon sputter targets with a diameter
by simultaneous X-ray measurements at grazing incidence, of 50 mm were mounted at a deposition angle of 18° and target−
which give information about phase formation, layer thickness, substrate distances of 325 and 129 mm, respectively. The a-Si buffer
roughness development, grain growth, and texture formation. layer was deposited with an RF power of 60 W, resulting in a deposition

39316 DOI: 10.1021/acsami.9b11492


ACS Appl. Mater. Interfaces 2019, 11, 39315−39323
ACS Applied Materials & Interfaces Research Article

rate of 0.032 nm/s. Pd was deposited at 40 W DC power, corresponding reproducibility of these changes was verified by monitoring the
to a deposition rate of 0.021 nm/s. The process gas, argon, had a flow deposition of several samples using the same growth condition.
rate of 0.8 sccm, corresponding to a pressure of 0.365 Pa. In the following, the observations will be analyzed in detail.
During sputter deposition, the samples were simultaneously studied The deposition of Si is accompanied by a compressive stress
by a combination of three in situ and real-time methods: X-ray
reflectivity (XRR), XRD, and optical stress measurements. A detailed
signal (Figure 1a), as already reported in ref 28. The deposition
description of the experimental approach can be found elsewhere.28 of Pd, however, leads to several rapid stress changes, which can
The measurement geometry is schematically shown in Figure 1 (left only be understood taking into account the phase evolution
column). For the here presented data, a photon energy, E, of 12 keV and during deposition probed by XRD (Figure 1b). The time-
an incident beam angle, αi, of 1.88° were employed. Under these dependent diffraction signal is constant during deposition of a-Si
conditions, the reflected X-ray beam is expected to oscillate with a but evolves rapidly during the initial 300 s of Pd deposition. Line
period of 1.6 nm during Pd deposition.30 The XRR signal was measured scans along q and t, shown in Figure 2, highlight the
with a Cyberstar YAP scintillation detector, mounted at 1 m distance to
the sample with a detector slit of 0.5 × 2 mm2 (vertical × horizontal).
The XRD signal was recorded with a 2D pixel detector (XPAD). Both
detectors were protected against overexposure by automatic filters.31 At
the optimized sample−detector distance 283 mm, the 2D detector
covered the 2θ range 20−32°. Due to the grazing incidence of the
incoming X-ray beam, the resulting momentum transfer q = 4π/
λ sin(θ) was tilted by ∼14° with respect to the sample surface normal.
The intrinsic (growth) stress was monitored using a multiple (3 × 3)
laser beam optical stress sensor (k-Space Associates) with a curvature
sensitivity better than 2 × 10−4 m−1. The film force per unit width (F/
w), equivalent to the stress × thickness product, σ × h, was derived from
the change in curvature Δκ using the Stoney equation.
Before and after deposition, wide-angle diffraction patterns were
collected by scanning the detector in horizontal and vertical direction.
Approximately 3000 images were taken during 10 continuous
horizontal detector scans at different vertical detector angles. From
this, reciprocal space maps were reconstructed, which show the
measured X-ray intensity as a function of q∥ and qz (the components of
q parallel and perpendicular to the sample surface). All presented maps
are corrected by the scattering background measured before deposition.
In addition to the real-time study, an in situ surface characterization
study was performed on a second sample series. The atomic ordering
close to the surface was studied by RHEED measurements with an
REG30 electron gun (Dr Gassler Electronic Devices) and an electron
energy of 20 keV. The chemical structure at the surface was
characterized by XPS using a nonmonochromated XR-50 Mg Kα X-
ray source and a Phoibos 150 analyzer (SPECS). The energy axis was Figure 2. Line scans extracted from the real-time diffraction data during
calibrated with a Ag reference sample. The XPS and AES peak positions deposition of Pd on a-Si (shown in Figure 1b), at (a) different Pd
were determined with an experimental error of ±0.1 eV. The estimated deposition times with an integration range of ±1 s and (b) different q
error for nominal composition (assuming a homogeneous material with an integration range of ±0.03 Å−1. The selected q values are
distribution) is 10%. indicated in (a) by dashed lines.
Sputter deposition is a nonequilibrium process, and the thin film
growth might be significantly influenced by the kinetic energy of the
deposited atoms. To exclude these effects, we performed ballistic characteristic features. During initial growth, a broad intensity
simulations of the entire transport process, from the ejection of the Pd distribution is observed at q = 2.76 Å−1 (black line in Figure 2a),
atoms at the target, simulated by SRIM,32 to their deposition on the indicating the formation of an amorphous or nanocrystalline (in
sample, using SIMTRA33 and TRIDYN34 (not shown here). The the following called “amorphous”) layer. At the Pd deposition
simulations confirmed that under the here used sputter conditions time t = tc ≈ 80 s (1.6 nm Pd), the broad peak suddenly
substrate heating and ballistic contributions to intermixing at the decreases. Simultaneously, narrow Pd2Si(111) and Pd2Si(201)
interface can be neglected. For the here discussed samples, the sputter
deposition process is dominated by surface and bulk diffusion peaks appear (Figure 2a, red line). For tc ≤ t ≤ 130 s, the intense
processes, similar to thermal evaporation and e-beam deposition. Pd2Si(111) Bragg peak reveals the dominant formation of Pd2Si
crystallites with [111] texture. A narrow size distribution of the
crystallites is confirmed by weak Kiessig fringes (Figure 1b, the
3. RESULTS inclined light-blue stripe marked by an arrow). The onset of
In the following, we will present real-time data measured during silicide crystallization, tc, is indicated by a dashed line in Figure 1.
sputter deposition of Pd on a-Si, complemented by post-growth During further Pd deposition, a Pd(111) peak appears in
RHEED and XPS measurements. between the silicide peaks (gray arrow). The intensity at the
3.1. Real-Time Observations during Sputter Deposi- Pd(111) position increases continuously up to t ≈ 300 s. Due to
tion. Figure 1 summarizes (a) the film force, (b) the XRD, and the peak overlap, it is very difficult to determine precisely the
(c) the XRR data measured simultaneously during deposition of onset of Pd growth. The line scan extracted at a slightly offset
Pd on a 9 nm thick a-Si buffer layer. The deposition time t = 0 s position, q = 3.1 Å−1, (Figure 2b, dashed green circle) shows first
corresponds to the onset of Pd deposition. The schematics show the reduction of the broad peak due to silicide crystallization,
the respective signal path (indicated by red arrows). During Pd then a clear intensity increase for t ≥ 100 s, indicating a change in
deposition, the characteristic features of all signals are highly structure formation already before t = 130 s. Possible
correlated and indicate significant structural changes. The explanations for this will be discussed in Section 3.4.
39317 DOI: 10.1021/acsami.9b11492
ACS Appl. Mater. Interfaces 2019, 11, 39315−39323
ACS Applied Materials & Interfaces Research Article

