• No results found

Substantial difference in target surface chemistry between reactive dc and high power impulse magnetron sputtering

N/A
N/A
Protected

Academic year: 2021

Share "Substantial difference in target surface chemistry between reactive dc and high power impulse magnetron sputtering"

Copied!
6
0
0

Loading.... (view fulltext now)

Full text

(1)

Journal of Physics D: Applied Physics

LETTER • OPEN ACCESS

Substantial difference in target surface chemistry

between reactive dc and high power impulse

magnetron sputtering

To cite this article: G Greczynski et al 2018 J. Phys. D: Appl. Phys. 51 05LT01

View the article online for updates and enhancements.

Related content

On reactive high power impulse magnetron sputtering

J T Gudmundsson

-Cross-field ion transport during high power impulse magnetron sputtering

Daniel Lundin, Petter Larsson, Erik Wallin et al.

-Process stabilization by peak current regulation in reactive high-power impulse magnetron sputtering of hafnium nitride

T Shimizu, M Villamayor, D Lundin et al.

(2)

1 © 2018 IOP Publishing Ltd Printed in the UK Journal of Physics D: Applied Physics

G Greczynski et al Printed in the UK 05LT01 JPAPBE © 2018 IOP Publishing Ltd 51

J. Phys. D: Appl. Phys.

JPD

10.1088/1361-6463/aaa0ee

5

Journal of Physics D: Applied Physics

Advanced functional coatings with applications ranging from wear-resistant layers on cutting tools [1, 2] to diffusion bar-riers in electronic devices [3, 4] are often produced with the advantage of reactive magnetron sputtering under high vac-uum conditions. The presence of a reactive gas, however, leads not only to the desired compound formation at the substrate, but also to a simultaneous chemisorption at the target that is covered with a compound layer by so-called target poisoning. This results in severe reduction of the sputtering rate and large off-sets in the process parameter operating points, because of the change in the secondary electron emission yield and the chemical composition of the sputtered flux, as well as the energy distribution thereof, [5, 6] ion implantation effects, [7] and process hysteresis [8].

As the latest development in the field of magnetron sput-tering, high power impulse magnetron sputtering (HIPIMS) [9] is mostly known for the ability to provide significant ioniz ation of the sputtered material flux [1014]. In contrast to conventional dc magnetron sputtering (DCMS), where the target state in the reactive process is to a large extent deter-mined by the applied power and the supply of the reactive gas, HIPIMS offers an additional means of control by tuning the pulse length and pulsing frequency, which can potentially lead to new phenomena at the target surface. For example, it was reported recently that a ~500 nm thick compound layer formed on a Ti target surface operated in an oxygen-contain-ing atmosphere by HIPIMS [15], while the typical thickness of the reacted film during DCMS is only a few nm [16, 17].

Experimental studies of the target surface chemistry [18, 19] have been limited by a lack of in situ analytical

Substantial difference in target surface

chemistry between reactive dc and high

power impulse magnetron sputtering

G Greczynski1,2,3 , S Mráz2, J M Schneider2 and L Hultman1

1 Thin Film Physics Division, Department of Physics (IFM), Linköping University, SE-581 83 Linköping, Sweden

2 Materials Chemistry, RWTH Aachen University, Kopernikusstr. 10, D-52074 Aachen, Germany E-mail: grzgr@ifm.liu.se

Received 4 October 2017, revised 2 December 2017 Accepted for publication 12 December 2017 Published 12 January 2018

Abstract

The nitride layer formed in the target race track during the deposition of stoichiometric TiN thin films is a factor 2.5 thicker for high power impulse magnetron sputtering (HIPIMS), compared to conventional dc processing (DCMS). The phenomenon is explained using x-ray photoelectron spectroscopy analysis of the as-operated Ti target surface chemistry supported by sputter depth profiles, dynamic Monte Carlo simulations employing the TRIDYN code, and plasma chemical investigations by ion mass spectrometry. The target chemistry and the thickness of the nitride layer are found to be determined by the implantation of nitrogen ions, predominantly N+ and N+

