atomic layer deposition of ruthenium films : properties and surface

171
Atomic layer deposition of ruthenium films : properties and surface reactions Leick-Marius, N. DOI: 10.6100/IR782932 Published: 01/01/2014 Document Version Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication: • A submitted manuscript is the author's version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website. • The final author version and the galley proof are versions of the publication after peer review. • The final published version features the final layout of the paper including the volume, issue and page numbers. Link to publication Citation for published version (APA): Leick-Marius, N. (2014). Atomic layer deposition of ruthenium films : properties and surface reactions Eindhoven: Technische Universiteit Eindhoven DOI: 10.6100/IR782932 General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal ? Take down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Download date: 17. Mar. 2018

Upload: haanh

Post on 10-Jan-2017

218 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: Atomic layer deposition of ruthenium films : properties and surface

Atomic layer deposition of ruthenium films : propertiesand surface reactionsLeick-Marius, N.

DOI:10.6100/IR782932

Published: 01/01/2014

Document VersionPublisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:

• A submitted manuscript is the author's version of the article upon submission and before peer-review. There can be important differencesbetween the submitted version and the official published version of record. People interested in the research are advised to contact theauthor for the final version of the publication, or visit the DOI to the publisher's website.• The final author version and the galley proof are versions of the publication after peer review.• The final published version features the final layout of the paper including the volume, issue and page numbers.

Link to publication

Citation for published version (APA):Leick-Marius, N. (2014). Atomic layer deposition of ruthenium films : properties and surface reactionsEindhoven: Technische Universiteit Eindhoven DOI: 10.6100/IR782932

General rightsCopyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright ownersand it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal ?

Take down policyIf you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediatelyand investigate your claim.

Download date: 17. Mar. 2018

Page 2: Atomic layer deposition of ruthenium films : properties and surface

Atomic Layer Deposition of Ruthenium Films

Properties and Surface Reactions

PROEFSCHRIFT

ter verkrijging van de graad van doctor aan de Technische Universiteit Eindhoven, op gezag van de rector magnificus prof.dr.ir. C.J. van Duijn, voor een

commissie aangewezen door het College voor Promoties, in het openbaar te verdedigen

op woensdag 3 december 2014 om 16.00 uur

door

Noémie Leick-Marius

geboren te Thionville, Frankrijk

Page 3: Atomic layer deposition of ruthenium films : properties and surface

Dit proefschrift is goedgekeurd door de promotoren: prof.dr.ir. W.M.M. Kessels en prof.dr. F. Roozeboom

Page 4: Atomic layer deposition of ruthenium films : properties and surface

This work was financially supported by the Dutch Technology Foundation STW (TFN 10018).

Cover design: Proefschriftmaken.nl || Uitgeverij BOXPress Printed & Lay Out by: Proefschriftmaken.nl || Uitgeverij BOXPress Published by: Uitgeverij BOXPress, ’s-Hertogenbosch A catalogue record is available from the Eindhoven University of Technology Library. ISBN: 978-94-6295-003-0

Page 5: Atomic layer deposition of ruthenium films : properties and surface
Page 6: Atomic layer deposition of ruthenium films : properties and surface

CONTENTS Chapter 1 Introduction and framework of the research 1

Chapter 2 Ruthenium and its surface reactions: an overview from surface science perspective

15

Chapter 3 CpRu(CO)2Et as the metal precursor during Ru ALD 31

Chapter 4 Atomic layer deposition of ruthenium from CpRu(CO)2Et using O2 gas and O2 plasma

Published as: Leick, N.; Verkuijlen, R. O. F.; Lamagna, L.; Langereis, E.; Rushworth, S.; Roozeboom, F.; van de Sanden, M. C. M.; Kessels, W. M. M., J. Vac. Sci. Technol. A 2011, 29, 021016.

45

Chapter 5 In situ spectroscopic ellipsometry during atomic layer deposition of Pt, Ru and Pd

To be published as authors: Leick, N., Weber, J. W.; Mackus, A. J. M.; Weber, M. J.; van de Sanden, M. C. M.; Kessels, W. M. M.

67

Chapter 6 Dehydrogenation reactions during atomic layer deposition of ruthenium using O2

Published as: Leick, N.; Agarwal, S. ; Mackus, A. J. M.; Kessels, W. M. M., Chem. Mater. 2012, 24, 3696.

95

Chapter 7 Catalytic combustion reactions during atomic layer deposition of Ru studied by 18O2 isotope labeling

Published as: Leick, N.; Agarwal, S. ; Mackus, A. J. M.; Potts, S. E.; Kessels, W. M. M., J. Phys. Chem. C 2013, 117, 21320.

111

Chapter 8 Conclusions and recommendations 143

Summary 151

List of publications related to this work 154

Acknowledgements 157

Curriculum Vitæ 163

Page 7: Atomic layer deposition of ruthenium films : properties and surface
Page 8: Atomic layer deposition of ruthenium films : properties and surface

1

Chapter 1 “Anybody who has been seriously engaged in scientific work of any kind realizes that over the entrance to the gates of the temple of science are written the words: Ye must

have faith. It is a quality which the scientist cannot dispense with. ” ‒ Max Planck Introduction and framework of the research

Over the last decade, the downscaling of electronic devices in semiconductor development has been the driving force to explore the deposition of nanometer-sized thin films and other structures. The microelectronics industrial roadmap has been dominated by Si-based materials, such as SiO2 and SiNx dielectrics, and conductors such as doped polycrystalline silicon and TiN, which are both highly Si-compatible.1,2 However, on the nanometer scale, these materials start to reach their physical limitations in terms of next-generation device requirements, especially leakage current and dielectric constant requirements.2‒4 Therefore, great efforts have been undertaken in the investigation of new materials. Here, atomic layer deposition (ALD) has emerged as a versatile deposition technique able to cope with the demand for atomically thin films and nanostructures.5,6 In addition to applications in semiconductor development, the need for nano-structured materials in general has significantly increased in a wide range of applications,7 such as heterogeneous catalysis,8‒10 energy storage and conversion11,12 or sensing13‒15.

Pt-group metals,a such as Pt, Ir, Pd and Ru, are catalytic in nature, which

means that they are able to break the chemical bonds of certain organic

molecules, as illustrated in Figure 1. They are predominantly used in

heterogeneous catalysis, but they also have the potential to be applied as

electrodes in microelectronic devices, such as in metal-oxide-semiconductor

structures or in metal-insulator-metal capacitors.17‒21 ALD of Pt-group metals

a The term Pt-group metals will be used in this manuscript rather than noble metal as this latter term is, strictly speaking, reserved to metals with filled d-orbitals (i.e. Au, Ag and Cu).

Page 9: Atomic layer deposition of ruthenium films : properties and surface

1. INTRODUCTION AND FRAMEWORK OF THE RESEARCH

2

not only yields high-quality films, but also benefits from a thickness control at

the atomic level, a superior thickness uniformity and an excellent conformality

in 3D topologies. However, the ALD process of Pt and the other Pt-group

metals still calls for further optimization and the film properties for further

tailoring. To this end, ruthenium ALD was chosen in this thesis to be studied as

a representative Pt-group metal to gain insight into the surface chemistry

leading to film growth. Before elaborating on Ru-ALD specifically, we will first

review the well-established ALD process of Al2O3.

Figure 1: Schematic overview of the ability of Ru, Pd, Ir and Pt to break the different bonds in volatile organic molecules and other species (based on Ref. 16).

Atomic layer deposition: the model system of Al2O3

ALD is a vapor-based deposition technique which is composed of two alternating and chemically self-limiting half-reactions, called half-cycles. In each half-cycle, gaseous species are introduced into the ALD reactor. In the specific case of Al2O3 ALD, trimethylaluminum (TMA or AlMe3, with Me = CH3) is generally used as the gaseous precursor in the first half-cycle and H2O vapor is typically the reactant used in the second half-cycle. The two half-cycles are separated by a step in which the excess gaseous species are purged (or pumped) from the reactor. This process is schematically summarized in Figure 2. Like most metal oxides, the Al2O3 surface is initially saturated with hydroxyl groups (OH). During the first half-cycle, AlMe3 reacts with the surface hydroxyl groups, and at the end of the AlMe3 pulse the surface is terminated with Me-ligands that are not reactive with the residual AlMe3 molecules in the gas-phase. During the second half-cyle, the H2O removes the remaining surface Me-ligands. At the end of the H2O pulse, the surface is fully restored to Al2O3

Ru PtPdRuC C

C OC HO OH H

C HO OH H

C CC O

Pd Pt

C HO OH H

C CC O

IrIrC C

C OC HO OH H

Page 10: Atomic layer deposition of ruthenium films : properties and surface

1. INTRODUCTION AND FRAMEWORK OF THE RESEARCH

3

saturated with OH-groups. Subsequently the ALD cycles can be repeated until the desired film thickness is reached. In such ALD process, the surface chemistry during both half-cycles is driven by Lewis acid-base reactions between Me-ligands and OH groups, releasing CH4 molecules into the gas phase. The reaction mechanism of the two Al2O3 ALD half-cycles can each be summarized with a reaction equation,6 where * denotes the surface species:

First half-cycle: AlOH* + Al(CH3)3 AlOAl(CH3)2* + CH4 (Eq. 1) Second half-cycle: AlCH3* + H2O AlOH* + CH4 (Eq. 2)

Figure 2: Schematic overview of the two half-cycles of Al2O3 ALD. (a) The first half-cycle consists of the self-limiting adsorption of the gaseous TMA-molecules on the available surface sites. (b)A purge step follows to remove un-reacted TMA molecules and gaseous CH4. (c) In the second half-cycle, the remaining C-containing surface groups are removed from the surface in a self-limiting manner using H2O as reactive gas. (d) At last, the gaseous by-products and excess gas are removed during the purge step.

The surface reactions in each ALD half-cycle are terminated “naturally” due to the finite number of available surface sites for AlMe3 and H2O to react with. This aspect is called self-limiting and is characteristic for ALD in general. As a result of the self-limiting behavior of ALD, up to one monolayer of material (~2 Å) can be deposited in one cycle. Typically, however, only 0.5-1 Å are deposited and consequently, the ALD film thickness can be controlled with Ångstrom accuracy. Another implication of ALD being ruled by the number of surface-sites is the excellent uniformity of the films on flat surfaces and outstanding conformality of the films in demanding 3D-topologies. On the other hand, the main limitation of ALD lies in the low growth rate compared to other deposition techniques. While chemical vapour deposition (CVD) or

CH3

CH3CH3

CH3HH

H

(a) (c)(b) (d)

TMA dose Purge PurgeH2O dose

ALD cyclefirst half-cycle second half-cycle

OHH

O O O O O O O OO O O OAl Al AlAl Al Al Al Al Al Al Al Al

Al Al Al

CH3

H

Al Al Al

CH3

CH3

CH3

CH3

Al Al Al

CH3

CH3

CH3

CH3

O O O O

H H

O O O O

H H

CH3

HAl

CH3CH3

CH3

H

Page 11: Atomic layer deposition of ruthenium films : properties and surface

1. INTRODUCTION AND FRAMEWORK OF THE RESEARCH

4

physical vapour deposition (PVD) can yield growth rates of micrometers-per-second, ALD typically achieves the deposition of ~0.1 Å /second in the case of Al2O3.22

To reach self-limiting surface reactions of ALD, the precursor and reactant dose have to be tuned each to ensure that the surfaces are saturated with surface groups. In the case of Al2O3, the surface groups are the Me-ligands in the first half-cycle and OH-groups in the second half-cycle. The surface saturation also depends on the substrate temperature and the process is usually called to be in the “ALD regime” when the growth per cycle stays relatively constant over a certain temperature range of typically one or several hundreds of degrees.

The activation barriers to be surmounted for the surface reactions of the Al2O3 ALD process based on TMA and H2O are provided by thermal energy from the substrate. Consequently, the process is known as the thermal ALD process of Al2O3. An alternative source of energy to deposit atomic layers can be a plasma, which consists of an ionized gas with radicals as the primary reactive species to drive surface reactions during the second ALD half-cycle. In this case, the ALD process is referred to as plasma-assisted ALD. Under specific conditions, plasma-assisted ALD can be more suitable over thermal ALD, and thereby deliver improved material properties, deposition at a reduced substrate temperature, good control of stoichiometry and film composition, increased growth rates, and it can provide an increased choice of precursors and materials, and more processing versatility in general.23 In the recent years, plasma-assisted ALD has gained in popularity for the deposition of a wide variety of materials, such as Al2O3, SiO2, TiO2, TiN, TaNx, Pt, and Ru.24‒31

Atomic layer deposition of Pt-group metals

The ability to synthesize structures or films of Pt-group metals on the nanometer scale has opened up a substantial number of applications for these metals, mostly in microelectronics and catalysis. The most common techniques to deposit these metal films for microelectronics applications have been CVD or PVD, while the synthesis of particles has mostly relied on wet chemistry in the field of catalysis. Recently, ALD has also shown potential in depositing nanometer thin Pt, Pd, Ir and Ru films,30‒33 as well as nanoparticles thereof.34‒37

Page 12: Atomic layer deposition of ruthenium films : properties and surface

1. INTRODUCTION AND FRAMEWORK OF THE RESEARCH

5

The surface reactions leading to the ALD growth of these metals differ considerably from those in ALD of Al2O3. Since this thesis focuses on Ru, the ALD process of Ru will be shortly described and its main differences with Al2O3 ALD will be evoked.

In comparison to precursors for ALD of Al2O3, there exists a very wide

selection of Ru-precursors for ALD (see Chapter 3).31,38‒43 In general, the choice

of reactants used in ALD is partly determined by the precursor used. Specifically

for Ru ALD, for example, a reactive H-containing reactant can be a preferred

choice when β-diketonate precursors are used, such as Ru(thd)3 (thd = 2,2,6,6,-

tetramethyl-3,5-heptanedione), since it can form stable reaction products with

the ligands, such as Hthd.6,38 These precursors, however, deliver very low

growth rates and it was shown that the resulting film properties can contain

high amounts of carbon residues.44 For this and other reasons,

cyclopentadienyl-based precursors have increasingly been studied for Pt-group

metal ALD. In the case of Pt and Ru ALD, Aaltonen et al.45 were the first to

propose the use of O2 gas as a reactant, thus the surface reactions would

mainly be driven by combustion reactions. The authors were inspired to use O2

on the basis of CVD processes of Ru that showed the successful deposition of

metallic Ru with O2 gas.46

Aaltonen et al. studied the combustion chemistry of Ru ALD from ruthenocene (RuCp2, with Cp = cyclopentadienyl,-C5H5) and O2 gas using quadrupole mass spectrometry (QMS).45 During both half-cycles, these authors measured CO2 and H2O as reaction products. This observation manifests the presence of reactive O-atoms chemisorbed at the Ru-surface reacting with the organic precursor ligands during the first half-cycle, via full combustion reactions. The following reaction equation gives the overall reaction for the complete oxidation of both Cp-ligands in one ALD cycle:

RuCp2 + 25 2 O2 Ru + 10 CO2 + 5 H2O (Eq. 3)

By making use of the results provided by dedicated surface science studies on the interaction of Ru with O2 gas, it was possible for Aaltonen et al. to speculate on a reaction mechanism during ALD illustrated in Figure 3. From these surface science studies, it is known that O2 chemisorbs dissociatively onto

Page 13: Atomic layer deposition of ruthenium films : properties and surface

1. INTRODUCTION AND FRAMEWORK OF THE RESEARCH

6

the Ru surface and that the saturation coverage of the reactive O atoms can reach values from less than a monolayer (ML) to many monolayers, depending on the substrate temperature and O2 exposure (= pressure × dosing time). From the growth per cycle that they measured for their Ru ALD process, the authors were able to deduce an O-coverage of ~3 ML on their Ru surface at the start of the ALD cycle.

Figure 3: Schematic of the Ru ALD reaction mechanism as proposed by Aaltonen et al.45 During the O2 pulse (second half-cycle), the O2 molecules chemisorb dissociatively onto the catalytic Ru surface. The reactive O atoms combust the hydrocarbon species remaining on the surface after precursor dosing, thereby liberating CO2 and H2O. At the end of the O2 pulse, the Ru surface is covered with O atoms. Upon the chemisorption of RuCp2 on the O-covered Ru surface (first half-cycle), the hydrocarbon ligands react with the surface O via combustion reactions, releasing CO2 and H2O into the gas-phase.

Furthermore, the successful use of O2 gas as a co-reactant in Ru ALD could be explained as follows.47 The oxidation state of the metal atom in the precursor describes the hypothetical charge of an atom in a molecule assuming that all the bonds are 100% ionic.48 Taking RuCp2 as an example, the Ru ion has an oxidation state of +2. It means that two electrons are required for the Ru atom to be in its pure metallic state (i.e. oxidation state is 0). Knowing that oxidation involves the loss of electrons whereas reduction involves the gain of electrons, it might, at first, seem counter-intuitive to use an oxidizing agent as the co-reactant. However, during the oxidation of incoming precursor ligands, electrons are released to reduce the metal ion Ru2+ to Ru0, and the surface O atoms desorb from the Ru surface in the form of CO2 and H2O.

From this mechanism, it becomes clear that the Ru ALD process relies on the dissociative chemisorption of O2 gas on the surface such that these reactive O atoms drive the surface reactions in the Ru ALD. However, at the start of the

metalorganicprecursor

Ru

chemisorbed O

O2

CO2

H2O

CO2

H2O

Ru

CxHy CxHy

Ru

CxHy CxHy

Ru

dissociative chemisorption of O2

self-limiting precursor adsorption self-limiting surface reactions

thermal ALD

O2 O chemisorbed O

combustionreactions

combustionreactions

Page 14: Atomic layer deposition of ruthenium films : properties and surface

1. INTRODUCTION AND FRAMEWORK OF THE RESEARCH

7

process, generally the substrate material is not able to catalyze the dissociation of O2. Therefore, the Ru-growth is initially inhibited and a nucleation delay is observed in the growth curve as opposed to Al2O3 ALD, as illustrated in Figure 4a. There are multiple ways of reducing this nucleation delay and in the case of Pt ALD, the use of an O2 plasma has shown to initiate Pt growth faster than the thermal process (Fig. 4b) as it provides the substrate with O-radicals directly from the gas-phase.30

Figure 4: (a) Thickness increase as measured by spectroscopic ellipsometry during ALD of Al2O3 and of Ru, (b) and in the case of Pt ALD. In (a) it is shown that there is a nucleation delay during Ru ALD from CpRu(CO)2(CH3CH2) and O2-gas compared to Al2O3 ALD which starts immediately. In (b) it is shown that Pt ALD growth from MeCpPtMe3 can be initiated using O2 plasma whereas the use of O2 gas did not lead to any substantial growth during the first 150 cycles.

Objectives and outline of the thesis

The catalytic activity of Ru has revealed itself as the most important factor to consider when the reaction mechanism of Ru ALD is investigated. In addition to characterizing a new Ru ALD process and the resulting film properties, the goal of this thesis work consisted of providing sufficient understanding of the Ru ALD reaction mechanism to be able to improve the deposition processes and to tailor the properties of the films. To this end, spectroscopic ellipsometry (SE) and quadrupole mass spectrometry (QMS) were used as in situ diagnostic tools. The data interpretation relied heavily on the wealth of information on Ru-surfaces provided by the extensive literature in the fields of surface science and catalysis. Aaltonen et al. had used pieces of

nucleation delay

(a) (b)

Page 15: Atomic layer deposition of ruthenium films : properties and surface

1. INTRODUCTION AND FRAMEWORK OF THE RESEARCH

8

information in their work on Ru ALD,45 but in this thesis work it is attempted to consider a more complete set of results from surface science to elucidate aspects of the surface chemistry taking place during Ru ALD. This approach is schematically illustrated in Figure 5.

Figure 5: Schematic of the approach adopted in this thesis to improve Ru ALD. Starting from a working ALD process, QMS and SE were utilized as in situ diagnostics to sample the gaseous species released and to monitor the film growth during each ALD half-cycle, respectively. To interpret the in situ studies it was mostly relied on the knowledge provided by surface science literature. From this combined approach, it was possible to unravel the most important reaction mechanisms taking place during Ru ALD. These can then be used to improve the ALD processes of Ru.

In Chapter 2, Ruthenium and its surface reactions: an overview from surface

science perspective, gives an overview of important results from surface science and catalysis. This chapter provides a theoretical background for the following chapters, in which these results are applied to the ALD process of Ru from (C5H5)Ru(CO)2(CH2CH3) and O2-gas.

Chapter 3, CpRu(CO)2Et as the metal precursor during Ru ALD, aims at giving a detailed chemical description of the Ru precursor, (C5H5)-Ru(CO)2(CH2CH3). In the second part of this chapter, the thermal stability of this compound is discussed and the precursor’s potential in processing is assessed.

In Chapter 4, Atomic layer deposition of ruthenium from CpRu(CO)2Et using O2 gas and O2 plasma, it is reported on a novel Ru ALD process that was developed using (C5H5)Ru(CO)2(CH2CH3) and O2 gas or O2 plasma. The material properties obtained with the two processes were investigated and compared.

Reaction mechanisms of Ru ALD

θO

Ru

CO

CO

C5H5

CH3CH2

Ru

CO2 COO2

CO2 H2OH2CO

Ru

C-rich speciesCxHy CxHy-1 CxHy-2

OO

Ru

C-rich species

Ru

Ru pulse

O2 pulse

100

101

102

103

104

105

106

107

0.0

0.5

1.0

1.5

2.0

O c

over

age

(ML)

O2 exposure (L)

Adsorbed OSubsurface O

1 ML p(1x1) O

Results from surface science

O CH

HH

OH

O CHH

OH

O CH

O H

H

HH

O CO H

HH

H

OHH

OH

OH

OHH

HH

COO

CO

O

CO2

CO 2 H2O

H2

CH4 + 2 O2Results from catalysis

SE

e-

QMS

H 2

(C5H5

)Ru(CO)2(CH2

CH3)

H2 O

CH2 CH

3CO

©JPvD

Ru

©JPvD

O2

CpRu(CO)2Et

In situ diagnostics

Development of ALD process

θO

Page 16: Atomic layer deposition of ruthenium films : properties and surface

1. INTRODUCTION AND FRAMEWORK OF THE RESEARCH

9

The thermal and plasma-assisted processes delivered high growth per cycle values (~1 Å/cycle) which were found to be a clear benefit of this ALD process compared to other existing processes. The plasma-assisted process was able to reduce the nucleation delay, however, at the expense of a high surface roughness of the films. Therefore, the thermal ALD process was chosen as the main research topic in this thesis.

Chapter 5, In situ spectroscopic ellipsometry during atomic layer deposition of Pt, Ru and Pd, treats of the film growth of these metals studied by SE. To take into account the change in the dielectric function of metal films thinner than 10 nm, the data were parametrized using polynomial base functions delivering accurate thickness values and dielectric functions of the films independently of each other. Our model also allowed the observation of the initial nucleation delay characteristic for ALD of metals.

In Chapter 6, Dehydrogenation reactions during atomic layer deposition of ruthenium using O2, we used QMS to probe the gas-phase products originating from the surface reactions during Ru thermal ALD. Among the reaction products measured during the Ru pulse H2, CO2 and H2O were observed. The combustion products CO2 and H2O manifest the presence of O on the metal surface at the start of the ALD cycle. This observation is in agreement with the indication reported previously by Aaltonen et al. that the process relied on the dissociative chemisorption of O2 on the catalytic surface during the oxygen exposure.45 In addition to breaking O-O bonds, we substantiated that the metal’s catalytic activity was also responsible for the scission of C-H bonds of the hydrocarbon ligands. Subsequently, the C- and H atoms are oxidized to CO2 and H2O or the H atoms desorb from the surface as H2.

Chapter 7, Catalytic combustion reactions during atomic layer deposition of Ru studied by 18O2 isotope labeling, comprises a QMS study in which an oxygen isotope 18O2 was used and the reaction products during Ru ALD were quantified. These results imply that the overall reaction mechanism of Ru ALD can be viewed as catalytic combustion reactions, occurring in each half-reaction. These reactions consist of the dissociative chemisorption of specific species on a catalytic surface (organic ligands during the Ru pulse, and oxygen in the O2 pulse) accompanied with the combustion of C- and H-rich species by the oxygen atoms at the surface.

Page 17: Atomic layer deposition of ruthenium films : properties and surface

1. INTRODUCTION AND FRAMEWORK OF THE RESEARCH

10

In the last Chapter, Conclusions and recommendations, the most important conclusions of this thesis are summarized and recommendations potential future studies are suggested.

The research subjects of this dissertation have taken place at the Plasma and Materials Processing (PMP) group at the Eindhoven University of Technology (TU/e) in the Netherlands. The entire sets of depositions and characterization were performed in the Department of Applied Physics of the TU/e, and predominantly in the “NanoLab@Tu/e” cleanroom facilities. The PMP group has a strong background in plasma chemistry and physics, but also in the synthesis of materials using plasma-assisted chemical vapor deposition and ALD; both thermal and plasma-assisted. The group also has a rich variety of diagnostic tools to be used during the deposition process as well as for studying material properties after deposition.

The research listed above was part of the project entitled “Atomic layer deposition as an enabling thin film technology” which was part of the “Thin Film Nanomanufacturing” program financially supported by the Dutch Technology Foundation STW (Stichting Technische Wetenschappen). Some of the aims of the project were to diversify the materials deposited by ALD for industrial applications and to understand the underlying reaction mechanisms using in situ diagnostic tools. An important aspect of the STW project is the interaction and possible cooperation between research institutes and industrial research groups in order to assess the feasibility and potential areas of interest for ALD processes.

The use of thin low-conductivity metal films together with high-κ dielectrics has also been a research interest of the MaxCaps project within the European Medea+ framework. This project focused on the fabrication of feasible metal-insulator-metal capacitors; one of which being a Ru-SrTiO3-Ru capacitor, preferably grown by ALD. Therefore, some of the work presented in this dissertation has been part of the results of the MaxCaps project as well.

Page 18: Atomic layer deposition of ruthenium films : properties and surface

1. INTRODUCTION AND FRAMEWORK OF THE RESEARCH

11

References

(1) Moore, G. E. In Electron Devices Meeting, 1975 International; 1975; pp. 11‒13.

(2) Clemens, J. T. Bell Labs Technical Journal 1997, 2, 76‒102.

(3) ITRS www.itrs.net - Process integration, devices, and structures 2010.

(4) Thompson, S. E.; Parthasarathy, S. Mater. Today 2006, 9, 20‒25.

(5) Sneh, O.; Clark-Phelps, R. B.; Londergan, A. R.; Winkler, J.; Seidel, T. E. Thin Solid Films 2002, 402, 248‒261.

(6) George, S. M. Chem. Rev. 2010, 110, 111‒131.

(7) Kim, H.; Lee, H. B. R.; Maeng, W. J. Thin Solid Films 2009, 517, 2563‒2580.

(8) Lu, J.; Fu, B.; Kung, M. C.; Xiao, G.; Elam, J. W.; Kung, H. H.; Stair, P. C. Science 2012, 335 , 1205‒1208.

(9) Chao, C.-C.; Motoyama, M.; Prinz, F. B. Adv. Eng. Mater. 2012, 2, 651‒654.

(10) Cassir, M.; Ringuede, A.; Niinisto, L. J. Mater. Chem. 2010, 20, 8987‒8993.

(11) Dingemans, G.; Kessels, W. M. M. J. Vac. Sci. Technol. A 2012, 30, 40802‒40827.

(12) Law, M.; Greene, L. E.; Radenovic, A.; Kuykendall, T.; Liphardt, J.; Yang, P. J. Phys. Chem. B 2006, 110, 22652‒22663.

(13) Marichy, C.; Donato, N.; Willinger, M.-G.; Latino, M.; Karpinsky, D.; Yu, S.-H.; Neri, G.; Pinna, N. Adv. Funct. Mater. 2011, 21, 658‒666.

(14) Niskanen, A. J.; Varpula, A.; Utriainen, M.; Natarajan, G.; Cameron, D. C.; Novikov, S.; Airaksinen, V.-M.; Sinkkonen, J.; Franssila, S. Sensor. Actuat. B - Chem. 2010, 148, 227‒232.

(15) Kim, W.-S.; Lee, B.-S.; Kim, D.-H.; Kim, H.-C.; Yu, W.-R.; Hong, S.-H. Nanotechnology 2010, 21, 245605.

(16) Freyschlag, C. G.; Madix, R. J. Mater. Today 2011, 14, 134‒142.

(17) Rathee, D.; Kumar, M.; Arya, S. K. Solid-State Electron. 2012, 76, 71‒76.

(18) Chen, X.; Wu, G.; Hu, W.; Zhou, H.; Bao, D. Appl. Phys. A Mater. Sci. Process. 2012, 108, 503‒508.

Page 19: Atomic layer deposition of ruthenium films : properties and surface

1. INTRODUCTION AND FRAMEWORK OF THE RESEARCH

12

(19) Pawlak, M. A.; Swerts, J.; Popovici, M.; Kaczer, B.; Kim, M.-S.; Wang, W.-C.; Tomida, K.; Govoreanu, B.; Delmotte, J.; Afanas’ev, V. V; Schaekers, M.; Vandervorst, W.; Kittl, J. A. Appl. Phys. Lett. 2012, 101, 42901‒42903.

(20) Kim, S.; Koo, J.; Shin, S.; Park, Y. Appl. Phys. Lett. 2005, 87, 212910‒212913.

(21) Nazarpour, S.; Langenberg, E.; Jambois, O.; Ferrater, C.; García-Cuenca, M. V; Polo, M. C.; Varela, M. App. Surf. Sci. 2009, 255, 3618‒3622.

(22) Dingemans, G.; Seguin, R.; Engelhart, P.; Sanden, M. C. M. van de; Kessels, W. M. M. Phys. Stat. Sol. (RRL) 2010, 4, 10‒12.

(23) Profijt, H. B.; Potts, S. E.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Vac. Sci. Technol. A 2011, 29, 50801‒50826.

(24) Kawamura, Y.; Tani, M.; Hattori, N.; Miyatake, N.; Horita, M.; Ishikawa, Y.; Uraoka, Y. Jpn. J. Appl. Phys. 2012, 51, 02BF04.

(25) Lei, W.; Li, X.; Chen, Q.; Wang, Z. Plasma Sci. Technol. 2012, 14, 129‒133.

(26) Dingemans, G.; Van Helvoirt, C. A. A.; Pierreux, D.; Keuning, W.; Kessels, W. M. M. J. Electrochem. Soc. 2012, 159, H277.

(27) Profijt, H. B.; Van de Sanden, M. C. M.; Kessels, W. M. M. Electrochem. Solid-State Lett. 2012, 15, G1.

(28) Knoops, H. C. M.; Langereis, E.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Vac. Sci. Technol. A 2012, 30, 01A1011‒01A10110.

(29) Black, K.; Werner, M.; Rowlands‒Jones, R.; Chalker, P. R.; Rosseinsky, M. J. Chem. Mater. 2011, 23, 2518‒2520.

(30) Knoops, H. C. M.; Mackus, A. J. M.; Donders, M. E.; Van de Sanden, M. C. M.; Notten, P. H. L.; Kessels, W. M. M. Electrochem. Solid-State Lett. 2009, 12, G34‒G36.

(31) Leick, N.; Verkuijlen, R. O. F.; Lamagna, L.; Langereis, E.; Rushworth, S.; Roozeboom, F.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Vac. Sci. Technol. A 2011, 29, 0210161‒0210167.

(32) Goldstein, D. N.; George, S. M. Thin Solid Films 2011, 519, 5339‒5347.

(33) Knapas, K.; Ritala, M. Chem. Mater. 2011, 23, 2766‒2771.

(34) Christensen, S. T.; Elam, J. W.; Rabuffetti, F. A.; Ma, Q.; Weigand, S. J.; Lee, B.; Seifert, S.; Stair, P. C.; Poeppelmeier, K. R.; Hersam, M. C.; Bedzyk, M. J. Small 2009, 5, 750‒757.

Page 20: Atomic layer deposition of ruthenium films : properties and surface

1. INTRODUCTION AND FRAMEWORK OF THE RESEARCH

13

(35) Feng, H.; Elam, J. W.; Libera, J. A.; Setthapun, W.; Stair, P. C. Chem. Mater. 2010, 22, 3133‒3142.

(36) Silvennoinen, R.; Jylhä, O.; Lindblad, M.; Österholm, H.; Krause, A. Catal. Lett. 2007, 114, 135‒144.

(37) Heo, J.; Eom, D.; Lee, S. Y.; Won, S.-J.; Park, S.; Hwang, C. S.; Kim, H. J. Chem. Mater. 2009, 21, 4006‒4011.

(38) Aaltonen, T.; Ritala, M.; Arstila, K.; Keinonen, J.; Leskelä, M. Chem. Vap. Deposition 2004, 10, 215‒219.

(39) Wang, H. T.; Gordon, R. G.; Alvis, R.; Ulfig, R. M. Chem. Vap. Deposition 2009, 15, 312‒319.

(40) Kukli, K.; Aarik, J.; Aidla, A.; Uustare, T.; Jogi, I.; Lu, J.; Tallarida, M.; Kemell, M.; Kiisler, A. A.; Ritala, M.; Leskelä, M. J. Cryst. Growth 2010, 312, 2025‒2032.

(41) Choi, S.-H.; Cheon, T.; Kim, S.-H.; Kang, D.-H.; Park, G.-S.; Kim, S. J. Electrochem. Soc. 2011, 158, D351‒D356.

(42) Kim, S. K.; Han, J. H.; Kim, G. H.; Hwang, C. S. Chem. Mater. 2010, 22, 2850‒2856.

(43) Park, S. K.; Kanjolia, R.; Anthis, J.; Odedra, R.; Boag, N.; Wielunski, L.; Chabal, Y. J. Chem. Mater. 2010, 22, 4867‒4878.

(44) Weber, M. J.; Mackus, A. J. M.; Verheijen, M. A.; Longo, V.; Bol, A. A.; Kessels, W. M. M. J. Phys. Chem. C 2014.

(45) Aaltonen, T.; Rahtu, A.; Ritala, M.; Leskelä, M. Electrochem. Solid-State Lett. 2003, 6, C130‒C133.

(46) Matsui, Y.; Hiratani, M.; Nabatame, T.; Shimamoto, Y.; Kimura, S. Electrochem. Solid-State Lett. 2001, 4, C9‒C12.

(47) Elliott, S. D. Langmuir 2010, 26, 9179‒9182.

(48) Jensen, W. B. J. Chem. Educ. 2007, 84, 1418.

Page 21: Atomic layer deposition of ruthenium films : properties and surface

1. INTRODUCTION AND FRAMEWORK OF THE RESEARCH

14

Page 22: Atomic layer deposition of ruthenium films : properties and surface

15

Chapter 2 “Creativity is just connecting things. When you ask creative people how they did some-thing, they feel a little guilty because they didn't really do it, they just saw something. It seemed obvious to them after a while. That's because they were able to connect experi-ences they've had and synthesize new things.” ‒ Steve Jobs

Ruthenium and its surface reactions: an overview overview from surface science perspective

Bulk properties of Ru

Ruthenium (Ru), whose name originates from the Latin name for Russia, Ruthenia, has the atomic number 44. It is a rare metal occurring as a minor component of Pt ores. Only 15 tons of Ru are extracted world-wide annually (compared to millions of tons in the case of iron).1,2 Ru belongs to the group of transition metals lying in the d -block called the Pt-group metals in the Periodic Table of Elements. This group refers to six metals (Ru, Rh, Pd, Os, Ir and Pt) hav-ing similar chemical properties occurring together in the same mineral depos-its.3 The primary application of Ru metal lies in wear-resistant electrical contacts, but as mentioned in Chapter 1, Ru finds myriads of other applications in fields such as heterogeneous catalysis,4‒6 energy storage and conversion,7,8 and sens-ing9‒11. They all rely on the metal’s properties such as its low electrical resistivity, its high work function and its ability to break bonds. The crystal structure, abil-ity to form a stable solid oxide (RuO2) and catalytic activity of Ru will be intro-duced in the following paragraphs, as these properties are significant for the understanding of this thesis.