For t ≤ 120 s, i.e., before the appearance of the significantly However, a steady-state regime is not observed with increasing
overlapping Pd(111) peak, the intensity distribution along q was film thickness, suggesting an evolving film microstructure, which

ÄÅ É2
yÑÑÑ
fitted by
ÅÅ i D
continuously alters the stress formation/relaxation processes.

I(q) = ∑ aiÅÅÅsincjjj (q − qi)zzzÑÑÑ e−(q − qi) σD


The real-time data indicate a competition between the silicide
ÅÇ k 2π {ÑÑÖ
i = 1,2 Å
2 2
formation at the interface and a Pd layer covering this interlayer.
For t ≳ 300 s, the XRD intensity saturates, whereas the XRR data
2a w clearly show that the Pd film is still growing. This can only be
+ 3 understood taking into account the texture formation during
π 4(q − q3)2 + w 2 (1) deposition, which is discussed in the following.
The first term describes the (111) and (201) silicide Bragg peaks 3.2. Texture Formation. The texture at different Pd
at the positions q1 and q2, using the same average crystallite size deposition times t was studied by post-growth XRD and
D and size distribution σD for both crystal orientations. The RHEED measurements. Figure 4a−c shows selected reciprocal
second term, a Lorentz peak at q3 with the width w, describes the space maps. Due to the measurement geometry, the reciprocal
broad intensity distribution observed during initial growth. space region around the surface normal (the qz axis) is not
Figure 3 shows D and σD (visualized as a gray region), plotted as accessible. The corresponding RHEED images (Figure 4e−g)
are presented in the right column. In agreement with the real-

Figure 3. Vertical size D of the silicide crystals, plotted as a function of


the nominal silicide thickness. For the fit, a size distribution (visualized
as a gray region) was taken into account. The red line corresponds to a
complete conversion of Pd into crystalline Pd2Si.

a function of the nominal silicide thickness, which is obtained by


converting the entire deposited Pd to Pd2Si. The error bar
indicates the fitting error of D. The fitting error of σD (not
shown) is below 0.1 nm for nominal thicknesses larger than 2.6
nm and increases up to 0.3 nm for lower values. Assuming bulk
densities, the formation of 1 nm Pd2Si requires 0.48 nm Si and
0.70 nm Pd. Up to the nominal Pd2Si thickness of 3.5 nm at t =
120 s (i.e., the entire plotted range), D is close to the red line
corresponding to a complete conversion of the deposited Pd to
crystalline silicide.
The silicide crystallization is accompanied by a sudden
increase of the force per unit width (Figure 1a). The tensile
stress observed for tc ≤ t ≲ 100 s increases up to 4.2 GPa and is
related to the volume reduction due to crystallization.28
However, already at t ≈ 100 s (2 nm Pd), the force per unit
width abruptly decreases and nearly the entire stress built up
during crystallization is relaxed. During further deposition, a low
compressive stress develops (−510 MPa at t = 350 s, −310 MPa
at t = 600 s), preceded by a characteristic and reproducible stress
change at t = 180 s (marked by a red circle).
The structural changes are also visible in the XRR signal.
Around t = 100 s, the oscillation period τ of the XRR signal
increases after two narrow oscillations from τ = 56 ± 3 s to τ = 79
± 2 s. These wider oscillations are related to the Pd growth. For t Figure 4. (a−c) XRD maps and (e−g) RHEED measurements for Pd
≳ 200 s, i.e., after the small stress feature, the oscillation deposition on a-Si with different deposition times. The red square in (e)
amplitude decreases continuously as expected for increasing corresponds to the XRD region. (d) and (h) show measurements of a
surface roughness. The Pd growth is accompanied by a slowly Pd film deposited on a Si substrate covered by native oxide. In (a) and
developing compressive stress (Figure 1a), which is typically (c), selected diffraction peaks of [111] textured Pd2Si (red dots) and Pd
observed for high-mobility polycrystalline metallic films.35 (white squares) are indicated.