2 for HIPIMS and DCMS, respectively. Knowledge of this

method-inherent difference enables robust processing of high quality functional coatings. Keywords: HIPIMS, HPPMS, TiN, XPS, magnetron sputtering, target poisoning, target chemistry

(Some figures may appear in colour only in the online journal)

Letter

IOP

2018

1361-6463

3 Author to whom any correspondence should be addressed.

1361-6463/18/05LT01+5$33.00

https://doi.org/10.1088/1361-6463/aaa0ee

J. Phys. D: Appl. Phys. 51 (2018) 05LT01 (5pp)

Made open access 13 February 2018

(3)

2

probes capable of operating under the harsh plasma process conditions used for film growth. An additional snag is that x-ray photoelectron spectroscopy (XPS), used for bonding assignments of magnetron sputtered thin films, is typically performed ex situ, which implies complications arising from the sample oxidation during transport to the spectrometer. Commonly used surface cleaning by Ar+-ion etching is not

a remedy, as it has other destructive effects [20]4, like

pref-erential sputtering, surface roughening, and redeposition of sputtered materials.

Here, we compare the chemical state of a Ti target oper-ating in Ar/N2 atmospheres using DCMS and HIPIMS, with

gas composition optimized to yield stoichiometric TiN films at the substrate. The targets are in situ capped immediately after each film deposition experiment with a few-nm-thick photoelectron-transparent metal-protective layers [21] and subsequently transferred from the growth chamber into the XPS instrument. This new method enables ex situ studies of native target chemistry [22, 23]. High-energy resolution core-level photoelectron spectra recorded from native surfaces, together with sputter-depth profiles reveal essential differ-ences between the two sputtering techniques. HIPIMS opera-tion results in a factor of 2.5× thicker nitride layer than in the case of conventional DCMS, despite the lower N2 partial

pressure pN2 used. Even if operated at pN2, too low to result

in stoichiometric TiN films on the substrate, the target surface region is fully nitrided during HIPIMS, which is in clear con-trast to DCMS [24].

Experiments are conducted in a multi-cathode CC800/9 CemeCon AG magnetron sputtering system [25], employing a rectangular 8.8 × 50 cm2 Ti target. A 2 × 4 cm2 section of

the target, positioned in the middle and across the race track, is made to be detachable to allow for transfer to the XPS sys-tem. The background pressure prior to sputtering is 0.2 mPa (1.5 × 10−6 Torr). The total pressure of Ar/N

2 gas mixture

is kept constant at 0.4 Pa (3 mTorr), with pN2 set at 92 mPa

for DCMS experiments, while pN2 = 45 mPa for HIPIMS. In

both cases, pN2 is the lowest N2 partial pressure to yield

stoi-chiometric TiN films at the substrate. DCMS is operated at a 2 kW average power, resulting in a target voltage of −351 V. The same target power is used for HIPIMS, together with the 200 µs pulse width and the frequency of 200 Hz. Such

experimental conditions result in the incorporation of oxygen from the residual gas [26] causing a bulk oxygen content in the DCMS and HIPIMS films of 0.2 and 0.9 at%, respectively, as assessed by time-of-flight elastic recoil detection analysis. Prior to each experiment, the target is conditioned for 5 min in pure Ar behind closed shutters to reset the target history5. For

each gas mixture studied, the target is operated for 10 min to assure steady-state conditions.

To prevent target surface oxidation by atmosphere expo-sure during transport to the XPS instrument [27], Al-caps are

deposited following the sputtering tests. First, the Al target, intended for cap layer deposition, is sputter-cleaned in Ar for 120 s at 2 kW behind closed shutters. We do not expect any chemical reactions on the target side during this time, as previous studies indicate that the post-deposition exposure of freshly-grown films to residual gases, including oxygen and water vapor, does not significantly affect the surface chemis-try [27]. After that, the Al target power is reduced to 0.4 kW and a capping overlayer is deposited on the Ti target with a thickness of 50 ± 5 Å, as estimated from the attenuation of core level signals [28]. Subsequently, the detachable Ti target fragment is transferred to the load-lock chamber of the UHV XPS system. The total air exposure time is shorter than 2 min.