As highlighted in Table I, the crystallographic structure of Ru is the hexago-nal closed packed (hcp) structure as schematically illustrated in Figure 1. For this structure, four Miller-Bravais direction indices (e.g. hkil) are used. It is there-fore useful to recall how they relate to the direction indices (e.g. h’k’l’). In sur-face science, single crystal Ru surfaces are generally used, of which Ru(0001) is

Page 23: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

16

the most extensively studied. From Eq. 1, the notations of Ru(0001) and Ru(001) are equivalent and both notations can be found in the literature.

h = (1/3) (2h’ ‒ k’) k = (1/3) (2k’ ‒ h’) i = -(h’ + k’) l = l’ (Eq. 1)

Figure 1: (a) Schematic representation of the hexagonally closed packed unit cell struc-ture. The lattice constants a and c are indicated (a = 2.71 Å and c = 4.28 Å), as well as the crystallographic direction [0001] along the unit cell vector z. (b) Schematic representa-tion of the rutile structure (Ru2O4 unit) as encountered in RuO2. The lattice constants a and c as well as the unit cell vectors are indicated (a = 4.49 Å and c = 3.11 Å).12

Table I: Material properties of bulk Ru and of RuO2, as reported by Refs.13‒15.

Property Ru RuO2 Atomic weight 101.07 amu -- Electrical resistivity (23 ºC) 7.4 μΩ.cm 35.2 μΩ.cm Electron configuration 2-8-18-15-1 or (Kr)4d75s1 -- Mass density (20 ºC) 12.2 g.cm-3 6.97 g.cm-3 Work function 4.7 eV 4.9 eV Crystal structure hexagonal closed packed tetragonal (rutile)

With its electrical resistivity of 7.4 μΩ.cm, Ru is a good conductor. The

configuration of the valence electrons is 4d7 5s1 (see Table I) and consequently, Ru can adopt various formal oxidation states in its chemical complexes and compounds, of which +2, +3 and +4 are the most common.12,16,17 The oxidation state is a number assigned to an element in a compound, representing a for-mal charge, making it convenient to use when balancing chemical reactions. The metal in its pure form has an oxidation state 0. Furthermore, the partially

(a) (b)

c

a

[0001]Ru

RuO

aa

c[001]

[110]

[110]

Page 24: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

17

occupied 4d and 5s orbitals can affect the reactivity of the metal. Generally speaking, surface atoms can be coordinatively unsaturated and they can form new chemisorptive bonds with species from the gas-phase, thereby changing existing bond configurations. In some cases, this process can even lead to the dissociation of the adsorbing species. On most surfaces, bond dissociation re-quires amounts of energy that can be very high. However, the d -electrons characteristic for the Pt-group metals provide more electron energy configura-tions and thus enable many reaction-pathway options unlike the valence elec-trons in the metals of the s-block or the metalloids of the p-block. With the ex-tra d ‒electron configuration the dissociation energy can be significantly low-ered. As a consequence, Ru surfaces specifically are able to induce the dissocia-tion of bonds such as O-O, C-O and C-H, which are of particular interest in this thesis work.

Ru is generally chemically stable, but if it is exposed to a sufficient O2 gas pressure and temperature, Ru can form stable oxide structures. RuO2 is the only solid oxide while RuO3 and RuO4 are (toxic) gaseous species. RuO2 is an impor-tant material on its own, and finds applications in heterogeneous catalysis18 and microelectronics19‒21. Its most important properties are its catalytic activity and the fact that it behaves as a conductor (resistivity of 35.2 μΩ.cm) with a work function close to that of other Pt-group metals (4.9 eV). RuO2 bulk gener-ally forms a rutile crystal structure, as illustrated in Fig. 1b. The main preparation method of RuO2 films (typically 1-2 nm thick) consists in exposing a Ru surface to large amounts of molecular oxygen at temperatures ranging from 600-750 K.12 This ability of Ru to form RuO2 has implications for certain chemical re-actions; some of which will be discussed in this Chapter.

Adsorption processes of atoms and molecules on transition metals

The adsorption of molecules onto a transition metal surface occurs via the interaction of the molecules’ outer electrons and the metals’ d -band. Generally spoken, when the bond of the molecule to the surface is strong, the internal molecular bond is weakened and ultimately, can dissociate. In an attempt to

Page 25: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

18

Figure 2: Schematic of the interaction of the valence electrons of an atom (a-b) or a molecule (c) with the d-bands of a metal. (a) If the d-band of the metal is half-filled, only the bonding states of the adsorbate are occupied upon adsorption of the atom, causing a strong adsorbate-metal bond. In (b) on the other hand, the metal d -band is more than half-filled and upon adsorption, the antibonding states are also (at least partially) occu-pied which weakens the adsorbate-metal bond. (c) In the case of a molecule adsorbing onto a d -metal, the orbitals of the adsorbed molecule hybridize further into σ and σ* orbitals. The interaction of the electrons in the highest occupied molecular orbital (HO-MO) with the d-band electrons is called donation, whereas the interaction of the d-band electrons with the lowest unoccupied molecular orbital (LUMO) is called back-donation. If the back-donation is strong, the internal molecular bonds are weakened and can ulti-mately break, as it is the case for simplest diatomic molecule H2 which dissociates on a d -metal surface. (Figures (a)-(c) are reproduced from Ref.22 with the author’s consent).

(a) (b)

EF

Evac

d-band

s-band

d-metal adsorbed atom

antibonding

bonding

freeatom

EF

Evac

d-metal adsorbed atom

freeatom

antibonding

bonding

’reactive’ d-metals, e.g. Cr less reactive d-metals, e.g. Rh

EF

Evac

d-metal adsorbed molecule free molecule

bonding

antibonding

bonding

antibonding

LUMO

HOMO

σ-orbitals σ*-orbitals

(c)

Atomic adsorption on a d-metal

Molecular adsorption on a d-metal

Page 26: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

19

discuss the dissociation in more details, basic concepts of molecular orbital theory are introduced.

First, the bonding between an adsorbing atom or molecule and a metal from the d -block in the Periodic Table of Elements will be discussed. Upon in-teraction of the ad-atom’s valence electrons with the metal’s d -electrons, the electron densities overlap, and bonding and anti-bonding states appear. As illustrated in Fig. 2a and b, if the metal d -band is half-filled (or less), the antibonding levels remain empty, which makes the adsorbate-metal bond strong. If the d -band is more than half-filled, the antibonding states are (at least) partially filled, thereby weakening the total adsorbate-metal binding en-ergy. The interaction of a metal with an adsorbing molecule occurs analogously to that with an atom. Figure 2c shows that the electrons in the highest occu-pied molecular orbital (HOMO) interact with the d -band of the metal, giving rise to the bonding and antibonding states of the adsorbate. This hybridization is caused by the electrons “donated” to the metal, and is therefore often called donation. On the other hand, the hybridization caused by the interaction of the lowest unoccupied molecular orbital (LUMO) with the metal d -band is caused by the electrons “back-donated” from the metal to the adsorbate, and is there-fore often called back-donation. If this back-donation is strong, e.g. in the case of H2 with most Pt-group metal surfaces, the internal molecular bonds are dras-tically weakened, eventually leading to the dissociation of the molecule, and two dissociation products are chemisorbed to the metal surface. In the case of adsorption onto a Ru-metal surface, molecular H2 will be chemisorbed undissociated with an enthalpy change of -1.5 eV. When these molecular spe-cies are further activated by dissociation into chemisorbed H-atoms the overall enthalpy change increases to -6.2 eV.22

The ability of Ru to break O-O and C-H bonds is of particular interest for this thesis work. Therefore, these processes on the Ru surface will be reviewed from surface science and catalysis studies in the following sections.

O2 dissociative chemisorption on Ru(0001)

To study the dissociative chemisorption of gas-phase O2 on Ru(0001), the metal surface is generally exposed to O2 gas of which the exposure values can be varied. The exposure is defined as the product of pressure and dosing time, and is measured in Langmuir (1 L = 1×10-6 Torr‧s). The main dissociative ad-

Page 27: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

20

sorption from the initial transient molecular chemisorption occurs via back-donation of charge from the metal into the anti-bonding 1πg and 3σu orbitals of O2.23‒25 However, not all of the impinging molecules become adsorbed at the surface, and the probability for a molecule to adsorb is determined by the stick-ing coefficient.a The values for the sticking coefficient lie between the two ex-treme values of 1 (all of the impinging molecules are adsorbed) and 0 (none of the impinging molecules are adsorbed). Under ultra-high vacuum (UHV) condi-tions (~10-9 Torr) the p (2×2)O and the p (2×1)O ordered overlayer configura-tionsb were experimentally observed on Ru(0001).26,27 These p (2×2)O and the p (2×1)O configurations correspond to the O coverage values (θO) of θO = 0.25 and 0.5 ML, whereby one monolayer (ML) corresponds to one adsorbate atom or molecule for each surface atom. At θO = 0.5 ML, the disso-ciative sticking coefficient of O2 drops from almost one to less than 10-3, such that under UHV conditions the p (2×1)O constitutes the saturation coverage.28,29 The use of NO2 has proven itself to achieve chemisorption with higher O coverage values under UHV conditions, but these O-Ru-phases showed to be catalytically inactive (e.g. in the surface-assisted CO oxidation reaction).30‒32

On the other hand, by increasing the O2 exposure to ≥103 L, coverage val-ues of θO = 0.75 and 1 ML can be reached,28,33‒37 corresponding to the p (2×2)3O and p (1×1)O overlayer structures, as summarized in Figure 3. Once the top layer of Ru(0001) is covered, the dissociative sticking coefficient is esti-mated to be less than 10-6.38 Nevertheless, under even higher oxidizing condi-tions (~105 L), the Ru(0001) surface can be further oxidized and subsurface oxy-

a The general expression for the sticking coefficient s in the case of activated adsorption is written as s=σ·f θ ·exp( -Eact kBT), and contains the following terms: σ, called the condensation coefficient, is responsible for steric factors of the incoming species. The coverage dependent function f θ describes the probability of finding an available ad-sorption site. This term f θ is determined by the type of the adsorption reaction (i.e. dissociative or non-dissociative), the statistics of site occupation, and the mobility of molecules in a precursor adsorption state. The temperature-dependent Boltzmann term (kBT) is associated with the energy of the activated adsorption (Eact)64. b These overlayer structures are denoted in accordance with Wood’s notation,64,65 which describes the adsorbate unit cell (m×n) as translations and rotations of the unit cell vec-tors of the substrate surface. The p preceding the unit cell stands for the so-called primi-tive structure of adsorbate unit cell.

Page 28: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

21

gen can be formed, i.e. θO > 1 ML, which has been validated experimentally and theoretically. For O to diffuse below the first Ru layer, it is required that the

Figure 3: Evolution of O coverage as a function of O2 exposure in Langmuirs (1 L = 1×10-6 Torr‧s) for a Ru(0001) surface at 600 K, and the expected surface configuration of metastable O at coverages of 0.25 ML, 0.5 ML, 0.75 ML, and 1 ML (ML=monolayer). Note, that p refers to the primitive adlayer unit cell. (based on Refs. 28,30,39).

first layer of Ru is fully covered with O and that the temperature T lies in the range 500 K < T < 800 K. Two mechanisms have been proposed for the for-mation of subsurface O in the top layers of Ru. According to Blume et al.,36 once the first ML of Ru is covered with O, these atoms permeate into the second lay-er of Ru and occupy a subsurface site, thereby creating new adsorption sites on the O-covered Ru surface for new O2 gas-phase molecules. On the other hand, Böttcher et al.30 claim that the adsorption of O2 at defects, still present when θO = 1 ML is reached on the Ru surface, triggers a structural distortion that spreads into the underlying Ru layer, thereby populating the subsurface region. By defect, we mean vacancies and adatoms, but also domain boundaries or surface roughness. Therefore, Böttcher et al. performed experiments on rough Ru surfaces and found the results to be in agreement with this proposed mechanism. Blume et al.40,41 later also confirmed that defects in the Ru layer facilitated not only the incorporation of O into the subsurface region but also

10 0 10 1 10 2 10 3 10 4 10 5 10 6 10 70.0

0.5

1.0

1.5

2.0

O c

over

age

(ML)

O2 exposure (L)

Adsorbed OAdsorbed OSubsurface OSubsurface O

1 ML p(1x1) ORu(0001) hcp

at 600K

0.5 ML p(2x1) O

0.25 ML p(2x2) O0.75 ML p(2x2) 3O

Page 29: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

22

the nucleation of RuO2 microregions. Experimental studies revealed that for an O coverage value of θO = 2-3 ML, the rates of CO oxidation were enhanced by more than two orders of magnitude with respect to the p (1×1)O pre-covered Ru(0001) surface. Under these circumstances, this O coverage value constitutes a critical amount of subsurface O inducing sufficient lattice deformations for an entire micro-region of Ru(0001)-O to unfold into a rutile RuO2(110) lattice struc-ture.42‒44 As soon as these oxide microregions are formed, the oxidation process proceeds in an autocatalytic manner,45 because the sticking coefficient of O2 on RuO2 is close to 1 (~0.7). This is significantly higher than 10-6 on the O-covered Ru(0001).38 The transition from oxygen adsorption to oxide formation on the Ru(0001) surface is structurally quite complex and leads to the co-existence of (1×1)O-Ru(0001) and RuO2(110) patches, which was schematically summarized by Over et al.45 and is reproduced in Figure 4. The RuO2 patches were identified as the most catalytically active phase for oxidation reactions of CO. These RuO2 domains provide coordinatively unsaturated (cus) Ru atoms at the RuO2 top layer, which were claimed to be very reactive and to represent the adsorption sites for impinging molecules. On the other hand, the Ru(0001) surface with 0 < θO < 3 ML does not play a significant role in the surface oxidation reactions

Coordinatively unsaturated RuChemisorbed O

O-)1x1(O-)1x1(

RuO2(110)

Fast O2dissociation

Slow O2dissociation

Buried RuO2

Dissolved/subsurface oxygen Ru

[110]

[110]–

[001]

RuO2(110)

Figure 4: Schematic overview of a heterogeneous Ru surface when it is exposed to large amounts of O2 (103‒106 L), as reproduced from Over et al.45 Microregions of chemisorbed oxygen, surface oxide, buried oxides, and subsurface oxygen can co-exist in the near-surface region of Ru.

Page 30: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

23

of CO.43,44 A dedicated set of experiments coupled with calculations on the CO oxidation reaction revealed the reason for this “inactivity” to lie in the strong Ru-O bond and the low desorption temperature of CO. Consequently, under UHV conditions, in which the O coverage is well below 3 ML, the chemisorbed CO desorbs without reacting with the surface O.25

(De)hydrogenation reactions of CxHy-species on Ru(0001)

In this section, the dissociative chemisorption of hydrocarbon species via hydrogen abstraction, more commonly called a dehydrogenation reaction will be discussed. This type of reaction is widely used in industrial processes. For example, in reforming reactions of gasoline, Pt is used to dehydrogenate al-kanes to olefins.46,47 Another important application is in the fat hardening in-dustry for food, in which noble metals are used to dehydrogenate fatty acids to increase the number of multiple unsaturated bonds (i.e. to make food healthi-er).48,49 Dehydrogenation reactions of hydrocarbons (typically C2H4) have re-cently also been investigated as a pathway to assist in the growth of graphene sheets on noble metals, especially on Ru.50‒52

In the pursuit to convert hydrocarbons (CxHy), mostly CH4, to higher hydro-carbons (Cn>xHm) hydrogenation and dehydrogenation reactions have been extensively studied in surface science. In a general manner, a dehydrogenation reaction can be written in the following way,53 where * denotes the surface species:

CxHy* CxHy-2* + H2 (gas) (Eq. 2)

Transition metal catalysts, and specifically Ru(0001), have been widely studied to decrease the activation energy of this reaction. However, these studies com-prise results from many different hydrocarbon species which makes it elaborate to review. Therefore, this section will only highlight the main results that the studies have in common. The abstraction of hydrogen in an adsorbed hydro-carbon is more complex than the dissociation of a diatomic molecule. There-fore, the molecular orbital model presented earlier is probably too simplistic to describe the H-abstraction from the hydrocarbon.

Surface dehydrogenation reactions on Ru(0001) involve many intermediate reactions, each leading to the abstraction and transfer of at least one H-atom to the Ru surface, as it is schematically represented in Fig. 5 in the case of CH4. The

Page 31: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

24

activation energy for each H-abstraction reaction is different and it seems to be largest for the final C-H scission.54,55 Under UHV conditions, at the end of surface dehydrogenation reactions on Ru(0001) surfaces, a layer of C atoms remains on the surface and all of the hydrogen is released into the gas-phase in the form of H2.54,56‒58 These surface reactions are terminated at temperatures of typically 700-800 K.54 It is also important to highlight that dehydrogenation reactions also seem to occur on Ru textures other than Ru(0001), but such reports are rare.59 In catalysis, the layer of C atoms is well-known as it leads to the effect of “surface poisoning” that prevents gas-phase species to adsorb on the metal surface. On the other hand, this disadvantage has also been exploited by re-searchers demonstrating that this surface poisoning could lead to graphene sheets on the Pt-group metals, such as Pt or Ru.50,51

Figure 5: Schematic representation of the dehydrogenation reaction of CH4 on a Ru surface. CH4 impinges onto the Ru surface and hydrogen atoms are abstracted from the CH4 molecule. The Ru metal atoms accept both the hydrocarbon fragment and the hy-drogen atoms. When H atoms are abstracted from the molecule to the surface, two H atoms recombine and H2 is released into the gas-phase.

Particularly for the current work, the dehydrogenation of larger hydrocar-

bons, such as ethylene (C2H4) or benzene (C6H6), are important to review be-cause they resemble the organic ligands used in the present precursor, (C5H5)Ru(CO)2(CH2CH3). The sticking probability of C2H4 on Ru(0001) surfaces was found to be in the order of one at 180 K.60 Furthermore, evidence for ethy- lene dehydrogenation at temperatures between 300 K and 600 K was reported,

CH

HHH

CHH

H

CH

H

H

HH

CH

HH

H

HH

H 2

CH4

C

Page 32: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

25

liberating H2 and leaving a characteristic carbon containing surface layer.60 In experiments on Ru(0001) pre-exposed to O2 gas, it was found that even a low O coverage was able to significantly weaken the chemisorption of ethylene on the Ru surface. Also in the case of benzene, it was shown that C6H6 had a stick-ing probability in the order of one.61 Benzene dehydrogenation on Ru(0001) was observed to start at ~320 K; and the activation temperature depended on the benzene coverage on the Ru surface.61 Recently, the dehydrogenation of CHx (0 x 4) and C2Hx (0 x 6) hydrocarbon fragments, was studied on a Ru(0001) surface using density functional theory based calculations. The results of this study confirm the dehydrogenation of these hydrocarbon species to be thermodynamically favorable on a Ru(0001) surface.62

Table II: Interaction of hydrocarbon with Ru vs. RuO2 surfaces. (Reproduced from Over et al.63, adapted for the hydrocarbons relevant in this work)

Interaction with Ru Interaction with RuO2 Hydrocarbons with conjugated dou-ble bonds (e.g. benzene): medium

Hydrocarbons with conjugated double bonds (e.g. benzene): weak

Hydrocarbons with carbonyl groups: medium

Hydrocarbons with carbonyl groups: medium

Forms easily a single graphite layer Cannot form a graphite layer Single C-atom: strong Single C-atom: can hardly be stabilized

Dehydrogenation reactions also occur on RuO2 surfaces, but the underlying

mechanism is different from that on Ru surfaces.63 On the RuO2 the undercoordinated metal atoms accommodate the hydrocarbon (fragment), while the undercoordinated O atom accepts the abstracted hydrogen atom. With RuO2, hydrocarbons can only be removed through total oxidation of hy-drocarbons into H2O and CO2, thereby reducing the RuO2 top layers. Further-more, the carbon layer remaining on the Ru surface as a result of hydrocarbon dehydrogenation could not be observed in the case of dehydrogenation reac-tions on RuO2. Table II highlights the differences of hydrocarbon interaction on Ru and RuO2 surfaces.

Page 33: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

26

Summary and Conclusion

In this Chapter, basic properties of Ru surfaces were introduced and specif-ic reactions were reviewed on the basis of surface science reports; both of which will be used in the following chapters. Surface reactions on Pt-group metal surfaces can be quite complex due to their unique electronic d -band structure which can lead to the dissociative chemisorption of adsorbing gase-ous species, such as O2 and CxHy. The interaction of O2 gas with Ru surfaces was summarized and the role of RuO2 as a reactive phase in surface reactions was addressed. Furthermore, the interaction of CxHy on Ru (and RuO2) surfaces was briefly reviewed and important results from surface science experiments were highlighted, such as the liberation of H2 as a result of surface dehydrogenation reactions and the carbon layer remaining on the Ru surface, once the dehydro-genation reactions are terminated. The temperature requirements for the de-hydrogenation reactions and the formation of RuO2 structures are schematical-ly summarized in Fig. 6 and are based on the literature reviewed earlier.

Figure 6: Schematic summary of the temperature requirements (in K) for dehydrogena-tion reactions of hydrocarbons on Ru(0001) surfaces and for the formation of RuO2 structures (microregions, thin films, etc.). This schematic is based on the literature re-ports discussed earlier in this Chapter.

References

(1) Salazar, K.; McNutt, M. K. U.S. Geological Survey, 2012, Mineral commodity summaries 2012: U.S. Geological Survey; 2012; p. 120.

(2) Jorgenson, J. D. Mineral Commodity Summaries: Iron ore; 2011; pp. 84‒85.

(3) Harris, D. C.; Cabri, L. J. Can. Mineral. 1991, 29, 231‒237.

(4) Lu, J.; Fu, B.; Kung, M. C.; Xiao, G.; Elam, J. W.; Kung, H. H.; Stair, P. C. Science 2012, 335 , 1205‒1208.

(5) Chao, C.-C.; Motoyama, M.; Prinz, F. B. Adv. Eng. Mater. 2012, 2, 651‒654.

100 K 500 K 1000 K

Temperature

dehydrogenation reactions of hydrocarbons on Ru(0001) surfaces

onset of RuO2 structures formation

Page 34: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

27

(6) Cassir, M.; Ringuede, A.; Niinisto, L. J. Mat. Chem. 2010, 20, 8987‒8993.

(7) Dingemans, G.; Kessels, W. M. M. J. Vac. Sci. Technol. A 2012, 30, 40802‒40827.

(8) Law, M.; Greene, L. E.; Radenovic, A.; Kuykendall, T.; Liphardt, J.; Yang, P. J. Phys. Chem. B 2006, 110, 22652‒22663.

(9) Marichy, C.; Donato, N.; Willinger, M.-G.; Latino, M.; Karpinsky, D.; Yu, S.-H.; Neri, G.; Pinna, N. Adv. Funct. Mater. 2011, 21, 658‒666.

(10) Niskanen, A. J.; Varpula, A.; Utriainen, M.; Natarajan, G.; Cameron, D. C.; Novikov, S.; Airaksinen, V.-M.; Sinkkonen, J.; Franssila, S. Sensor. Actuat. B - Chem. 2010, 148, 227‒232.

(11) Kim, W.-S.; Lee, B.-S.; Kim, D.-H.; Kim, H.-C.; Yu, W.-R.; Hong, S.-H. Nanotechnology 2010, 21, 245605.

(12) Over, H. Chem. Rev. 2012, 112, 3356‒426.

(13) Lide, D. R. CRC Handbook of Chemistry and Physics: A Ready-reference Book of Chemical and Physical Data; CRC Press,: Boca Raton, USA, 2004.

(14) Meaden, G. T. Electrical resistance of metals; Plenum Press: New York, 1965.

(15) Ryden, W.; Lawson, A.; Sartain, C. Phys. Rev. B 1970, 1, 1494‒1500.

(16) Rard, J. A. Chem. Rev. 1985, 85, 1‒39.

(17) Seddon, E. A.; Seddon, K. R. The Chemistry of Ruthenium; Elsevier: Amsterdam, 1984.

(18) Over, H.; Kim, Y. D.; Seitsonen, A. P.; Wendt, S.; Lundgren, E.; Schmid, M.; Varga, P.; Morgante, A.; Ertl, G. Science 2000, 287 , 1474‒1476.

(19) Manke, C.; Boissière, O.; Weber, U.; Barbar, G.; Baumann, P. K.; Lindner, J.; Tapajna, M.; Fröhlich, K. Microelectron. Eng. 2011, 83, 2277‒2281.

(20) Kwon, S. H.; Kwon, O. K.; Kim, J. H.; Jeong, S. J.; Kim, S. W.; Kang, S. W. J. Electrochem. Soc. 2007, 154, H773‒H777.

(21) Fröhlich, K.; Husekova, K.; Machajdik, D.; Hooker, J. C.; Perez, N.; Fanciulli, M.; Ferrari, S.; Wiemer, C.; Dimoulas, A.; Vellianitis, G.; Roozeboom, F. Mat. Sci. Eng. B-Solid 2004, 109, 117‒121.

(22) Chorkendorff, I.; Niemantsverdriet, J. W. Concepts of Modern Catalysis and Kinetics; Wiley-VCH: Weinheim, Germany, 2003.

Page 35: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

28

(23) Wheeler, M. C.; Seets, D. C.; Mullins, C. B. In The 42nd national symposium of the American Vacuum Society; AVS: Mineapolis, Minnesota (USA), 1996; Vol. 14, pp. 1572‒1577.

(24) Mitchell, W. J.; Xie, J.; Lyons, K. J.; Weinberg, W. H. J. Vac. Sci. Technol. A 1994, 12, 2250‒2254.

(25) Bonn, M.; Funk, S.; Hess, C.; Denzler, D. N.; Stampfl, C.; Scheffler, M.; Wolf, M.; Ertl, G. Science 1999, 285, 1042‒1045.

(26) Lindroos, M.; Pfnür, H.; Held, G.; Menzel, D. Surf. Sci. 1989, 222, 451‒463.

(27) Pfnür, H.; Held, G.; Lindroos, M.; Menzel, D. Surf. Sci. 1989, 220, 43‒58.

(28) Stampfl, C.; Schwegmann, S.; Over, H.; Scheffler, M.; Ertl, G. Phys. Rev. Lett. 1996, 77, 3371‒3374.

(29) Praline, G.; Koel, B. E.; Lee, H.-I.; White, J. M. Appl. Surf. Sci. 1980, 5, 296‒312.

(30) Böttcher, A.; Niehus, H. J. Chem. Phys. 1999, 110, 3186‒3195.

(31) Schwegmann, S.; Seitsonen, A. P.; De Renzi, V.; Dietrich, H.; Bludau, H.; Gierer, M.; Over, H.; Jacobi, K.; Scheffler, M.; Ertl, G. Phys. Rev. B 1998, 57, 15487‒15495.

(32) He, P.; Jacobi, K. Phys. Rev. B 1997, 55, 4751‒4754.

(33) Peden, C. H. F.; Goodman, D. W. J. Phys. Chem. 1986, 90, 1360‒1365.

(34) Goodman, D. W.; Peden, C. H. F. J. Phys. Chem. 1986, 90, 4839‒4843.

(35) Jakob, P.; Gsell, M.; Menzel, D. J. Chem. Phys. 2001, 114, 10075‒10085.

(36) Blume, R.; Niehus, H.; Conrad, H.; Böttcher, A. J. Chem. Phys. 2004, 120, 3871‒3879.

(37) Kostov, K. L.; Gsell, M.; Jakob, P.; Moritz, T.; Widdra, W.; Menzel, D. Surf. Sci. 1997, 394, L138‒L144.

(38) Böttcher, A.; Niehus, H. Phys. Rev. B 1999, 60, 14396‒14404.

(39) Meinel, K.; Wolter, H.; Ammer, C.; Beckmann, A.; Neddermeyer, H. J. Phys. Cond. Matter. 1997, 9, 4611‒4619.

(40) Blume, R.; Niehus, H.; Conrad, H.; Böttcher, A.; Aballe, L.; Gregoratti, L.; Barinov, A.; Kiskinova, M. J. Phys. Chem. B 2005, 109, 14052‒14058.

(41) Blume, R.; Havecker, M.; Zafeiratos, S.; Teschner, D.; Kleimenov, E.; Knop-Gericke, A.; Schlogl, R.; Barinov, A.; Dudin, P.; Kiskinova, M. J. Catal. 2006, 239, 354‒361.

Page 36: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

29

(42) Reuter, K.; Ganduglia-Pirovano, M. V.; Stampfl, C.; Scheffler, M. Phys. Rev. B. 2002, 65, 165403.

(43) Kim, Y. D.; Seitsonen, A. P.; Over, H. Surf. Sci. 2000, 465, 1‒8.

(44) Böttcher, A.; Niehus, H.; Schwegmann, S.; Over, H.; Ertl, G. J. Phys. Chem. B 1997, 101, 11185‒11191.

(45) Over, H.; Seitsonen, A. P. Science 2002, 297, 2003‒2005.

(46) Wei, J.; Iglesia, E. J. Phys. Chem. B 2004, 108, 4094‒4103.

(47) Rioux, R. M.; Marsh, A. L.; Gaughan, J. S.; Somorjai, G. A. Catal. Today 2007, 123, 265‒275.

(48) Fu, J.; Lu, X.; Savage, P. E. Chem. Sus. Chem. 2011, 4, 481‒6.

(49) Piqueras, C.; Bottini, S.; Damiani, D. Appl. Catal. A 2006, 313, 177‒188.

(50) Loginova, E.; Bartelt, N. C.; Feibelman, P. J.; McCarty, K. F. New J. Phys. 2009, 11, 63046.

(51) Land, T. A.; Michely, T.; Behm, R. J.; Hemminger, J. C.; Comsa, G. Surf. Sci. 1992, 264, 261‒270.

(52) Sutter, E.; Albrecht, P.; Sutter, P. Appl. Phys. Lett. 2009, 95, 133109.

(53) Huff, M.; Schmidt, L. D. J. Phys. Chem. 1993, 97, 11815‒11822.

(54) George, P. M.; Avery, N. R.; Weinberg, W. H.; Tebbe, F. N. J. Am. Chem. Soc. 1983, 105, 1393‒1394.

(55) Ciobica, I. M.; Frechard, F.; Van Santen, R. A.; Kleyn, A. W.; Hafner, J. Chem. Phys. Lett. 1999, 311, 185‒192.

(56) Ilharco, L. M.; Garcia, A. R.; Hargreaves, E. C.; Chesters, M. A. Surf. Sci. 2000, 459, 115‒123.

(57) Garcia, A. R.; Barros, R. B.; Ilharco, L. M. Surf. Sci. 2009, 603, 380‒386.

(58) Kis, A.; Kiss, J.; Olasz, D.; Solymosi, F. J. Phys. Chem. B 2002, 106, 5221‒5229.

(59) Wu, M. C.; Xu, Q.; Goodman, D. W. J. Phys. Chem. 1994, 98, 5104‒5110.

(60) Barteau, M. A.; Broughton, J. Q.; Menzel, D. Appl. Surf. Sci. 1984, 19, 92‒115.

(61) Jakob, P.; Menzel, D. Surf. Sci. 1988, 201, 503‒530.

(62) Ande, C. K.; Elliott, S. D.; Kessels, W. M. M. accepted for publication in J. Phys. Chem.

Page 37: Atomic layer deposition of ruthenium films : properties and surface

2. RUTHENIUM AND ITS SURFACE REACTIONS: A PERSPECTIVE OVERVIEW...

30

(63) Over, H.; He, Y. B.; Farkas, A.; Mellau, G.; Korte, C.; Knapp, M.; Chandhok, M.; Fang, M. J. Vac. Sci. Technol. B 2007, 25, 1123‒1138.

(64) Oura, K.; Lifshits, V. G.; Saranin, A. A.; Zotov, A. V; Katayama, M. Surface Science: An Introduction; Springer, 2003; p. 440.

(65) Nix, R. M. An Introduction to Surface Chemistry http://www.chem.qmul.ac.uk/surfaces/scc/scat6_1.htm.

Page 38: Atomic layer deposition of ruthenium films : properties and surface

31

Chapter 3 “For the things we have to learn before we can do, we learn by doing.” ‒ Aristotle

CpRu(CO)2Et as the metal precursor during Ru ALD

Ru ALD precursors and their requirements

As mentioned in Chapter 1, the metal precursor used throughout this the-sis has been cyclopentadienylethyl(dicarbonyl)ruthenium, (C5H5)Ru(CO)2-(CH2CH3) or CpRu(CO)2Et, which was supplied by SAFC Hitech®. Since Ru ALD started to be investigated, a wide range of Ru precursors has been studied. In this first part of the chapter, a brief classification of precursors will be made and some of the requirements on ALD precursors will be highlighted.

Recently, the requirements to be met by precursors used in ALD have been reviewed;1 next to trivial requirements such as high purity and volatility the most crucial requirement consists of a trade-off between a high reactivity in ALD surface reactions and a good thermal stability of the compound. A high reactivity toward the substrate surface leads to fast saturation reactions, while the thermal stability contributes to the self-limiting behavior of the first precur-sor half-cycle. Finding the “perfect” combination of the two properties has proven to be very challenging in the design and synthesis of Ru precursors.

The majority of Ru precursors that have been used in the last decade can be categorized into the following groups:2 the β-diketonates, the metal-locenes (i.e. cyclopentadienyl-based metalorganics) and other highly reactive compounds (such as RuO4 and zero-valent Ru metal complex compounds). The detailed table in Appendix A.1 shows most of the Ru precursors mentioned in the literature as reviewed by Hämäläinen et al.2 One of the differences between the precursor groups lies in the oxidation state which is +3 in the case of β-diketonates and +2 for metallocenes. In many cases, the Ru film growth per cycle values reported were higher when metallocenes were used than for β-diketonates and the levels of O-, C- and H impurities are slightly lower when metallocene precursors are involved in the Ru ALD process. For these reasons metallocene-based precursors have been more developed than β -

Page 39: Atomic layer deposition of ruthenium films : properties and surface

3. CpRu(CO)2Et AS THE METAL PRECURSOR DURING Ru ALD

32

diketonates. In the pursuit of yet higher growth per cycle values (> 1Å/cycle), more reactive metallocenes have been developed as well as highly reactive compounds, such as RuO4 or other zero-valent precursors.3,4

The latest trend is the development of so-called heteroleptic precurors, i.e. precursors with different ligands around their metallic center, synthesized to optimize the molecular structure in terms of physical properties (volatility, liqu-id state and improved thermal stability) and chemical reactivity towards the substrate surface. Our precursor, CpRu(CO)2Et, is such an example.

Cracking pattern of CpRu(CO)2Et and considerations on the molecular

binding

The precursor of choice for this thesis, CpRu(CO)2Et, falls in the category of heteroleptic metallocenes. The reaction mechanism involved in Ru ALD using this precursor with O2 gas will be discussed in Chapters 6 and 7. This section aims at giving chemical information of the compound.