39318 DOI: 10.1021/acsami.9b11492


ACS Appl. Mater. Interfaces 2019, 11, 39315−39323
ACS Applied Materials & Interfaces Research Article

time measurements, all Bragg peaks can be attributed to [111]


textured Pd and Pd2Si. The calculated Bragg peak positions of
Pd2Si are indicated by red dots, and the Pd peaks are shown as
white squares (Figure 4a,c). Selected Pd2Si Bragg peaks are
labeled in Figure 4a. The labels in Figure 4c correspond to [111]
textured Pd. At t = 150 s (3 nm Pd), only Pd2Si peaks with a
mosaicity of 10° are observed (Figure 4a). The diffraction rings
are very narrow along q∥ but much broader along qz. This is
related to the anisotropic shape of the silicide crystallites.
Neglecting small strain effects, a disk shape with a diameter of
9.0 ± 1.0 nm and a height of 3.9 ± 0.3 nm was determined from
the peak width. At t = 225 s (4.5 nm Pd), additional Pd peaks can
be distinguished (Figure 4b). The Pd peaks replicate the
interlayer mosaicity, indicating an orientation relationship
between the Pd film and the Pd2Si interlayer. After t = 800 s
(16 nm Pd), however, the orientation distribution of the Pd layer
reduces to 3° mosaicity (Figure 4c). Note that the deposition of
800 s Pd on SiOx (Figure 4d) results in a much larger mosaicity
of 25°.
For t ≥ 225 s, the crystallization of Pd was also confirmed by
RHEED measurements. Examples are shown in Figure 4e−g.
The streaky RHEED images observed for t ≥ 225 s are quite Figure 5. (a) Time-dependent layer thickness and (b) surface
distinct from the typical transmission image of a textured film, as roughness employed for the simulation of the real-time XRR data
found, e.g., for the Pd film on native oxide (Figure 4h). The during deposition of Pd/a-Si.
reason for this is the comparatively narrow texture, which is not
common for polycrystalline thin films. For t ≤ 180 s, however,
the RHEED intensity is very diffuse, and only one intensity ring Figure 6a shows exemplary measurements of the Pd 3d5/2
is visible. The Pd2Si formation, which is unequivocally proven by photoelectron peak. For t ≤ 120 s, the binding energy 336.6 ±
XRD, is not detectable by RHEED. 0.1 eV is characteristic for the Pd−Si bonds in Pd2Si36,37 and is
The XRD and RHEED measurements show that the Pd attributed to a Si 3p−Pd 4d hybridization.38 For increasing
texture narrows during deposition. The real-time measurements deposition time, the Pd 3d5/2 peak shifts to lower binding
are performed at grazing incidence and cannot access the energies, and at t = 360 s, Pd is in a metallic phase, indicated by
angular region close to the surface normal. This explains why the an asymmetric line profile with the maximum intensity at 335.0
time-dependent XRD signal (Figure 1b) does not change ± 0.1 eV.36,39 In between, the Pd 3d5/2 peak shape and position
significantly for t > 300 s. (Figure 6b, black dots, left scale) change continuously, and the
3.3. Simulation of the XRR Data. The real-time XRD data peak becomes slightly broader. The gray background in Figure
6b,c indicates regions with dominant covalent or metallic bonds,
suggest the following growth scenario: For t ≤ tc = 80 s, an
respectively.
amorphous phase is formed. For t > tc, this phase is converted to
Figure 7 shows the Si 2p XPS peak as a function of the Pd
crystalline Pd2Si with [111] texture and continues to grow.
deposition time. For t = 40 s, the most intense peak is observed.
Above t ≈ 130 s, a [111] textured Pd layer is deposited. A
The peak intensity decreases with increasing deposition time.
simplified fit model for the XRR data was developed, assuming a Excluding t = 40 s (1.1 nm Pd2Si) where the signal is dominated
sequential formation of Si, Pd2Si, and Pd layers with bulk by the a-Si contribution, Si 2p is found at 99.8 ± 0.1 eV, which is
densities. The manually optimized layer model is summarized in in agreement with the Pd−Si bond formation.36
Figure 5, and the calculated XRR is shown as a red line in Figure A peak shift with increasing Pd deposition time was also found
1c. For the here presented data, the deposition of a-Si and Pd is for the Pd MVV Auger signal, which has a higher surface
described by constant deposition rates FPd = 0.0207 nm/s and sensitivity. The position of the main peak is plotted in Figure 6b
FSi = 0.032 nm/s. For t ≤ 127 s, the formation of Pd2Si is (red stars, right scale). The strongest changes of the XPS and
described by a complete conversion of the deposited Pd to AES peak position occur in the regions 120−150 and 200−260
Pd2Si, reducing the Si layer accordingly. The time-dependent s, respectively. As a guide to the eye, the nominal composition
surface roughness σ is shown in Figure 5b. For t ≤ 210 s, the Pd (assuming a homogeneous material distribution) is shown in
roughness is in the range 0.25 ± 0.05 nm. For t ≥ 210 s, however, Figure 6c. The composition was determined from high-
the roughness increases continuously up to 0.55 nm at t = 600 s. resolution scans of the Pd 3d5/2 and Si 1s peaks, excluding a
The fitting error for the deposition rates is ±0.0005 nm; the small Si−O contribution of less than 2 %. At t = 40 s, the nominal
fitting error of the roughness values is ±0.05 nm. composition is still dominated by the a-Si buffer layer. For 100 ≤
Combining the XRR model and the real-time XRD data, we t ≤ 200 s, the slightly increasing nominal composition is close to
propose that the initially forming amorphous phase has already a Pd2Si. The increase of the Pd content accelerates for t ≥ 200 s
composition close to stoichiometric Pd2Si. and saturates after t ≈ 360 s at ≈100% Pd.
3.4. XPS Study of the Pd/a-Si Interface. For a more The XPS and AES peak shifts for Pd on a-Si are in agreement
detailed picture of the interface formation, post-growth in situ with results reported for Pd evaporation on crystalline Si(100)
XPS measurements were done at different Pd deposition times. and Si(111) surfaces.16,36,40 As already observed by Grunthaner
To avoid changes induced by repeated growth interruptions, et al., the XPS peaks cannot be reproduced by a simple
each measurement was performed on a new sample. superposition of Pd−Pd and Pd−Si reference peaks. The peak
39319 DOI: 10.1021/acsami.9b11492
ACS Appl. Mater. Interfaces 2019, 11, 39315−39323
ACS Applied Materials & Interfaces Research Article