XPS is performed in an Axis Ultra DLD instrument from Kratos Analytical (UK) employing a monochromatic Al Kα

source (hν = 1486.6 eV). The base pressure during spectra

acqui-sition is better than 1.5 × 10−7 Pa (1.1 × 10−9 Torr). All spectra

are collected at a normal emission angle from a 0.3 × 0.7 mm2 Figure 2. Ti 2p XPS spectra obtained from Al-capped Ti target operated with DCMS at pN2 = 92 mPa, as well as, with HIPIMS

at pN2 = 45 mPa. N2 partial pressures are optimized to yield

stoichiometric films at the substrate. Included for the reference is the Ti 2p spectrum from TiN thin film sputter-cleaned with 0.5 keV Ar+ ions incident at 70° from the surface normal. Satellite peaks on the high BE side of the primary signals are commonly observed, however, beyond the scope of this paper.

Figure 1. Target voltage and target current waveforms recorded during HIPIMS of Ti target in Ar/N2 mixture with N2 partial pressure of 45 mPa.

4 For more detailed treatment see, e.g. [20] and references therein. 5 It was established in a control experiment that the sputtering rate from the

Ti target is not lower than 54 nm min−1 × kW. Hence, we estimate that the

target cleaning procedure applied in this work results in removing the surface layer which is about 2 orders of magnitude thicker than the compound layer formed during reactive operation.

(4)

area centered in the middle of the target race track. The BE scale is calibrated against the Fermi level cut-off [29]6, using the

procedure described in detail elsewhere [23] in order to avoid BE referencing problems resulting from the fact that C 1s BE depends on the type of surface oxides formed during the venting procedure [27], and to remove ambiguities related to the use of C 1s as the BE [30]. For depth profile experiments, Ar+ ions with

an energy of 0.5 keV incident at a 70° angle from the surface normal are used to clean a 3 × 3 mm2 area in the middle of the

race track. The sputter rate is calibrated by etching through a 300 nm thick TiN film grown on Si(0 0 1) substrates.

Figure 1 shows voltage and current waveforms recorded during HIPIMS of Ti target, which are used below to derive the energy-averaged nitrogen implantation profile. The target voltage is −700 V at t = 0 µs and decreases rapidly with time, due to the size of the capacitor bank with respect to the target area, to reach −140 V with t = 70 µs. After plasma ignition at t = 5 µs, the target current increases to a maximum value 1280 A at 32 µs and decays to zero within the next ~45 µs.

Normalized Ti 2p spectra recorded from Al-capped Ti tar-gets after sputtering in (i) DCMS mode with pN2 = 92 mPa,

and (ii) HIPIMS mode with pN2 = 45 mPa are shown in

fig-ure 2. In both cases, pN2 is adjusted to grow stoichiometric

TiN films. In addition, the Ti 2p spectrum obtained from a polycrystalline TiN thin film surface, deposited in the HIPIMS mode with the exact same settings as specified above, and pre-viously sputter-etched with 0.5 keV Ar+ ions, is included for

reference. There is a very clear difference in the appearance of Ti 2p spectra depending on whether the target was oper-ated in DCMS or in HIPIMS. In the former case, the Ti 2p spectra possess two distinct contributions, with stronger 2p3/2

spin-split components at 454.0 and 455.0 eV, corresponding to Ti and TiN, respectively [31]. The relative signal variation revealed by the angle-dependent XPS (not shown) indicated a layer-over-layer structure. The thickness of the top nitride layer was estimated to be 29 Å [22], based on the relative peak

intensities of metal and nitride components in the fitted Ti 2p XPS spectra supported by TRIDYN simulations [32, 33].

In contrast to the DCMS case, the Ti 2p spectrum acquired after HIPIMS operation exhibits only one 2p3/2 peak, assigned

to TiN. The lack of a metallic contribution is direct evidence for a thicker surface TiN layer than during DCMS, clearly exceeding 54 Å, which is the estimated XPS probing depth7.