A fragmentation pattern of the precursor CpRu(CO)2Et due to electron-impact was first measured to gain knowledge about the most abundant frag-ments of the molecule. From this precursor fingerprint, it is possible to select the mass that can be used to monitor the precursor during the ALD process by mass spectrometry (QMS). It was measured by injecting CpRu(CO)2Et vapor into the deposition chamber with closed top and bottom valve, which maintained the chamber pressure at ~10 mTorr. The molecular weight of the intact CpRu(CO)2Et precursor molecule is 252 amu. Since the intensities of the lower mass-to-charge-ratios (m/z = 1-50) were relatively high compared to the higher masses, the m/z range of 50 <m/z < 300 was scanned with different settings and the signals were normalized to the intensity of m/z = 50. Figure 1 shows the fragmentation pattern of CpRu(CO)2Et (m/z = 1-107) obtained from direct electronic ionization and dissociative ionization of the precursor molecule in the QMS.5 All of the precursor ligands could be identified, as well as a small sig-nal from the intact precursor molecule: CO+ and CH2CH2

+ (m/z = 28), C5H5+

(m/z = 65), Ru+ and its stable isotopes (99 < m/z < 101)6, and CpRu(CO)2Et+ (m/z = 252, not shown here). Because the direct ionization cross-section of CH2CH2 and of CO are in the same order of magnitude,7 the peak at m/z = 28 was assigned to a convolution of the signals from CO+ and CH2CH2

+. This is in

Page 40: Atomic layer deposition of ruthenium films : properties and surface

3. CpRu(CO)2Et AS THE METAL PRECURSOR DURING Ru ALD

33

agreement with the fragmentation pattern reported by NIST8 of ethylene (C2H6) and of the carbonyl ligand.

Figure 1: Electron impact fragmentation pattern of CpRu(CO)2Et (m/z = 1-107) at 40 eV measured by QMS. The fragmentation pattern for RuCp2 and ethylene (C2H6) from the NIST database are shown for comparison purposes, and were normalized to m/z = 28 for the ethyl ligand and to m/z = 39 for RuCp2. The highlighted groups of m/z values are assigned to known species part of the parent molecule.

Ru can form complexes with heteronuclear molecules, such as CO and/or

ionic cyclopentadienyl species, like in the heteroleptic precursor that we stud-ied, i.e. (C5H5)Ru(CO)2(CH2CH3). In the case of CO ligands, Ru will not only form σ-bonds by donation of the lone electron pair on the carbonyl’s carbon atom but it will also interact through the formation of a bond with so-called π-type symmetry by donation in the opposite direction of electrons from a filled metal d -orbital back into the empty π* antibonding orbital on the CO molecule. This double-donation mechanism takes place as illustrated in Figure 2. Note, that the double donation prevents the build-up of charge in the molecule. In heteroleptic precursors the bond order of the dissimilar metal-ligand bonds can vary, depending on the ligand’s donor or acceptor nature. In their turn the metal-ligand bonds will be differently polarized by surface groups (e.g. electro-negative O-atoms) on a catalytic surface, leading to chemisorption or partial dissociation into fragments that can be different from a homoleptic precursor.

m/z

Page 41: Atomic layer deposition of ruthenium films : properties and surface

3. CpRu(CO)2Et AS THE METAL PRECURSOR DURING Ru ALD

34

Figure 2. Schematic illustration of σ-donation and pπ-dπ back bonding and electron transfer from filled metal d -orbitals into the π*-orbital of a CO ligand. (Illustration re-produced from Ref. 9)

Temperature-dependent infrared studies on CpRu(CO)2Et

As mentioned above, the thermal stability is one of the critical aspects of an ALD precursor. In the scope of the European MaxCaps project, Dr. S. Haukka of ASM Microchemistry studied the thermal stability of “fresh” (i.e. not stored) CpRu(CO)2Et. She performed Fourier transform infrared absorption (FT-IR) spec-troscopy of CpRu(CO)2Et adsorbed onto porous high surface-area silica sub-strates heated at different temperatures.10 From the samples shown in Fig. 3a, it appears that CpRu(CO)2Et becomes darker with increasing annealing tempera-ture. Furthermore, Fig. 3b depicts the FTIR spectra of the CpRu(CO)2Et-silica sys-tem annealed at temperatures varying from 125 to 325 ºC. In this figure, the FT-IR spectra from 300 ºC on are clearly different from the others as the FT-IR sig-natures of the Cp- and the ethyl ligands are no longer present. This observation suggests that the compound has started to decompose at 300 ºC. Furthermore, the FT-IR spectra also show that the feature of C=O is largely removed at all the temperatures investigated, suggesting that this ligand is loosely bound to the Ru center. This observation is consistent with the mass loss data from thermogravimetric analysis (TGA) observed at 190 ºC (Fig. 1b in Chapter 4) which was attributed to the removal of the carbonyl ligand.

It is also interesting to note that the CpRu(CO)2Et precursor can react with a

hydroxylated Si surface. This suggests that the incubation time in Ru film

growth can be mostly attributed to the difficult removal of the ligands rather

than the adsorption of the precursor onto the surface. According to the reac

Page 42: Atomic layer deposition of ruthenium films : properties and surface

3. CpRu(CO)2Et AS THE METAL PRECURSOR DURING Ru ALD

35

Figure 3: (a) Photograph of four samples of CpRu(CO)2Et adsorbed on silica beads that have been heated to different temperatures between 125-300 ºC. (b) Infrared absorp-tion spectra of CpRu(CO)2Et adsorbed on silica beads during the thermal treatment at 125, 150, 175, 200, 250, 300 and 325 ºC. The reference spectra of CpRu(CO)2Et in ambient air and of the silica beads heated at 350 ºC are shown above and below the measure-ments, respectively. (with courtesy of S. Haukka10 for (a) and (b)).

tion mechanism proposed for Pt,11 the initial limiting factor is the lack of catalyt-

ic material present on the surface to break down the precursor ligands. In the

case of Pt, it was shown that initially the O2 gas assists the surface diffusion of

Pt-containing nuclei to form larger Pt islands which become catalytically active

125ºC150ºC175ºC200ºC250ºC300ºC325ºC

(b)

(a)

Page 43: Atomic layer deposition of ruthenium films : properties and surface

3. CpRu(CO)2Et AS THE METAL PRECURSOR DURING Ru ALD

36

once a certain island size is reached. Once these Pt islands are able to break

down the organic ligands, also the O2 molecules are dissociated on these is-

lands and the reactive O-atoms combust the organic surface species. If the

same nucleation mechanism holds for Ru ALD, the presence of catalytically ac-

tive Ru-islands and reactive O-atoms on the Si surface are both limiting factors

to Ru’s initial growth.

Thermal stability of CpRu(CO)2Et during usage (shelf-life time)

In the last five years, myriads of new Ru precursors have been developed for ALD of Ru.2 Most of these precursors are heteroleptic, i.e. made by combin-ing highly reactive ligands with other specific ligands to stabilize the precursor. The cyclopentadienyl ligand was the one used most widely. Therefore, the CpRu(CO)2Et precursor developed by SAFC Hitech® appeared promising in terms of both, reactivity and stability requirements. Our initial depositions of Ru ALD with this compound delivered Ru films with satisfactory properties, as will be pointed out in Chapter 4. However, we also observed challenges concern-ing the shelf-life time of the precursor. It appeared that the precursor starts to decompose when stored at temperatures ≤ 60 ºC over a certain period of time; in our case weeks or months depending on the temperature. Nevertheless, thanks to this result SAFC Hitech® could revisit the choice of Ru precursor for ALD in terms of its suitability to industrial standards. Yet, this thesis reports pre-dominantly results of Ru ALD films produced with this precursor and in this Section we provide support that these films were deposited with CpRu(CO)2Et that was not decomposed.

First of all, when CpRu(CO)2Et is “completely” decomposed it has turned in-to a brown-reddish solid powder (same as when heated to 325 ºC, shown in Fig. 3a) from which no Ru growth can be achieved; at least under the standard experimental conditions. A dedicated study using TGA and nuclear magnetic resonance experiments was performed by SAFC Hitech® and the results are shown in Appendix A.2. The exact decomposition pathway could not be eluci-dated, but from this temperature-dependent study it seems likely that the pre-cursor decomposition involves the loss of the CO ligand. Moreover, the Et lig-and could also play a role for the change of the decomposed compound to react with the surface under ALD conditions.

Page 44: Atomic layer deposition of ruthenium films : properties and surface

3. CpRu(CO)2Et AS THE METAL PRECURSOR DURING Ru ALD

37

Our empirical observations suggest that there is an intermediate regime between the onset of decomposition and the “complete” decomposition. This intermediate regime might last for a different period of time, as it probably de-pends on the amount of precursor present in a bubbler at once and on the temperature it is stored and used at. While the “completely” decomposed pre-cursor yields (almost) no film growth, the precursor at the start of its decompo-sition can still deliver Ru film growth. However, Ru films grown in this interme-diate regime might be detrimental to their film properties. Therefore, in situ monitoring is necessary to assess whether the precursor has started to decom-pose during an ALD process.

Figure 4: Electron impact fragmentation pattern of CpRu(CO)2Et after prolonged stor-age (m/z = 1-100) at 40 eV measured by QMS during two consecutive mass scans. The fragmentation pattern for RuCp2 and the ethyl ligand from the NIST database are shown for comparison purposes, and were normalized to m/z = 28 for the ethyl ligand and to m/z = 39 for RuCp2. The highlighted groups of m/z values are assigned to known species part of the parent molecule.

Quadrupole mass spectrometry (QMS) enabled us to monitor the thermal

stability of the compound when it was heated over a significant period of time. The optimum conditions to obtain Ru growth were found for a bubbler tem-perature of 90 ºC. Thus the data were collected at that temperature and in the same way as for the data shown in Fig. 1. The fragmentation pattern of CpRu(CO)2Et during this unstable transition regime is shown in Fig. 4. The most

m/z

Page 45: Atomic layer deposition of ruthenium films : properties and surface

3. CpRu(CO)2Et AS THE METAL PRECURSOR DURING Ru ALD

38

striking change in cracking pattern between Fig. 1 and Fig. 4 can be made out for the CO- and Et ligands. This suggests that these ligands might be responsi-ble for the lower reactivity toward the Ru surface.

Recording a cracking pattern of a precursor consumes not only time but al-so precursor, since multiple injections are required. Furthermore, the onset of decomposition might occur during a Ru deposition of several hundreds of ALD cycles. Therefore, it is also important to assess whether the signal of the m/z

Figure 5: m/z = 2, 14, 18, 28 and 44 signal intensity measured by QMS during (a) a typical CpRu(CO)2Et dose only (no ALD) when the precursor is “intact”, (b) a typical CpRu(CO)2Et dose during ALD (O2 gas is pulsed into the deposition chamber during the second ALD half-cycle) when the precursor is “intact” and (c) a typical CpRu(CO)2Et dose during ALD when CpRu(CO)2Et has started to decompose.

used to monitor the Ru ALD process is affected by the decomposition process of the precursor. Figure 5 shows the QMS signals at m/z = 2, 14, 18, 28 and 44 during “unaffected” CpRu(CO)2Et pulses only (Fig. 5a) that are due to precursor fragmentation in the QMS ionizer. These signals serve as a reference level for the signals during the first half-cycle of Ru ALD (Fig. 5b and c, for example). Note that Fig. 5b and Fig. 5c are part of a Ru deposition run of a few hundred ALD cycles, and each deposition started like the one in Fig. 5b. In the case of Fig. 5c, at some point during the deposition the m/z-signals started to change

m/z = 28

m/z = 2

m/z = 44

m/z = 18

m/z = 14

no ALD

m/z = 28

m/z = 2

m/z = 44

m/z = 18

m/z = 14

m/z = 28m/z = 2

m/z = 44

m/z = 18

m/z = 14

ALD, 1st half-cycle ALD, 1st half-cycle

precursor intact (a) precursor intact (b) precursor decomposing (c)

Page 46: Atomic layer deposition of ruthenium films : properties and surface

3. CpRu(CO)2Et AS THE METAL PRECURSOR DURING Ru ALD

39

gradually, until it became the one depicted. The most striking difference in sig-nal during an ALD process with “intact” (Fig. 5b) and “decomposing” (Fig. 5c) precursor is the one at m/z = 14 (CH2

+). This signal was attributed to an ioniza-tion product of the ethyl (Et) ligand and its QMS signal during an ALD process helps to identify the thermal stability of the precursor. During a typical ALD cy-cle (Fig. 5b) the signal at m/z = 14 shows a very short increase and then drops rapidly, indicating that the Et ligand of incoming precursor reacts with the Ru surface. On the other hand, when the precursor has started to decompose (Fig. 5c) it rises and settles to a level similar to the reference level (Fig. 5a), sug-gesting that the Et ligand no longer reacts with the Ru surface.

Conclusion

As a conclusion from the data, it is evident that the CpRu(CO)2Et com-pound is not thermally stable when heated over a prolonged period of time, even at moderate temperatures, and the decomposition pathway probably involves the loss of the CO ligand. It seems that the decomposition process can take a significant amount of time. When the precursor is completely decom-posed, no Ru growth was achieved under the standard ALD conditions. It was therefore crucial to monitor the onset of decomposition of the precursor. We have been able to monitor the state of the precursor using QMS, prior to a deposition by recording a fragmentation pattern of the precursor, and during Ru ALD by monitoring m/z = 14, specifically. Using this procedure, we were able to differentiate between Ru films deposited using intact CpRu(CO)2Et and those deposited with decomposed precursor. Only the films and ALD process-es using intact CpRu(CO)2Et were checked for reproducibility and investigated for this thesis.

Page 47: Atomic layer deposition of ruthenium films : properties and surface

3. CpRu(CO)2Et AS THE METAL PRECURSOR DURING Ru ALD

40

Appendix A A.1. Detailed overview of the Ru precursors used in ALD (based on

Hämäläinen et al.)2. The hyperscripts **, #, ‡, * link a particular process proper-ty to the precursor with the corresponding hyperscript.

group of

metal precursor

Tvap(oC) reactant Tdep (oC) growth rate

(Å cycle‒1)

metallocenes

(RuCp2, Ru(EtCp)2, Ru(EtCp)2**,

(EtCp)Ru(MeCp),

(Me2NEtCp)RuCp #,

(EtCp)Ru(DMPD),

(EtCp)Ru(DMPD) in ECH*,

Ru(DMPD)2, (EtCp)Ru(Py),

(MeCp)Ru(Py), Ru(Me2Py)2,

Ru(Cp)(CO)2Et

50-90,

200,230*

O2 (mostly),

O2 (air), O3

210-230

240-250

270-280

300

325

350-375

400-500

0.2-0.9

0.2-0.5, 1.1**

0.5-1.5

0.3-1.2, 1.8**

0.2 ‒ 1.0

0.4-0.5

0.8#

β-diketonates

(Ru(thd)3‡, Ru(thd)3 in ECH,

Ru(od)3 ** in n-butylacetate,

(iPr-Me-Be)Ru(CHD), (Et-

Be)Ru(CHD), (Et-Be)Ru(Et-CHD),

Ru(Me-Me2-CHD)2 *,

Ru(CO)3(CHD) #, Ru(tBu-Me-

amd)2(CO)2 #

100-140,

60 *,

200 **

O2, O2 (air) ‡,

NH3#

185

200, 225

250-275

325

350-400

400-450

0.6

0.1-1.0, 3.5-4#

0.2-0.9

0.2-1.0

0.3-1.5

0.5, 1.8-2**

0-oxidation state

(ToRuS)

25 H2 100-175

200-250

1.3

1.4-1.5

Page 48: Atomic layer deposition of ruthenium films : properties and surface

3. CpRu(CO)2Et AS THE METAL PRECURSOR DURING Ru ALD

41

Figure 6: Ruthenium precursors reported for Ru ALD processes, reproduced from

Hämäläinen et al.2 A.2. Thermogravimetric analysis of CpRu(CO)2Et

A dedicated study using TGA experiments were performed by SAFC Hitech® and results are shown in Figure 7. Reference TGA measurements of CpRu(CO)2Et were performed after synthesis, and after 1, 3 and 7 weeks of thermal treatment at 60 ºC. The TGA curves show that the residual mass in-creases with the duration of the thermal treatment. Figure 7a shows a residual mass of ~ 0% after synthesis, whereas after being heated at 60 ºC for 3 weeks it

Page 49: Atomic layer deposition of ruthenium films : properties and surface

3. CpRu(CO)2Et AS THE METAL PRECURSOR DURING Ru ALD

42

Figure 7: Thermogravimetric analysis as a function of temperature for CpRu(CO)2Et (a)

directly after synthesis, (b) after heating during 3 weeks at 60 ºC and (c) after thermal

treatment of 7 weeks at 60 ºC. The insets are photographs of the corresponding sam-

ples. (Courtesy of SAFC Hitech®)

increased to 0.4% and after 7 weeks at 60 ºC it reached 1.3%. This clearly indi-cates the onset of decomposition. The photographs shown in the insets show

after synthesis (a)

after 3 weeks at 60 ºC (b)

after 7 weeks at 60 ºC (c)

Page 50: Atomic layer deposition of ruthenium films : properties and surface

3. CpRu(CO)2Et AS THE METAL PRECURSOR DURING Ru ALD

43

the darkening of the compound over time, when heated. This was also shown in Fig. 3a. Moreover, the mass loss at ~190 ºC (Fig. 7) also increases over time which was attributed to the loss of the CO ligand (see Experimental section of Chapter 4). This could suggest that the decomposition pathway occurs via the loss of this ligand.

References

(1) Hatanpää, T.; Ritala, M.; Leskelä, M. Coord. Chem. Rev. 2013, 257, 3297‒3322.

(2) Hämäläinen, J.; Ritala, M.; Leskelä, M. Chem. Mater. 2013, 26, 786‒801.

(3) Hong, T. E.; Choi, S.-H.; Yeo, S.; Park, J.-Y.; Kim, S.-H.; Cheon, T.; Kim, H.; Kim, M.-K.; Kim, H. ECS J. Solid-State Sci. Technol. 2013, 2 , P47‒P53.

(4) Eom, T. K.; Sari, W.; Choi, K. J.; Shin, W. C.; Kim, J. H.; Lee, D. J.; Kim, K. B.; Sohn, H.; Kim, S. H. Electrochem. Solid-State Lett. 2009, 12, D85‒D88.

(5) Knoops, H. C. M.; Langereis, E.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Vac. Sci. Technol. A 12AD, 30, 01A101.

(6) De Laeter, J. R.; Bohlke, J. K.; De Bievre, P.; Hidaka, H.; Peiser, H. S.; Rosman, K. J. R.; Taylor, P. D. P. Pure Appl. Chem. 2003, 75, 683‒800.

(7) Hwang, W.; Kim, Y. K.; Rudd, M. E. J. Chem. Phys. 1996, 104, 2956‒2966.

(8) NIST Chemistry WebBook,. NIST Standard Reference Database Number 69.

(9) Robert H. Crabtree (Yale University, New Haven, C.) The organometallic chemistry of the transition metals; Fourth Edi.; John Wiley & Sons, Inc., Publication: Hoboken, New Jersey, 2005; p. 16.

(10) Haukka, S. Personal communication 2010.

(11) Mackus, A. J. M.; Verheijen, M. A.; Leick, N.; Bol, A. A.; Kessels, W. M. M. Chem. Mater. 2013, 25, 1905‒1911.

Page 51: Atomic layer deposition of ruthenium films : properties and surface

3. CpRu(CO)2Et AS THE METAL PRECURSOR DURING Ru ALD

44

Page 52: Atomic layer deposition of ruthenium films : properties and surface

45

Chapter 4 “By believing passionately in something that still does not exist, we create it. The non-

existent is whatever we have not sufficiently desired.” ‒ Franz Kafka

Atomic layer deposition of ruthenium from CpRu(CO)2Et using O2 gas and O2 plasma* Abstract

The organometallic precursor cyclopentadienylethyl(dicarbonyl)ruthenium [CpRu(CO)2Et] was used to develop an atomic layer deposition (ALD) process for ruthenium. O2 gas and O2 plasma were employed as reactants. For both processes, thermal and plasma-assisted ALD, a relatively high growth per cycle of ~1 Å was obtained. The Ru films were dense and polycrystalline, regardless of the reactant, yielding a resistivity of ~16 μΩ.cm. The O2 plasma enhanced the Ru nucleation on the TiN substrates but also led to an increased roughness compared to thermal ALD.

* This chapter was published as Leick, N.; Verkuijlen, R. O. F.; Lamagna, L.; Langereis, E.;Rushworth, S.; Roozeboom, F.; van de Sanden, M. C. M.; Kessels, W. M. M. J. Vac. Sci. Technol. A 2011, 29, 021016.

Page 53: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

46

Introduction

High density metal-insulator-metal (MIM) capacitors are required for next generation dynamic random access memories (DRAM) with equivalent oxide thickness (EOT) values of 0.35 nm.1,2 MIM-capacitors integrated in silicon have also emerged as feasible candidates to reach the capacitance densities >5 nF/mm2 required for automotive and decoupling applications.3,4 To meet the technical specifications strontium titanate SrTiO3 (STO) is currently one of the most investigated dielectric materials for these capacitor applications.2 Crystalline STO has an ultra-high dielectric constant (theoretically, κ≥ 300)5 at room temperature and amorphous films have a low crystallization temperature (T ~ 500-600 ºC).6 As electrode material in the MIM capacitors noble metals such as Pt and Ru are preferentially considered.7,8 In addition to a high work function, these metals have a sufficiently low resistivity such that ultra-thin films can be employed. Moreover, these noble metals seem able to minimize the device’s leakage current,9 and they appear to have a better chemical com-patibility with STO compared to more conventional electrode materials such as TiN. Ru, having a bulk resistivity of 7.1 μΩ.cm and a work function of 4.7 eV, is preferred from a device-integration perspective as it can be dry etched relative-ly easily, unlike platinum (Pt).2,10 Ru is also chemically stable towards oxygen, forming a stable, conductive oxide.

The work described in this Chapter focuses on atomic layer deposition (ALD) of Ru, which is a technique able to fulfil the requirements of ultra-thin films (thickness ~5 nm) and conformality for the high-aspect ratio structures to be employed (electrode conformality >70% for an aspect ratio of ~30).3,4 For STO, the method has already shown promise to keep a low leakage current (~1×10-8 A/cm2) and a low EOT value (1.5 nm).11 In the past years, numerous precursors with organic ligands, such as Ru(thd)3,12 Ru(acac)3,13,14 RuCp2 13,15‒18 and Ru(EtCp)2 18‒20 have been used with O2 gas for Ru ALD. Employing RuCp2, Aaltonen et al. showed that these processes rely on the ability of Ru to dissocia-tively chemisorb O2, providing atomic O necessary for the oxidation of the pre-cursor ligands.21 However, due to their relatively long nucleation delays (70-250 cycles)15,22,23 and relatively low growth rates (0.2-0.8 Å/cycle)13,15,18,20,24,25, al-ternatives to the aforementioned ALD processes of Ru have been considered. To address these issues the choice of both, the Ru precursor and the co-

Page 54: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

47

reactant, have been re-evaluated. For example, the nucleation delay has been successfully addressed using NH3 plasmas18,20,25 which easily initiated the Ru growth and delivered smooth films. However, the growth per cycle remained low (0.25-0.8 Å/cycle) 18,20,25 and NH3 is also likely to chemically affect the STO during its integration in MIM capacitors.26‒29 Regarding Ru precursors, consider-able attention has been devoted to new compounds which have an improved reactivity and volatility. In light of this, several precursors, such as 2,4-(dimethylpentadienyl)-(ethylcyclopentadienyl)ruthenium (DER),22 RuO4,30 (η6-1-isopropyl-4-methyl-benzene)(η4-cyclohexa-1,3-diene)ruthenium (C16H22Ru), 31 1-ethyl-1’-methyl-ruthenocene (EMR),32 or bis(N,N’-di-tert-butylacetamidinato) ruthenium(II)dicarbonyl (Ru(tBu-Me-amd)2(CO)2),33 have become available for Ru ALD.

In this Chapter, ALD of Ru was investigated using the recently developed precursor cyclopentadienylethyl(dicarbonyl)ruthenium (CpRu(CO)2Et).34 O2 gas, as well as an O2 plasma, was employed as the co-reactant. Providing O radicals directly from the gas phase, an O2 plasma might facilitate nucleation of metals by ALD as reported by Knoops et al.35 in the case of Pt. Moreover, an O2 plasma is expected to be compatible with the STO, as mentioned before.

Experimental details

CpRu(CO)2Et (SAFC Hitech®, purity > 99.999%) is a heteroleptic precursor that is liquid at room temperature, enhancing its evaporation over solid precur-sors. Figure 1a shows its high vapor pressure (0.4 Torr at 60 ºC) compared to several other Ru precursors. Results from thermogravimetric analysis (TGA) in Fig. 1b show a first weight loss due to volatility and a feature at T ~ 190 ºC due to decomposition. This could indicate the loss of the carbonyl groups since the same feature was observed for the related precursor CpRu(CO)2Me (Me = -CH3). At higher temperatures (200-250ºC) a further weight loss took place, leading to a constant mass of 1.2%. During deposition, the precursor was heated to 90 ºC and the precursor lines to 110 ºC to prevent precursor condensation during transport to the deposition chamber. The precursor was vapor-drawn into the reactor. Additional insight into the thermal decomposition behaviour of CpRu(CO)2Et was shown in Chapter 3.

Page 55: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

48

Figure 1: (a) Vapor pressure of cyclopentadienylethyl(dicarbonyl)ruthenium, compared with other heteroleptic metalorganic Ru precursors, and its (b) thermogravimetric analy-sis as a function of temperature. Inset: molecular structure of CpRu(CO)2Et.

The Ru depositions were carried out in an Oxford Instruments FlexALTM re-

actor shown schematically and physically in Figure 2.36 This setup consists of a deposition chamber connected to a turbo pump (base pressure of 10-6 Torr) and a remote inductively coupled plasma source allowing a maximum power of 500 W. The O2 gas (purity > 99.999%) was delivered to the reactor through the plasma source which was only switched on for the plasma-assisted ALD experiments. The pressure inside the deposition chamber was kept at 30 mTorr during each oxidant exposure. The temperature of the substrate carrier was set

30 40 50 60 70 801

10

100

1000

Temperature (oC)

Vap

or P

ress

ure

(mTo

rr)

CpRu(CO)2Et

(MeCp)Ru(CO)2Me

CpRu(CO)2Me

(iPrCp)2Ru

(MeCp)2Ru

(tBuCp)2Ru

Cp2Ru

(a)

50 100 150 200 250 300 350 400 4500

10

20

30

40

50

60

70

80

90

100

Wei

ght (

%)

1.2 %

(b)

Temperature (oC)

(EtCp)2Ru

RuCO

COH3CH2C

Page 56: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

49

to 400 ºC which corresponds to a wafer temperature ≲325 ºC, as deduced from calibration measurements shown in Appendix A.1.

Figure 2: (a) Schematic overview of the FlexALTM tool with the in situ diagnostics used in this thesis: in situ spectroscopic ellipsometry and quadrupole mass spectrometry. (b) Photograph of the FlexALTM tool situated in the “NanoLab@TU/e” cleanroom facilities of the TU/e. The spectroscopic ellipsometer can also be seen.

The Ru films were deposited on Si wafers with native oxide, subsequently covered by a ~8 nm thick TiN film because Ru/TiN is expected to be a more cost-effective electrode stack than Ru only. The TiN film was deposited by plasma-assisted ALD in the same FlexAL reactor and at the same substrate temperature, using the process based on TiCl4 and H2-N2 plasma developed by Knoops et al..37

The Ru film thickness and mass density were extracted from a Philips X’Pert MPD X-ray reflectometer (XRR) and the crystal structure from a grazing inci-dence X-ray diffractometer (GI-XRD). Rutherford backscattering spectroscopy (RBS) using a 2 MeV 4He+ beam at the singletron facility of the Eindhoven Uni-versity of Technology was employed to measure the areal densities of the Ru films. From the thickness values obtained by XRR, the mass density of the films was also extracted from the RBS measurements. The sheet resistance of the Ru films was measured at room temperature by a four-point probe (FPP). Time of flight secondary ion mass spectrometry (ToF-SIMS) was performed by a dual

(a) (b)

Page 57: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

50

beam ION-TOF IV instrument using Cs+ ions of 1 keV for sputtering and Ga+ ions of 25 keV for analysis. The roughness of the Ru films was investigated by atomic force microscopy (AFM) using a Veeco Dimension 3100 scanning probe micro-scope by scanning areas of 2 × 2 μm2 on two different locations on each sam-ple. The nucleation of the Ru films on TiN was studied by transmission electron microscopy (TEM) using the FEI Tecnai F30ST microscope operated at 300 kV. The measurements were performed on silicon nitride TEM membranes, on which a ~8 nm TiN layer was deposited by the same plasma-assisted ALD process as for the Si wafers.

Results and discussion

Ru films were deposited for various numbers of cycles by thermal and plasma-assisted ALD. Figure 2 shows the results obtained from XRR, RBS, FPP and AFM analysis as a function of the number of cycles. The unit cycle of the two ALD processes consisted of the following steps: CpRu(CO)2Et dose ‒ Ar purge ‒ O2 dose ‒ O2 plasma excitation ‒ Ar purge with the timings of 2 ‒ 5 ‒ 5 ‒ 0 ‒ 5 s for thermal ALD and 2 ‒ 5 ‒ 2 ‒ 0.8 ‒ 5 s for plasma-assisted ALD. For the plasma-based process the O2 dosing was inserted to stabilize the gas pres-sure in the reactor before igniting the plasma. The plasma exposure chosen was short (800 ms) to make sure that Ru was deposited, since for longer plasma exposures the films tended to become RuO2. Furthermore, when operating the plasma source at relatively high powers, no film growth was observed. There-fore, a plasma power of 100 W was chosen. It is likely that at higher powers, the flux of reactive oxygen species is of such magnitude that etching of the Ru film38 is dominating over deposition. This was also corroborated by a brief ex-periment in which Ru films were exposed to O2 plasma and after few (<10 s) seconds 0.3-0.5 nm had been etched off the films.

The aforementioned cycle timings resulted in saturated ALD conditions as verified by increasing the precursor dose time. We also note that the substrate temperature (wafer temperature is ~325 ºC) appears to be relatively high for the CpRu(CO)2Et precursor having carbonyl groups (see Chapter 3). Our results also showed good agreement with the results obtained internally by SAFC Hi-tech® and those recently reported by Park et al.34 (both shown in Appendix A.2) and at lower temperatures we observed reduced growth rates for thermal ALD.

Page 58: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

51

Also no growth occurred when only the precursor was dosed into the reactor (heated to 325 ºC) without oxidant step and for the ALD conditions no sign of CVD-like growth was observed from film uniformity measurements (uniformity of thickness and sheet resistance < 4% on 100 mm wafers). Investigations of the cracking pattern by quadrupole mass spectrometry did not reveal thermal decomposition products either. However, it still appears that the wafer tem-perature is such that the precursor is close to the onset of decomposition. Pre-liminary infrared studies monitoring the initial reaction of the precursor on high surface area silica, shown in Chapter 3, revealed that the precursor starts de-composing at 325 ºC.34,39

Figure 3a shows that the Ru thickness increase measured by XRR is linear with the number of cycles for both ALD processes. The growth per cycle has been determined from a linear fit to the data. Although within the experimen-tal error, the growth per cycle of the plasma-based process of 1.1 Å seems to be slightly higher than the growth per cycle of 1.0 Å for thermal ALD. The latter is also confirmed by the RBS results in Fig. 3b from which the growth per cycle, in terms of Ru atoms deposited, can be determined. For thermal ALD a value of (7.1±0.5)1014 at/cm2.cycle can be deduced whereas for plasma-assisted ALD it is (7.6±0.1)1014 at/cm2.cycle. The slightly higher growth per cycle for the plasma-assisted process can possibly be attributed to the higher surface roughness of the films compared to those prepared by thermal ALD, as will be discussed below.

The Ru growth per cycle of ~1 Å for the CpRu(CO)2Et precursor is relatively high. Making the crude assumption of a Ru(0001) crystal structure (XRD results are discussed below), this growth rate corresponds roughly to ~0.5 ML/cycle. This is about twice as high as the growth obtained with other metalorganic precursors, such as Ru(EtCp)2, DER or EMR.22,25,32 The growth per cycle is, how-ever, quite similar to the one of a new Ru ALD process recently reported by Eom et al..31 Using the open-ring compound C16H22Ru a similarly high growth per cycle of 0.9 Å was obtained at 220 ºC.

From XRR the mass density of the films was determined which suffered from a relatively large uncertainty (much more than the XRR thickness) due to a relatively large surface roughness, and values of 11±1 g/cm3 and 10±1 g/cm3 could be estimated for the thermal and plasma-assisted ALD films, respectively.

Page 59: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

52

Figure 3: (a) Thickness determined by XRR as a function of the number of ALD cycles, (b) Ru areal density determined by RBS, (c) Ru resistivity measured by FPP and (d) the root-mean-square (RMS) roughness of the Ru films for plasma-assisted and thermal ALD. The growth per cycle in (a, Å/cycle) and (b, at/cm2.cycle) are given by the slope of the fitted lines and the nucleation delay by the extrapolation of the fits to zero thickness.

Thic

knes

s (n

m)

Are

al d

ensi

ty (1

016 a

t/cm

2)

0

5

10

15

20

0

5

10

15

20

25

30

350 50 100 150 200 250 300

(b) RBS

(a) XRR

1.1±0.1 1.1±0.1 Å/cycle

/cycle

1.0±0.1 1.0±0.1 Å/cycle

/cycle

(7.6 ±0.1)x10

(7.6 ±0.1)x101414 at.cm at.cm

-2-2 /cycle/cycle

(7.1 ±0.5)x10

(7.1 ±0.5)x101414 at.cm at.cm

-2-2 /cycle/cycle

FPP

resi

stiv

ity

( μΩ. c

m)

0

2

4

6

8

10

12 (d) AFM

RM

S ro

ug

hn

ess

(nm

)

Number of ALD cycles

0

5

10

15

20

25

30

(c) FPP

Ru - bulk resistivity

0 50 100 150 200 250 300

plasma-assisted ALDthermal ALD

Page 60: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

53

From the combination of RBS results and thickness from XRR, more accurate values yielded a mass density of 12.2±0.5 g/cm3 for thermal ALD and 11.7±0.5 g/cm3 for plasma-assisted ALD. Despite the results obtained for both ALD processes being in agreement within the error, both measurement meth-ods show a lower mass density for the films prepared by plasma-assisted ALD. The difference might be explained by the presence of low levels of O in the films prepared by plasma-assisted ALD as will be discussed below. The mass densities of the films are lower than the bulk density (12.4 g.cm-3) of Ru but within the same range of previously reported ALD films (8.7-12.0 g/cm3).24,25,32

The resistivity of the films was calculated from the sheet resistance and the XRR thickness and is shown in Fig. 3c as a function of ALD cycles. The resistivity decreased with thickness and the lowest value of 16 μΩ.cm is reached for the thickest films (25-30 nm). This decrease is most likely related to the increasing grain size with thickness, decreasing the influence of grain boundary scattering. The resistivity is apparently independent of the oxidant and the value of 16 μΩ.cm is in the range (12-35 μΩ.cm) of other Ru films deposited by ALD.13,15,16,19,20 Moreover, 5 nm Ru films reach a resistivity < 30 μΩ.cm, fulfilling the resistivity specifications (< 300 μΩ.cm for 5 nm) imposed on MIM capaci-tor electrodes.40

Figure 4 shows GI-XRD θ-2θ scans of Ru films prepared by 300 cycles of thermal and plasma-assisted ALD. For the films prepared by both ALD methods, several diffraction peaks are observed from which it can be concluded that the films are polycrystalline with a hexagonal closed packed structure. The peaks for the plasma-assisted ALD Ru film have a higher intensity due to the higher thickness of that film. Similar polycrystalline films have been obtained by sev-eral other Ru ALD processes using precursors with organic ligands and O2 gas.13,32,41 Since the relative peak intensities resemble those of the reference spectrum of a powder sample the crystallites in the Ru layer have no dominant crystallographic orientation, although it cannot be excluded that the film pre-pared by plasma-assisted ALD might be slightly preferentially oriented in the (002) and (101) planes. In a comparison between Ru ALD with O2 gas and NH3 plasma, Kwon et al.25 also observed a (002) preferential orientation when a NH3 plasma was used. The (002) orientation allows a more compact configuration

Page 61: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

54

Figure 4. Grazing incidence XRD spectra of a 35 nm and of a 22 nm thick Ru film depos-ited by plasma-assisted and thermal ALD, respectively. The spectra are offset vertically for clarity. For comparison, the diffraction pattern of a Ru powder sample is also shown.

of the Ru atoms than the other orientations and it is therefore possible that the crystallographic arrangement of the Ru atoms is influenced by additional en-ergy delivered to the substrate by the plasma. The peak positions are the same for both samples, indicating no difference in lattice constants which could be caused by differences in stress or composition. Furthermore, the peak widths of the plasma-assisted ALD film are slightly narrower than those of the thermal ALD film, indicating a slightly larger crystallite size or smaller crystal defect con-centration for the film prepared by plasma-assisted ALD. This can possibly again be attributed to the higher energy flux to the substrate during the plas-ma exposure.