shape was fixed using a Gaussian/Lorentzian product form for


Pd−Si and an asymmetric variation of this function for Pd−Pd.
The width of both components was limited to the range 1.2−
1.25 eV estimated from the Pd−Si peak at t < 130 s and the Pd−
Pd peak for t = 360 s. Examples for the peak fit are shown in
Figure 8. The fitted peak positions and the Pd−Pd fraction are

Figure 6. Results of X-ray electron spectroscopy measurements. (a)


Selected Pd 3d5/2 photoelectron peaks after subtraction of a Shirley
background. (b) Position of the Pd 3d5/2 XPS peak and the main MVV
Auger peak as a function of the Pd deposition time. For the XPS peak,
the binding energy is plotted and for the AES peak the kinetic energy.
The composition (assuming homogeneous material distribution) is
shown in (c). The red line indicates the composition of stoichiometric
Pd2Si. The XPS data were fitted with a two component model. The two
components were attributed to Pd−Pd and Pd−Si. The corresponding
peak positions are shown in (b) and the Pd−Pd fraction in (c).

Figure 8. Two-component fit of the Pd 3d5/2 photoelectron peaks


shown in Figure 6a. The data are shown as dots, the Pd−Si component
as a thick blue line, the Pd−Pd component as a thin blue line, the
background as a black line, and the fit result as a red line. The dashed
lines indicate the peak positions at the Pd deposition times 120 and 360
s.

plotted in Figure 6b,c (blue symbols). The core-level peak shift


of both peaks is ≈0.7 eV, which is similar to the interface-
induced peak shifts observed for elemental metals.41
The XRD signal suggests structural changes already for 100 < t
< 130 s (Figure 2b, green circle), but the XPS and AES signals
exclude the formation of a stable Pd layer for t < 120 s. A likely
explanation is the surface nucleation of less ordered Pd2Si
crystallites in addition to the continued growth of the [111]
oriented Pd2Si crystals. Alternatively, the formation of transient
Figure 7. Si 2p XPS peak as a function of the Pd deposition time. The Pd clusters might occur, which is not anymore detectable by
dashed lines indicate the two observed peak positions. post-growth measurements. For 130 ≤ t ≤ 200 s (1.5 nm or 6.5
monolayers of fcc Pd), two shifted components are required to
shift was attributed to a chemical gradient. However, ab initio fit the XPS peak. For t > 200 s, the Pd−Pd component moves to
calculations show similar peak shifts for elemental metal/metal the position expected for bulk Pd and increases with film
interfaces, affecting a thickness range of ±5−6 monolayers thickness. This is in agreement with the formation of a sharp or
around the interface. Assuming intermixing, this region is further slightly intermixed Pd/Pd2Si interface at t ≈ 130 s, followed by
broadened.41 Taking into account possible peak shifts, the data six monolayers Pd where the electronic state is still influenced by
were fitted assuming a Shirley background and two peaks the interface. For t ≥ 200 s, an increasing layer of unperturbed
corresponding to the Pd−Pd and the Pd−Si bond. The peak metallic Pd grows.
39320 DOI: 10.1021/acsami.9b11492
ACS Appl. Mater. Interfaces 2019, 11, 39315−39323
ACS Applied Materials & Interfaces Research Article