The overall spectral envelope is essentially identical to that from the sputter-etched TiN-film reference. The agreement is exceptionally good, given that the Ti 2p spectrum of TiN is prone to exhibit the destructive effects of an Ar+ ion etch

[21]. This can be rationalized by the fact that, in both cases, for the target sputtered in Ar/N2 atmosphere, as well as the

Ar+-etched TiN film, the surface is exposed to Ar+ ions with

similar energy. Good agreement also indicates that the poten-tial influence of a higher surface roughness of the Ti target on the quality of the XPS spectra is negligibly small.

To verify the estimates of compound layer thickness based on Al-capped samples, we performed sputter depth profiles on Ti targets after (a) DCMS and (b) HIPIMS operation. Figure 3

shows the Ti 2p3/2 spectra evolution as a function of

sputter-ing depth. The startsputter-ing point denoted as ‘0 Å’ corresponds to the sample state after removing the Al capping layer. In the case of the target operated under DCMS conditions, the Ti 2p3/2 spectrum changes rapidly with increasing depth.

The TiN component at 455.0 eV decreases, while the metal-lic peak at 454.0 eV increases in intensity, such that after first removing12 Å, both contributions are approximately equal. However, the steady state is not reached until the top 28–32 Å are removed and the spectrum contains one metallic peak at 454.0 eV with an asymmetric tail on the high BE side charac-teristic of a metallic Ti [34]. Hence, there is a very good agree-ment between the present result and the nitride thickness of 29

Figure 3. Ti 2p3/2 XPS depth profiles obtained from Ti target surface after (a) DCMS at pN2 = 92 mPa, and (b) HIPIMS at pN2 = 45 mPa.

N2 partial pressures are optimized to yield stoichiometric films at the substrate.

6 See for instance chapter 1 in [29].

7 The probing depth is defined as the surface layer that yields 95% of the

signal intensity, and is equal to 3 × λ, in which λ stands for the inelastic electron mean free path for Ti 2p electrons excited with Al Kα radiation and

(5)

4

Å, estimated above from the Ti 2p spectrum of the Al-capped Ti target.

The corresponding depth profile performed under the exact same conditions on the target previously operated in HIPIMS fully confirms that a much thicker compound layer is formed (see figure 3(b)). First, after removing 12 Å, which in the case of the DCMS target results in a Ti 2p3/2 spectrum with

equal Ti and TiN peak intensities, the core level signal from the HIPIMS target is still completely dominated by the TiN contrib ution, with only a small indication of a metallic peak at 454.0 eV. It is not until the 36 Å material is removed that both the Ti and TiN peaks become equally intense. Interestingly, the gradual change between two consecutive spectra is signifi-cantly smaller in the case of a HIPIMS-exposed target, indica-tive of a more diffuse nitride/metal interface than observed for a DCMS target, which likely results from the fact that the amplitude of the target voltage (hence, incident ion energy) varies in a wide range during the HIPIMS pulse (see figure 1). The steady-state condition, corresponding to the metallic Ti spectrum, is reached at the depth of 72–76 Å, which is ~2.5 times deeper than for the DCMS target. Thus, at the reactive gas partial pressures that ensure stoichiometric film growth at the substrate, compound layer formation is more efficient dur-ing HIPIMS as compared to DC processdur-ing.

To explain the observed differences between the surface composition of Ti targets operated in DCMS and HIPIMS, we performed dynamic Monte Carlo collisional computer simula-tions using the TRIDYN code [35, 36] of the stationary nitro-gen depth profiles. During DCMS with Ar/N2 gas mixtures

optimized to obtain stoichiometric films, the majority of the ion species incident at the growing film are Ar+, N+

2, and N+,

with the N+

2/N+ ratio increasing with increasing pN2 [37].