For both thermal and plasma-assisted ALD, the XRD spectra only revealed crystalline peaks related to Ru and no peaks from RuO2 could be identified, in contrast to Park et al. who used the same precursor.34 The presence of O in the films was examined by ToF-SIMS. As shown in Fig. 5 for the plasma-assisted ALD film the RuO+ signal is more pronounced than for thermal ALD. Moreover TiO+ is detected throughout the TiN film, as opposed to thermal ALD for which TiO+ is predominantly detected at the Ru/TiN and TiN/SiO2 interfaces. Although no evidence for RuO2 was provided by XRD and RBS (within its rather high de-tection limit of < 20% for O), more oxygen was detected by ToF-SIMS in the

30 40 50 60 70 80 90

plasma-assisted ALD

thermal ALD

Inte

nsity

(a.u

.)

Angle (2θ)

(100)(002)

(101)

(102) (110) (103)(200)

(112)

Page 62: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

55

films deposited by the O2 plasma, which is a stronger oxidant. This higher oxy-gen content is also a likely explanation for the slightly lower mass density of the Ru films deposited by plasma-assisted ALD.

Figure 5: ToF-SIMS depth profiles of a 35 nm and of a 22 nm thick Ru film deposited by (a) plasma-assisted and (b) thermal ALD, respectively.

XPS results on the thermal ALD film corroborated that the Ru was in the metallic state with only a small amount of oxygen present at the top surface of the film. The TiN was oxidized at the Ru interface but this could be caused by the oxidant step during ALD as well as by oxidation of the TiN when the film was exposed to ambient air after the TiN deposition. By XPS only adventitious carbon at the top surface was detected, i.e., no significant amount of C in the film could be observed.

The Ru film properties discussed so far were quite similar for thermal and plasma-assisted ALD. However, larger differences were observed in terms of film nucleation and surface roughness. As shown in Fig. 3a and 3b, the O2 plasma leads to faster nucleation of the Ru film than thermal ALD. Extrapolating the thickness and areal density data reveals that film nucleation takes place after ~45 cycles for plasma-assisted ALD whereas it is delayed to ~85 cycles for thermal ALD. As mentioned in the Introduction faster nucleation compared to thermal ALD was observed earlier for Ru films deposited by plasma-assisted ALD using a NH3 plasma as the co-reactant.18 Also in a comparison between O2

Page 63: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

56

gas and O2 plasma as oxidants for ALD of Pt, a much faster nucleation was ob-served for the plasma-based process.35 The higher reactivity of the plasma, pro-viding reactive oxygen atoms from the gas phase, apparently promotes the nucleation of the Ru, potentially also through oxidation of the TiN surface form-ing TiO2-like surface species.7 However, as opposed to plasma-assisted ALD of Pt, a significant nucleation delay is observed in the case of Ru ALD. This could be related to the etching of Ru atoms by the O2 plasma releasing volatile RuOx species which can compete with the initial growth. More research is necessary to substantiate these conclusions. Furthermore, Fig. 3d shows that the Ru films deposited by plasma-assisted ALD have a significantly higher roughness than the films prepared by thermal ALD. Moreover, all Ru films are relatively rough and the roughness increased considerably with the number of ALD cycles. The AFM scan data, shown in Fig. 6 for two almost equivalently thick samples (16.3 and 14.8 nm for plasma-assisted and thermal ALD, respectively) reveal a surface morphology consisting of densely packed Ru agglomerates. When O2 gas was used, the agglomerates observed in the AFM scan have similar heights whereas the film deposited using an O2 plasma consists of several dominant agglomer-ates having different heights and which are randomly distributed over the sur-face. The root-mean-square (RMS) roughness values of these ALD films, 5 nm and 9.6 nm for the thermal and plasma-assisted ALD films respectively, are very high for ALD films compared to the surface roughness values reported so far.

Figure 6: AFM scans (2 × 2 μm2) for plasma-assisted and thermal ALD Ru films with a thickness of (a) 16.3 nm and (b) 14.8 nm, respectively. Note the difference in height scale.

(a)(b)

Page 64: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

57

For thermal ALD of Ru, RMS surface roughness values are typically within the range of 0.3-3.8 nm (for film thicknesses ranging from 5-80 nm)19,32,33,42 while for plasma-assisted ALD films even values as low as 0.7 nm have been reported for 50 nm thick films.25 However, the surface roughness seems to be correlated to the substrate temperature and the surrounding ambient.13,19,32,33 Kukli et al. reported high roughness values (1.7 nm for a 5 nm thin film) for Ru prepared from EMR at high substrate temperatures (T ~ 400 ºC)32 and Park et al.34 reported a surface roughness of 1.4 nm for a film of only ~3.1 nm thickness prepared from CpRu(CO)2Et at 300 ºC. Furthermore, Kwon et al.19 dedicated an experiment to heating smooth Ru films (3.82 nm roughness for 80 nm film thickness) at 600 ºC and 750 ºC in O2 and Ar atmosphere. They observed a clear surface roughening using O2 depending on temperature: 27.3 nm and 42.3 nm surface roughness at 600 ºC and 750 ºC, respectively. On the other hand, an-nealing the same film at 750 ºC in an Ar atmosphere, barely affected the surface roughness (3.94 nm).

Figure 7: TEM images of (a) a TiN-covered Si3N4-membrane, and of such membranes on which 50 and 100 cycles thermal (b and c) and plasma-assisted ALD (d and e) of Ru were carried out. The estimated thickness of the Ru layer is indicated in white in the corre-sponding image. The arrows indicate large Ru islands in (e).

The results from the TEM measurements on the TiN-covered membranes are shown in Fig. 7. On these membranes 50 and 100 cycles thermal and plas-ma-assisted ALD was carried out. Due to the polycrystalline nature of the 8 nm

TiN Thermal Plasma

2 nm 5 nm

2 nm0 nm

100 nm 100 nm 100 nm

100 nm100 nm

(a) (b)

(c)

(d)

(e)

Page 65: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

58

thick TiN layer the contrast is poor and no clear differences can be observed for the two depositions with thermal ALD. However differences with the TiN-covered membrane and the thermal ALD results are clearer for the plasma-assisted ALD experiments after 50 and 100 cycles. Already after 50 cycles of plasma-assisted ALD, Ru islands can be observed and after 100 cycles it is clear that several large islands are present. The TEM images corroborate therefore the thickness and AFM results in that the nucleation of the plasma-assisted ALD films is enhanced and that the roughness increases with thickness through the development of large islands.

A short nucleation delay and a low surface roughness of films prepared by ALD usually indicate a uniform film growth; whereas an enhanced film rough-ness and pronounced nucleation delay is characteristic for island-like growth.7,18,43,44 Interestingly, unlike with a NH3 plasma,18 it turns out that the length of the nucleation period of our films prepared by the O2 plasma based ALD process is reduced whereas the island-like growth is more pronounced. The situation is therefore more complex and it is possible that several phe-nomena are playing a role simultaneously. The lower film roughness obtained with a NH3 plasma than with O2 gas has so far been attributed to the presence of ions in the plasma. For the plasma used in this work the presence of non-thermal ions (mean ion energy is ~10 eV for ions arriving at the substrate) has also been established.45 The choice of the reactant itself, however, seems to affect the film morphology as well. Indeed, Wang et al.33 reported a change in morphology (larger islands and non-uniform island size distribution) when O2 gas was used to grow Ru thin films instead of N2 gas. They also observed that a higher O2 exposure increased the Ru film roughness. In our case, the O2 plasma provides a high flux of atomic oxygen species to the surface which could trig-ger the same increase in roughness as a longer O2 gas exposure. Another phe-nomenon to take into account is ripening. A recent TEM study that we per-formed during the nucleation of Pt ALD using MeCpPtMe3 and O2 gas clearly demonstrated the important role of the O2 gas on the ripening mechanism of these islands.46 We showed that an increase in O2 pressure and/or exposure time accelerated the coarsening of the Pt islands, thereby significantly reducing the nucleation delay of the process. In the case of Ru ALD, data depicted in Ap-pendix A.3 show that an increased O2 gas pressure also leads to an accelerated Ru nucleation. The AFM and TEM data (Figs. 6 and 7) for plasma-assisted ALD

Page 66: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

59

suggest that the density of nucleation sites is larger and that the islands be-come larger than for thermal ALD. Island ripening is known to be enhanced by higher energy fluxes,47,48 as is likely in the case of plasma-assisted ALD com-pared to thermal ALD. The combination of these effects, the additional flux of atomic oxygen and the energy provided by the plasma, might enhance the film nucleation on one hand and the surface roughness on the other hand. The underlying mechanisms and the role of the Ru film roughness in the perform-ance of the metal-dielectric stack yet remain to be addressed in future studies.

Conclusions

Thermal and plasma-assisted ALD using O2 and CpRu(CO)2Et both resulted in Ru films with similar properties. The Ru films obtained were dense, polycrys-talline with a hexagonal closed packed crystal structure and had a low resistivi-ty (16 μΩ.cm). These properties make the ALD process and the Ru films suita-ble to be used as electrode layers in MIM capacitors. The merits of these new ALD processes lie in the relatively high growth (1 Å/cycle) and in the reduction of the nucleation delay down to ~ 45 ALD cycles by the O2 plasma-based pro-cess. However, despite the improved nucleation, the plasma-assisted ALD Ru films had a higher surface roughness than the thermal ALD films which appears to be correlated to the nucleation of the films during the initial growth.

Under our operating conditions, no major benefits have emerged for the plasma-assisted ALD process of Ru over the thermal process, although the plasma might be beneficial over thermal ALD at lower substrate temperatures, for example. From the perspective of process integration, the thermal ALD pro-cess with the CpRu(CO)2Et precursor and O2 gas can deliver high-quality films in terms of material and process properties, and most likely also in terms of film conformality and performance as an electrode material in MIM capacitors. This will be addressed in a future study.

Acknowledgments

Dr. P.C.J. Graat, R. Beerends (Philips MiPlaza Materials Analysis) and Dr. M.A. Verheijen (Technical University Eindhoven) are thanked for performing the ex situ characterization (XRR, XRD, RBS, AFM and TEM) and M. Perego and C. Wie-mer (MDM) for the ToF-SIMS analysis and discussions. C. van Helvoirt and W.

Page 67: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

60

Keuning are acknowledged for their technical support throughout this study. This work is funded by the Dutch Technology Foundation STW and is carried out in cooperation with the MaxCaps project (2T210) within the European Medea+ framework.

Appendix ‒ Additional data:

A.1 Relationship between set and actual temperature of the substrate car-rier during Ru ALD

As mentioned in the Experimental section of this Chapter, the actual tem-perature of the substrate carrier deviates from the temperature set in the ALD reactor (400 ºC for the Ru process). The relationship between these tempera-ture values can be seen in Figure A.1. During this calibration, the temperature at a 4” Si wafer was measured by a thermocouple external to the FlexALTM reac-tor and after pumping down the chamber to the base pressure level of ~ 1×10-6 Torr. This procedure was repeated for two pressure values of

Figure A.1: Relationship between the set temperature and the corresponding actual temperature at the substrate (4” Si-wafer) measured with an external thermocouple. The denomination “vacuum” for the blue squares means that the measurement was performed at the base pressure of the chamber, which is ~ 1×10-6 Torr.

25 mTorr and 200 mTorr to see the effect of convective heating. The 25 mTorr value represents the partial pressure during the CpRu(CO)2Et pulse, which is ~30 mTorr. In this thesis work, the O2 pressure values used during the O2 pulse

Page 68: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

61

ranged from 30 mTorr to 180 mTorr, such that the temperature value at 200 mTorr represents a reasonable upper value for the temperature during the ALD processes using elevated O2 pressure values. From this data, we estimated the temperature at the substrate to lie ≲ 325 ºC during all of the Ru ALD pro-cesses described in this thesis.

A.2 Growth per cycle of ALD using CpRu(CO)2Et and O2 gas

Growth per cycle values around 1 Å/cycle for the Ru ALD process using the same precursor, CpRu(CO)2Et, and O2 gas were reported on two earlier occa-sions, and the reproduced data can be found in Figure A.2.

Figure A.2: Ru thickness as a function of the number of ALD cycles reported by (a) SAFC Hitech®,40 on a Ta substrate and (b) Park et al.34 on a H/Si(111) substrate.

SAFC Hitech® performed Ru depositions at 265 ºC on a Ta-substrate with a

2 s-CpRu(CO)2Et exposure followed by 2 s-air exposure. The Ru thickness in-crease due to this ALD process is shown in Fig. A.2a and was measured by X-ray fluorescence. The measurements were first calibrated using the thickness of a thick Ru sample determined by scanning electron microscopy.40 The resulting growth per cycle value of their process lies at ~ 0.94 Å/cycle. Park et al.34 meas-ured the Ru film thickness shown in Fig. A.2b by RBS and obtained a growth per cycle value of 1.3 Å/cycle from Fig. A.2b. That Ru process was conducted at 300 ºC by repeating 11 ALD cycles of a 2 s-CpRu(CO)2Et exposure followed by 1 s-O2 gas exposure.

0 2 4 6 8 10 120

2

4

6

8

10

12

14

16

Thic

knes

s (Å

)

Cycles

(b)(a)

Page 69: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

62

A.3 O2-pressure dependent nucleation delay

Ru ALD was also performed on Si substrates with native SiO2 (~2nm) at O2 pressure values of ~30, 90 and 180 mTorr. The Ru thickness as a function of ALD cycles shown in Fig. A.3 was measured by in situ ellipsometry and the reader is referred to Chapter 5 for more information on the data analysis. The data from Fig. A.4 show no measurable Ru growth even after 150 ALD cycles when the O2 pressure is 30 mTorr, whereas at 90 mTorr and above, the nucleation delay is drastically decreased. This observation was first made for Pt ALD, and further TEM experiments showed that a higher O2 exposure (pressure × time) led to a higher level of Pt island ripening. The O2 dependence of the Pt nucleation was explained by the enhanced diffusion of Pt species in the presence of oxygen, and the resulting faster aggregation of Pt atoms in metal clusters that catalyze the surface reactions of ALD growth.46

Figure A.3: Ru film thickness deposited on a Si with native oxide substrate as a func-tion of ALD cycles for various O2 gas pressure. The thickness was determined by in situ spectroscopic ellipsometry. From this data, the nucleation delay can be extracted: ~15 cycles at 180 mTorr and ~25 cycles at 90 mTorr.

In addition to the nucleation delay, the growth per cycle of Ru also increas-

es as a function of O2 pressure. The reason for this observation is in agreement with the reaction mechanism presented in the Chapters 6 and 7. Surface sci-ence studies show that several monolayers of O can be accommodated into a Ru(0001) lattice structure for similar experimental conditions to those used in

on Si with native oxide (~ 2 nm)

Page 70: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

63

this thesis. Based on these studies, a higher O coverage is achieved on the Ru surface when the O2 pressure is increased in Ru ALD. The more abundant sur-face O at the start of the CpRu(CO)2Et pulse is probably responsible for a great-er removal of precursor ligands and therefore the increased growth per cycle observed.

References

(1) ITRS www.itrs.net - Process integration, devices, and structures 2010.

(2) Kittl, J. A.; Opsomer, K.; Popovici, M.; Menou, N.; Kaczer, B.; Wang, X. P.; Adelmann, C.; Pawlak, M. A.; Tomida, K.; Rothschild, A.; Govoreanu, B.; Degraeve, R.; Schaekers, M.; Zahid, M.; Delabie, A.; Meersschaut, J.; Polspoel, W.; Clima, S.; Pourtois, G.; Knaepen, W.; Detavernier, C.; Afanas’ev, V. V; Blomberg, T.; Pierreux, D.; Swerts, J.; Fischer, P.; Maes, J. W.; Manger, D.; Vandervorst, W.; Conard, T.; Franquet, A.; Favia, P.; Bender, H.; Brijs, B.; Van Elshocht, S.; Jurczak, M.; Van Houdt, J.; Wouters, D. J. Microelectron. Eng. 2009, 86, 1789‒1795.

(3) Klootwijk, J. H.; Jinesh, K. B.; Dekkers, W.; Verhoeven, J. F.; Van den Heuvel, F. C.; Kim, H. D.; Blin, D.; Verheijen, M. A.; Weemaes, R. G. R.; Kaiser, M.; Ruigrok, J. J. M.; Roozeboom, F. IEEE Electron Device Letters 2008, 29, 740‒742.

(4) Roozeboom, F.; Kemmeren, A. L. A. M.; Verhoeven, J. F. C.; Van den Heuvel, F. C.; Klootwijk, J.; Kretschman, H.; Fri , T.; Van Grunsven, E. C. E.; Bardy, S.; Bunel, C.; Chevrie, D.; LeCornec, F.; Ledain, S.; Murray, F.; Philippe, P. Thin Solid Films 2006, 504, 391‒396.

(5) Van Elshocht, S.; Adelmann, C.; Clima, S.; Pourtois, G.; Conard, T.; Delabie, A.; Franquet, A.; Lehnen, P.; Meersschaut, J.; Menou, N.; Popovici, M.; Richard, O.; Schram, T.; Wang, X. P.; Hardy, A.; Dewulf, D.; Van Bael, M. K.; Blomberg, T.; Pierreux, D.; Swerts, J.; Maes, J. W.; Wouters, D. J.; De Gendt, S.; Kittl, J. A. J. Vac. Sci. Tech. B 2009, 27, 209‒213.

(6) Popovici, M.; Van Elshocht, S.; Menou, N.; Swerts, J.; Pierreux, D.; Delabie, A.; Brijs, B.; Conard, T.; Opsomer, K.; Maes, J. W.; Wouters, D. J.; Kittl, J. A. J. Electrochem. Soc. 2010, 157, G1‒G6.

(7) Kang, S. Y.; Hwang, C. S.; Kim, H. J. J. Electrochem. Soc. 2005, 152, C15‒C19.

Page 71: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

64

(8) Matsui, Y.; Hiratani, M.; Nabatame, T.; Shimamoto, Y.; Kimura, S. Electrochem. Solid-State Lett. 2001, 4, C9‒C12.

(9) Vallee, C.; Gonon, P.; Jorel, C.; El Kamel, F.; Mougenot, M.; Jousseaume, V. Microelectron. Eng. 2009, 86, 1774‒1776.

(10) Yunogami, T.; Kumihashi, T. Jpn. J. Appl. Phys. 1998, 37, 6934‒6938.

(11) Menou, N.; Popovici, M.; Clima, S.; Opsomer, K.; Polspoel, W.; Kaczer, B.; Rampelberg, G.; Tomida, K.; Pawlak, M. A.; Detavernier, C.; Pierreux, D.; Swerts, J.; Maes, J. W.; Manger, D.; Badylevich, M.; Afanas’ev, V. V; Conard, T.; Favia, P.; Bender, H.; Brijs, B.; Vandervorst, W.; Van Elshocht, S.; Pourtois, G.; Wouters, D. J.; Biesemans, S.; Kittl, J. A. J. Appl. Phys. 2009, 106, 94101.

(12) Aaltonen, T.; Ritala, M.; Arstila, K.; Keinonen, J.; Leskelä, M. Chem. Vap. Deposition 2004, 10, 215‒219.

(13) Aaltonen, T.; Ritala, M.; Tung, Y. L.; Chi, Y.; Arstila, K.; Meinander, K.; Leskelä, M. J. Mat. Res. 2004, 19, 3353‒3358.

(14) Igumenov, I. K.; Sernyannikov, P. P.; Trubin, S. V; Morozova, N. B.; Gelfond, N. V; Mischenko, A. V; Norman, J. A. Surf. Coat. Tech. 2007, 201, 9003‒9008.

(15) Aaltonen, T.; Alen, P.; Ritala, M.; Leskelä, M. Chem. Vap. Deposition 2003, 9, 45‒49.

(16) Park, K. J.; Doub, J. M.; Gougousi, T.; Parsons, G. N. Appl. Phys. Lett. 2005, 86, 51903.

(17) Park, S. J.; Kim, W. H.; Maeng, W. J.; Yang, Y. S.; Park, C. G.; Kim, H.; Lee, K. N.; Jung, S. W.; Seong, W. K. Thin Solid Films 2008, 516, 7345‒7349.

(18) Park, S. J.; Kim, W. H.; Lee, H. B. R.; Maeng, W. J.; Kim, H. Microelectron. Eng. 2008, 85, 39‒44.

(19) Kwon, S. H.; Kwon, O. K.; Kim, J. H.; Jeong, S. J.; Kim, S. W.; Kang, S. W. J. Electrochem. Soc. 2007, 154, H773‒H777.

(20) Yim, S. S.; Lee, M. S.; Kim, K. S.; Kim, K. B. Appl. Phys. Lett. 2006, 89, 93115.

(21) Aaltonen, T.; Rahtu, A.; Ritala, M.; Leskelä, M. Electrochem. Solid-State Lett. 2003, 6, C130‒C133.

(22) Kim, S. K.; Lee, S. Y.; Lee, S. W.; Hwang, G. W.; Hwang, C. S.; Lee, J. W.; Jeong, J. J. Electrochem. Soc. 2007, 154, D95‒D101.

(23) Kim, S. K.; Hoffmann-Eifert, S.; Waser, R. J. Phys. Chem. C 2009, 113, 11329‒11335.

Page 72: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

65

(24) Choi, J.; Choi, Y.; Hong, J.; Tian, H.; Roh, J. S.; Kim, Y.; Chung, T. M.; Oh, Y. W.; Kim, C. G.; No, K. Jpn. J. Appl. Phys. 2002, 41, 6852‒6856.

(25) Kwon, O. K.; Kwon, S. H.; Park, H. S.; Kang, S. W. Electrochem. Solid-State Lett. 2004, 7, C46‒C48.

(26) Marozau, I.; Shkabko, A.; Dinescu, G.; Dobeli, M.; Lippert, T.; Logvinovich, D.; Mallepell, M.; Schneider, C. W.; Weidenkaff, A.; Wokaun, A. Appl. Surf. Sci. 2009, 255, 5252‒5255.

(27) Shkabko, A.; Aguirre, M. H.; Marozau, I.; Lippert, T.; Chou, Y. H.; Douthwaite, R. E.; Weidenkaff, A. J. Phys. D-Appl. Phys. 2009, 42, 145202.

(28) Shkabko, A.; Aguirre, M. H.; Marozau, I.; Doebeli, M.; Mallepell, M.; Lippert, T.; Weidenkaff, A. Mat. Chem. Phys. 2009, 115, 86‒92.

(29) Chambers, J. J.; Busch, B. W.; Schulte, W. H.; Gustafsson, T.; Garfunkel, E.; Wang, S.; Maher, D. M.; Klein, T. M.; Parsons, G. N. Appl. Surf. Sci. 2001, 181, 78‒93.

(30) Gatineau, J.; Yanagita, K.; Dussarrat, C. Microelectron. Eng. 2006, 83, 2248‒2252.

(31) Eom, T. K.; Sari, W.; Choi, K. J.; Shin, W. C.; Kim, J. H.; Lee, D. J.; Kim, K. B.; Sohn, H.; Kim, S. H. Electrochem. Solid-State Lett. 2009, 12, D85‒D88.

(32) Kukli, K.; Ritala, M.; Kemell, M.; Leskelä, M. J. Electrochem. Soc. 2010, 157, D35‒D40.

(33) Wang, H. T.; Gordon, R. G.; Alvis, R.; Ulfig, R. M. Chem. Vap. Deposition 2009, 15, 312‒319.

(34) Park, S. K.; Kanjolia, R.; Anthis, J.; Odedra, R.; Boag, N.; Wielunski, L.; Chabal, Y. J. Chem. Mater. 2010, 22, 4867‒4878.

(35) Knoops, H. C. M.; Mackus, A. J. M.; Donders, M. E.; Van de Sanden, M. C. M.; Notten, P. H. L.; Kessels, W. M. M. Electrochem. Solid-State Lett. 2009, 12, G34‒G36.

(36) Van Hemmen, J. L.; Heil, S. B. S.; Klootwijk, J. H.; Roozeboom, F.; Hodson, C. J.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Electrochem. Soc. 2007, 154, G165‒G169.

(37) Knoops, H. C. M.; Baggetto, L.; Langereis, E.; Van de Sanden, M. C. M.; Klootwijk, J. H.; Roozeboom, F.; Niessen, R. A. H.; Notten, P. H. L.; Kessels, W. M. M. J. Electrochem. Soc. 2008, 155, G287‒G294.

Page 73: Atomic layer deposition of ruthenium films : properties and surface

4. ATOMIC LAYER DEPOSTION OF RUTHENIUM FROM CpRu(CO)2Et USING O2 GAS...

66

(38) Shamiryan, D.; Baklanov, M. R.; Boullart, W. Electrochem. Solid-State Lett. 2005, 8, G176‒G178.

(39) Haukka, S. Personal communication 2010.

(40) MaxCaps consortium - technical report, project number 2T210 2009.

(41) Fröhlich, K.; Husekova, K.; Machajdik, D.; Hooker, J. C.; Perez, N.; Fanciulli, M.; Ferrari, S.; Wiemer, C.; Dimoulas, A.; Vellianitis, G.; Roozeboom, F. Mat. Sci. Eng. B-Solid 2004, 109, 117‒121.

(42) Kukli, K.; Aarik, J.; Aidla, A.; Uustare, T.; Jogi, I.; Lu, J.; Tallarida, M.; Kemell, M.; Kiisler, A. A.; Ritala, M.; Leskelä, M. J. Cryst. Growth 2010, 312, 2025‒2032.

(43) George, S. M. Chem. Rev. 2010, 110, 111‒131.

(44) Puurunen, R. L.; Vandervorst, W. J. Appl. Phys. 2004, 96, 7686‒7695.

(45) Profijt, H. B.; Kudlacek, P.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Electrochem. Soc. 2011, 158, G88.

(46) Mackus, A. J. M.; Verheijen, M. A.; Leick, N.; Bol, A. A.; Kessels, W. M. . Chem. Mater. 2013, 25, 1905‒1911.

(47) Tsoukalas, J. T. and P. P. and A. T. and D. Nanotechnology 2011, 22, 235306.

(48) Lucas, S.; Moskovkin, P. Thin Solid Films 2010, 518, 5355‒5361.

Page 74: Atomic layer deposition of ruthenium films : properties and surface

67

Chapter 5 “Whatever it is you’re expecting won’t come in the form you’re expecting.” ‒ Haruki Mu-

rakami

In situ spectroscopic ellipsometry during atomic layer deposition of Pt, Ru and Pd* Abstract

The thickness of ultra-thin platinum-group metal films, such as Pt, Ru and Pd,

prepared by atomic layer deposition (ALD) was monitored in situ using spec-troscopic ellipsometry in the photon energy range of 0.75-5 eV. The metals’

dielectric function was parametrized using a “flexible” Kramers-Kronig con-

sistent dielectric function because it was able to provide accurate curve shape

control over the optical response of the metals. From this dielectric function, it

was possible to extract the film thickness values during the ALD process. The

important ALD process parameters, such as the nucleation period and growth

per cycle of Pt, Ru and Pd could be determined from the thickness evolution.

Spectroscopic ellipsometry thereby revealed itself as a feasible and valuable

technique to be used in research and development applications, as well as for

process monitoring during ALD.

* This chapter is to be submitted with the following authors: Leick, N.; Weber, J. W.; Mackus, A. J. M.; Weber, M. J.; van de Sanden, M. C. M. and Kessels, W. M. M.

Page 75: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

68

Introduction

In the last decade, nanometer-thin films have found myriads of applica-tions for which atomic layer deposition (ALD) has become the prevalent tech-nique to deposit these films conformally on complex 3-dimensional surfaces with sub-monolayer thickness accuracy.1 In order to control and monitor the film thickness during the ALD process, the optical and non-invasive technique of spectroscopic ellipsometry (SE) has been increasingly used,2‒9 since it is able to detect changes in nominal film thickness as small as 0.01 monolayers.2 This powerful tool has been widely employed in situ to extract thin film properties of metal oxide and metal nitride thin films, as previously reviewed by Langereis et al.4 One of important parameters obtained by SE is the thickness increase per ALD cycle, known as growth per cycle, from which also saturation and nuclea-tion behavior can be derived during the characterization of an ALD process. Furthermore, optical film properties such as refractive index, extinction coeffi-cient, and band gap can be extracted from the SE data; and for certain films also material crystallinity, composition and electrical properties.2,4,5,10‒17

Recently, thin films (5-50 nm) of Pt-group metals, such as Pt, Ru and Pd, have attracted interest for a wide range of emerging applications mostly in nano-electronics,18‒22 for which precise thickness control and conformality of ALD are also required.1,23 Several studies have shown that the initial growth of Pt, Ru and Pd on non-metallic substrates suffers initially from a substrate inhib-ited growth mode24 and is characterized by island formation.25,26 The initial metal-island formation during ALD is so pronounced that ALD has also become a method to synthesize Pt-group metallic nanoparticles.27‒32 This gives rise to additional considerations in measuring and modeling the obtained SE data in order to be able to extract some of the above-mentioned parameters from the data.

ALD is relatively new for Pt-group metals compared to other deposition techniques, such as physical vapor deposition (PVD), including sputtering and evaporation. Therefore, such films deposited by ALD have hardly been charac-terized optically.3,6,33,34 The material properties of Pt-group metal films are de-pendent on the deposition method which in turn affects the dielectric function of the metal, as was demonstrated in a comparative study of Pt deposited by evaporation and sputtering.35 Therefore, it is important to study the optical

Page 76: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

69

properties of Pt, Ru and Pd films deposited specifically with ALD to obtain ref-erence data for future ALD studies.

Furthermore, most PVD metal films studied by SE had thickness values at which the metals were opaque and showed bulk-like optical properties.36,37 Zollner showed that 50 nm Pt films exhibited optical properties that were al-most the same as bulk Pt.37 However, ALD generally yields films with thickness-es well below 50 nm of which the optical properties can be very different from those of bulk metals.38 This holds especially for films with thicknesses below 10 nm.39,40

The thickness of a thin film is generally extracted from the SE data via an optical model which typically consists of a stack of layers, and each layer repre-sents a material having its own dielectric function and thickness value. To parametrize the dielectric function of conductive materials, Drude-Lorentz os-cillators are the most frequently used.4 The Drude oscillator describes the ab-sorption of light due to free carriers in the metal film, while the Lorentz term(s) describe the absorption of light due to interband transitions of bound charges in the material.41‒43 The absorption of light in the infrared region (< 2 eV) is generally due to free carriers; however, in the case of Pt, Ru and Pd, interband absorptions also occur in this region. The absorption at low energies can there-fore not solely be attributed to the absorption due to free carriers.44 Therefore, the Drude-Lorentz parametrization of ultra-thin Pt-group metal films in the op-tical range studied (0.75-5 eV) cannot be applied unambiguously.

This work aims at extending the review by Langereis et al.4 with findings on Pt, Ru and Pd as representatives for Pt-group metals. To this end, we monitored the film thickness in situ during Pt, Ru and Pd ALD. In the optical model for the dielectric function of the film, a “flexible” parametrization was investigated that accounts for the above-mentioned challenges: material properties which are significantly different from those obtained by

deposition techniques other than ALD; effects due to the initial island-growth during ALD; thickness-dependent dielectric functions for films thinner than ~10 nm; light absorption due to both, free and bound, charges in the spectral re-

gion below 2 eV. This “flexible” parametrization was introduced as a variational analysis

method of optical spectra by Kuzmenko.45 It can be regarded as a limiting case

Page 77: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

70

of a Drude-Lorentz parametrization, since it also uses pre-established shapes to fit optical data. The main difference lies in the number of these shapes, which can vary depending on the level of fitting accuracy desired. Due to a high number of these shapes, this variational fitting procedure allows for an accu-rate shape control of the dielectric function, while at the same time being con-strained to the Kramers-Kronig relationship, ensuring the physical meaning of the dielectric function.46 This kind of variational parametrization has been used previously: (i) under the name of B-spline parametrization (see Results and dis-cussion),46 to obtain the dielectric function of “optically challenging” materials, such as SrTiO3 films or Au particles,5,9,47 and (ii) under the name of variational dielectric function (VDF; see Results and discussion) for superconductor materi-als.48‒51

This Chapter presents general accomplishments of SE which can be ap-plied to any of the three metals (Pt, Pd and Ru) treated, such as the determina-tion of film thickness and growth per cycle using a B-spline model, and film resistivity by extending the VDF into a Drude-Lorentz model.

Preparation of the Pt, Pd and Ru films

The Pt and Pd thin films were deposited by ALD in a home-built ALD reac-tor10 and the Ru films in an Oxford Instruments FlexAL™ reactor.52 The films were deposited by plasma-assisted ALD using MeCpPtMe3 and an O2 plasma for Pt, and Pd(hfac)2 and an H2 plasma for Pd. A thick spacer film of 500 nm ALD Al2O3 was used to increase the optical contrast between the metallic films and the Si wafer, thereby enhancing the sensitivity of SE during the growth of Pt and Pd.53 The Ru films were deposited by thermal ALD using CpRu(CO)2Et and O2 gas and directly deposited on Si wafers with native oxide, i.e. without the thick dielectric spacer film in-between. The set substrate temperature was 100 ºC, 300 ºC and 400 ºC during Pt, Pd and Ru ALD, respectively. More details about the Pt and Pd ALD processes can be found elsewhere.3,26 With respect to the Ru process described in detail in Chapter 4, the O2 gas pressure was in-creased to ~120 mTorr for 5 s to enhance the Ru nucleation on Si substrates. As a consequence, the overall growth per cycle increased to ~1.6 Å/cycle, com-pared to the ~1 Å/cycle reported in Chapter 4.

Page 78: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

71

The film properties obtained by ellipsometry were compared with the properties obtained by ex situ measurement techniques. The Pt and Pd film thickness values were studied by X-ray reflectometry (XRR) and the Ru thick-ness was measured by scanning electron spectroscopy (SEM) due to its high surface roughness. The resistivity of all the films was determined using a four point probe (FPP). Being still subjects of research, the processes did not yield optimized material properties in all cases. The Pt films were pure and smooth (0.4 nm roughness for a 30 nm film), but the Pd films contained ~ 14% C impu-rities whereas the Ru films suffered from a high surface roughness (9 nm roughness for a 22 nm film) and delamination. The delamination is claimed to be due to the difficulty of Ru to wet non-metallic surfaces.33 These effects are not uncommon for ALD metal films and can affect the film resistivity as will be discussed in later Sections.

Optical characterization of the Pt, Pd and Ru films

in situ spectroscopic ellipsometry

SE is based on the change in polarization of incident polarized light upon

its reflection from a sample material, as it is illustrated in Figure 1. From the po-

larization of the reflected light the film thickness and dielectric function of the

material can be deduced from the amplitude ratio ψ and the phase

difference Δ. These two parameters are related to the complex Fresnel

reflection coefficients, rp and rs, by the ellipsometry equation:

rprs

=Erp Eip⁄

Ers Eis⁄=

Erp Eip

|Ers| |Eis|⁄exp[i δp δs ]≡tanψ·exp(iΔ),

where E represents the complex field amplitudes, the subscripts i and r repre-sent the incident and reflected electric field, respectively, whereas s and p refer to the polarization of the electric field.