4. DISCUSSION on a-Si.28 For Mo, the bcc metal crystallizes, whereas for Pd,
4.1. Structure Formation Mechanisms and Stress significant intermixing occurs during deposition, favoring
Development. The relation between structural development crystallization of silicide. The tensile stress occurring during
and stress evolution is summarized in Figure 9. Combining real- crystallization of Mo is propagated via homoepitaxy into the
continuously growing metal film. For the deposition of Pd on a-
Si, the limited thickness of the silicide prevents the propagation
of tensile stress over larger thicknesses but does not explain the
complete relaxation directly after stress buildup. It is likely that
the grain boundary diffusion, which distinguishes the Mo/a-Si
and Pd/a-Si interface formation, facilitates the relaxation of the
tensile stress, which is maintained only by the contact area
between neighboring grains.
4.2. Relation between Interface Formation and Pd
Texture Evolution. Comparing the structure formation during
deposition of Pd on SiOx and Pd on a-Si, it is obvious that the
silicide formation at the interface is crucial for the texture
formation of the Pd layer. For both substrates, the [111]
orientation of Pd (i.e., the stacking direction of the close-packed
lattice planes) is favored. On native oxide, the weak interaction
Figure 9. Schematic of the structural development during Pd with the substrate surface results in a nearly random orientation
deposition on a-Si, visualizing the relation between structure and stress distribution of the crystallites. On a-Si, the well-oriented Pd2Si
buildup. Before crystallization, the diffusion through the amorphous interlayer serves as a template for the Pd nuclei. This can be
silicide is dominant; after crystallization, the silicide formation is imagined as a “head start” for the overgrowth process and
controlled by grain boundary diffusion (indicated by arrows). explains the surprisingly narrow orientation distribution of thick
Pd films deposited on a-Si.
time and post-growth studies of the Pd/a-Si interface, we found The question that remains is why Pd2Si improves the texture
that for t < tc ≈ 80 s, the deposited Pd is instantly converted to of the Pd film. Figure 10 compares the atomic ordering of
amorphous Pd2Si [Figure 9(1)]. At tc, the amorphous layer with
a nominal thickness of d(Pd2Si) = 2.3 nm begins to crystallize
[Figure 9(2)]. The volume reduction due to crystallization
results in a sudden tensile stress, building up during tc ≤ t ≤ 100
s. At t ≈ 100 s, the entire silicide film is crystallized [Figure
9(3)]. The [111] oriented Pd2Si crystallites continue to grow up
to d(Pd2Si) ≈ 3.7 nm at t ≈ 130 s, accompanied by the
nucleation of randomly oriented Pd2Si crystallites and/or
transient Pd clusters. Simultaneously, the tensile stress is
released [Figure 9(4)]. For t ≳ 130 s, the Pd growth dominates.
The Pd/Pd2Si interface is formed, and at t ≈ 200 s, a
homogeneous [111] textured Pd layer covers the surface. The
subsequent Pd growth is accompanied by the development of
compressive stress (∼−310 MPa) [Figure 9(5)]. During silicide
growth and Pd/Pd2Si interface formation, the surface roughness
Figure 10. Epitaxial relationship between Pd2Si(111) and Pd(111).
is very low, comparable to the a-Si buffer layer. During further The atomic positions of Pd2Si(111) are represented by balls, and the
deposition, the roughness of the Pd layer increases and the crossing points of the black lines correspond to the atomic positions of
texture narrows. Pd(111). The green triangles indicate tetrahedrons formed by three Pd
It is well known that the grain boundary diffusion plays a and a central Si atom.
major role for the silicide formation during annealing.42 During
the early stages of thin film growth, however, these grain
boundaries do not yet exist. As long as the surface is only Pd2Si(111) and Pd(111). The [11̅0] directions of both lattices
partially covered by Pd2Si, the required material transport might are aligned. The lattice spacing of Pd(11̅1) is slightly larger,
occur via surface diffusion. Once a continuous silicide layer is resulting in a misfit of 2.2%. In the orthogonal direction, the
formed [Figure 9(1)], Pd2Si growth can only proceed via Pd Pd2Si(111̅) and Pd(112̅) lattice planes are nearly parallel and
and/or Si diffusion through a-Pd2Si. For t > 100 s, the entire have a projected mismatch of −0.8%. Several authors report
layer is crystallized and the grain boundary diffusion known from [001] textured Pd 2Si after annealing of Pd/a-Si layer
annealing processes can take place [indicated by arrows in systems.14,15 Compared to Pd2Si(001)/Pd(111), where an
Figure 9(4)]. The onset of metallic Pd formation indicates that isotropic lattice mismatch of 2.2% is found, the strain energy of
the material transport via grain boundaries is significantly Pd2Si(111)/Pd(111) is slightly lower. This explains why in our
reduced compared to the transport through a-Pd2Si. The bulk case the [111] texture of Pd2Si is favored.
diffusion within c-Pd2Si is not expected to contribute to the The silicide texture can be explained by its small lattice
interlayer formation since it is even lower than the Pd2Si grain mismatch with Pd(111), but experimentally, Pd was detected
boundary diffusion.42 only after silicide crystallization. This might be due to the fast
A thickness-dependent crystallization at tc, accompanied by a reaction of the deposited Pd but might also be related to the
tensile stress buildup, was also observed during deposition of Mo detection limit of the here used measurement techniques. A
39321 DOI: 10.1021/acsami.9b11492
ACS Appl. Mater. Interfaces 2019, 11, 39315−39323
ACS Applied Materials & Interfaces Research Article

surface-induced crystallization, triggered by transient Pd mobility and reactivity of the deposited metal atoms determine
clusters, would be in agreement with the comparatively narrow the crystallizing phase (metal or silicide) and controls thus the
Pd2Si texture. A candidate for the nucleation sites might be structure formation during further metal deposition. This is
triangular Pd clusters. As shown in Figure 10, these clusters are crucial for the stress buildup during deposition: if the metal
subunits of the close-packed Pd(111) layers evidenced during phase crystallizes, the tensile strain due to crystallization of the
later deposition. With a much lower density, they also occur in initially deposited amorphous phase can propagate to the
the Pd2Si(111) plane (green triangles). Buckley et al. already subsequently deposited metal layer via coherent crystal growth.
suggested that these triangular configurations play an important The crystalline silicide layer, however, grows via grain boundary
role for the epitaxial growth of Pd2Si(001) on Si(111).4 diffusion, facilitating the relaxation of the tensile stress. The
4.3. Crystalline versus Amorphous Si Surfaces. The detailed understanding of these structure formation mechanisms
deposition of Pd on a-Si and c-Si [referring mainly to the well- during deposition is essential for the miniaturization of silicide
studied Si(111) and Si(100) surfaces] shows many similarities: interconnects and offers new paths for controlling the texture of
in both cases, Pd reacts instantly with the Si surface atoms, metal thin films.
reducing thus the thickness of the underlying silicon layer.16,20
Independent of the substrate crystallinity, the minimum
thickness of the resulting silicide interlayer is around 3−5
■ AUTHOR INFORMATION
Corresponding Author
nm,6,16,18,25 and XPS/AES measurements confirm that the *E-mail: baerbel.krause@kit.edu.
electronic environment at the Pd/c-Si16,36 and Pd/a-Si
interfaces is similar. ORCID
For Pd on a-Si, the crystallization of Pd2Si occurs at a critical Bärbel Krause: 0000-0001-6288-0283
thickness. The crystallization of Pd2Si at the Pd/c-Si interface, Dirk Hauschild: 0000-0001-9088-8944
however, is less well understood. STM measurements on these Notes
layer systems, with up to 8 ML deposited Pd, indicate the The authors declare no competing financial interest.