Our previously reported in situ ion mass spectrometry measurements of Ti-DCMS plasma with pN2 = 80 mPa

showed that N+

2 population is ~50 times higher than N+ [38],

thus, in the TRIDYN simulations of DCMS operation, the lat-ter species can be neglected. The first approximation to the N+

2 ion content in the plasma (relative to that of Ar+) can be

obtained from the N2/Ar partial pressure ratio, since under

DCMS conditions, both gases have similar electron impact ionization cross sections [39], while the ionization potentials are very similar [40]. The cathode sheaths are collision-less, thus, with the plasma potential not higher than a few volts, the incident ion energy can be approximated by the target volt-age [41], given that the voltage drop in the magnetic sheath is minor [42]. Therefore, in simulations, we use 350 eV Ar+,

while each N+

2 ion, first neutralized by electron pickup from

the target, dissociates upon impact and the incident energy splits equally between the two N atoms. The steady-state con-ditions require 2 × 105 pseudoparticles, representing the total

fluence of 5 × 1017 ions/cm2. The simulated profile indicates

close-to-stoichiometric TiNx with x = 0.94, which forms a

relatively sharp interface to metallic Ti at 32 ± 2 Å, in excel-lent agreement with XPS estimates from Al-capped samples (29 Å) and the results of sputter depth profile (28–32 Å).

The TRIDYN simulations in the case of HIPIMS are more complex since the energy of incident ions is not constant, due to the target voltage VT varying during the pulse (see figure 1).

To handle this effect, we performed TRIDYN for an appli-cable range of incident ion energies from 100 to 700 eV, in steps of 100 eV. As the ion flux to the target is proportional to the target current, the I–V plot can be considered a first-order approximation to the ion energy distribution function, which can in turn be used to derive the energy-averaged N implant-ation profile based on TRIDYN results obtained for specific ion energies.

In the first case, similar to DCMS, we assumed Ar+

(EAr+ = 350 eV) and N+2 ions that dissociate upon impact and

give rise to two N atoms, each with EN = 175 eV. As shown

in figure 4, the results indicated that the N implantation depths are too short to explain the XPS data. Essentially, no N is found deeper than ~45 Å, which is inconsistent with the XPS findings (see figure 3(b)).

Importantly, previous ion mass spectrometry measure-ments, conducted during Ti HIPIMS in Ar/N2 atmos phere in

the same deposition system as used for the present experiments, revealed that gas ion fluxes at the substrate are dominated by N+ rather than N+

2, with Ar+, N+2, and N+ contrib utions

amounting to 36, 17, and 47%, respectively [38]. This is also consistent with other reports on reactive HIPIMS of trans-ition metal targets [43, 44]. Therefore, in the following simu-lations we assumed that the population of reactive gas ions during HIPIMS is dominated by N+ implying E

Ar+=EN+.

Figure 4 shows the calculated N/Ti implantation profiles for a N+ fraction in the Ar+/N+ flux incident on the Ti target

varying between 20 and 100%. The lower limit corresponds to the N/Ar atom ratio for the N2 partial pressure of 45 mPa.

Interestingly, the minimum N+ fraction necessary to explain

the XPS results is 50%, which corresponds very well to the relative N+ population during HIPIMS discharge analyzed in [40]. The source of N+ can be (a) electron-impact induced

dissociation of N2, and (b) N sputtered from the poiso ned

tar-get surface. An additional effect, which may further increase the N+/Ar+ ratio in the ion flux incident on the Ti target is Ar

rarefaction [45]. With a better mass match to Ti atoms, result-ing in a more effective momentum transfer, Ar is diluted more than N2 and N.

Figure 4. The energy-averaged N implantation profiles based on TRIDYN results obtained for specific ion energies assuming: (open circles) EAr+= 350 eV together with EN= 175 eV, and (filled circles)

EAr+=EN+. In the latter case, results are shown as a function of N+

fraction in the ion flux incident on the Ti target.

(6)

To complement the above results, we also analyzed the HIPIMS target chemistry for pN2 lower than 45 mPa. The N/

Ti ratio, as determined by the XPS (not shown), hence reflect-ing the average N and Ti concentrations within the 54 Å thick surface layer [28], was at 0.98 and 0.88 for pN2 = 27 and

9 mPa, respectively. Thus, in the HIPIMS mode, the Ti tar-get surface region remained severely nitrided, while corre-sponding films were understoichiometric in nitrogen.