Page 79: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

72

Figure 1: (a) Schematic representation of linearly polarized light incident onto a sample and reflected as elliptically polarized light. The subscripts i and r represent the incident and reflected electric field E, respectively, whereas s and p refer to the polarization of the electric field. (b) Schematic configuration of the spectroscopic ellipsometer mounted on the ALD reactor. The reactor is equipped with gate valves to protect the SE windows from film deposition. The light from the light source of the ellipsometer passes through the fixed polarizer (P) and the rotating compensator (RC) and is reflected from the sub-strate through a fixed analyzer (A) into the detector.

The ALD setups used in this work are equipped with optical viewports to monitor film growth by in situ SE. The quartz windows are protected from film deposition by gate-valves, as illustrated in Figure 1b. The view port is designed to have an angle of θ = 70º with respect to the substrate normal. This angle is close to the Brewster angle of Si (our most-commonly used substrate) to en-sure the highest sensitivity for the measurements for thin films. The data acqui-sition during the deposition was carried out by a J.A. Woollam Co., Inc. M2000U near-infrared (NIR) ellipsometer with an optical range of 0.75-5.0 eV, using the CompleteEASE® 4.48 software. The ellipsometer, equipped with a fixed polarizer, a rotating compensator and a fixed analyzer, is able to measure the four Stokes

turbopump

SElight source SE

detector

RC

heated substrate

P A

metalprecursor

(MeCpPtMe3, CpRu(CO)2Et, Pd(hfac)2)

co-reactant(O2, H2 plasma,

O2 gas)

gate valve

silica window

to spectro-graph

Erp

Δ

ψ Ers

ErEi

Eip

Eis sample

θ

(a)

(b)

Page 80: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

73

parameters in a single acquisition. The Stokes parameters relate the reflected polarized light intensity to the ellipsometric angles, ψ and Δ. The Stokes pa-rameters are important for the evaluation of the fit and will be discussed sepa-rately in the Results and discussion.

The dynamic measurement procedure starts with the automatic acquisi-tion of ( ψ, Δ) spectra as soon as the gate-valves are opened. The dynamic da-ta were acquired every 5 ALD cycles with a measurement time of 40 s during Ru and Pd ALD and during the first 200 cycles of Pt ALD. For the remaining 800 cycles of Pt ALD, the data were collected every 10 ALD cycles. For each deposi-tion all the spectra were appended into a single file. In doing so, the starting model of each ( ψ, Δ) spectrum during the analysis (see below) was based on the model of the previously modeled spectrum. This increases the accuracy of the fit and makes it possible to obtain an accurate growth per cycle value di-rectly from the fit.

Figure 2: Schematic representation of the film stacks used (in the case of Ru ALD no Al2O3 was present) and the associated optical model used. The film thickness of each layer was extracted from the data using a parametrization of the dielectric function of the layer. The Cauchy parametrization of the Al2O3 was numerically inverted prior to modeling the metal films.

Prior to modeling the dynamic data of the metal films, the optical response of the substrates was carefully characterized in order to minimize the fitting error of very thin films. The film stacks and associated optical model used in this study are schematically described in Figure 2. The single-sided-polished Si wa-fer (200 μm thick) was fitted by a model including the temperature-dependent response of the silicon54 to account properly for the substrate tem-perature used during the ALD process. The native SiO2 layer of ~ 2 nm on top of the Si and the alumina films were described by the Cauchy parametrization, yielding an Al2O3 thickness of 498 ± 1 nm for the experiments with Pt and Pd. All of the fit parameters of the Cauchy parametrizations were then fixed for

Pt, Ru, PdSiO2 / Al2O3

Si substrate

B-splineCauchy

Si-bulk (semi-infinite)

Film stack Optical model

Page 81: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

74

each metal film, and ψ and Δ were numerically inverted to ε1 and ε2, as de-scribed in Ref. 39. Due to this precise substrate characterization, any changes in ψ and Δ from subsequent data sets can be ascribed to adlayers only; in this case to the deposition of the metal films.

ex situ Fourier transform infrared reflectance

In order to obtain an optical resistivity value of the films, it is desirable to extract the Drude contribution from the dielectric function. As it was men-tioned in the Introduction, this is not straightforward from data in the NIR range. Therefore, it was necessary to extend the measured range in the infrared spectrum, where the absorption by free carriers in the metal is most promi-nent.55 The combination of SE data and Fourier transform infrared (FT-IR) reflec-tance data has successfully been applied previously to metal films deposited by PVD and chemical vapor deposition to obtain a uniquely-defined Drude-Lorentz description of the dielectric constant.56‒58 As a proof-of-principle for metal films deposited by ALD, we applied this procedure to the Pt, Ru and Pd films after deposition, because they were sufficiently thick to show bulk-like properties.

To this end, a Bruker Tensor 27 spectrometer was employed ex situ in re-flection mode at an angle of 12° covering the spectral range of 0.04-0.86 eV (6997-368 cm-1) with a resolution of 2×10-4 eV (2 cm-1). The reflectance data and the SE data were modeled simultaneously using the software RefFit45. The met-al films were first modeled using a VDF which was further parametrized with a Drude-term and Lorentz-oscillators over the range of 0.04-5 eV.

Each reflectance spectrum is the average spectrum of 1000 scans and the systematic error assigned to the reflectance data in the RefFit software was a fixed error of 1%. The reflectance data are, therefore, given a fitting weight of 10,000 during the simultaneous fit of reflectance and SE data. Due to this fixed systematic error and this high fitting weight, the χred2 of the Drude-Lorentz parametrization also has a weight of at least 10,000 and therefore defeats its meaning. For this reason, the χred2 values will not be discussed in this Chapter for the Drude-Lorentz parametrization.

Page 82: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

75

Figure 3: The imaginary part of the dielectric function, ε2, determined by the B-spline parametrization at different stages of the (a) Pt, (b) Ru and (c) Pd deposition. For each dielectric function, the thickness and number of cycles (abbreviated by “c.”) are given. The number of knots used for the B-spline parametrization (see Eq. 1) was 11, 6 and 6 for Pt, Ru and Pd, respectively. The graphs show that the dielectric function significantly changes in the infrared region, especially during the initial cycles (up to ~10 nm film thickness).

3.5 nm (70 c.)

5 nm (100 c.)

10 nm (200 c.)

53 nm (1000 c.)

3.5 nm (40 c.)5.5 nm (55 c.)

8.3 nm (70 c.)

28 nm (200 c.)

11 nm (100 c.)

3.5 nm (265 c.)

7.5 nm (455 c.)

10 nm (645 c.)

41 nm (500 c.)

Page 83: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

76

Table I: Models used for the substrates in the fitting software packages used in this study, RefFit and CompleteEASE®, and the reference thereof. KK is the acronym for Kramers-Kronig consistent, VDF stands for variational dielectric function. The abbrevia-tion Ntve stands for the native silicon oxide layer and Si_Temp is a temperature-dependant silicon optical model; and JAW is the acronym of J. A. Woollam Co., Inc..

RefFit CompleteEASE

Spectral range 0.04 ‒ 0.86 eV 0.75 ‒ 5 eV 0.75 ‒ 5 eV

Al2O3 (500 nm) KK VDF 45 KK VDF 45 Cauchy 40

SiO2 (1-2 nm) -- -- Ntve_JAW 54

Si (> 200 μm) KK VDF 45 KK VDF 45 Si_Temp_JAW 54

In the RefFit software, it is not possible to load tabulated optical constant;

therefore, the Si substrate and the Al2O3 film were fitted using the variational dielectric function in each spectroscopic range used. In the case of the native SiO2, we assumed that the thickness of 2 nm was negligible with respect to the thickness of the metal film. The substrate characterization is summarized in Ta-ble I for each software and spectroscopic range used.

From the ex situ SE measurements on a variable angle stage, it became ap-parent that the SE was most sensitive to the metal films deposited on Si wafers for an angle of incidence of 85º. Therefore, the SE data appended to the reflec-tance data of the thick Pt, Ru and Pd films was taken at this angle of incidence.

Results and discussion

ALD film growth of Pt, Ru and Pd studied by in situ SE

Variational analysis of the in situ optical spectra

To accommodate the optical properties changing with metal film thickness in the optical range of 0.75-5 eV, we chose to apply the variational dielectric function available in the Complete-EASE® software called B-spline.46 In that case, the imaginary part of the dielectric function as a function of photon energy, ,

Page 84: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

77

is represented by the linear sum of B-spline polynomials over a given number of knots, n:

ε2=∑ ci Bi ( )ni=1 (Eq. 1)

where ci are the B-spline coefficients and Bi are the B-spline polynomials of de-gree 3. As mentioned in the Introduction, the B-spline dielectric function can be “physical” by enforcing that ε2 ≥ 0 and that ε2 and ε1 fulfill the causality requirement of Kramers-Kronig consistency.46 Recently, the dielectric function of amorphous carbon films and graphene were parameterized accurately with-out making assumptions on the physical properties of the layers using a Kramers-Kronig consistent B-spline parametrization.59,60 For ALD prepared films, SrTiO3 was also parameterized by B-splines.5,9

Figure 3 shows the imaginary part of the dielectric function, ε2, obtained by the B-spline parametrization of spectra selected from the dynamic data ac-quired during Pt (a), Ru (b) and Pd (c) ALD. For each metal, the dielectric func-tion ε2 can accurately describe the thickness dependent optical response. Clearly, for all three metals the absorption behavior in the near-infrared region is sensitive to changes in film thickness, especially below 10 nm thickness. From previous studies using non-metallic substrate materials,28,31,32,61 it is known that metals deposited by ALD initially grow as islands rather than films. For the particular case of Au particles (not prepared by ALD),47 the thickness values ob-tained by the B-spline parameterization was compared with transmission elec-tron microscopy images. This comparison suggested that the variational parametrization delivered an effective dielectric function and thickness value.

For Pt and Pd, ε2 in Fig. 3 decreases monotonously with increasing pho-ton energy, whereas for Ru an absorption peak ~ 4 eV can be observed, notably for very thin films. Since the optical response of the substrate was well de-scribed prior to deposition, it is not likely that this absorption peak is related to the underlying Si, but instead is a real absorption peak of the Ru. Furthermore, this feature at ~ 4 eV loses amplitude with increasing Ru thickness, but never vanishes completely, indicating that it is a feature intrinsic to the Ru metal. In a later section of this Chapter, we show that the most adequate oscillator-based model for the bulk-like Ru film includes a Lorentz oscillator at ~4 eV, very close to the one reported by Choi et al..36 A similar absorption feature at ~2 eV has

Page 85: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

78

been previously reported for thin Au films, even during the initial growth of the films.58

Determination of film thickness and growth per cycle

As shown in the previous section, the film thickness can be extracted for each data set using a B-spline parametrization. These thickness values are plot-ted in Figure 4 as a function of ALD cycles for Pt (a), Ru (b) and Pd (c). Initially, the thickness remains (almost) zero which represents the nucleation delay of the ALD processes. This nucleation delay is approximately 20, 20 and 100 ALD cycles for Pt, Ru and Pd, respectively. The film thicknesses determined at the end of the Pt and Pd deposition were 53.3 ± 0.4 nm and 40.9 ± 0.1 nm, re-spectively. These thickness values are in fair agreement with the values of 50 ± 2 nm for Pt and 36± 2 nm for Pd obtained by ex situ XRR. Between the nucleation delay and the final thickness value, the three plots exhibit a linear thickness increase which is characteristic for ALD film growth from which the growth per cycle (GPC) can be determined. For the Pt and Pd ALD processes, the GPC values are 0.55 ± 0.01 and 0.17 ± 0.01 Å/cycle, respectively. In this particular case, the error values represent the accuracy of the linear regression analysis of the thickness data points obtained by optical modeling. The Pt GPC is somewhat higher than the one reported previously (0.45 Å/cycle),3 but such fluctuations in growth per cycle have been observed occasionally for the plas-ma-assisted ALD process of Pt. For Pt films approaching 50 nm, it can be ob-served that the “noise” in the thickness value increases. This can be attributed to the loss of sensitivity as the films become almost opaque.

Page 86: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

79

Figure 4: Thickness increase of (a) Pt, (b) Ru and (c) Pd films determined by the B-spline parametrization of the in situ SE data. After the nucleation period, the thickness increases linearly with the number of cycles. The GPC of each deposition process is obtained by linear regression of the linear part of the plot.

Pt

0.55 Å/cycle

(a)

(b)Ru

1.6 Å/cycle

(c) Pd

0.17 Å/cycle

Page 87: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

80

In the case of Ru, the comparison is somewhat more complicated since Ru thickness values > 26 nm (i.e. for >165 ALD cycles) cannot be extracted from the SE optical model due to the absence of the thick transparent (Al2O3) spacer in the film stack that can enhance the optical sensitivity. The slope of the linear thickness increase (Fig. 4b) delivers a GPC value of 1.6 ± 0.2 Å/cycle for Ru. From this value it is possible to extrapolate the linear thickness increase and estimate the thickness value of the Ru film after 500 ALD cycles. On the basis of the GPC from the SE data, an expected thickness value at the end of the Ru deposition of ~77 nm could be estimated. The thickness was estimated to be ~90 nm using SEM, although in the case of a rough, non-coalesced film, the “film thickness” measurement using SEM in cross-section mode is always over-rated by the surface roughness as well as the possibility of additional blister formation. These two effects add to image contrast, thus suggesting a too high film thickness. Thus we may conclude that the extrapolated SE-thickness value is semi-quantitatively in line, especially since there is no thick transparent opti-cally enhancing Al2O3 spacer under the Ru-films.

Drude-Lorentz parametrization for bulk-like films of Pt, Ru and Pd from ex situ optical data

A variational parametrization is a very useful tool to extract the dielectric function of the Pt-group metals with proper curve shape control. Furthermore, it is much easier to add an oscillator-based description to features present in a dielectric function extracted by a variational dielectric function. Therefore, a VDF over the optical range of 0.04-5 eV was performed in the RefFit software, prior to the Drude-Lorentz parametrization, using the final film thickness ob-tained from the B-spline parametrization. A VDF, like the B-spline parametrization, can be described by Eq. 1, in which case Bi are B-spline poly-nomials of degree 1.

Since ε1 and ε2 are correlated by the Kramers-Kronig relation, we mainly show and discuss ε2. For metals, it is common to describe the dielectric func-tion of metals as containing an intraband (D, for Drude) and interband (i) con-tribution(s), i.e. ε2 E =(εD+εi)2 ,41‒43 which is represented as shown in Eq. 2 in the case of a Drude-Lorentz model with n Lorentz oscillators.

Page 88: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

81

ε2 ω =ε∞ - ωp2

ω2-iω τD⁄+ ∑

fjω0 j2

ω0 j2 -ω2+iγjω

nj=1 (Eq. 2)

In this equation, ε∞ accounts for contributions to ε2 outside the measured range when ε∞>1. The resistivity of the metal, ρ, can be extracted from the Drude parameters, i.e. from the unscreened plasma energy ℏωp and the aver-age lifetime of the free carriers τD by the expression ρ= 1 ε0ωp

2τD⁄ . Lorentz os cillators are located at the energy position ℏω0 j, with amplitude fj and damping factor γj.

In the case of Pt, Ru or Pd, strong absorption from bound and free carriers can both occur below 2 eV.44 To separate these contributions, reflectance measurements in the optical range from 0.04-0.85 eV were performed for the films after deposition and modeled simultaneously with the SE data in the RefFit software as described in the Experimental details. The imaginary part of the dielectric function resulting from the Drude-Lorentz model in the range of 0.04-5 eV is shown in Figure 5 for the three metals. The thickness was fixed to the value found by the B-spline model, and for Ru at the SEM thickness of 90 nm. The parameters of the fit are listed in Table II using the fitting terminol-ogy position, amplitude and broadening.

To verify the validity of these models, the parameters of the Lorentz oscilla-tors can be compared with those of other studies, as also summarized in Ta-ble II. Overall, the results of this study agree fairly well with previous work on Pt-group metals not prepared by ALD. Choi et al.36 specifically investigated interband transitions experimentally and theoretically. Amongst others, they considered Pt and Ru, deposited by PVD, for a spectral range of 1 ‒ 10 eV, hence exceeding our measurement range. The Lorentz oscillators from our work are in relatively good agreement with theirs. In the case of Pt, the oscilla-tor positioned at 3.1 eV shows the biggest discrepancy. Instead of an oscillator at ~ 3 eV, Choi et al. found two oscillators, one at 2.2 eV and one at 4.2 eV. Re-placing the 3.1 eV oscillator from our model with two oscillators at 2.2 eV and 4.2 eV led to a high fitting correlation of the two oscillators, i.e. the fit became worse. For Pt, a second difference with the work of Choi et al. lies in the value of ε∞, which is likely due to the different spectral range covered in the two stud-ies. The remaining small differences can possibly be assigned to the different material properties obtained by ALD and PVD.

Page 89: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

82

Figure 5: The imaginary part of the dielectric function, ε2, determined by the Drude-Lorentz dielectric function in the spectral range of 0.04-5.0 eV of the nearly bulk (a) Pt, (b) Ru and (c) Pd films. The dotted lines represent the Drude contribution whereas the Lorentz-oscillators are shown in the inset of each figure.

Pt (a)

Ru (b)

Pd (c)

Page 90: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

83

Table II: Position, amplitude and broadening for the Drude-Lorentz parametrization of the Pt, Ru and Pd films shown in Fig. 5 compared to the parameters of earlier studies (indicated in italic). In the case of the Drude parameters, ℏωp is represented by the amplitude and ℏτD by the broadening; whereas in the case of the Lorentz oscillators, ℏω0 is represented by the position, f by the amplitude and ℏγ by the broadening. The errors give the 90% confidence interval of the fitted value (which will be discussed below). When not stated specifically, the 90% confidence interval is equal or less than 1% of the value.

Position

[eV]

Amplitude

[eV]

Broadening

[eV]

Position

[eV]

Amplitude

[eV]

Broadening

[eV] Ref.

Pt 0 7.18 0.09 0 7.54 n/a 63

0.78 5.2 ± 0.1 0.6 0.7; 0.8;

0.9 5 0.9 3,36,63,64

1.2 6.5 ± 0.2 1.4 1.5; 1.7 4 2.2 3,36

3.1 12 6.7 2.2 10 4.3 36

-- -- -- 4.2 2.5 2.5 36

ε∞ = 3.5 ε∞ = 1.9 36

Ru 0 3.0 0.04 0 n/a n/a

0.4 3.3 1.04 0.6 4.1 1.1 36

1.6 5.2 1.7 1.8; 1.5 9.3 1.7 36,62

-- -- -- 2.9; 2.7 2.2 0.8 36,62

4.0 5.9 2.9 4.1 9.1 4.5 36

ε∞ = 1.50 ε∞ = 2 36

Pd 0 4.8 0.21 0 3; 6.99 0.12; 0.21 64,65

1.27 7.5 2.76 1.16;

1.15 10 2.88 65,66

5.12 9.6 6.84 2.94 2.1 1.33 65

ε∞ = 2.50 ε∞ = 1 65

Page 91: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

84

In the case of Ru, our oscillator positions are shifted ~0.2-0.3 eV with re-spect to Choi’s oscillators at 0.6 and 1.8 eV. The most striking difference is the absence of an oscillator ~ 3eV. Adding such an oscillator at 2.9 eV to our data resulted in a high fitting correlation of the fit parameter. Henrion et al.62 report-ed a dielectric function for Ru whose shape is very similar to the one we report. Moreover, they also found a Lorentz oscillator at 1.5 eV, as well as one at 2.7 eV.

The Drude parameters and the Lorentz oscillator at 1.2 eV of the Pd film are in good agreement with reports in the literature, most particularly with the work of Yamada et al..65 However, we find an oscillator at 5.1 eV. The amplitude of this oscillator is much higher than the one reported at 2.94 eV in their work. No good fit to our data could be obtained with an oscillator at a position of ~3 eV. The reason for the discrepancy in oscillator positions is not yet clear.

The resistivity values for the 53 nm Pt, 90 nm Ru and 40 nm Pd films calcu-lated from the Drude parameters extracted from the model are in good agreement with the values obtained by ex situ FPP measurements, as shown in Table III. For Pt, the optical resistivity and FPP values coincide very well, and the optical resistivity is very close to the Pt bulk resistivity of single crystal. Since the ALD Pt films are polycrystalline, this slight discrepancy can mainly be attributed to scattering effects of conduction electrons at grain boundaries of the poly-crystalline film 67,68 and at the film surface.69,70

We expect that this agreement is potentially also true for ALD Pd and Ru films, but not for the films of this study. As mentioned in the Experimental sec-tion, the Pd films of this study contained 14% of C impurities, potentially lead-ing to scattering of the conduction electrons at these C impurities, 71 thereby increasing the measured resistivity values compared to the bulk resistivity of Pd. The optical resistivity and FPP values, however, show a very good agreement. In the case of Ru, film delamination and the high film roughness deteriorate electronic transport by phenomena which are complex to take into account in the SE model.

Page 92: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

85

Table III: Comparison of the optical resistivity (in μΩ·cm) with the FPP-resistivity of the Pt, Ru and Pd films, shown in Fig. 5. The error in the optical resistivity values is represent-ed by the 90% confidence interval which is equal or less than 1% of the value for the three metals, and is therefore not shown. The bulk resistivity values are given at 27 ºC for Pt and Pd and at 20 ºC for Ru.72

Optical resistivity

[μΩ.cm] FPP resistivity [μΩ.cm]

Bulk resistivity [μΩ.cm]

Pt (53 nm) 12.6 13.0 ± 0.2 10.4 Ru (90 nm) 32.8 18.0 ± 0.6 6.7 Pd (42 nm) 67.5 67 ± 1 10.5

Fit reproducibility and goodness of the fit for Pt, Ru and Pd

For clarity, the previous sections did not include a discussion on the good-ness of the fit, which is crucial in order to select a given parametrization over another. To evaluate the goodness of the fit, the most commonly used parame-ters include mean squared error, the error bars, the uniqueness of a fit parame-ter and the correlation of fit parameters. These parameters will be presented and discussed for the fits of the Pt, Ru and Pd films prepared by ALD in this sec-tion. As mentioned in Experimental section, the χred2 values obtained during the simultaneous fitting of the FT-IR and SE data are affected by the high fitting weight attributed to the FT-IR data. Therefore, the data of the SE data only will be discussed, i.e. over the spectral range of 0.7-5 eV.

The experimental data used to extract the ellipsometric angles, ψ and Δ, are the Stokes parameters mentioned in the Experimental details; defined as follows: S0=1; S1=- cos 2ψ ; S2= sin 2ψcosΔ ;S3=- sin 2ψsinΔ. To obtain the best possible match between the experimental data and the model, CompleteEASE® and RefFit use the iterative Marquardt-Levenberg regression algorithm to minimize χred2 , expressed in Eq. 3.45,73

χred2 = ∑ 1

Nj-M∑ ωi

Xiexp-Xi

mod

σiexp

j

2Ni=1

pj=1 (Eq. 3)

where N is the number of data points, M is the number of fit parameters, p is the number of datasets fitted simultaneously, X a data set (ψ, Δ, reflectance,

Page 93: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

86

Figure 6: The value (left y-axis, light) corresponding to the fit of the experimental data and the confidence interval (right y-axis, dark) for each thickness value of the (a) Pt, (b) Ru and (c) Pd metal films. The graphs in (d) ‒ (f) represent the uniqueness plot for a low and a high value of the thickness.

etc.) and σ the standard deviation of the normally dispersed N datapoints.45 As ellipsometry uses the three Stokes parameters to extract ψ and Δ , CompleteEASE® uses S1-S3 as datasets with a fixed standard deviation of ~ 0.001. In CompleteEASE®, the parameter χred2 in Eq. 3 is commonly called mean squared error (MSE) for historical reasons, but strictly speaking an MSE does not

Pd 10 nm Pd 41 nm

(d) (e)

(f )

(b) Ru

(c) Pd

Pt 11 nm

0

100

2010Thickness (nm)

Pt 53 nm

Thickness (nm)30 50 700

10 Ru 26 nm

Thickness (nm)0

10

10 30 50

Ru 7 nm

Thickness (nm)

80

0 10 20

Pt(a)

Thickness (nm) Thickness (nm)0 10 20

120

0

70

20 40 60

00

Χ2 red

Χ2 red

Χ2 red

Χ2 red

Χ2 red

Χ2 red

Page 94: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

87

include the first summation, i.e. the division by the number of degrees of free-dom.74

In general, there are three main parameters used to assess the quality and validity of a fit: (i) the reduced chi-square χred2 (ideally ~1), (ii) the 90% confi-dence interval which in our study has been represented by the error bars of the fit parameter value obtained and (iii) the uniqueness of the value of a fit pa-rameter.75 In Figure 6a-c, each (grey) data point shows the χred2 value obtained for the Pt, Ru and Pd thickness values extracted from the in situ SE measure-ments. Starting from the thickest Pt film (Fig. 6a) to the thinnest, χred2 is close to unity, remains (almost) constant until it starts to increase below ~20 nm film thickness, and finally peaks at a value of χred2 ~ 6 for a thickness value of ~6 nm. A similar behavior can be observed for Pd (Fig. 6c). Starting from the thickest Pd film to the thinnest, χred2 increases until a thickness of ~7.5 nm is reached when it peaks at a value of χred2 ~ 13. According to Tompkins et al.40, a χred2 value around 10 is acceptable in the case of metal films. The rise in χred2 as the films become thinner may indicate that around these thickness values, phenomena caused by the increasingly granular nature of the metal films due to the island formation might play a role.76,77 In the B-spline parametrization these nuclea-tion effects (i.e. island formation) are accounted for and it delivers an effective dielectric function and thickness value, as recently shown in the case of Au par-ticles.47 Nevertheless, the thickness values extracted from the dielectric function are accurate from a modeling perspective, and this will be discussed in more details below by scrutinizing the 90% confidence interval and the uniqueness of the thickness value.

The 90% confidence interval is a useful parameter to assess how reliable the Pt, Ru and Pd thickness values are. By definition, it means that the fit pro-vides a 90% confidence that the true value of the fit parameter lies within the given interval. Based on this definition, if the interval is small, the value of the fit parameter is close to the true value and can therefore be regarded as accurate.77,78 The confidence intervals for each thickness value of Pt, Ru and Pd are also shown in Fig. 6a-c. For all of the three metals, the confidence interval is small, but strikingly small for all of the Pd thickness values. It is also worthwhile mentioning that the 90% confidence interval is smallest for low thickness val-ues in the case of the three metals.

Page 95: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

88

The uniqueness plots for Pt, Ru and Pd for a thinner and a thicker film are shown in Fig. 6d-f. They were obtained by varying the thickness values, and for each the thickness value the other fit parameters are fitted. The extracted thickness is regarded as unique when it clearly has the lowest χred2 value of all the thickness values scanned. As indicated by the narrow confidence interval for Pd, the uniqueness of the thickness value is well defined for all thickness values, even for thicker films. The Pt thickness is uniquely defined for thin films, not for thicker ones, which is related to the decrease in sensitivity as the films are almost opaque. In the case of Ru, the evolution of χred2 and the uniqueness plots as a function of thickness behave differently from those of Pt and Pd. The reason does not lie in the material itself, but in the poorer optical contrast be-tween Si and Ru due to the absence of a thick transparent spacer material. Nevertheless, the values of χred2 remain well under the value of 10 which is re-garded as acceptable for metal films. In addition, while the uniqueness of the 7-nm-thin Ru film is not well-defined, the χred2 values for films of < 8 nm are low (~ 2) and the 90% confidence interval is small. Overall, on the basis of this eval-uation, the thickness values for Ru films are also regarded as well-defined.

From Figure 6, it is possible to conclude that the thickness values and the fit are reliable. However, it is also important to test the reproducibility of a fit. The uniqueness of a fit consists of the uniqueness of every individual fit param-eter. For each fit parameter, the regression algorithm can converge to a local minimum of χred2 , and yet yield incorrect parameter values, even if the fit is good overall. It is therefore essential to probe the reproducibility of the graphs shown in Fig. 4 by employing different starting values. For example, the dy-namic fit can once be started from the data at cycle number 200, and another time at cycle 500, etc. Selecting different starting values after the nucleation delay reproduced the graphs shown in Fig. 6d-f quite well with only minor dif-ferences (i.e. < 2nm in final film thickness). This and the fact that none of the fit parameters were correlated yield strong support for the fact that the thickness determination using the B-spline approach is unique in the case of Pt, Ru and Pd films deposited by ALD.

The thickness values retrieved from the fit during the metal nucleation ex-hibit the largest χred2 in the case of Pt and Pd, and in the case of Ru the thickness values are not unique in this region. Therefore, the thickness values below the linear growth increase are indicative rather than absolute. Optical phenomena

Page 96: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

89

due to the initial island growth of the metal can arise and make the interpreta-tion of the data very challenging, such as capacitive behavior of the islands, scattering due to size effects or phonon emission to name a few.71,79‒81

Conclusions

Atomic layer deposition (ALD) has gained considerable interest over the last decade as a technique able to deposit nanometer thin films of platinum-group metals, such as Pt, Pd and Ru. We monitored the evolution of the thick-ness and the dielectric function of the films using in situ spectroscopic ellipsometry (SE) in the range of 0.75-5 eV. Using a Kramers-Kronig consistent variational parametrization to account for the thickness-dependent dielectric functions, we were able to obtain accurate film thickness values at almost any stage of the ALD process. Based on this model, we found the nucleation delay and the growth per cycle values to be 40, 20 and 100 cycles and 0.55 Å/cycle, 1.6 Å/cycle, 0.17 Å/cycle for Pt, Ru, and Pd, respectively. To study the films’ elec-trical properties specifically, we performed Fourier-transform infrared spectros-copy and modeled this data simultaneously with ex situ SE data. The variational parametrization of the metal films over the large photon energy range not only provided important process-related information but also served as a basis for the more traditional Drude-Lorentz parametrization of the metals in the range of 0.04 eV-5 eV. The resistivity values from the Drude parameter and four-point-probe values were in good agreement, especially in the case of the Pt films, and the parameters of the Lorentz oscillators were in line with previous ellipsometry reports on Pt, Ru and Pd.

Acknowledgments

C.A.A. van Helvoirt is acknowledged for his technical support throughout this study. We thank W. Keuning for performing the XRR measurements on the films. This work was financially supported by the Dutch Technology Founda-tion STW which is part of the Netherlands Organization for Scientific Research (NWO). The work of J.W. Weber was financially supported by the Foundation for Fundamental Research on Matter (FOM).

Page 97: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

90

References

(1) George, S. M. Chem. Rev. 2010, 110, 111‒131.

(2) Langereis, E.; Heil, S. B. S.; Van De Sanden, M. C. M.; Kessels, W. M. M. J. Appl. Phys. 2006, 100, 023534.

(3) Knoops, H. C. M.; Mackus, A. J. M.; Donders, M. E.; Van de Sanden, M. C. M.; Notten, P. H. L.; Kessels, W. M. M. Electrochem. Solid-State Lett. 2009, 12, G34‒G36.

(4) Langereis, E.; Heil, S.; Knoops, H.; Keuning, W.; Van de Sanden, M.; Kessels, W. J. Phys. D-Appli. Phys. 2009, 42, 073001‒1/19.

(5) Langereis, E.; Roijmans, R.; Roozeboom, F.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Electrochem. Soc. 2011, 158, G34‒G38.

(6) Junige, M.; Geidel, M.; Knaut, M.; Albert, M.; Bartha, J. W. 2011 Semiconductor Conference Dresden 2011, 1‒4.

(7) Park, S.-J.; Lee, J.-P.; Jang, J. S.; Rhu, H.; Yu, H.; You, B. Y.; Kim, C. S.; Kim, K. J.; Cho, Y. J.; Baik, S.; Lee, W. Nanotechnology 2013, 24, 295202.

(8) Hägglund, C.; Zeltzer, G.; Ruiz, R.; Thomann, I.; Lee, H.-B.-R.; Brongersma, M. L.; Bent, S. F. Nano letters 2013, 13, 3352‒3357.

(9) Longo, V.; Leick, N.; Roozeboom, F.; Kessels, W. M. M. ECS J. Solid-State Sci. Technol. 2013, 2, N15‒N22.

(10) Heil, S. B. S.; Langereis, E.; Kemmeren, A.; Roozeboom, F.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Vac. Sci. Technol. A 2005, 23, L5‒ L8.

(11) Langereis, E.; Creatore, M.; Heil, S. B. S.; Van de Sanden, M. C. M.; Kessels, W. M. M. Appl. Phys. Lett. 2006, 89, 81913‒81915.

(12) Langereis, E.; Heil, S. B. S.; Van de Sanden, M. C. M.; Kessels, W. M. M. Phys. Stat. Sol. C 2005, 2, 3958‒3962.

(13) Heil, S. B. S.; Langereis, E.; Roozeboom, F.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Electrochem. Soc. 2006, 153 , G956‒G965.

(14) Potts, S. E.; Keuning, W.; Langereis, E.; Dingemans, G.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Electrochem. Soc. 2010, 157, P66‒P74.

(15) Langereis, E.; Knoops, H. C. M.; Mackus, A. J. M.; Roozeboom, F.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. App. Phys. 2007, 102, 083517.

(16) Heil, S. B. S.; Langereis, E.; Kemmeren, A.; Roozeboom, F.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Vac. Sci. Technol. A 2005, 23, L5.

Page 98: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

91

(17) Langereis, E.; Heil, S. B. S.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Appl. Phys. 2006, 100, 023534.

(18) Kittl, J. A.; Opsomer, K.; Popovici, M.; Menou, N.; Kaczer, B.; Wang, X. P.; Adelmann, C.; Pawlak, M. A.; Tomida, K.; Rothschild, A.; Govoreanu, B.; Degraeve, R.; Schaekers, M.; Zahid, M.; Delabie, A.; Meersschaut, J.; Polspoel, W.; Clima, S.; Pourtois, G.; Knaepen, W.; Detavernier, C.; Afanas’ev, V. V; Blomberg, T.; Pierreux, D.; Swerts, J.; Fischer, P.; Maes, J. W.; Manger, D.; Vandervorst, W.; Conard, T.; Franquet, A.; Favia, P.; Bender, H.; Brijs, B.; Van Elshocht, S.; Jurczak, M.; Van Houdt, J.; Wouters, D. J. Microelectron. Eng. 2009, 86, 1789‒1795.

(19) Kim, S. K.; Lee, S. Y.; Lee, S. W.; Hwang, G. W.; Hwang, C. S.; Lee, J. W.; Jeong, J. J. Electrochem. Soc. 2007, 154, D95‒D101.

(20) Kwon, O. K.; Kwon, S. H.; Park, H. S.; Kang, S. W. J. Electrochem. Soc. 2004, 151, C753‒C756.

(21) Hämäläinen, J.; Hatanpää, T.; Puukilainen, E.; Costelle, L.; Pilvi, T.; Ritala, M.; Leskelä, M. J. Mater. Chem. 2010, 20, 7669‒7675.

(22) Zhu, Y.; Dunn, K. A.; Kaloyeros, A. E. J. Mater. Res. 2007, 22, 1292‒1298.

(23) Bernal Ramos, K.; Saly, M. J.; Chabal, Y. J. Coord. Chem. Rev. 2013, 257, 3271‒3281.

(24) Puurunen, R. L.; Vandervorst, W. J. Appl. Phys. 2004, 96, 7686‒7695.