formation of amorphous silicide.19,21,43 This is in agreement
with the initial growth of Pd on a-Si. Epitaxial silicide was found ACKNOWLEDGMENTS
after deposition of thicker Pd layers on c-Si4,6 but also during
real-time XRD measurements of Pd/Si(111): a relaxed silicide We thank the synchrotron radiation source SOLEIL for the
peak appeared after deposition of 1.6 nm Pd.20,44 The structural provision of beamtime within the standard proposal 20170440
transition observed for Pd/Si(111) occurs at the same critical and acknowledge the support of the synchrotron staff. We are
thickness as the here observed amorphous/crystalline tran- especially grateful for the technical support at the SIXS beamline
sition.44 The transition was explained with the relaxation of by Benjamin Voisin. For the technical assistance during
epitaxially strained silicide. However, in the case of a-Si surfaces, preparation of the synchrotron experiment we thank Annette
the transition cannot be driven by relaxation of an epitaxial strain Weißhardt, Tobias Hilverling, and Andreas Breitenstein.
between film and substrate. The similar observations for Pd/c-Si Furthermore, we acknowledge the general support provided
by the staff of the KIT institutes IBPT and IPS.


and Pd/a-Si interface formation suggest a general structure
formation mechanism, independent of the crystalline order of
the substrate. For a similar system, Ni deposited on Si(111), the REFERENCES
commensurate structure forming during initial deposition was (1) Knaepen, W.; Detavernier, C.; Meirhaeghe, R. V.; Sweet, J. J.;
explained by long-range ordered interfacial silicide with large Lavoie, C. In-situ X-ray Diffraction Study of Metal Induced
atomic displacements with respect to the lattice positions, Crystallization of Amorphous Silicon. Thin Solid Films 2008, 516,
4946−4952.
covered by an amorphous silicide layer.45 Such a displacement (2) Bennett, P.; He, Z.; Smith, D. J.; Ross, F. Endotaxial Silicide
structure can be seen as an intermediate state between Nanowires: A Review. Thin Solid Films 2011, 519, 8434−8440.
amorphous and perfectly crystalline phase. Assuming a similar Selected Papers from the Asia-Pacific Conference on Semiconducting
growth model for Pd on c-Si, it is likely that in both cases the Silicides Science and Technology Towards Sustainable Optoelectronics
structural transition at a critical thickness is mainly driven by the (APAC-SILICIDE 2010), Tsukuba, July 24−26, 2010.
disorder/order transition and the relaxation of epitaxial strain is (3) Alberti, A.; La Magna, A. Role of the Early Stages of Ni-Si
only a secondary effect. Interaction on the Structural Properties of the Reaction Products. J.
Appl. Phys. 2013, 114, No. 121301.
5. SUMMARY AND CONCLUSIONS (4) Buckley, W.; Moss, S. Structure and Electrical Characteristics of
Epitaxial Palladium Silicide Contacts on Single Crystal Silicon and
We have shown that detailed information about complex growth Diffused P-N Diodes. Solid-State Electron. 1972, 15, 1331−1337.
processes on the nanoscale is directly accessible for non-ideal, (5) Rivero, C.; Gergaud, P.; Gailhanou, M.; Thomas, O.; Froment, B.;
comparatively weakly scattering polycrystalline layer systems. Jaouen, H.; Carron, V. Combined Synchrotron X-Ray Diffraction and
Combining real-time XRD, XRR, and optical stress measure- Wafer Curvature Measurements during Ni-Si Reactive Film Formation.
ments with in situ RHEED and XPS, the reactive interface Appl. Phys. Lett. 2005, 87, No. 041904.
formation during deposition of Pd on a-Si was studied. It was (6) Richard, M.-I.; Fouet, J.; Texier, M.; Mocuta, C.; Guichet, C.;
found that [111] textured Pd2Si crystallizes at a critical thickness Thomas, O. Continuous and Collective Grain Rotation in Nanoscale
of 2.3 nm and continues to grow up to ≈3.7 nm, followed by Pd Thin Films during Silicidation. Phys. Rev. Lett. 2015, 115, No. 266101.
(7) Putero, M.; Mangelinck, D. Effect of Pd on the Ni2Si Stress
formation with a surprisingly narrow [111] texture. The Relaxation during the Ni-Silicide Formation at Low Temperature. Appl.
structure formation on a-Si and c-Si is very similar, indicating Phys. Lett. 2012, 101, No. 111910.
that the disorder/order transition is the driving force for the (8) Gergaud, P.; Thomas, O.; Chenevier, B. Stresses Arising from a
formation of a crystalline interlayer. Solid State Reaction between Palladium Films and Si(001) Investigated
The crystallization at a critical thickness seems to be a by In Situ Combined X-Ray Diffraction and Curvature Measurements.
fundamental aspect of metal/a-Si interface formation. The J. Appl. Phys. 2003, 94, 1584−1591.