In summary, during the deposition of stoichiometric TiN thin films by HIPIMS, the formation of a 2.5× thicker nitride layer in the center of the race track was observed, as compared to conventional dc processing. The ~50 Å thick target surface region is nitrided during HIPIMS, even at pN2 values too low

to yield stoichiometric TiN films on the substrate, which is in clear contrast to conventional DCMS. The differences between the nitride layer thickness during reactive DCMS and HIPIMS can be understood based on correlative plasma chemical invest-igations and TRIDYN simulations. N+ ions that dominate

nitrogen ion flux during HIPIMS arrive at the target with sig-nificantly higher energy than molecular N+

2 present in DCMS

discharge, resulting in much longer implant ation depths. Based on the excellent agreement between TRIDYN simulations and XPS analyses, we conclude that the formation of the compound layer is entirely caused by the implantation of nitrogen ions. Acknowledgments

The authors want to thank Dr. Andre Anders for helpful dis-cussions. The authors gratefully acknowledge the financial support of the German Research Foundation (DFG) within SFB-TR 87, the Knut and Alice Wallenberg Foundation Scholar Grant KAW2016.0358, the VINN Excellence Center Functional Nanoscale Materials (FunMat-2) Grant 2016-05156, the Åforsk Foundation Grant 16-359, and Carl Tryggers Stiftelse contracts CTS 15:219 and CTS 14:431. ORCID iDs

G Greczynski https://orcid.org/0000-0002-4898-5115

References

[1] Knotek O, Böhmer M and Leyendecker T 1986 J. Vac. Sci. Technol. A 4 2695

[2] Knutsson A, Johansson M P, Karlsson L and Odén M 2010 J. Appl. Phys.108 044312

[3] Holloway K, Fryer P M, Cabral C, Harper J M E, Bailey P J and Kelleher K H 1992 J. Appl. Phys. 71 5433

[4] Chun J S, Desjardins P, Lavoie C, Shin C-S, Cabral C Jr, Petrov I and Greene J E 2001 J. Appl. Phys. 89 7841 [5] Mráz S and Schneider J M 2006 J. Appl. Phys. 100 023503 [6] Mráz S and Schneider J M 2006 Appl. Phys. Lett. 89 051502 [7] Depla D and De Gryse R 2004 Surf. Coat. Technol. 183 184

Depla D and De Gryse R 2004 Surf. Coat. Technol. 183 190 Depla D and De Gryse R 2004 Surf. Coat. Technol. 183 196 [8] Berg S, Blom H-O, Larsson T and Nender C 1987 J. Vac. Sci.

Technol. A 5 202

[9] Kouznetsov V, Macak K, Schneider J M, Helmersson U and Petrov I 1999 Surf. Coat. Technol. 122 290

[10] Alami J, Sarakinos K, Mark G and Wuttig M 2006 Appl. Phys. Lett.89 154104

[11] Desecures M, de Poucques L, Easwarakhanthan T and Bougdira J 2014 Appl. Phys. Lett. 105 181120

[12] Loquai S, Zabeida O, Klemberg-Sapieha J E and Martinu L 2016 Appl. Phys. Lett. 109 114101

[13] Hecimovic A, Britun N, Konstantinidis S and Snyders R 2017 Appl. Phys. Lett.110 014103

[14] Raman P, Weberski J, Cheng M, Shchelkanov I and Ruzic D N 2016 J. Appl. Phys. 120 163301

[15] Audronis M, Abrasonis G, Munnik F, Heller R, Chapon P and Bellido-Gonzalez V 2012 J. Phys. D: Appl. Phys. 45 375203

[16] Güttler D, Abendroth B, Grötzschel R, Möller W and Depla D 2004 Appl. Phys. Lett. 85 6134

[17] Abe Y, Takisawa T, Kawamura M and Sasaki K 2007 Japan. J. Appl. Phys.46 6778

[18] Combadiere L and Machet J 1996 Surf. Coat. Technol. 82 145 [19] Depla D, Haemers J and De Gryse R 2002 Plasma Sources

Sci. Technol.11 91

[20] Wagner T, Wang J Y and Hofmann S 2003 Sputter depth profiling in AES and XPS Surface Analysis by Auger and X-Ray Photoelectron Spectroscopy ed D Briggs and J T Grant (Manchester: IM Publications) p 618