(25) Baker, L.; Cavanagh, A. S.; Seghete, D.; George, S. M.; Mackus, A. J. M.; Kessels, W. M. M.; Liu, Z. Y.; Wagner, F. T. J. Appl. Phys. 2011, 109, 84310‒84333.

(26) Weber, M. J.; Mackus, A. J. M.; Verheijen, M. A.; Longo, V.; Bol, A. A.; Kessels, W. M. M. J. Phys. Chem. C 2014, 16, 8702‒8711.

(27) Biener, J.; Baumann, T. F.; Wang, Y.; Nelson, E. J.; Kucheyev, S. O.; Hamza, A. V.; Kemel, M.; Ritala, M.; Leskelä, M. Nanotechnology 2007, 18, 55303.

(28) Christensen, S. T.; Feng, H.; Libera, J. L.; Guo, N.; Miller, J. T.; Stair, P. C.; Elam, J. W. Nano Lett. 2010, 10, 3047‒3051.

(29) Christensen, S. T.; Elam, J. W.; Rabuffetti, F. A.; Ma, Q.; Weigand, S. J.; Lee, B.; Seifert, S.; Stair, P. C.; Poeppelmeier, K. R.; Hersam, M. C.; Bedzyk, M. J. Small 2009, 5, 750‒757.

(30) Feng, H.; Libera, J. A.; Stair, P. C.; Miller, J. T.; Elam, J. W. ACS Catal. 2011, 1, 665‒673.

Page 99: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

92

(31) Goldstein, D. N.; George, S. M. Thin Solid Films 2011, 519, 5339‒5347.

(32) Weber, M. J.; Mackus, A. J. M.; Verheijen, M. A.; Van der Marel, C.; Kessels, W. M. M. Chem. Mater. 2012, 24, 2973‒2977.

(33) Knaut, M.; Junige, M.; Albert, M.; Bartha, J. W. J. Vac. Sci. Technol. A 2012, 30, 01A151‒9.

(34) Geidel, M.; Junige, M.; Albert, M.; Bartha, J. W. Microelectron. Eng. 2013, 107, 151‒155.

(35) Tompkins, H. G.; Tasic, S.; Baker, J.; Convey, D. Surf. Interface Anal. 2000, 29, 179‒187.

(36) Choi, W. S.; Seo, S. S. A.; Kim, K. W.; Noh, T. W.; Kim, M. Y.; Shin, S. Phys. Rev. B 2006, 74, 26.

(37) Zollner, S. Phys. Stat. Sol. A 2000, 177, R7‒R8.

(38) Guo, S.; Hedborg, E.; Lundström, I.; Arwin, H. Thin Solid Films 1997, 293, 179‒184.

(39) Fujiwara, H. Spectroscopic Ellipsometry: Principles and Applications; John Wiley & Sons Ltd.: Chichester, UK, 2007.

(40) Tompkins, H. G.; McGahan, W. A. Spectroscopic ellipsometry and reflectometry - a user’s guide; John Wiley & Sons Ltd., 1999.

(41) Chambers, R. G. Electrons in metals and semiconductors; Chapman and Hall, 1990.

(42) Monachesi, P.; Capelli, R.; Del Sole, R. Eur. Phys. J. B 2002, 26, 159‒166.

(43) Romaniello, P.; De Boeij, P. L.; Carbone, F.; Van der Marel, D. Phys. Rev. B 2006, 73, 75115.

(44) Johnson, P. B.; Christy, R. W. Phys. Rev. B 1974, 9, 5056‒5070.

(45) Kuzmenko, A. B. Rev. Sci. Instrum. 2005, 76, 83108‒83109.

(46) Johs, B.; Hale, J. S. Phys. Stat. Sol. A 2008, 205, 715‒719.

(47) Beyene, H. T.; Weber, J. W.; Jongen, J. C. B.; Verheijen, M. A.; Van de Sanden, M. C. .; Creatore, M. Nano Res. 2012, 5, 513‒520.

(48) Crandles, D. A.; Eftekhari, F.; Faust, R.; Rao, G. S.; Reedyk, M.; Razavi, F. S. Appl. Opt. 2008, 47, 4205‒4211.

Page 100: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

93

(49) Van Heumen, E.; Lortz, R.; Kuzmenko, A. B.; Carbone, F.; Van der Marel, D.; Zhao, X.; Yu, G.; Cho, Y.; Barisic, N.; Greven, M.; Homes, C. C.; Dordevic, S. V Phys. Rev. B 2007, 75, 54522.

(50) Teyssier, J.; Kuzmenko, A. B.; Van der Marel, D.; Marsiglio, F.; Liashchenko, A. B.; Shitsevalova, N.; Filippov, V. Phys. Rev. B 2007, 75, 134503.

(51) Kuzmenko, A. B.; Van Heumen, E.; Carbone, F.; Van der Marel, D. Phys. Rev. Lett. 2008, 100, 117401.

(52) Van Hemmen, J. L.; Heil, S. B. S.; Klootwijk, J. H.; Roozeboom, F.; Hodson, C. J.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Electrochem. Soc. 2007, 154, G165‒G169.

(53) McGahan, W. A.; Johs, B.; Woollam, J. A. Thin Solid Films 1993, 234, 443‒446.

(54) Herzinger, C. M.; Johs, B.; McGahan, W. A.; Woollam, J. A.; Paulson, W. J. Appl. Phys. 1998, 83, 3323‒3336.

(55) Ashcroft, N. W.; Mermin, N. D. Solid State Physics; Saunders College: Orlando, FL, 1976.

(56) Pribil, G. K.; Johs, B.; Ianno, N. J.; Huang, H. Thin Solid Films 2004, 455‒456, 443‒449.

(57) An, I.; Nguyen, H. V; Heyd, A. R.; Collins, R. W. Rev. Sci. Instrum. 1994, 65, 3489‒3500.

(58) Hövel, M.; Alws, M.; Gompf, B.; Dressel, M. Phys. Rev. B 2009, 81, 7.

(59) Weber, J.; Hansen, T.; Van de Sanden, M.; Engeln, R. J. App. Phys. 2009, 106.

(60) Weber, J. W.; Calado, V. E.; Van de Sanden, M. C. M. Appl. Phys. Lett. 2010, 97, 91903‒91904.

(61) Mackus, A. J. M.; Verheijen, M. A.; Leick, N.; Bol, A. A.; Kessels, W. M. . Chem. Mater. 2013, 25, 1905‒1911.

(62) Henrion, W.; Rebien, M.; Antonov, V. N.; Jepsen, O. Sol. St. Phen. 1999, 67-8, 471‒476.

(63) Glantschnig, K.; Ambrosch-Draxl, C. New J. Phys. 2010, 12, 103048.

(64) Weaver, J. H.; Benbow, R. L. Phys. Rev. B 1975, 12, 3509‒3510.

(65) Yamada, Y.; Tajima, K.; Bao, S.; Okada, M.; Roos, A.; Yoshimura, K. J. Appl. Phys. 2009, 106.

(66) Weaver, J. Phys. Rev. B 1975, 11, 1416‒1425.

Page 101: Atomic layer deposition of ruthenium films : properties and surface

5. IN SITU SPECTROSCOPIC ELLIPSOMETRY DURING ALD OF Pt, Ru AND Pd

94

(67) Mayadas, A. F.; Shatzkes, M. Phys. Rev. B 1970, 1, 1382‒1389.

(68) Mayadas, A. F.; Shatzkes, M.; Janak, J. F. Appl. Phys. Lett. 1969, 14, 345‒347.

(69) Fuchs, K. Mathematical Proceedings of the Cambridge Philosophical Society 1938, 34, 100‒108.

(70) Sondheimer, E. H. Adv. Phys. 1952, 1, 1‒42.

(71) Sun, T.; Yao, B.; Warren, A. P.; Barmak, K.; Toney, M. F.; Peale, R. E.; Coffey, K. R. Phys. Rev. B 2010, 81, 155454.

(72) Gale, W. F.; Totemeier, T. C. Smithells Metals Reference Book (8th Edition).

(73) J. A. Woollam Co., I. CompleteEASE manual; 2011.

(74) J. A. Woollam Co., I. Guide to using WVASE32; p. 70.

(75) Press, W. H.; Teukolsky, S. A.; Vetterling, W. T.; Flannery, B. P. Numerical Recipes in Fortran; Teukolsky, S. A.; Vetterling, W. T.; Flannery, B. P., Eds.; Cambridge University Press, 1992; Vol. vol. 1, p. 1486.

(76) Kreibig, U.; Vollmer, M. Optical properties of metal clusters; Springer, 1995.

(77) Mann, P. S.; Lacke, C. J. Introductory Statistics; John Wiley & Sons, 2010; pp. 342‒350.

(78) Waner, S.; Costenoble, S. R. Confidence Intervals Miscellaneous on-line topics for Finite Mathematics 2e http://www.zweigmedia.com/RealWorld/finitetopic1/confint.html.

(79) Hövel, M.; Gompf, B.; Dressel, M. Phys. Rev. B 2010, 81, 35402.

(80) Soffer, S. B. J. Appl. Phys. 1967, 38, 1710‒1715.

(81) Xu, P.; Li, Z.-Y. Physica B 2004, 348, 101‒107.

Page 102: Atomic layer deposition of ruthenium films : properties and surface

95

Chapter 6 “It would be so nice if something made sense for a change.” ‒ Lewis Carroll

Dehydrogenation reactions during atomic layer

deposition of ruthenium using O2*

Abstract

The surface reactions during the thermal ALD of Ru using an metalorganic pre-cursor [CpRu(CO)2Et] and O2 gas were studied using quadrupole mass spec-trometry. The analysis of the gas phase species showed that the surface chemi-sorption of CpRu(CO)2Et resulted in the formation of combustion products such as CO2, CO, and H2O. Additionally, H2 was detected as a major surface re-action product during the metal pulse. Strikingly, during the O2 pulse virtually no H2O or other H-containing reaction products were measured. The key con-clusion of these observations is that dehydrogenation reactions occur on the Ru surface, resulting in the release of a large fraction of the precursor’s hydro-gen, predominantly in the form of H2. Based on the catalysis literature, we infer that dehydrogenation reactions can also occur for other Pt-group metal ALD processes.

* This chapter was published as Leick, N.; Agarwal, S. ; Mackus, A. J. M.; Kessels, W. M. M. Chem. Mater. 2012, 24, 3696.

Page 103: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

96

Introduction

Ultrathin conformal films of Pt-group metals such as Ru, Pt, Pd and Ir are of wide interest in a variety of applications such as electronic devices,1 catalytic converters,2 and fuel cells.3 Atomic layer deposition (ALD) of Pt-group metals has been studied to grow such films with precise thickness control even on complex nanostructured surfaces.4 A majority of the Pt-group metal ALD pro-cesses consists of cycles based on two half-reactions involving a metalorganic precursor and O2 gas.5‒9 However, to date, several questions regarding the sur-face reaction mechanisms during Pt-group metal ALD remains unanswered. According to most studies,10‒12 the Pt-group metal precursor reacts with the O atoms present in the metal’s top few monolayers. These O atoms combust a large fraction of the precursor ligands, and this reaction releases the electrons that reduce the metal atoms. Any remaining surface ligands are combusted in the subsequent O2 pulse upon dissociative chemisorption of O2, which also restores the saturated surface O-atom coverage. In addition to combustion-like reactions that produce CO2 and H2O, the catalysis literature on Pt-group metal surfaces shows that dehydrogenation of hydrocarbons can also occur on the catalytic Ru surface via C-H bond scission.13‒18 It is therefore likely that the hy-drocarbon ligands of an metalorganic precursor for ALD of Pt-group metals also undergo such dehydrogenation reactions releasing H-containing species into the gas phase. A previous study conducted within the same research group as ours19 discussed the characteristics of Pt ALD using MeCpPtMe3 and O2. In particular, the CH4 observed in the gas phase during the metal pulse could be explained by the hydrogenation of a methyl ligand by the hydrogen atoms released from dehydrogenation reactions of other precursor ligands.

In this Chapter, we report evidence for surface dehydrogenation reactions during the thermal ALD of Ru using cyclopentadienylethyldi(carbonyl)-ruthenium [CpRu(CO)2Et, shown in Fig. 1a] and O2 gas as precursors. Analysis of the gas phase species with quadrupole mass spectrometry (QMS) showed that surface chemisorption of CpRu(CO)2Et at 325 ºC resulted in the formation of combustion products such as CO2, CO, and H2O. Additionally, H2 was detected as a major surface reaction product during the metal pulse. However, during the O2 pulse, no H2O or other H-containing reaction product was detected,

Page 104: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

97

Figure 1: (a) Molecular structure of CpRu(CO)2Et [C5H5Ru(CO)2CH2CH3] (b) Schematic overview of the ALD reactor. The top and bottom gate valves are used to isolate the chamber from the pump and the O2 inlet, thereby trapping the surface reaction products. A differentially-pumped QMS is used to detect the gas phase species. (c) Schematic of the pulsing sequence for the modified Ru ALD process. The modified ALD process comprised additional steps in which the top and bottom valves were mostly closed throughout the cycle (also during CpRu(CO)2Et dosing). After each half-cycle, the gas-phase products were pumped out of the reactor for 5 min ‒ a duration that was sufficient to reach the same QMS background signal intensity as immediately before each pulse.

top valveopen

time

©JPvD

O2

substrate

CpRu(CO)2Et

top valve

bottom valve

QMS

turbopump

bottomvalve open

(b)

(c)

O2 dosing

CpRu(CO)2Etdosing

(a)

RuCO

COH3CH2C

Page 105: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

98

which is different from what was previously inferred for Ru ALD.10 Thus, the key

conclusion of these observations is that dehydrogenation reactions can occur

on the Ru surface resulting in the release of most of the metal precursor’s hy-

drogen, predominantly in the form of H2.

Experimental details

The Ru films were deposited at 325 C onto polished 150-mm-diameter Si(001) wafers in an Oxford Instruments FlexAL™ ALD reactor (see Fig. 1b).20 The deposition chamber was evacuated by a turbomolecular pump, which pro-vided a base pressure of ~10-6 Torr. The Ru precursor was kept at 90 C and va-pour-drawn into the chamber through a stainless steel gas line heated to 110 C. The Ru precursor dose was controlled by a fast-acting ALD valve. In this chamber, which is also designed for plasma-assisted ALD, O2 was delivered via a mass flow controller into the alumina tube of the plasma source. The pressure in the chamber increased to ~30 and ~120 mTorr during the metal and the O2 pulses, respectively. The chamber walls were heated to 120 ºC to avoid precur-sor condensation. The reactor surfaces and the Si wafer were conditioned prior to the QMS study by depositing at least 30 nm of Ru. After the initial nucleation phase, we obtained a constant growth per cycle of ~1 Å as measured by in situ spectroscopic ellipsometry and corroborated by X-ray reflectivity (see Chap-ter 5). No film deposition was observed at substrate temperatures below 200 C, which eliminated the possibility of significant growth on the reactor walls. A differentially-pumped QMS (MKS Vision 2000C) with a mass range of 300 amu sampled the gases in the process chamber via an orifice. The unit cycle of the ALD process reported in Chapter 4 was modified to accurately identify and quantify the surface reaction products using the QMS, which did not affect the growth per cycle and the material properties of the Ru films. A 5-min pump step separated the Ru- and the O2-pulse, during which all of the remaining gas-phase species were removed from the chamber. After the 2 s-CpRu(CO)2Et pulse, the top and bottom gate valves were closed for 5 s, during which the reaction products were trapped in the chamber (see Fig. 1c). In this way, the QMS signal at a given mass-to-charge ratio (m/z ) is proportional to the mole fraction of the corresponding neutral species in the reaction chamber being sampled. To balance the sensitivity with the fragmentation of the molecules in

Page 106: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

99

the QMS ionizer, we selected an electron energy of 40 eV. Specifically in this study, we report the QMS signals for m/z = 2, 18, 28, 32, and 44. Changes in the QMS signal were also observed for other m/z ratios between 1 and 50. Howev-er, these changes were not considered because they originated from the disso-ciative ionization of parent molecules represented by the selected m/z ratios or because their QMS intensity was at least two orders of magnitude less than that for the selected m/z ratios.

Figure 2: m/z = 2, 18, 28, 32 and 44 signal intensity measured by QMS during (a) a typical CpRu(CO)2Et dose obtained after pulsing the precursor several times without pulsing O2 in between, and during a typical ALD cycle: (b) the CpRu(CO)2Et dose followed by (c) O2 gas exposure. The top and bottom gate valves are closed throughout the scan in (a) and (b) and the top valve is open from t = 3-5 s in (c). The signals shown in (a) are assigned to ionic fragments created by dissociative ionization of CpRu(CO)2Et, whereas in (b) and (c) the signals at m/z = 2, 18, 28, 32 and 44 are attributed to predominantly H2+, H2O+, CO+, O2+ and CO2+, respectively, and they are related to the reaction products H2, H2O, CO and CO2. N.B.: The data of five continuous ALD cycles can be found in Appendix 1. 

CpRu(CO)2Et dose O2 dose(a) (c)

10-4

10-5

10-6

10-7

5 10 15 20 25 5 10 15 200Time (s) Time (s)

CpRu(CO)2Etdose

QM

S io

n cu

rren

t (a.

u.)

10-4

10-5

10-6

10-7

4 8 12 16 20 24Time (s)

(b)no ALD ALD, 1st half-cycle ALD, 2nd half-cycle

m/z = 28

m/z = 2

m/z = 18

m/z = 44

m/z = 32

m/z = 28

m/z = 28

m/z = 44

m/z = 44

m/z = 2

m/z = 18

m/z = 18

m/z = 2

Page 107: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

100

Results and discussion

During ALD, the gas-phase concentration of the surface reaction products is generally overwhelmed by the excess precursor in the chamber. The precur-sor is dissociatively ionized by the electrons in the QMS ionizer leading to ionic fragments whose m/z values can overlap with those of the surface reaction products. To separate the contributions of the reaction products formed during the ALD cycles, the m/z values of interest were recorded during Ru precursor pulsing only (Fig. 2a), and during a normal ALD cycle with alternating exposure of the surface to CpRu(CO)2Et (Fig. 2b) and O2 gas (see Fig. 2c). In this proce-dure,21 the QMS ion current intensity of the m/z ratios in Fig. 2a constitutes the signal originating from CpRu(CO)2Et in the absence of surface reactions as the spectrum in Fig. 2a is recorded after pulsing the precursor several times with-out pulsing O2 in between. This precursor signal is used as a reference for the spectrum in Fig. 2b. As discussed in Chapter 3, the substrate temperature of 325 ºC is at the onset of CpRu(CO)2Et decomposition but no impact of possible decomposition reactions on the ALD process could be observed. Nevertheless, any potential decomposition product would contribute to the precursor signal shown in Fig. 2a and would therefore be corrected for.

The QMS signals for m/z = 2, 18 and 44 in Fig. 2b are much higher than the precursor signal in Fig. 2a, which indicates that these ions originate from mole-cules produced during the surface reactions in the metal pulse, i.e. the surface reaction products. On the other hand, the signal at m/z = 28 is lower in Fig. 2b than in Fig. 2a. This can be attributed to the fact that the signal at m/z = 28 mainly arises from cracking products of the precursor (mainly CO+ and C2H4

+) which get consumed by surface reactions during the ALD cycle. The contribu-tion of reaction products to m/z = 28 in Fig. 2b is apparently not high enough to compensate for this precursor consumption, and therefore the overall signal at m/z = 28 is lower during the ALD process. In Fig. 2c, the QMS signals for m/z = 28 and 44 were attributed to the main surface reaction products.

The QMS signals at m/z = 18 and 44 during the Ru precursor pulse (Fig. 2b) were assigned to the parent molecules H2O and CO2, respectively. These com-bustion-like reaction products originate from the interaction of CpRu(CO)2Et with chemisorbed O present on the Ru surface at the beginning of the metal pulse. Comparing Figure 2b with Figure 2c, it is striking that the signal at

Page 108: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

101

m/z = 44 is much more intense during the metal pulse. In fact, the quantifica-tion of the QMS signal over the two half-cycles revealed that ~70% of the car-bon contained in CpRu(CO)2Et reacted with surface O and was released into the gas phase during the metal pulse. This observation is similar to the one made by Aaltonen et al.10 who used RuCp2 and O2 gas, and it highlights the par-ticular ability of Ru to adsorb high amounts of O on the surface.22 More details on the quantification of the QMS signal will be given in Chapter 7.

The highest QMS signal intensity in Fig. 2b was measured for m/z = 2, which was assigned to H2

+ originating predominantly from the direct ionization of H2. We also carefully examined other pathways that could result in H2

+ ions. First, we considered H2 originating from H2O dissociation. To this end, we intro-duced pure H2O vapor in the chamber and measured the QMS signal intensity at m/z = 2 and 18. The QMS signal at m/z = 2 was < 5% of the intensity at m/z = 18. The cracking pattern of H2O also matched well with the one reported by NIST.23 Besides potentially dissociating in the mass spectrometer, H2O could also dissociate on the Pt-group metal surface. However, the surface dissociation of H2O onto Ru(0001) has been debated extensively,24‒26 and most studies con-clude that H2O desorbs from these metal surfaces at temperatures between -93 and -43º C, which is well below the substrate temperature of our experiments. Therefore, we can conclude that the H2O produced in the combustion-like re-action during the Ru precursor pulse does not dissociate and cannot account for the H2

+ ions detected by the QMS. Second, we examined the reported cracking patterns of several hydrocarbon molecules with m/z ratios over the range of 16-60 amu, but the cracking of these hydrocarbons in a QMS pro-duced an intensity at m/z = 2 that is < 5% of the highest intensity for the par-ent ionization. Since m/z = 2 is the most intense QMS peak over m/z = 1-60, the cracking of hydrocarbons cannot account for the high signal intensity of H2

+ observed during the CpRu(CO)2Et cycle either. Finally, the fragmentation pat-tern of H2 measured in our QMS and the pattern observed during the Ru pulse in the ALD process for m/z = 1-3 (H+, H2

+, H3+ and isotope species) coincide,

which confirms the assignment of m/z = 2 to H2+ produced primarily from the

direct ionization of H2. The data in Fig. 2c correspond to the O2 pulse. The strong signal at

m/z = 32 is due to the direct ionization of O2, and the rise and fall of this signal indicate the duration of the O2 pulse. The QMS signal at m/z = 28 arises in part

Page 109: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

102

from the electron-impact dissociative ionization of CO2 to CO+. In the cracking pattern of CO2 measured in our QMS, the ratio of the signals at m/z = 44 and 28 was ~10. We detected a QMS signal at m/z = 28 that was well above 10% of the intensity at m/z = 44 and, therefore, conclude that CO must be produced in partial combustion reactions along with the complete combustion product CO2. Strikingly for the O2 pulse, no significant signal from H2 and H2O were ob-served at m/z = 2 and 18, respectively. This implies that most of the H atoms in the Ru precursor were released during the Ru pulse of the ALD process.

We further quantified the amount of H released during the metal pulse in the form of H2 and H2O. After background correction and a careful calibration of the QMS signal of any parent molecule XY, SXY, it was possible to determine the gas-phase number density, nXY, using the expression SXY→XY+ = k·σXY→XY+·nXY. Here, the factor k is a calibration constant that accounts for the m/z-dependent sensitivity of the QMS and σXY→XY+ is the direct ionization cross-section of the molecule at a given electron energy. Using reference measurements for H2 and H2O, we determined n

ALDH and n

ALDH2O as 1.9 × 1014 and 0.41 × 1014 cm-3, respectively,

during the Ru pulse. The experimental error in the calculation of these densities was estimated at ±10% based on the reproducibility of our measurements. However, there may be additional systematic errors, especially in the estima-tion of n

ALDH2O due to the high sticking coefficient of H2O on the unheated surfaces

of the reactor and the QMS housing, which may have led to a systematic un-derestimation. But nevertheless, these estimates of the gas phase number den-sities clearly show that > 50% of the H-containing species produced during the surface reaction of CpRu(CO)2Et were released as H2 molecules. A careful analy-sis of all m/z ratios from 1-60 amu enabled us to identify H2 as a major hydro-gen-containing reaction product during the metal pulse of our process. Along with H2O, CO and CO2, the formation of H2 during the Ru pulse was also clearly observed in the QMS study performed by Aaltonen et al.27 for Ru ALD with Ru(thd)3 and O2 gas.

In the catalysis literature, it is known that hydrocarbon species undergo dehydrogenation reactions on a Ru surface. These reactions become even more complex in the presence of surface chemisorbed O or H, which is repre-sentative of our growth surfaces during Ru ALD. Combustion, hydrogenation, and dehydrogenation reactions of linear18,28,29 and cyclic13,14 hydrocarbon spe-cies are known to occur on Ru over a wide range of temperatures. Using sur-

Page 110: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

103

face infrared spectroscopy, Ilharco et al. reported the dehydrogenation of 1-hexyne, CH3(CH2)3C≡CH, on Ru(0001) to form methylidyne, ≡CH, at ~300 K.15 Using a combination of high-resolution electron energy loss spectroscopy (HREELS) and temperature-programmed desorption (TPD), George et al.29 showed that at ~700 K, H2 had desorbed into the gas phase from a Ru(0001) surface covered with methylene, =CH2, and surface carbon was the only dehy-drogenation product. Furthermore, these authors reported that all the surface hydrogen desorbed as H2. Since our films are polycrystalline, it is also important to highlight that Wu et al.18 reported dehydrogenation of CH4 on Ru(0001) and Ru(1120) at 800 K, also using HREELS. In a subsequent study using surface infra-red spectroscopy, Garcia et al.16 showed that 0.25-0.50 ML of surface chemi-sorbed O on Ru(0001) lowered the temperature for the dehydrogenation of 3-hexyne, C2H5C≡CC2H5. This means that, unlike on Pt,15 the presence of chemi-sorbed O at the beginning of the metal pulse might slightly enhance the de-hydrogenation reactions of the precursor ligands. Initial experiments conduct-ed at different O2 gas pressure values (shown in Appendix A.2) suggest that the initial O coverage on the Ru surface only has a minor effect on the formation of gas-phase H2. Further, through TPD experiments, Hrbeck showed that an even lower O coverage, 0.15 ML, lowered the desorption temperature of H2 on Ru.17 The water-gas-shift reaction, CO2 + H2 ↔ CO + H2O, on Ru surfaces could also result in the formation of H2; however, it is unlikely to occur under our experi-mental pressure conditions.30 Therefore, we conclude that the H2 observed dur-ing the metal pulse of Ru ALD results from dehydrogenation reactions of the Et- and Cp ligands on the catalytic Ru surface.

Surface science literature comprises reports on dehydrogenation reactions occurring on Ru surfaces as well as on other Pt-group metal surfaces, such as Pt31‒33 and Ir34,35. We therefore infer that dehydrogenation reactions can also occur during other Pt-group metal ALD processes using precursors with organ-ic ligands. During the ALD process, the dehydrogenation reactions can mani-fest themselves in different ways depending on the material deposited, the precursor used as well as on other experimental conditions. As a result, differ-ent gas-phase products can be released during the metal pulse in other metal ALD processes. A good example for that is the Pt ALD process using MeCpPtMe3 and O2 gas, mentioned in the Introduction, during which CH4 was observed as a (de)hydrogenation product.11,12,19

Page 111: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

104

Conclusions

In conclusion, we propose a reaction mechanism for Ru ALD as illustrated in Fig. 3. In the metal pulse, the precursor molecules adsorb on the O-covered Ru surface. Subsequently, the hydrocarbon ligands of the adsorbed precursor

Figure 3: Schematic of the proposed reaction mechanism of Ru ALD. The large grey spheres indicate Ru surface atoms.

molecules dehydrogenate on the Ru surface followed by a release of almost all surface H atoms into the gas phase, predominantly as H2. A fraction of the lig-ands also reacts with surface O to form H2O and CO2, while some carbon re-mains on the Ru surface (~30% of the C atoms in the precursor considered in this study). During the subsequent oxygen pulse, O2 molecules dissociatively chemisorb onto Ru. The O atoms react with the surface C primarily to form CO2 and CO. At the end of the O2 pulse, all surface C is removed and the O coverage on the Ru surface is restored for the next metal pulse.

O2CO2

CO

Ru Ru

O2-pulse:O2 dissociation and combustion

of remaining carbon species

C

CpRu(CO)2Et

Ru

CO2

H2O

Ru

Ru-pulse: dehydrogenation and combustion reactions

CO

CO

H2

HC CnHmCn’Hm’C

C C

H

C5H5

CH3CH2

O O OOOOOOO O

O O OOOOOOO O OC C C C C

Page 112: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

105

The main conclusion of this work is that dehydrogenation reactions, re-ported extensively in the catalysis literature, also play a significant role in the overall reaction mechanism during Ru ALD with CpRu(CO)2Et as a precursor. From the many surface science reports it can be inferred that surface dehydro-genation reactions can also occur during ALD with other Ru-precursors and during ALD of other Pt-group metals when metalorganic precursors and an oxygen source are used.

Acknowledgments

C. A. A. van Helvoirt and W. Keuning are acknowledged for their technical support throughout this study. We thank Dr. H. C. M. Knoops from the Eindho-ven University of Technology, and Dr. L. Baker from the University of Colorado for fruitful discussions. This work was financially supported by the Dutch Tech-nology Foundation STW, which is part of the Netherlands Organisation for Sci-entific Research (NWO). Sumit Agarwal gratefully acknowledges the financial support from the NSF CAREER program (Grant No. CBET-0846923) and from the Dutch Science Foundation (FOM) for his sabbatical leave.

Appendix:

A - Additional data:

A.1 QMS scan during five consecutive Ru ALD cycles

The QMS ion currents in Fig. 2 of this chapter were only shown during one representative ALD cycle for brevity and clarity. To show the reproducibility of the measurements, the QMS signals of m/z = 2, 18, 28, 32 and 44 were record-ed during five consecutive ALD cycles and are shown in Figure A.1, for the same experimental conditions and using the same procedure as described ear-lier in this Chapter.

Furthermore, Figure A.1 also shows the signals during the 5-minute pump-ing step separating the two ALD half-cycles. During this step, the signal of the m/z recorded level off to a constant level before the CpRu(CO)2Et and O2 dose. Especially m/z = 2 and 18 settle quite fast, whereas the signals at m/z = 28 and 44 depend more on the history of the ALD reactor/gas pulsing and take more time to settle.

Page 113: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

106

Figure A.1: QMS ion current of the m/z = 2, 18, 28, 32 and 44 during five consecutive Ru ALD cycles which were attributed to H2+, H2O+, CO+, O2+ and CO2+, respectively.

A.2 Reaction products released during Ru ALD at different O2 gas pressure values

To investigate the influence of the O2 exposure on the release of H2 into the gas phase during the first half-cycle of Ru ALD (i.e. the CpRu(CO)2Et dose), we performed an experiment on a Ru seed layer in which the O2 pressure used during the second half-cycle was varied between 50 and 270 mTorr while the exposure time was maintained at 2 s. For each O2 pressure, the m/z = 2, 18, 28, 32 and 44 were scanned during 20 consecutive ALD cycles. As it was explained in the Chapter, proper reference measurements were carried out to isolate the contributions of the reaction products from the QMS signal, which we call “cor-rected” QMS signal. The peak intensity of each m/z was averaged over the 20 CpRu(CO)2Et (see Fig. A.2a) and O2 doses for each O2 pressure under investiga-tion (Fig. A.2b). According to surface science reports, increasing the O2 pressure is expected to lead to a higher O coverage on the Ru surface. At the beginning of the CpRu(CO)2Et pulse, we therefore expect that the O coverage increases as a function of O2 pressure such that more CpRu(CO)2Et molecules can react with the Ru surface, thereby releasing more CO2 and H2O via combustion reactions.

5 minCpRu(CO)2Et O2

Page 114: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

107

Figure A.2: The “corrected” QMS ion current intensity of the m/z = 2, 18 and 44 as a function of the O2 pressure used (a) during the Ru-pulse and (b) during the O2-pulse of the ALD cycle. These m/z ratios were attributed to H2+, H2O+ and CO2+, respectively. The O2 exposure time was fixed to 2 s for each measurement. The error bars represent the systematic errors in the measurements.

In agreement with this model, we observed an increased growth per cycle of Ru ALD as a function of O2 pressure (see Chapter 4 ‒ Appendix A.3). Conse-quently, during the Ru-pulse we expect an increase in O-containing reaction products and a decrease in H2 as a function of O2 pressure, which is in agree-ment with our measurements shown in Fig.A.2a. The signal at m/z = 2 (H2

+) during the Ru-pulse (Fig. A.2a) slightly decreases as a function of O2 pressure, whereas the signals of m/z = 18 (H2O+) and 44 (CO2

+) increase. In addition, the signal at m/z = 2 is clearly observed over the entire O2 pressure range investi-

Ru - pulse (a)

O2 - pulse (b)

“cor

rect

ed” Q

MS

ion

curr

ent (

a.u.

)“c

orre

cted

” QM

S io

n cu

rren

t (a.

u.)

Page 115: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

108

gated suggesting that the formation of gas-phase H2 due to surface dehydro-genation reactions is not sensitive to the surface O coverage in this study.

The surface science experiments dedicated to the study of dehydrogena-tion reactions on Ru(0001) surfaces reveal that carbon-rich species remain on the surface once the dehydrogenation is terminated. On the basis of these conclusions, we expect that under ALD conditions the top layer on the Ru sur-face at the beginning of the O2 pulse mainly contains carbon-rich species and that these block the available surface sites. Thus, the carbon-rich layer originat-ing from the surface dehydrogenation reactions should not vary much as a function of O2 pressure. Figure A.2b, however, shows that the signal for m/z = 44 (CO2

+) increases by a factor ~5 over the pressure range studied. These results are preliminary and therefore ask for further investigation, but they sug-gest that our current understanding may not be complete, yet.

References

(1) Kim, S. K.; Lee, S. W.; Han, J. H.; Lee, B.; Han, S.; Hwang, C. S. Adv. Funct. Mater. 2010, 20, 2989‒3003.

(2) Christensen, S. T.; Feng, H.; Libera, J. L.; Guo, N.; Miller, J. T.; Stair, P. C.; Elam, J. W. Nano Lett. 2010, 10, 3047‒3051.

(3) Liu, C.; Wang, C.-C.; Kei, C.-C.; Hsueh, Y.-C.; Perng, T.-P. Small 2009, 5, 1535‒1538.

(4) Comstock, D. J.; Christensen, S. T.; Elam, J. W.; Pellin, M. J.; Hersam, M. C. Adv. Funct. Mater. 2010, 20, 3099‒3105.

(5) Eom, T. K.; Sari, W.; Choi, K. J.; Shin, W. C.; Kim, J. H.; Lee, D. J.; Kim, K. B.; Sohn, H.; Kim, S. H. Electrochem. Solid-State Lett. 2009, 12, D85‒D88.

(6) Kim, S. K.; Lee, S. Y.; Lee, S. W.; Hwang, G. W.; Hwang, C. S.; Lee, J. W.; Jeong, J. J. Electrochem. Soc. 2007, 154, D95‒D101.

(7) Kukli, K.; Ritala, M.; Kemell, M.; Leskelä, M. J. Electrochem. Soc. 2010, 157, D35‒D40.

(8) Kukli, K.; Aarik, J.; Aidla, A.; Uustare, T.; Jogi, I.; Lu, J.; Tallarida, M.; Kemell, M.; Kiisler, A. A.; Ritala, M.; Leskelä, M. J. Cryst. Growth 2010, 312, 2025‒2032.

Page 116: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

109

(9) Wang, H. T.; Gordon, R. G.; Alvis, R.; Ulfig, R. M. Chem. Vap. Deposition 2009, 15, 312‒319.

(10) Aaltonen, T.; Rahtu, A.; Ritala, M.; Leskelä, M. Electrochem. Solid-State Lett. 2003, 6, C130‒C133.