39322 DOI: 10.1021/acsami.9b11492


ACS Appl. Mater. Interfaces 2019, 11, 39315−39323
ACS Applied Materials & Interfaces Research Article

(9) Molina-Ruiz, M.; Lopeandía, A. F.; González-Silveira, M.; Crystallization and Stress Build-Up during Sputter-Deposition of
Anahory, Y.; Guihard, M.; Garcia, G.; Clavaguera-Mora, M. T.; Nanoscale Silicide Films. ACS Appl. Mater. Interfaces 2016, 8, 34888−
Schiettekatte, F.; Rodríguez-Viejo, J. Formation of Pd2Si on Single- 34895. PMID: 27998117.
Crystalline Si (100) at Ultrafast Heating Rates: An In-Situ Analysis by (29) Krause, B.; Darma, S.; Kaufholz, M.; Gräfe, H.-H.; Ulrich, S.;
Nanocalorimetry. Appl. Phys. Lett. 2013, 102, No. 143111. Mantilla, M.; Weigel, R.; Rembold, S.; Baumbach, T. Modular
(10) Rao, C. N.; Rao, K. K. Effect of Temperature on the Lattice Deposition Chamber for In Situ X-Ray Experiments during RF and
Parameters of some Silver-Palladium Alloys. Can. J. Phys. 1964, 42, DC Magnetron Sputtering. J. Synchrotron Radiat. 2012, 19, 216−222.
1336−1342. (30) Kaufholz, M.; Krause, B.; Kotapati, S.; Köhl, M.; Mantilla, M. F.;
(11) Nylund, A. Some Notes on the Palladium-Silicon System. Acta Stüber, M.; Ulrich, S.; Schneider, R.; Gerthsen, D.; Baumbach, T.
Chem. Scand. 1966, 20, 2381−2386. Monitoring the Thin Film Formation during Sputter Deposition of
(12) Hutchins, G.; Shepela, A. The Growth and Transformation of Vanadium Carbide. J. Synchrotron Radiat. 2015, 22, 76−85.
Pd2Si on (111), (110) and (100) Si. Thin Solid Films 1973, 18, 343− (31) Dawiec, A.; Garreau, Y.; Bisou, J.; Hustache, S.; Kanoute, B.;
363. Picca, F.; Renaud, G.; Coati, A. Real-Time Control of the Beam
(13) Vidal, R.; Ferrón, J. Application of Auger Electron Spectroscopy Attenuation with XPAD Hybrid Pixel Detector. J. Instrum. 2016, 11,
and Principal Component Analysis to the Study of the Pd/c-Si and Pd/ P12018.
a-Si Interfaces. Appl. Surf. Sci. 1988, 31, 263−276. (32) Ziegler, J. F.; Ziegler, M.; Biersack, J. SRIM − The Stopping and
(14) Edelman, F.; Cytermann, C.; Brener, R.; Eizenberg, M.; Weil, R.; Range of Ions in Matter (2010). Nucl. Instrum. Methods Phys. Res., Sect.
Beyer, W. Interfacial Reactions in the Pd/a Si/c Si System. J. Appl. Phys. B 2010, 268, 1818−1823. 19th International Conference on Ion Beam
1992, 71, 289−295. Analysis.
(15) Molina-Ruiz, M.; Ferrando-Villalba, P.; Rodríguez-Tinoco, C.; (33) van Aeken, K.; Mahieu, S.; Depla, D. The Metal Flux from a
Garcia, G.; Rodríguez-Viejo, J.; Peral, I.; Lopeandía, A. F. Simultaneous Rotating Cylindrical Magnetron: a Monte Carlo Simulation. J. Phys. D:
Nanocalorimetry and Fast XRD Measurements to Study the Silicide Appl. Phys. 2008, 41, No. 205307.
Formation in Pd/a-Si Bilayers. J. Synchrotron Radiat. 2015, 22, 717− (34) Möller, W.; Eckstein, W.; Biersack, J. Tridyn-Binary Collision
722. Simulation of Atomic Collisions and Dynamic Composition Changes in
(16) Rubloff, G. W.; Ho, P. S.; Freeouf, J. F.; Lewis, J. E. Chemical Solids. Comput. Phys. Commun. 1988, 51, 355−368.
Bonding and Reactions at the Pd/Si Interface. Phys. Rev. B 1981, 23, (35) Abadias, G.; Chason, E.; Keckes, J.; Sebastiani, M.; Thompson,
4183−4196. G. B.; Barthel, E.; Doll, G. L.; Murray, C. E.; Stoessel, C. H.; Martinu, L.
(17) Okado, H.; Hirono, S.; Mori, H. Formation of Palladium Silicide Review Article: Stress in Thin Films and Coatings: Current Status,
Films on Silicon(111) 7x7 Surface at ∼150 K. Jpn. J. Appl. Phys. 2005, Challenges, and Prospects. J. Vac. Sci. Technol., A 2018, 36, No. 020801.
44, 1393−1396. (36) Grunthaner, P. J.; Grunthaner, F. J.; Madhukar, A.; Mayer, J. W.
(18) Anton, R.; Neukirch, U. Comparative AES and RHEED Study of Metal/Silicon Interface Formation: The Ni/Si and Pd/Si Systems. J.
the Formation of Pd Silicide on Clean and Oxide Covered Si(100) and Vac. Sci. Technol. 1981, 19, 649−656.