[21] Greczynski G, Petrov I, Greene J E and Hultman L 2015 J. Vac. Sci. Technol. A 33 05E101

[22] Greczynski G, Mráz S, Hultman L and Schneider J M 2017 Appl. Phys. Lett.111 021604

[23] Greczynski G, Primetzhofer D, Lu J and Hultman L 2017 Appl. Surf. Sci.396 347

[24] Sproul W D, Christie D J and Carter D C 2005 Thin Solid Films491 1

[25] www.cemecon.de/en/coating-plants/cc-800-hipims (Accessed: 9th January 2018)

[26] Schneider J M, Hjörvarsson B, Wang X and Hultman L 1999 Appl. Phys. Lett.75 3476

[27] Greczynski G, Mráz S, Hultman L and Schneider J M 2016 Appl. Phys. Lett.108 041603

[28] Tanuma S, Powell C J and Penn D R 2010 Surf. Interface Anal.43 689

[29] Hufner S 2003 Photoelectron Spectroscopy: Principles and Applications (Berlin: Springer)

[30] Greczynski G and Hultman L 2017 Chem. Phys. Chem. 18 1507 [31] Greczynski G and Hultman L 2016 Appl. Surf. Sci. 387 294 [32] Carlson T A 1982 Surf. Interface Anal. 4 125

[33] Strohmeier B R 1990 Surf. Interface Anal. 15 51

[34] Moulder J F, Stickle W F, Sobol P E and Bomben K D 1992 Handbook of X-Ray Photoelectron Spectroscopy (Eden Prairie, MN: Perkin-Elmer Corporation)

[35] Möller W and Eckstein W 1984 Nucl. Instrum. Methods Phys. Res. B 2 814

[36] Möller W, Eckstein W and Biersack J P 1988 Comput. Phys. Commun.51 355

[37] Petrov I, Myers A, Greene J E and Abelson J R 1994 J. Vac. Sci. Technol. A 12 2846

[38] Greczynski G, Lu J, Johansson M, Jensen J, Petrov I, Greene J E and Hultman L 2012 Surf. Coat. Technol. 206 4202

[39] Massey H S W and Burhop E H S 1956 Electronic and Ionic Impact Phenomena (Oxford: Oxford University Press) [40] Lide D R (ed) 2003 Atomic, molecular, and optical physics;

ionization potentials of atoms and atomic ions CRC Handbook of Chemistry and Physics 84th edn (Boca Raton, FL: CRC Press)

[41] Anders A 2002 Appl. Phys. Lett. 80 1100

[42] Panjan M and Anders A 2017 J. Appl. Phys. 121 063302 [43] Breilmann W, Maszl C, Hecimovic A and von Keudell A 2017

J. Phys. D: Appl. Phys.50 135203

[44] Greczynski G and Hultman L 2010 Vacuum 84 1159 [45] Horwat D and Anders A 2010 J. Appl. Phys. 108 123306

References

Related documents

- gruppdata tillämpas på individ inom eller utom gruppen (om t.ex. 60% av en grupps medlemmar har beteende X, så kan detta inte hävdas om en enskild medlem som kan tillhöra

Åsa poängterar att om man inom handeln sparar pengar på att bli duktigare och mer kompetenta i sitt ar- bete med att spara energi så sparar man ju faktiskt pengar på något som

”Vilka motiverande faktorer finns för klienten?” ”Hur rangordnas de av klienten?” ”Vilka förstärkningar är mest effektiva för klienten?”, ”Vilka stimuli är aversiva

Samma förhörsledare, som ensidigt sökt bekräfta att sexuella handlingar ägt rum mellan Rasmus och barnet, används. Om syftet med det andra förhöret är att kontrollera resultaten

gemensamt. Detta är ett genomfört samtal som talar emot att P skulle gjort sig skyldig till sexuellt övergrepp på M en knapp månad tidigare. Det framstår som psykologiskt

The results of this study can neither confirm nor contradict any correlation between a defect gene coding for UCMA and patients with stage 5 CKD since the population of our

ALP från njure och humant serum band inte in alls till kollagentäckta plattor men band in ganska starkt till både bovint kollagen typ I och II samt humant kollagen typ I vid