(11) Christensen, S. T.; Elam, J. W. Chem. Mater. 2010, 22, 2517‒2525.

(12) Kessels, W. M. M.; Knoops, H. C. M.; Dielissen, S. A. F.; Mackus, A. J. M.; van de Sanden, M. C. M. Appl. Phys. Lett. 2009, 95.

(13) Rauscher, H.; Jakob, P.; Menzel, D.; Lloyd, D. R. Surf. Sci. 1991, 256, 27‒41.

(14) Weiss, M. J.; Hagedorn, C. J.; Weinberg, W. H. J. Am. Chem. Soc. 1999, 121, 5047‒5055.

(15) Ilharco, L. M.; Garcia, A. R.; Hargreaves, E. C.; Chesters, M. A. Surf. Sci. 2000, 459, 115‒123.

(16) Garcia, A. R.; Barros, R. B.; Ilharco, L. M. Surf. Sci. 2009, 603, 380‒386.

(17) Hrbek, J. J. Catal. 1986, 100, 523‒525.

(18) Wu, M. C.; Xu, Q.; Goodman, D. W. J. Phys. Chem. 1994, 98, 5104‒5110.

(19) Mackus, A. J. M.; Leick, N.; Baker, L.; Kessels, W. M. M. Chem. Mater. 2012, 24, 1752‒1761.

(20) Van Hemmen, J. L.; Heil, S. B. S.; Klootwijk, J. H.; Roozeboom, F.; Hodson, C. J.; van de Sanden, M. C. M.; Kessels, W. M. M. J. Electrochem. Soc. 2007, 154, G165‒G169.

(21) Knoops, H. C. M.; Langereis, E.; van de Sanden, M. C. M.; Kessels, W. M. M. J. Vac. Sci. Technol. A 12AD, 30, 01A101.

(22) Böttcher, A.; Niehus, H. J. Chem. Phys. 1999, 110, 3186‒3195.

(23) NIST Chemistry WebBook,. NIST Standard Reference Database Number 69.

(24) Mugarza, A.; Shimizu, T. K.; Cabrera-Sanfelix, P.; Sánchez-Portal, D.; Arnau, A.; Salmeron, M. J. Phys. Chem. C 2008, 112, 14052‒14057.

(25) Feibelman, P. J. Science 2002, 295 , 99‒102.

(26) Tatarkhanov, M.; Ogletree, D. F.; Rose, F.; Mitsui, T.; Fomin, E.; Maier, S.; Rose, M.; Cerdá, J. I.; Salmeron, M. J. Am. Chem. Soc. 2009, 131, 18425‒18434.

(27) Aaltonen, T.; Ritala, M.; Arstila, K.; Keinonen, J.; Leskelä, M. Chem. Vap. De-position 2004, 10, 215‒219.

Page 117: Atomic layer deposition of ruthenium films : properties and surface

6. DEHYDROGENATION REACTIONS DURING ALD OF Ru USING O2

110

(28) Ciobica, I. M.; Frechard, F.; van Santen, R. A.; Kleyn, A. W.; Hafner, J. Chem. Phys. Lett. 1999, 311, 185‒192.

(29) George, P. M.; Avery, N. R.; Weinberg, W. H.; Tebbe, F. N. J. Am. Chem. Soc. 1983, 105, 1393‒1394.

(30) Ollis, D. F. J. Catal. 1981, 102, 90‒102.

(31) Zaera, F. Surf. Sci. 1992, 262, 335‒350.

(32) Fairbrother, D. H.; Peng, X. D.; Trenary, M.; Stair, P. C. J. Chem. Soc., Faraday Trans. 1995, 91, 3619‒3625.

(33) Avery, N. R. Surf. Sci. 1984, 146, 363‒381.

(34) Colindres, S. C.; García, J. R. V.; Antonio, J. A. T.; Chavez, C. A. J. Alloy Compd. 2009, 483, 406‒409.

(35) Johnson, D. F.; Weinberg, W. H. J. Chem. Soc., Faraday Trans. 1995, 91, 3695‒3702.

Page 118: Atomic layer deposition of ruthenium films : properties and surface

111

Chapter 7 “At some point in life the world’s beauty becomes enough. You don’t need to photo-

graph, paint or even remember it. It is enough.” ‒ Toni Morrison

Catalytic combustion reactions during atomic layer deposition of Ru studied by 18O2 isotope labeling* Abstract

The surface reaction products liberated during the atomic layer deposition (ALD) of Ru from (C5H5)Ru(CO)2(C2H5) and 18O2 were quantitatively analyzed us-ing quadrupole mass spectrometry (QMS). The gas-phase reaction products during the Ru precursor pulse were CO2, CO, H2, and H2O, while during the O2 pulse primarily CO2 and CO were produced. Approximately 70% of the C atoms and ~100% of the H atoms contained in the Ru precursor were released during the Ru pulse of the ALD process. From these observations, and on the basis of the surface science and catalysis literature, we conclude that the complex sur-face chemistry during Ru ALD can be described by catalytic combustion reac-tions. These reactions consist of the dissociative chemisorption of the Ru pre-cursor-ligands on the Ru surface during the metal pulse. These hydrocarbon ligands undergo dehydrogenation and combustion reactions on the catalytic metal surface in the presence of surface O formed due to the dissociation of O2 molecules in the previous O2 pulse. We postulate that the carbon-rich species produced due to the dehydrogenation of the Ru-precursor hydrocarbon-ligands self-limit precursor chemisorption by blocking adsorption sites. The use of the 18O2 isotope also unambiguously shows that a certain fraction of the pre-cursor’s carbonyl ligands dissociate on the Ru surface.

* This chapter was published as Leick, N.; Agarwal, S. ; Mackus, A. J. M.; Potts, S. E.; Kessels, W. M. M. J. Phys. Chem. C 2013, 117, 21320.

Page 119: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

112

Introduction

Thin films and nanostructures of Pt-group metals, such as Pt, Ru, Ir, and Pd, have been widely studied over the last decade for a variety of applications, such as power generation,1 electricity and information storage,2‒4 sensing,5‒7 and control of exhaust pollution.8‒10 Atomic layer deposition (ALD) has emerged as a promising technique to deposit these metallic thin films and nanostructures with accurate thickness control onto demanding substrate to-pologies.11‒13 In particular, Ru ALD is of interest in applications such as hetero-geneous catalysis14,15 and memory devices.4,16‒20 Similar to most other metal ALD processes, nucleation delays of 70-250 cycles have been reported during ALD of Ru on a variety of substrate surfaces.21‒23 Novel Ru ALD precursors24‒29 have been developed to reduce this long nucleation delay, and to enhance the growth per cycle (GPC), which is typically <1 Å/cycle.19,21,30‒33 Regardless of the type of precursor employed, a better understanding of the relevant surface re-action mechanisms during metal ALD is crucial for improving the deposition processes, and tailoring the resulting properties of Ru and other Pt-group metal films.

In one of the first studies of Ru ALD with ruthenocene (RuCp2) and O2 gas, Aaltonen et al.,34 using quadrupole mass spectrometry (QMS), identified CO2 and H2O as the primary surface reaction products in both half cycles. Later, similar observations were made for Pt ALD using (methyl-clopentadienyl)trimethylplatinum (MeCpPtMe3) and O2 gas.34‒36 Based on these and several subsequent studies,37,38 it has now been accepted that combustion-like reactions govern the surface chemistry in both half cycles during ALD of Pt-group metals. Due to their catalytic nature, these Pt-group metal surfaces are able to break the bonds of adsorbed species,39‒44 such as hydrocarbons and O2: the surface O, produced due to the dissociative chemisorption of O2 gas, reacts with the precursor’s organic ligands to produce combustion-like products, such as CO2 and H2O.34 Such oxidation reactions of hydrocarbons are very important in the overall surface reaction mechanism since the electrons that are released in these reactions are necessary to reduce the Ru or Pt atoms in the +2 and +4 oxidation state, respectively, to their metallic state (oxidation state 0).45

More recently, it has been shown that CO2 and H2O are not the only surface reaction products during ALD of Pt-group metals.35,37,46,47 In Chapter 6, H2 was

Page 120: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

113

identified as a primary reaction product during the Ru precursor pulse, besides CO2 and H2O. Based on the surface science and catalysis literature,43,44,48 we concluded that H2 was released in the Ru precursor cycle due to surface dehy-drogenation reactions of the precursor’s hydrocarbon ligands. In addition, dur-ing the ALD of Pt with MeCpPtMe3 and O2 gas, using gas-phase infrared spec-troscopy and QMS, CH4 was identified as a surface reaction product during the metal precursor pulse.35,36,47 Similarly, based on the catalysis literature, it was concluded that surface dehydrogenation of the Me and MeCp ligands in the Pt precursor was the most likely source of surface H, which reacts with other Me ligands to form CH4.46,47 Thus, these observations suggest that additional reac-tions of hydrocarbon ligands, other than complete combustion to CO2 and H2O, must occur on Pt-group metal surfaces due to their catalytic nature. Since most Pt-group ALD processes, including Ru, utilize metalorganic precursors and O2 gas,22,25,26,49 it is important to gain a deeper insight into the interaction of these precursors with catalytic metal surfaces.

The surface interactions of adsorbates on Pt-group metals has been exten-sively studied in the context of heterogeneous catalysis and fuel cell applica-tions.41,50‒52 For example, the interaction of CO and O2 with the Ru(0001) surface has been studied over a wide range of conditions from ultra-high vacuum (~10-10-10-9 Torr) to high pressure (~103 Torr), and up to temperatures >500 K.53‒55 The results of these studies are summarized below, which are rele-vant to the pressure (~1 Torr) and temperature regime (~600 K) used for Ru ALD processes.53‒60 In particular, these studies suggest that the surface oxygen coverage after the O2 pulse can significantly affect the surface reactions of metalorganic Ru precursors. Figure 1 shows a summary of the literature on Ru(0001) surface O coverage, θO, as a function of O2 exposure.59 It has been reported that O-chemisorption extends over four different ordered overlayer phases on Ru(0001): p(22)-O at θO = 0.25 monolayer (ML), p(21)-O at θO = 0.5 ML, p(22)-3O at θO = 0.75 ML, and p(11)-O at θO = 1 ML, where p refers to the primitive adlayer phase.59‒62 According to these studies, Ru also has the ability to adsorb atomic O beyond 1 ML, where atomic O dissolves into the

Page 121: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

114

Figure 1: O coverage as a function of O2 exposure in Langmuirs (1 L = 1×10-6 Torr‧s) for a Ru(0001) surface at 600 K, and the expected surface configurations of metastable O at coverages of 0.25 ML, 0.5 ML, 0.75 ML, and 1 ML. Note, that p refers to the primi-tive adlayer unit cell. The values have been compiled from Refs.59‒61.

subsurface region.54,58,59,63,64 At coverages beyond θO = 3 ML, RuO2(110) micro-regions begin to form, which can coexist with p(11)-O domains on Ru(0001).56,57,65 In fact, RuO2 has been identified as the active phase in the oxida-tion of CO at high pressures and temperatures.55,66,67 It is also known from ca-talysis that dehydrogenation reactions occur on catalytic metal surfaces in which C‒H bonds are broken, and the atomic H from the hydrocarbon is trans-ferred onto the metal surface. Eventually, surface H atoms recombine and de-sorb as H2, leaving carbon-rich surface dehydrogenation products on the metal surface.40,41,68 This overall reaction can be written as, CxHy

* CxHy-2* + H2 (gas),

where * denotes surface species.69 Thus, based on previous literature, we expect that, depending on the surface O coverage prior to the Ru precursor pulse, dif-ferent surface reactions could occur: combustion-like products would be dom-inant for a high surface O coverage and dehydrogenation reaction products would be more likely for a low surface O coverage.

In this Chapter, we present our study of the surface reaction mechanisms during ALD of Ru from the heteroleptic precursor cyclopentadienylethyl-di(carbonyl)ruthenium (CpRu(CO)2Et, see Chapter 6, Fig. 1a) and 18O2 using QMS.

10 0 10 1 10 2 10 3 10 4 10 5 10 6 10 70.0

0.5

1.0

1.5

2.0

O c

over

age

(ML)

O2 exposure (L)

Adsorbed OAdsorbed OSubsurface OSubsurface O

1 ML p(1x1) ORu(0001) hcp

at 600K

0.5 ML p(2x1) O

0.25 ML p(2x2) O0.75 ML p(2x2) 3O

Page 122: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

115

Isotope labeling enabled us to differentiate between the 16O in the Ru precur-sor’s carbonyl ligands from the 18O delivered in the O2 half-reaction cycle. In addition, we were also able to distinguish between C18O and C2H4 in the QMS, since both species can be released during the surface reactions of the Ru pre-cursor. To quantify the gas-phase concentration of the surface reaction prod-ucts, we have carefully calibrated the QMS ion current for different mass-to-charge (m/z) ratios. Based on this analysis, we conclude that the Ru surface cat-alyzes the dissociation of adsorbed species ‒ hydrocarbons in the Ru-precursor half-cycle, and 18O2 in the subsequent half-cycle ‒ thereby facilitating the com-bustion of the fragmented precursor ligands. Such catalytically assisted oxida-tion reactions of hydrocarbons are generally referred to as catalytic combustion reactions in the field of heterogeneous catalysis:70‒72 herein, we show that they also play a significant role under the experimental conditions for Ru ALD.

Experimental details

ALD reactor and film deposition

The Ru films were deposited at ~325 C onto polished 150-mm-diameter Si(001) wafers in an Oxford Instruments FlexAL™ ALD reactor73 (see reactor schematic in Chapter 6, Fig. 2b) suited for thermal and plasma-assisted ALD. The deposition chamber was evacuated by a turbomolecular pump, which provided a base pressure of ~10-6 Torr. The CpRu(CO)2Et (SAFC Hitech®, purity > 99.999%) precursor was kept at 90 C and was vapor-drawn into the chamber through a stainless steel tube heated to 110 C, and pulsed by an ALD valve. In this chamber, 18O2 (Linde Gas, purity > 99.8%) was delivered via a mass flow controller into the alumina tube of the inductively coupled plasma source that was not operated in the experiments described here. The pressure in the chamber increased to ~30 and ~120 mTorr during Ru and 18O2 pulsing, respec-tively. The chamber walls were heated to 120 C to avoid precursor condensa-tion. The reactor surfaces and the Si wafer were conditioned prior to the QMS study by depositing at least 30 nm of Ru. Since almost no film deposition was detected at substrate temperatures < 200 C, we expected negligible Ru depo-sition on the reactor walls. However, the 225-mm-diameter substrate heater can radiatively heat the lower walls to a temperature at which Ru growth could occur and, therefore, it was difficult to accurately estimate the total surface area

Page 123: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

116

on which Ru was actually deposited. In addition, a significant part of the sub-strate-stage was not covered by the relatively smooth Si wafer and directly ex-posed to the ALD precursors, which could significantly increase the area for deposition due to surface roughness.

The Ru film properties deposited by this ALD process have been reported in Chapter 4, and only the properties relevant for this study are briefly summa-rized below. The Ru films had a mass density of 12.2 ± 0.5 g.cm-3, which is close to the Ru bulk density (12.4 g.cm-3). From Rutherford back-scattering (RBS) and X-ray reflectivity, 7.1×1014 cm-2 of Ru atoms were deposited in one ALD cycle. Furthermore, the average areal density of Ru atoms on the surface based on this RBS data is 1.7×1015 cm-2, which is very similar to the value reported on a Ru(0001) surface.59 X-ray diffraction (XRD) showed that the films were polycrys-talline with no preferred orientation. RBS and time-of-flight secondary ion mass spectrometry detected traces of oxygen in the films; however, no RuOx peak was observed in XRD. The electrical resistivity of the films was as low as 16 µΩ.cm for a film thickness of 30 nm. The surface roughness measured by atomic force microscopy was ~5 nm for a ~15-nm-thick film. This roughness value is characteristic for Ru ALD films using O2 gas,25,74,75 although it may be too high for several applications of Ru ALD.

Quadrupole mass spectrometry

A differentially-pumped QMS (MKS Vision 2000C) with a mass range of 300 amu sampled the gases in the process chamber via an orifice. To balance the sensitivity with the level of fragmentation of the molecules in the QMS ion-izer, an electron energy of 40 eV was selected. As in Chapter 6, the unit cycle of the ALD process reported in Chapter 4 was modified to accurately identify and quantify the surface reaction products using the QMS. These modifications did not affect the GPC and the material properties of the Ru films. Specifically, in the modified cycles, a 5-min evacuation step separated the Ru and the 18O2 pulses, during which all of the remaining gas-phase species were pumped out of the chamber. This was necessary to establish a constant baseline for the QMS ion current at any given m/z value. In addition, the top and bottom gate valves of the vacuum chamber were closed for 7 s (see Chapter 6, Fig. 1c) dur-ing which time we introduced 2-s CpRu(CO)2Et and 18O2 pulses. Since the gate

Page 124: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

117

valves were closed the reaction products were trapped within the chamber. The QMS signal at a given m/z was thus proportional to the mole fraction of the corresponding neutral species in the reaction chamber.

During ALD, the gas-phase concentration of the surface reaction products can be significantly lower than that of the excess precursor in the chamber. In the QMS, the precursor is dissociatively ionized by the electrons in the ionizer leading to ionic fragments whose m/z values can overlap with those of the re-action products. To select the relevant m/z values that result from fragmenta-tion of the reaction products, and not the parent precursor molecule, we first performed a mass scan over the range of m/z = 1-60 with a low time resolution during a reference measurement (CpRu(CO)2Et pulsing only) and during the ALD process.76 Based on the difference in intensity in these mass spectra (see Figure A.1 in the Appendix A.1), no signal from reaction products could be de-tected above the noise level beyond m/z = 48. By comparing the fragmenta-tion patterns of the measured m/z values with those published in the NIST da-tabase,77 we could assign the different peaks to specific parent molecules pro-duced during surface reactions. We also checked for potential artifacts in the signal intensity due to the limited resolution of the QMS and, if present, took this into consideration in the assignment of the m/z values to particular gas-phase species. For example, the QMS signal at a given m/z value has an asym-metric shape with a tail towards the lower m/z value.78 Thus, when the peak intensity at a certain m/z is very high compared to the intensity at the neigh-boring lower m/z, it can result in an erroneous determination of the QMS signal intensity at this lower m/z value. In such cases, we carefully analyzed the peak shape to ascertain the true origin of the signal. Following the identification of the relevant m/z values to track different surface reaction products, we record-ed the data in a multiplexing mode where the QMS signal intensity for the rele-vant m/z values was tracked simultaneously during the ALD cycles.

To perform a quantitative analysis of the QMS data, it was necessary to cal-ibrate the QMS signal intensities. The QMS signal S for any parent molecule XY is related to its gas-phase number density nXY as,79

S = k·σXY→XY+·nXY (Eq. 1)

where S is the measured ion current after subtracting the background (see in Appendix A.2), σXY→XY+ is the electron-energy dependent direct ionization

Page 125: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

118

cross-section of the molecule XY available in the literature,80 and the factor k accounts for the m/z-dependent sensitivity of the QMS. To eliminate k and σXY→XY+, the QMS signal intensity can be directly related to the number densi-ty of the species using the following expression,

SALD

Sref = nXY

ALD

nXYref (Eq. 2)

with Sref being the QMS reference signal for species XY at a known reference pressure, and nXY

ref the associated number density of that species. However, Eq. 2 requires QMS reference measurements for every species identified as a reaction product. This was not possible for practical reasons, especially with the use of 18O2, which results in reactions products that contain the isotope. Therefore, the factor k was determined by calibrating the QMS signal over the range of 18 < m/z < 40 using several gases (see Figure A.2 in the Appendix A.3) and tak-ing the appropriate cross-sections into account80. The factor k for any m/z value within or slightly outside this range was interpolated or extrapolated, and Eq. 1 was used to relate the QMS signal of each reaction product to its gas-phase number density. However, H2 generally does not follow this trend due to the inefficient pumping of H2 by turbomolecular pumps and, therefore Eq. 2 was used for H2.

To check our quantification procedure for the QMS signal, first, we estimat-ed the number of gas-phase C and H atoms per unit volume, Nx (x = C or H). Assuming that all C and H contained in CpRu(CO)2Et is converted to gas-phase reaction products after one complete ALD cycle, Nx can be expressed as,

NX= mX·σRu·Ad

V (Eq. 3)

where mx is the number of atoms of x (x = C or H) in CpRu(CO)2Et (mC = 9, mH = 10), σRu= 7.1×1014 cm-2 is the number of Ru atoms deposited in one ALD cycle, Ad is the surface area for Ru deposition (Ad = 398 cm2 based on a 225-mm-diameter heated area), and V is the volume of the ALD chamber with the gate valve to the pump closed (V = 21 ± 1 l ) . From these values, we obtain NC = 1.2×1014 cm-3 and NH = 1.3×1014 cm-3. Second, we accounted for the num-ber density of all C- and H-containing reaction products identified in the gas-phase with the QMS, and using the corresponding stoichiometry of these spe-cies, we determined Nx (x = C or H): the values obtained via the quantification

Page 126: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

119

of the QMS signals (see Table I) might be overestimated by a factor of ~4. The largest error probably lies in the underestimation of the active surface area used in Eq. 3 above, which was simply based on the area of the heated sub-strate. The actual surface area in the Ru ALD process was probably greater than that calculated from the diameter of the substrate heater due to two reasons ‒radiative heating of surrounding surfaces and the surface roughness of the substrate heater.

Results

Identification and quantification of the reaction products

For clarity, the QMS data shown in Figure 2 were separated into two panels to distinguish between the C- and H-containing reaction products. During the Ru pulse, the contributions of the reaction products formed due to ALD surface reactions were isolated by first recording the m/z values of interest during the pulsing of CpRu(CO)2Et several times, without pulsing 18O2 in between. Conse-quently, the data shown in Fig. 2a-b provides a reference signal level for each of the m/z values in the absence of ALD surface reactions. This reference signal was recorded by interrupting a sequence of normal ALD cycles consisting of alternating exposure of the surface to CpRu(CO)2Et and 18O2 to ensure that the surface during the reference measurement was representative of ALD condi-tions.

The QMS signal for a regular ALD cycle during the Ru pulse is shown in Fig. 2c-d. For m/z = 2, 18, 20, 30, 44, 46 and 48, the signal is much higher than the corresponding reference level in Fig. 2a-b, indicating that these ions originate from molecules produced via surface reactions during the Ru pulse. Conse-quently, the difference between the signals in Fig. 2a-b and Fig. 2c-d for each m/z value directly provides the QMS ion current measured due to the surface reaction products produced during the Ru pulse.76 The signal at m/z = 2 was assigned to H2

+, and m/z = 18 and 20 represent H216O+ and H2

18O+, respectively.

Page 127: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

120

Figure 2: QMS ion currents for the m/z ratios relevant to Ru ALD. Figure (a)-(b) corre-spond to a typical CpRu(CO)2Et dose with no preceding O2 dose; (c)-(d) correspond to the CpRu(CO)2Et dose, and (e)-(f) correspond to the O2 dose during a typical ALD cycle. The figures in the upper panel (a), (c) and (e), show H-containing species: m/z = 2 (H2+), 18 (H216O+ or 18O+), and 20 (H218O+). The lower panels, (b), (d) and (f), show the C-containing species: m/z = 28 (C2H4+ and C16O+), 30 (C18O+), 36 (18O2+), 44 (C16O2+), 46 (C16O18O+), and 48 (C18O2+). The reference signals for the m/z values measured during the Ru precursor pulse in (c) and (d) were taken from (a) and (b), respectively.

CpRu(CO)2Et dose O2 doseCpRu(CO)2Et dose

CpRu(CO)2Et dose O2 doseCpRu(CO)2Et dose

(a) (c) (e)

(b) (d) (f )

CpRu(CO)2Et onlyno ALD

Ru-pulseALD

O2-pulseALD

H - containing reaction products

C - containing reaction products

m/z = 28

m/z = 2

m/z = 18

m/z = 20

m/z = 2

m/z = 18

m/z = 20

m/z = 2

m/z = 18

m/z = 20

m/z = 36

m/z = 30

m/z = 28m/z = 48m/z = 46m/z = 30

m/z = 44

m/z = 36

m/z = 48

m/z = 30m/z = 46m/z = 28

m/z = 44

m/z = 44

m/z = 46

10-4

10-5

10-6

10-7

10-4

10-7

10-6

10-5

10-5

10-4

10-7

10-4

10-5

10-7

10-6 10-6

Page 128: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

121

We also considered the possibility that m/z = 18 corresponds to 18O+ produced from the dissociative ionization of 18O2, 18O16O, or H2

18O during the Ru pulse (Fig. 2c-d). However, the concentration of these species is expected to be neg-ligible since no QMS signal for m/z = 36 (18O2

+) or 34 (18O16O+) was detected, and 18O+ only accounts for < 1% of the cracking pattern of H2

18O.81 The signal at m/z = 30 was assigned to C18O+, and m/z = 44, 46 and 48 were assigned to C16O2

+, C16O18O+ and C18O2+, respectively. The ionized carbonyl and ethyl ligands

can both contribute to m/z = 28 (C16O+ or C2H4+). In Fig. 2d, m/z = 28 is lower

than the reference level in Fig. 2b. This can be attributed to the fact that the signal at m/z = 28 mainly arises from cracking products of the precursor which are consumed by surface reactions during the ALD cycle. The contribution of reaction products to m/z = 28 in Fig. 2d is apparently not high enough to compensate for this precursor consumption, and therefore, the overall signal at m/z = 28 is lower during the ALD process. Because m/z = 28 could not be at-tributed to a single parent molecule, no clear conclusion concerning C16O could be drawn as a reaction product during the Ru pulse. However, the iso-tope labeling enables the distinction between the C2H4

+ and C18O+, as m/z = 30 can clearly be assigned to C18O+ only.

The reference measurement for the 18O2 gas was performed using the same procedure as for the CpRu(CO)2Et half-cycle, but this reference measure-ment did not lead to any significant signal for the m/z values scanned, other than 18O2

+ and 18O+. Therefore, the QMS signal prior to 18O2 injection (i.e. t < 4s in Fig. 2e-f) was taken as the reference level for each m/z value measured, and m/z = 20, 28, 30, 44, 46, and 48 were identified as the species contributing to the reaction products due to ALD surface reactions. During the 18O2 pulse, the signal at m/z = 20 represents H2

18O+ (similar to the Ru pulse), but unlike during the Ru pulse, the signal at m/z = 18 can now be predominantly attributed to 18O+ originating from the dissociative ionization of 18O2, rather than H2

16O+. The signal at m/z = 28 and 30 can be attributed to C16O+ and C18O+, respectively, and m/z = 44, 46 and 48 to C16O2

+, C16O18O+ and C18O2+, respectively. We will

show later that, since nearly all H is released during the Ru pulse, detection of C2H4

+ at m/z = 28 is unlikely.

Page 129: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

122

Table I: Gas-phase number densities of CO2, CO, H2, and H2O in the ALD reactor calcu-lated from the QMS data measured during the Ru precursor- and 18O2 pulse. The uncer-tainty in the number densities is estimated as ~10%. The signal at m/z = 28 for the Ru pulse during ALD was lower than during the Ru precursor-pulsing-only-condition and, therefore, no number density is included in the table.

Densities ( 1014 cm-3)

Ru pulse 18O2 pulse Total

CO2 C18O2 m/z = 48 1.9 0.6 2.5 C16O18O m/z = 46 1.6 0.2 1.8 C16O2 m/z = 44 negligible 0.3 0.3

CO C18O m/z = 30 0.27 0.18 0.45 C16O m/z = 28 unknown 0.11 0.11 Total C 3.77 1.39 5.16

H2 m/z = 2 2 negligible 2

H2O H218O m/z = 20 0.1 negligible 0.1

H216O m/z = 18 0.26 negligible 0.26

Total H 4.72 negligible 4.72

Summarizing, we can state that from the assigned m/z values it is possible to identify the main gas-phase reaction products during each ALD half-cycle: during the Ru pulse, H2, H2

18O, H220O, C16O2, C18O2, and C18O were released as the

gas-phase reaction products, whereas, C16O2, C16O18O, C18O2, C16O, and C18O were measured as the most dominant reaction products during the O2 pulse. Table I gives an overview of the gas-phase number densities of each reaction product identified above, including isotopes, using the method described in the Experimental Section.

On the basis of this overview in Table I, we can compare the type and amount of reaction products liberated during the Ru- and O2 pulses. Approxi-mately 70% of the carbon in the metal precursor was released during the Ru pulse in the form of CO2 and CO, and ~ 30% during the O2 pulse. In contrast, ~100% of the hydrogen was released during the Ru precursor pulse as H2 (85%) and H2O (15%). Previously, Aaltonen et al.34 studied the ALD of Ru using RuCp2 and O2 gas, and reported that during the RuCp2 pulse 50% of the carbon was

Page 130: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

123

released in the form of CO2 and ~90% of hydrogen was liberated in the form of H2O. Thus, the reaction products observed in our ALD process are in contrast with those reported by Aaltonen et al..

Further analysis of the QMS data reveals that a certain fraction of the CO ligands in the Ru precursor dissociated on the Ru surface during the metal pre-cursor pulse. Specifically, species containing 16O, such as H2

16O (m/z = 18) and C16O2 (m/z = 44) were measured during the Ru pulse. Due to the use of the 18O2 isotope, the main source for 16O in this ALD process is the Ru precursor’s car-bonyl ligands (C16O). Thus, the most likely pathway for the formation of H2

16O and C16O2 during the Ru precursor cycle is via combustion-like reaction of the hydrocarbon ligands with surface 16O atoms produced via the dissociative chemisorption of the C16O ligands.

We have also estimated the surface O coverage (θO) on Ru after the 18O2 pulse. To obtain θO, first, we assume all 18O atoms incorporated on the Ru sur-face after the 18O2 pulse react with the Ru precursor ligands in the subsequent Ru precursor pulse. Second, we have determined the gas phase number densi-ty of each 18O-containing species (C18O, C18O2, C18O16O, and H2

18O) released dur-ing the Ru precursor cycle using the same procedure described in the Experi-mental Section (see Eq. 1). Third, using the stoichiometry of each 18O-containing species, we can also obtain the number of 18O atoms per unit vol-ume released into the gas phase, NO = 5.77 × 1014 cm-3. Finally, using a similar expression as Eq. 3, we obtained an 18O areal density on Ru, σO ~30.4 × 1015 cm-2. However, as mentioned in the Experimental Section, the values obtained via the quantification of the QMS signals might be overesti-mated by a factor of up to ~4. Taking this correction factor and the areal densi-ty of our Ru films (1.7 × 1015 cm-2), the estimated 18O coverage on Ru was ~4.5 ML.

Discussion

Dehydrogenation and combustion reactions

In Chapter 6, we showed that surface dehydrogenation reactions of the Et and Cp ligands in the Ru precursor occur on the catalytic Ru surface, and nearly all of the precursor’s hydrogen was released during the Ru pulse as H2. This ob-

Page 131: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

124

servation was in agreement with experimentally observed dehydrogenation reactions of hydrocarbon molecules conducted on Ru surfaces under similar conditions.40,41,43,48,82

The fraction of CO2 and H2O produced during the Ru precursor pulse from the combustion of the organic precursor ligands with the surface 18O atoms is higher than that reported for the metal precursor pulses in Pt and Ir ALD pro-cesses. During Pt ALD from MeCpPtMe3, the CO2 produced during the Pt pulse was only ~6% of the total amount of CO2 produced, with the remaining ~94% being released during the O2 pulse.36 Similarly, for Ir ALD using Ir(acac)3 and O2 gas, Knapas et al. measured 14% of the CO2 and 57% of the H2O being released into the gas phase during the Ir pulse.38 Using the same precursors, Christensen et al. measured 31% of the CO2 during the Ir pulse.35 According to surface sci-ence investigations, the saturation coverage of O on Pt(111) and on Ir(110) sur-faces is ~0.25 ML (at 300 ºC) and ~1 ML (at 300 ºC), respectively.83,84 As men-tioned in the Introduction, Ru is able to accommodate a higher O coverage, even > 1 ML, under the ALD conditions used in this study. Therefore, in this Ru ALD process, it is likely that the high amounts of CO2 and H2O formed during the Ru pulse are due to the higher surface O coverage on Ru compared to Pt and Ir. In line with these findings, Aaltonen et al.34 showed that the CO2 and H2O measured in the gas-phase during the RuCp2 pulse was equivalent to an O coverage of ~3 ML on their Ru surface which is close to our established 4.5 ML of O coverage.

Catalytic combustion reactions

Dehydrogenation and combustion reactions can be regarded as two dif-ferent kinds of chemical reactions independent of each other, which is very illustrative, but in reality (catalysis or ALD) these processes may be harder to separate. According to the catalysis literature, dehydrogenation and combus-tion reactions are both intermediate steps in so-called catalytic combustion reactions. This term, generally used in the field of heterogeneous catalysis, re-fers to catalytically-assisted oxidation reactions of gaseous hydrocarbon species on Ru surfaces.70‒72 In practical catalysis applications, the temperature (300 K < T < 1000 K) and pressure (750 Torr < p < 7500 Torr) ranges used are higher than those used in ALD, but laboratory-scale experiments nevertheless

Page 132: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

125

CO2

H2O

H2 CO

CpRu(CO)2Et pulseadsorption of the precursor

and dehydrogenation of the ligands at the surfaceCO

CO

C5H5

CH3CH2

Ru

CO2

CO O2 pulse

complete and partial oxidation of the dehydrogenation products

dissociative chemisorption of O2 molecules at the surface

complete and partial combustionof C-rich surface species

Ru

O OOOOOO

Ru

C-rich species COC CCC C C

Ru

C-rich speciesC-rich species COC CCC C C

Ru

O OOOOOO

Figure 3: Schematic of the catalytic combustion reactions driving the surface process-es during CpRu(CO)2Et and O2 pulses of the Ru ALD process. At the start of the CpRu(CO)2Et pulse, the initial surface consists of a Ru layer with an O coverage, θO = 0.75-3 ML, determined from our experiments. As CpRu(CO)2Et adsorbs onto the sur-face, the organic ligands are dehydrogenated, and the dehydrogenation products react with each other or with surface O, releasing H2, CO2, H2O, and CO into the gas phase. As a result of the dehydrogenation reactions, a layer of carbon-rich species re-mains on the surface. As the O2 dissociatively chemisorbs onto this surface during the O2 pulse, it combusts the carbon-rich species, releasing CO2 and CO into the gas phase. It also restores the O coverage on the Ru surface before the next Ru precursor pulse.

cover the pressure and temperature regime relevant to ALD. The catalytic combustion reactions comprise two processes that can occur simultaneously: 1) the dissociation of incoming species on the catalytic surface, and 2) the oxida-tion of C- and H-rich surface species. Usually, the catalytic combustion reac-tions are divided into two distinct regimes: the fuel-rich combustion regime and the fuel-lean combustion regime. Analogous to these regimes, the metal precursor pulse during Pt-group-metal ALD is characterized by an excess of metal precursor (with its organic ligands), and can therefore be regarded as the fuel-rich regime, whereas the second half-cycle is best represented by the fuel-

Page 133: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

126

lean conditions since O2 is the predominant species compared to the carbon-rich species on the metal surface.