(111) Surfaces. Appl. Surf. Sci. 1987, 29, 287−299. (37) Dai, D.-X.; Davoli, I. An Experimental Study of an Interface
(19) Kö hler, U. K.; Demuth, J. E.; Hamers, R. J. Surface Reaction at the Practical Pd/Si Interface by XPS. Vacuum 1995, 46,
Reconstruction and the Nucleation of Palladium Silicide on Si(111). 139−142.
Phys. Rev. Lett. 1988, 60, 2499−2502. (38) Tanaka, K.; Furui, K.; Yamada, M. Electronic Structure and
(20) Bennett, P. A.; DeVries, B.; Robinson, I. K.; Eng, P. J. Layerwise Metal-Insulator Transition in Amorphous Pd-Si Alloy Films. J. Phys.
Reaction at a Buried Interface. Phys. Rev. Lett. 1992, 69, 2539−2542. Soc. Jpn 1995, 64, 4790−4798.
(21) Tanaka, M.; Takeguchi, M.; Furuya, K. In-Situ Observation of (39) Zhang, Y.; Gajjala, G.; Hofmann, T.; Weinhardt, L.; Bär, M.;
Pd2Si Islands on Si by UHV-TEM/STM. J. Cryst. Growth 2002, 237− Heske, C.; Seelmann-Eggebert, M.; Meisen, P. X-Ray Photoelectron
239, 254−258. The Thirteenth International Conference on Crystal Spectroscopy Study of the Chemical Interaction at the Pd/SiC
Growth in Conjunction with the Eleventh International Conference on Interface. J. Appl. Phys. 2010, 108, No. 093702.
Vapor Growth and Epitaxy. (40) Schmid, P. E.; Ho, P. S.; Föll, H.; Rubloff, G. W. Electronic States
(22) Chen, J.; Chen, L. Epitaxial Growth and Thermal Stability of and Atomic Structure at the Pd2Si-Si Interface. J. Vac. Sci. Technol.
Thin Pd2Si Films on (001), (011) and (111) Si. Thin Solid Films 1995, 1981, 18, 937−943.
261, 107−114. (41) Olovsson, W.; Holmström, E.; Marten, T.; Abrikosov, I. A.;
(23) Okado, H.; Sakata, T.; Ukezono, Y.; Hirono, S.; Mori, H. Room Niklasson, A. M. N. Interface Core-Level Shifts as a Probe of Embedded
Temperature Formation of Crystallized Palladium Silicide on Si(111) Thin-Film Quality. Phys. Rev. B 2011, 84, No. 085431.
Observed by Ultrahigh Voltage Electron Microscopy. Jpn. J. Appl. Phys. (42) Portavoce, A.; Hoummada, K.; Dahlem, F. Influence of
2002, 41, L344. Interfacial Reaction upon Atomic Diffusion Studied by In Situ Auger
(24) Anton, R.; Neukirch, U.; Harsdorff, M. Auger-Electron- Electron Spectroscopy. Surf. Sci. 2014, 624, 135−144.
Spectroscopy Analysis of a Plasmon Loss in Palladium Silicide Formed (43) Casalis, L.; Casati, C.; Rosei, R.; Kiskinova, M. Surface Structures
from Pd Deposits on Silicon. Phys. Rev. B 1987, 36, 7422−7427. of Some Stages Preceding Formation of a Pd2Si Phase. Surf. Sci. 1995,
331−333, 381−388.
(25) Lee, R. Y.; Whang, C. N.; Kim, H. K.; Smith, R. J. Silicide
(44) Robinson, I.; Eng, P.; Bennett, P.; DeVries, B. Interfacial X-Ray
Formation by Ar+ Ion Bombardment of Pd/Si. J. Mater. Sci. 1988, 23,
Oscillations during Growth of Pd2Si on Si(111). Appl. Surf. Sci. 1992,
2740−2744.
60−61, 498−504.
(26) Richard, M.-I.; Fouet, J.; Guichet, C.; Mocuta, C.; Thomas, O.
(45) Bennett, P. A.; Lee, M. Y.; Yang, P.; Schuster, R.; Eng, P. J.;
Exploring Pd−Si(001) and Pd−Si(111) Thin-Film Reactions by
Robinson, I. K. Template Structure at the Silicon/Amorphous-Silicide
Simultaneous Synchrotron X-Ray Diffraction and Substrate Curvature
Interface. Phys. Rev. Lett. 1995, 75, 2726−2729.
Measurements. Thin Solid Films 2013, 530, 100−104. 6th Size-Strain
International Conference Diffraction Analysis of the Microstructure of
Materials.
(27) Coati, A.; Chavas, L. M. G.; Fontaine, P.; Foos, N.; Guimaraes,
B.; Gourhant, P.; Legrand, P.; Itie, J. P.; Fertey, P.; Shepard, W.; Isabet,
T.; Sirigu, S.; Solari, P. L.; Thiaudiere, D.; Thompson, A. Status of the
Crystallography Beamlines at Synchrotron SOLEIL. Eur. Phys. J. Plus
2017, 132, 174.
(28) Krause, B.; Abadias, G.; Michel, A.; Wochner, P.; Ibrahimkutty,
S.; Baumbach, T. Direct Observation of the Thickness-Induced

39323 DOI: 10.1021/acsami.9b11492


ACS Appl. Mater. Interfaces 2019, 11, 39315−39323

You might also like