Applying the concept of catalytic combustion reactions to Ru ALD means that in each ALD half-cycle, species dissociatively chemisorb onto the surface, thereby facilitating the combustion reactions. The reaction mechanism for Ru ALD presented in this study is preliminary, and is based solely on our QMS measurements combined with literature reports from catalysis and surface sci-ence. Figure 3 aims at simplifying the complex Ru ALD chemistry using the concept of catalytic combustion reactions, which we envision to proceed as follows. During the Ru precursor pulse, the precursor’s hydrocarbon ligands are largely dehydrogenated on the O-covered Ru surface, producing chemisorbed C-rich species while most of the H is released into the gas-phase as H2. The pre-cursor’s carbonyl ligands may be directly released as gas-phase C16O, or molec-ularly chemisorb onto the Ru surface as C16O, or dissociatively chemisorb (see p. 124) as C and 16O. Subsequently, a fraction of the surface dehydrogenation products and the carbonyl ligands can react with the surface 18O or 16O to form CO2 and H2O, and thereby the surface Ru atoms are reduced to its pure metallic state during the combustion process. When the surface 18/16O surface reservoir is depleted, predominantly partial combustion reactions occur, producing mainly gaseous C16O or C18O. At the end of the Ru pulse, carbon-rich species remain on the Ru surface that correspond to the ~30% of the total carbon eventually released from the Ru surface in one complete ALD cycle. In the 18O2 pulse, the 18O2 chemisorbs dissociatively onto the Ru surface and the reactive 18O atoms (and some residual 16O from the Ru pulse) react with the remaining surface species, liberating CO and CO2. In both half-cycles, combustion prod-ucts containing both oxygen isotopes were detected, which confirms the dis-sociative chemisoprtion of C16O ligands from CpRu(C16O)2Et.

Additional considerations on the basis of surface science investigations

Ru vs. RuO2

From additional information reported in the surface science literature, more subtle speculations can be made concerning the formation of CO2 and H2O during the Ru ALD process. In a simplistic view, the origin of CO2 and H2O is a result of complete combustion reactions of the Ru precursor’s ligands on O-

Page 134: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

127

rich Ru surfaces. From surface science studies it is possible to refine this view by providing evidence that, in fact, these reactions are most efficient on RuO2 re-gions. Depending on the O2 pressure and dose, and the surface temperature, RuO2 domains can exist on a Ru surface, which can have a very different catalyt-ic activity. Based on the reaction products measured, we speculate that com-bustion-like reactions that lead to the release of CO2 and H2O occur primarily on RuO2 domains, whereas the dehydrogenation reactions that lead to the re-lease of H2 occur primarily on Ru surfaces.

One pathway for the formation of CO2 is via the oxidation of CO,55,85 which is either released thermally from the Ru precursor or generated due to partial combustion of the precursor’s hydrocarbon ligands. Below we speculate on the possible pathways for CO oxidation. As mentioned in the Introduction, the CO oxidation reaction on Ru(0001) surfaces has been well explored, and it has been reported that O-chemisorbed Ru(0001) surfaces with 0 <θO< 3 ML do not play a significant role in the CO oxidation reaction under UHV conditions55,86,87 because of the strong Ru-O bond, and the low desorption temperature for adsorbed CO.66 Consequently, under UHV conditions the chemisorbed CO desorbs without reacting with the chemisorbed O on the Ru surface. Other experiments suggest that the RuO2 microregions formed at a coverage value of 3 ML are significantly more active in combusting C-containing species than Ru(0001), because the coordinatively unsaturated Ru atoms in RuO2 act as “dangling bonds” and can provide the adsorption sites for CO.88‒90 In previous studies it was shown that the CO oxidation reaction pro-ceeds via the Mars-Van Krevelen (reduction-reoxidation) mechanism91 in which an O atom is abstracted from the RuO2 lattice to form gaseous CO2 leaving be-hind an O vacancy. If present, gaseous O2 is dissociated on the surface, repopu-lating the O vacancies.90 In the absence of gas-phase O2, the surface Ru atoms are reduced to metallic Ru as a result of CO oxidation on the RuO2 surface.90

Most surface science studies are based on single-crystal Ru(0001) surfaces, and this is an important difference compared to ALD grown films, which are polycrystalline. In terms of O coverage, the few reports available show that Ru(1010) surfaces can also take up chemisorbed O beyond 1 ML.92,93 The poly-crystalline nature of the films seems to become more important at high O coverages, because the crystal orientation of the developing RuO2 patches was claimed to be determined by that of the underlying Ru domains.90,94,95 The RuO2

Page 135: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

128

domain orientation, on the other hand, was reported to be decisive in the cata-lytic activity of RuO2 in the specific surface reactions.94 Concerning dehydro-genation reactions, little is known about Ru crystal orientations other than Ru(0001). These dehydrogenation reactions have been reported to occur for CH4 on Ru(1120) surfaces,44 and they are known to occur on polycrystalline Ru catalysts.

Both Ru and RuO2 are known to be good dehydrogenation catalysts, but the proposed microscopic dehydrogenation mechanisms are different for each of these surfaces.96 The dehydrogenation of hydrocarbon species on RuO2 sur-faces leads to the preferential formation and liberation of CO2 and H2O,96,97 whereas on Ru surfaces, H2 is the main gas-phase dehydrogenation product due to the lack of surface O.48 In the catalytic oxidation of hydrocarbons on supported Ru-catalysts at low temperatures (300-600 K), the formation of CO2 and H2O was also attributed to the Mars-Van Krevelen mechanism occurring on thin RuO2 layers or domains.98‒100

In Ru ALD, the substrate temperature is ~300 ºC (~573 K) and O2 exposure is ~105 L, which coincides with the typical experimental conditions at which these RuO2 patches develop.101 In addition, we observed that ~70% of the C16O ligands were oxidized on the surface to CO2 during the Ru precursor pulse. From these facts, it is plausible that RuO2 patches are present on the surface during ALD. It is unlikely that the entire surface consists of RuO2 due to the ob-servation that high amounts of H2 are liberated, which is a fingerprint for Ru surfaces. The transition from oxygen adsorption to oxide formation on the Ru(0001) surface is quite complex and leads to the co-existence of (1×1)O-Ru(0001) and RuO2(110) patches.54 It is therefore possible that our Ru surface consists of a heterogeneous system of RuO2 and O-rich Ru patches at the start of the ALD cycle. Upon the interaction of the organic precursor ligands, the oxide surface would be entirely reduced to metallic Ru by the end of the Ru pulse. However, it will be necessary to study the Ru surface during the ALD cy-cle by means other than QMS in order to confirm this view and draw final con-clusions.

Page 136: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

129

Saturation mechanism during Ru ALD

In general, the saturation of surface half-reactions in ALD occurs when ad-sorption of new molecules is inhibited. This can occur either due to steric hin-drance provided by the chemisorbed surface species or due to the occupation of surface reactive sites. From Chapter 6, previous literature46,47 and the QMS data during Ru ALD of this study suggest that during the metal pulse, the sur-face reactions are terminated due to the C-rich species remaining on the sur-face, which inhibit further precursor adsorption. On the other hand, during the O2 pulse, the surface reactions are terminated when all of the remaining C-rich species are eliminated: the relevant parameters include surface temperature, and O2 pressure and exposure time. Specifically, we postulate that during the Ru precursor pulse, hydrocarbon dehydrogenation reactions play an important role. Based on the surface science literature, it is known that on Ru(0001) sur-faces dehydrogenation reactions can occur at low temperatures (105 K),102 but are complete only at high temperatures (700-800 K), leaving a layer of C atoms on the Ru surface.44,48,82 The deposition temperature (600 K) for the Ru ALD pro-cess discussed herein is somewhat lower than the temperature for complete dehydrogenation and, therefore, we expect that the species remaining on the surface at the end of the Ru pulse are primarily C atoms with only a small frac-tion of H atoms. This is in agreement with the fact that we observed almost no H-containing species during the following 18O2 pulse. In fact, we were able to estimate the coverage of these C-rich surface species after the Ru precursor cycle to be ~1 ML: this coverage was calculated using the number density of CO2 and CO released into the gas phase during the 18O2 cycle (see Table I), and relating this number density to the amount of carbon on the Ru surface using the same procedure described earlier for determining the surface O coverage (see p. 125). Furthermore, a study on ethylene dehydrogenation on Ru(0001) surfaces (for the purpose of depositing graphene) showed that a monolayer of C-atoms was sufficient to deactivate the dehydrogenation of incoming eth-ylene molecules on the Ru surface.103 As a consequence, we speculate that eventually most of the available Ru sites are occupied by C species, thereby blocking the adsorption of incoming CpRu(CO)2Et molecules, and leading to the saturation of the ALD surface reactions during the Ru precursor pulse.

Page 137: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

130

Gro

wth

-per

-cyc

le (Å

/cyc

le)

Substrate temperature (ºC)

Temperature dependence of the Ru ALD process

Figure 4 shows that over the range of process conditions studied for Ru ALD, the GPC was low at temperatures below 200 ºC. Several other Ru ALD processes using metalorganic precursors and O2 gas show a very similar trend,26,28,37,74,75,104 and in most cases, film deposition was reported at ~300 ºC. A possible surface reaction that would require these elevated temperatures is the dehydrogenation of the Ru precursor’s hydrocarbon ligands. However, these dehydrogenation reactions can be excluded as the rate limiting step as they have been reported to occur on pristine Ru(0001) surfaces at temperatures as low as 105 K,102 and at even lower temperatures on O-covered Ru(0001).41 Therefore, we postulate that the rate limiting step for Ru ALD is during the O2 pulse where O2 gas has to dissociate on the carbon-covered Ru surface ob-tained after the Ru precursor pulse. These carbon atoms have to be removed via combustion-like reactions by the dissociated O2 gas. Previous surface sci-ence literature on Ru shows that the transition from Ru to RuO2 during O2 ex-posure occurs over the temperature range of 500-550 K.105 Interestingly, com-bustion reactions of C-containing species on fully oxidized RuO2 surfaces have also been shown to require a threshold temperature of 500-600 K.96,101,106 Since Ru ALD occurs precisely over the temperature range of 500-550 K, we speculate that RuO2 domains may contribute to the elimination of surface carbon.

Figure 4: Growth per cycle during Ru ALD as a function of the substrate temperature

as measured by spectroscopic ellipsometry.

Page 138: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

131

Conclusions

In this work, QMS was used to obtain quantitative information on the sur-face reactions involved in Ru ALD from CpRu(CO)2Et and oxygen isotope 18O2. The gas-phase reaction products were CO2, CO, H2, and H2O during the Ru pre-cursor pulse, and mainly CO2 and CO during the O2 pulse. Most of the reaction products were formed during the Ru pulse of the ALD process: Approximately 70% of the carbon contained in the precursor was released during this pulse of which ~90% was in the form of CO2 and the rest as CO. During the same metal precursor pulse, ~70% of the carbonyl ligands reacting with the surface were oxidized, and the remaining ~30% were liberated during the O2 pulse in the form of C16O and C16O18O. Furthermore, during the Ru pulse, ~100% of the hy-drogen contained in the precursor was released as H2 (85%) and H2O (15%). From these observations and on the basis of surface science and catalysis litera-ture, we concluded that the complex chemistry taking place on the Ru surface during ALD can be described by catalytic combustion reactions, which were originally introduced in the field of catalysis. These reactions consist of the dis-sociative chemisorption of specific species on a catalytic surface (organic lig-ands during the Ru pulse, and oxygen in the O2 pulse) accompanied with the oxidation of the C- and H-rich species with the adsorbed oxygen. On the basis of the surface science and catalysis literature on Ru surfaces, we have speculat-ed on specific properties of the Ru ALD process: the role of RuO2 and Ru due to the different catalytic chemistry of both materials, the deposition temperature dependence, and on the saturation mechanism. These speculations give op-portunities to test and refine the mechanism such that Ru ALD processes can be tuned further.

Acknowledgments

We thank C.A.A. van Helvoirt for his technical support throughout this study and Dr. H.C.M. Knoops from the Eindhoven University of Technology for very fruitful discussions. This work was financially supported by the Dutch Technology Foundation STW, which is part of the Netherlands Organization for Scientific Research (NWO), and specifically the work of one of the authors (W.M.M. Kessels) is supported through the VICI program on “Nanomanufacturing”. S. A. gratefully acknowledges financial support from the

Page 139: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

132

NSF CAREER program (Grant No. CBET-0846923), and from the Foundation for Fundamental Research on Matter (FOM) for his sabbatical leave.

Appendix ‒ Additional data:

A.1 Identification of the reaction products

To select the m/z values to be monitored in time during the Ru ALD half-cycles, we first performed a mass scan over the range of m/z = 1-60 with a low time resolution during (i) a reference measurement with CpRu(CO)2Et pulsing only (no ALD), and (ii) during a Ru precursor pulse during the ALD process. The reference measurement was performed on a saturated Ru surface, and without pulsing 18O2 gas in between the CpRu(CO)2Et pulses. The reference measure-ment in Figure A.1 is during one precursor dose, whereas the data for the Ru precursor pulse during ALD was averaged over 18 consecutive half-cycles. The-se data indicate that during the Ru pulse, m/z = 1-3, 18, 20, 44, 46, and 48 rep-resent species produced during surface reactions. The signals observed during ALD at m/z = 12 and m/z = 16 can be attributed to C+ and 16O+, respectively, originating from dissociative ionization of parent molecules already accounted for as reaction products.

Figure A.1: QMS ion current during a mass scan over the range of m/z = 1-60 for a reference measurement during CpRu(CO)2Et pulsing only, and of the Ru precursor pulse during thermal ALD. The mass scan during ALD was averaged over 18 consecu-tive ALD cycles, whereas the reference measurement was performed for one precursor dose.

m/z

Page 140: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

133

A.2 Corrected QMS intensity used for the quantification of the reaction products

To utilize Eq. 1 in the main text, the QMS signal S has to be corrected for possible contributions that are not directly related to surface reactions during ALD. To illustrate this, appropriate corrections for C18O2

+ (m/z = 48) and C18O+ (m/z = 30) are presented for the Ru precursor pulse based on the signals in Figure 2c-d. The signal measured during the Ru pulse is a combination of thermally cracked CpRu(CO)2Et in the QMS ionizer and surface reaction prod-ucts formed during the ALD cycle. Therefore, the value of Sm/z=48 that is actually related to surface reactions during ALD is 2 2

48 48 48

. , .

/ / /Fig d ref Fig bRu pulse

m z m z m zS S S . In the case of C18O, there is an additional contribution due to the dissociative ionization of CO2. Based on the cracking pattern of CO2, CO+ signal was predicted to be ~10% of the CO2

+ signal,107,108 which was also confirmed for our QMS. Therefore, the appropriate value for Sm/z=30 to be used in Eq. 1 is

.2 , .2 .2 .2

/ 30 / 30 / 30 / 48 / 46( 0.1 0.05 )Fig d ref Fig b Fig d Fig dRu pulsem z m z m z m z m zS S S S S ,

2 2

48 48 48

. , .

/ / /Fig d ref Fig bRu pulse

m z m z m zS S S , and 2 2

46 46 46

. , .

/ / /Fig d ref Fig bRu pulse

m z m z m zS S S ,

where 2

48

./

Fig dm zS (C18O2

+) and 2

46

./

Fig dm zS (C18O16O+) correct for precursor cracking in

the QMS ionizer. The factor of 0.05 preceding Sm/z=46 accounts for the fact that C18O16O+ only contains one 18O atom to potentially contribute to Sm/z=30.

A.3 Calibration of the QMS

Ideally, the factor k in Eq. 1 can be determined for a QMS using the parent ionization signal for several gases at a known pressure over a range of atomic weights. Specifically, to calibrate our QMS, we determined the signal for the direct ionization of H2, H2O, N2, 16O2, and 18O2 over the pressure range of 5-50 mTorr.79 Figure A.2a shows that the QMS signal for H2

+ (m/z = 2), H2O+ (m/z = 18), N2

+ (m/z = 28), 16O2+ (m/z = 32), 18O2

+ (m/z = 36), and Ar+ (m/z = 40) varies linearly with the gas partial pressure in the chamber. Using nXY→XY+

ref cal-culated from the ideal gas law (in cm-3), published values80 for σXY→XY+ (in cm-2), and the measured QMS signal S (in arbitrary units [a.u.]), we determined k for each gas. Figure A.2b that k approximately proportional to the inverse of the

Page 141: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

134

m/z ratio. From a linear regression fit over m/z = 18-40, we obtained the follow-ing expression for the factor k for our QMS,

k = [-0.035 + 57.8/(m/z)] × 10-3 a.u.cm . (Eq. A.1)

From the above expression, we could interpolate k for any species with a m/z value contained within or close to the range calibrated (18 < m/z < 40). Note that H2 generally does not follow this trend due to the inefficient pumping, and the slope shown in Fig. A.2a was used directly in Eq. 2 to calibrate the H2 densi-ty.

Figure A.2: The QMS signal intensity due to parent ionization of different gases is shown in (a) as a function of the partial pressure for m/z = 2 (H2+), 18 (H2O+), 28 (N2+), 32 (16O2+), 36 (18O2+), and 40 (Ar+). Instrument calibration of the m/z-dependent factor k is shown in (b). Due to the inefficient pumping of H2, k for m/z = 2 did not follow the general trend and, therefore, the slope of m/z = 2 from (a) was used as input value in Eq. 2.

References

(1) Liu, C.; Wang, C.-C.; Kei, C.-C.; Hsueh, Y.-C.; Perng, T.-P. Small 2009, 5, 1535‒1538.

(2) ITRS www.itrs.net - Process integration, devices, and structures 2010.

(3) Kim, S. K.; Lee, S. W.; Han, J. H.; Lee, B.; Han, S.; Hwang, C. S. Adv. Funct. Mater. 2010, 20, 2989‒3003.

(4) Kittl, J. A.; Opsomer, K.; Popovici, M.; Menou, N.; Kaczer, B.; Wang, X. P.; Adelmann, C.; Pawlak, M. A.; Tomida, K.; Rothschild, A.; Govoreanu, B.;

(a) (b)H2

16O

N2

16O2

18O2Ar

Page 142: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

135

Degraeve, R.; Schaekers, M.; Zahid, M.; Delabie, A.; Meersschaut, J.; Polspoel, W.; Clima, S.; Pourtois, G.; Knaepen, W.; Detavernier, C.; Afanas’ev, V. V; Blomberg, T.; Pierreux, D.; Swerts, J.; Fischer, P.; Maes, J. W.; Manger, D.; Vandervorst, W.; Conard, T.; Franquet, A.; Favia, P.; Bender, H.; Brijs, B.; Van Elshocht, S.; Jurczak, M.; Van Houdt, J.; Wouters, D. J. Microelectron. Eng. 2009, 86, 1789‒1795.

(5) Kohl, D. Sensors Actuat. B - Chem. 1990, 1, 158‒165.

(6) Patil, L. A.; Bari, A. R.; Shinde, M. D.; Deo, V.; Kaushik, M. P. Sensors and Actuat. B-Chem. 2012, 161, 372‒380.

(7) Bartlett, P. N.; Guerin, S. Anal. Chem. 2002, 75, 126‒132.

(8) Ukisu, Y.; Miyadera, T. Chemosphere 2002, 46, 507‒510.

(9) Wang, J.; Zhong, J.; Gong, M.; Liu, Z.; Zhao, M.; Chen, Y. J. Environ. Sci. 2007, 19, 644‒646.

(10) Bhattacharyya, S.; Das, R. K. Int. J. Energ. Res. 1999, 23, 351‒369.

(11) George, S. M. Chem. Rev. 2010, 110, 111‒131.

(12) Comstock, D. J.; Christensen, S. T.; Elam, J. W.; Pellin, M. J.; Hersam, M. C. Adv. Funct. Mater. 2010, 20, 3099‒3105.

(13) Knoops, H. C. M.; Mackus, A. J. M.; Donders, M. E.; Van de Sanden, M. C. M.; Notten, P. H. L.; Kessels, W. M. M. Electrochem. Solid-State Lett. 2009, 12, G34‒G36.

(14) Christensen, S. T.; Feng, H.; Libera, J. L.; Guo, N.; Miller, J. T.; Stair, P. C.; Elam, J. W. Nano Lett. 2010, 10, 3047‒3051.

(15) Biener, J.; Baumann, T. F.; Wang, Y.; Nelson, E. J.; Kucheyev, S. O.; Hamza, A. V.; Kemel, M.; Ritala, M.; Leskelä, M. Nanotechnology 2007, 18, 55303.

(16) Joo, J. H.; Seon, J. M.; Jeon, Y. C.; Oh, K. Y.; Roh, J. S.; Kim, J. J.; Choi, J. T. Japn. J. Appl. Phys. 1998, 37, 3396‒3401.

(17) Swerts, J.; Armini, S.; Carbonell, L.; Delabie, A.; Franquet, A.; Mertens, S.; Popovici, M.; Schaekers, M.; Witters, T.; Tokei, Z.; Beyer, G.; Van Elshocht, S.; Gravey, V.; Cockburn, A.; Shah, K.; Aubuchon, J. J. Vac. Sci. Technol. A 2012, 30, 01A103‒6.

(18) Kim, J.-Y.; Kil, D.-S.; Kim, J.-H.; Kwon, S.-H.; Ahn, J.-H.; Roh, J.-S.; Park, S.-K. J. Electrochem. Soc. 2012, 159, H560‒H564.

Page 143: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

136

(19) Park, S. J.; Kim, W. H.; Lee, H. B. R.; Maeng, W. J.; Kim, H. Microelectron. Eng. 2008, 85, 39‒44.

(20) Kwon, O. K.; Kwon, S. H.; Park, H. S.; Kang, S. W. J. Electrochem. Soc. 2004, 151, C753‒C756.

(21) Aaltonen, T.; Alen, P.; Ritala, M.; Leskelä, M. Chem. Vap. Deposition 2003, 9, 45‒49.

(22) Kim, S. K.; Lee, S. Y.; Lee, S. W.; Hwang, G. W.; Hwang, C. S.; Lee, J. W.; Jeong, J. J. Electrochem. Soc. 2007, 154, D95‒D101.

(23) Kim, S. K.; Hoffmann-Eifert, S.; Waser, R. J. Phys. Chem. C 2009, 113, 11329‒11335.

(24) Kukli, K.; Kemell, M.; Puukilainen, E.; Aarik, J.; Aidla, A.; Sajavaara, T.; Laitinen, M.; Tallarida, M.; Sundqvist, J.; Ritala, M.; Leskelä, M. J. Electrochem. Soc. 2011, 158, D158‒D165.

(25) Kukli, K.; Ritala, M.; Kemell, M.; Leskelä, M. J. Electrochem. Soc. 2010, 157, D35‒D40.

(26) Kukli, K.; Aarik, J.; Aidla, A.; Uustare, T.; Jogi, I.; Lu, J.; Tallarida, M.; Kemell, M.; Kiisler, A. A.; Ritala, M.; Leskelä, M. J. Cryst. Growth 2010, 312, 2025‒2032.

(27) Gregorczyk, K.; Henn-Lecordier, L.; Gatineau, J.; Dussarrat, C.; Rubloff, G. Chem. Mater. 2011, 23, 2650‒2656.

(28) Choi, S.-H.; Cheon, T.; Kim, S.-H.; Kang, D.-H.; Park, G.-S.; Kim, S. J. Electrochem. Soc. 2011, 158, D351‒D356.

(29) Geidel, M.; Junige, M.; Albert, M.; Bartha, J. W. Microelectron. Eng. 2013, 107, 151‒155.

(30) Aaltonen, T.; Ritala, M.; Tung, Y. L.; Chi, Y.; Arstila, K.; Meinander, K.; Leskelä, M. J. Mat. Res. 2004, 19, 3353‒3358.

(31) Choi, J.; Choi, Y.; Hong, J.; Tian, H.; Roh, J. S.; Kim, Y.; Chung, T. M.; Oh, Y. W.; Kim, C. G.; No, K. Jpn. J. Appl. Phys. 2002, 41, 6852‒6856.

(32) Kwon, O. K.; Kwon, S. H.; Park, H. S.; Kang, S. W. Electrochem. Solid-State Lett. 2004, 7, C46‒C48.

(33) Yim, S. S.; Lee, M. S.; Kim, K. S.; Kim, K. B. Appl. Phys. Lett. 2006, 89, 93115.

(34) Aaltonen, T.; Rahtu, A.; Ritala, M.; Leskelä, M. Electrochem. Solid-State Lett. 2003, 6, C130‒C133.

Page 144: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

137

(35) Christensen, S. T.; Elam, J. W. Chem. Mater. 2010, 22, 2517‒2525.

(36) Kessels, W. M. M.; Knoops, H. C. M.; Dielissen, S. A. F.; Mackus, A. J. M.; Van de Sanden, M. C. M. Appl. Phys. Lett. 2009, 95, 013114.

(37) Aaltonen, T.; Ritala, M.; Arstila, K.; Keinonen, J.; Leskelä, M. Chem. Vap. Deposition 2004, 10, 215‒219.

(38) Knapas, K.; Ritala, M. Chem. Mater. 2011, 23, 2766‒2771.

(39) Rauscher, H.; Jakob, P.; Menzel, D.; Lloyd, D. R. Surf. Sci. 1991, 256, 27‒41.

(40) Weiss, M. J.; Hagedorn, C. J.; Weinberg, W. H. J. Am. Chem. Soc. 1999, 121, 5047‒5055.

(41) Garcia, A. R.; Barros, R. B.; Ilharco, L. M. Surf. Sci. 2009, 603, 380‒386.

(42) Ilharco, L. M.; Garcia, A. R.; Hargreaves, E. C.; Chesters, M. A. Surf. Sci. 2000, 459, 115‒123.

(43) Hrbek, J. J. Catal. 1986, 100, 523‒525.

(44) Wu, M. C.; Xu, Q.; Goodman, D. W. J. Phys. Chem. 1994, 98, 5104‒5110.

(45) Elliott, S. D. Langmuir 2010, 26, 9179‒9182.

(46) Mackus, A. J. M.; Leick, N.; Baker, L.; Kessels, W. M. M. Chem. Mater. 2012, 24, 1752‒1761.

(47) Erkens, I. J. M.; Mackus, A. J. M.; Knoops, H. C. M.; Smits, P.; Van de Ven, T. H. M.; Roozeboom, F.; Kessels, W. M. M. J. Electrochem. Soc. 2012, 6, P255‒P262.

(48) George, P. M.; Avery, N. R.; Weinberg, W. H.; Tebbe, F. N. J. Am. Chem. Soc. 1983, 105, 1393‒1394.

(49) Eom, T. K.; Sari, W.; Choi, K. J.; Shin, W. C.; Kim, J. H.; Lee, D. J.; Kim, K. B.; Sohn, H.; Kim, S. H. Electrochem. Solid-State Lett. 2009, 12, D85‒D88.

(50) Bradford, M. C. J. J. Catal. 2000, 189, 238‒243.

(51) Choudhary, T. V; Goodman, D. W. Top. Catal. 2002, 20, 35‒42.

(52) Ditlevsen, P. D.; Vanhove, M. A.; Somorjai, G. A. Surf. Sci. 1993, 292, 267‒275.

(53) Ertl, G. Angewandte Chemie-International Edition 2008, 47, 3524‒3535.

(54) Over, H.; Seitsonen, A. P. Science 2002, 297, 2003‒2005.

Page 145: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

138

(55) Böttcher, A.; Niehus, H.; Schwegmann, S.; Over, H.; Ertl, G. J. Phys. Chem. B 1997, 101, 11185‒11191.

(56) Blume, R.; Niehus, H.; Conrad, H.; Böttcher, A.; Aballe, L.; Gregoratti, L.; Barinov, A.; Kiskinova, M. J. Phys. Chem. B 2005, 109, 14052‒14058.

(57) Blume, R.; Havecker, M.; Zafeiratos, S.; Teschner, D.; Kleimenov, E.; Knop-Gericke, A.; Schlogl, R.; Barinov, A.; Dudin, P.; Kiskinova, M. J. Catal. 2006, 239, 354‒361.

(58) Blume, R.; Niehus, H.; Conrad, H.; Böttcher, A. J. Chem. Phys. 2004, 120, 3871‒3879.

(59) Böttcher, A.; Niehus, H. J. Chem. Phys. 1999, 110, 3186‒3195.

(60) Stampfl, C.; Schwegmann, S.; Over, H.; Scheffler, M.; Ertl, G. Phys. Rev. Lett. 1996, 77, 3371‒3374.

(61) Meinel, K.; Wolter, H.; Ammer, C.; Beckmann, A.; Neddermeyer, H. J. Phys. Cond. Matter. 1997, 9, 4611‒4619.

(62) Kim, Y. D.; Seitsonen, A. P.; Wendt, S.; Wang, J.; Fan, C.; Jacobi, K.; Over, H.; Ertl, G. J. Phys. Chem. B 2001, 105, 3752‒3758.

(63) Reuter, K.; Ganduglia-Pirovano, M. V.; Stampfl, C.; Scheffler, M. Phys. Rev. B. 2002, 65, 165403.

(64) Todorova, M.; Reuter, K.; Scheffler, M. Phys. Rev. B. 2005, 71, 195403.

(65) Peden, C. H. F.; Goodman, D. W.; Weisel, M. D.; Hoffmann, F. M. Surf. Sci. 1991, 253, 44‒58.

(66) Bonn, M.; Funk, S.; Hess, C.; Denzler, D. N.; Stampfl, C.; Scheffler, M.; Wolf, M.; Ertl, G. Science 1999, 285, 1042‒1045.

(67) Kim, Y. D.; Seitsonen, A. P.; Over, H.; Kim, D.; Seitsonen, P. Surf. Sci. 2000, 465, 1‒8.

(68) Somorjai, G. A.; Li, Y. Top. Catal. 2010, 53, 832‒847.

(69) Huff, M.; Schmidt, L. D. J. Phys. Chem. 1993, 97, 11815‒11822.

(70) Mantzaras, J. In Handbook of Combustion; Wiley-VCH Verlag GmbH & Co. KGaA, 2010.

(71) Lee, J. H.; Trimm, D. L. Fuel Process. Technol. 1995, 42, 339‒359.

(72) Reinke, M.; Mantzaras, J.; Schaeren, R.; Bombach, R.; Inauen, A.; Schenker, S. Combusti. Flame 2004, 136, 217‒240.

Page 146: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

139

(73) Van Hemmen, J. L.; Heil, S. B. S.; Klootwijk, J. H.; Roozeboom, F.; Hodson, C. J.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Electrochem. Soc. 2007, 154, G165‒G169.

(74) Wang, H. T.; Gordon, R. G.; Alvis, R.; Ulfig, R. M. Chem. Vap. Deposition 2009, 15, 312‒319.

(75) Park, S. K.; Kanjolia, R.; Anthis, J.; Odedra, R.; Boag, N.; Wielunski, L.; Chabal, Y. J. Chem. Mater. 2010, 22, 4867‒4878.

(76) Knoops, H. C. M.; Langereis, E.; Van de Sanden, M. C. M.; Kessels, W. M. M. J. Vac. Sci. Technol. A 2012, 30, 01A1011‒01A10110.

(77) NIST Chemistry WebBook,. NIST Standard Reference Database Number 69.

(78) Austin, W. E.; Holme, A. E.; Leck, J. H. In Quadrupole Mass Spectrometry and Its Applications; Dawson, P. H., Ed.; American Institute of Physics: Amsterdam, 1997.

(79) Singh, H.; Coburn, J. W.; Graves, D. B. J. Vac. Sci. Technol. A 2000, 18, 299‒305.

(80) Hwang, W.; Kim, Y. K.; Rudd, M. E. J. Chem. Phys. 1996, 104, 2956‒2966.

(81) Morrow, B. A.; Ramamurthy, P. J. Phys. Chem. 1973, 77, 3052‒3058.

(82) Hills, M. M.; Parmeter, J. E.; Mullins, C. B.; Weinberg, W. H. J. Am. Chem. Soc. 1986, 108, 3554‒3562.

(83) Getman, R. B.; Xu, Y.; Schneider, W. F. J. Phys. Chem. B 2008, 112, 9559‒9572.

(84) Taylor, J. L.; Ibbotson, D. E.; Weinberg, W. H. Surf. Sci. 1979, 79, 349‒384.

(85) Gao, F.; Wang, Y.; Cai, Y.; Goodman, D. W. Surf. Sci. 2009, 603, 1126‒1134.

(86) Kim, Y. D.; Seitsonen, A. P.; Over, H. Surf. Sci. 2000, 465, 1‒8.

(87) Lee, H.-I.; White, J. M. J. Catal. 1980, 63, 261‒264.

(88) Over, H.; Seitsonen, A. P. Science 2002, 297, 2003.

(89) Over, H.; Kim, Y. D.; Seitsonen, A. P.; Wendt, S.; Lundgren, E.; Schmid, M.; Varga, P.; Morgante, A.; Ertl, G. Science 2000, 287 , 1474‒1476.

(90) Over, H. Appl. Phys. A: Mater. Sci. Process. 2002, 75, 37‒44.

(91) Mars, P.; Van Krevelen, D. W. Chem. Eng. Sci. 1954, 3, Supplem, 41‒59.

(92) Baraldi, A.; Lizzit, S.; Paolucci, G. Surf. Sci. 2000, 457, L354‒L360.

Page 147: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

140

(93) Schwegmann, S.; Seitsonen, A. P.; De Renzi, V.; Dietrich, H.; Bludau, H.; Gierer, M.; Over, H.; Jacobi, K.; Scheffler, M.; Ertl, G. Phys. Rev. B 1998, 57, 15487‒15495.

(94) Assmann, J.; Narkhede, V.; Breuer, N. A.; Muhler, M.; Seitsonen, A. P.; Knapp, M.; Crihan, D.; Farkas, A.; Mellau, G.; Over, H. J. Phys. Cond. Matter. 2008, 20, 184017.

(95) Kim, Y. D.; Schwegmann, S.; Seitsonen, A. P.; Over, H. J. Phys. Chem B 2001, 105, 2205‒2211.

(96) Over, H.; He, Y. B.; Farkas, A.; Mellau, G.; Korte, C.; Knapp, M.; Chandhok, M.; Fang, M. J. Vac. Sci. Technol. B 2007, 25, 1123‒1138.

(97) Knapp, M.; Crihan, D.; Seitsonen, A. P.; Resta, A.; Lundgren, E.; Andersen, J. N.; Schmid, M.; Varga, P.; Over, H. J. Phys. Chem B 2006, 110, 14007‒14010.

(98) Liu, H.; Iglesia, E. J. Phys. Chem B 2005, 109, 2155‒63.

(99) Okal, J.; Zawadzki, M. Appl. Catal. B: Environ. 2009, 89, 22‒32.

(100) Knapp, M.; Crihan, D.; Seitsonen, A. P.; Lundgren, E.; Resta, A.; Andersen, J. N. J. Phys. Chem. C 2007, 111, 5363‒5373.

(101) Villani, K.; Kirschhock, C. E. A.; Liang, D.; Van Tendeloo, G.; Martens, J. A. Angewandte Chemie International Edition 2006, 45, 3106‒3109.

(102) Henderson, M. A.; Radloff, P. L.; Greenlief, C. M.; White, J. M.; Mims, C. A. J. Phys. Chem. 1988, 92, 4120‒4127.

(103) Loginova, E.; Bartelt, N. C.; Feibelman, P. J.; McCarty, K. F. New J. Phys. 2009, 11, 63046.

(104) Kim, S. K.; Han, J. H.; Kim, G. H.; Hwang, C. S. Chem. Mater. 2010, 22, 2850‒2856.

(105) He, Y. B.; Knapp, M.; Lundgren, E.; Over, H. J. Phys. Chem. B 2005, 109, 21825‒21830.

(106) Paulus, U. A.; Wang, Y.; Bonzel, H. P.; Jacobi, K.; Ertl, G. J. Phys. Chem. B 2004, 109, 2139‒2148.

(107) Crowe, A.; Mcconkey, J. W. J. Phys. B 1974, 7, 349‒361.

(108) Hudson, J. E.; Vallance, C.; Harland, P. W. J. Phys. B 2004, 37, 445‒455.

Page 148: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

141

Page 149: Atomic layer deposition of ruthenium films : properties and surface

7. CATALYTIC COMBUSTION REACTIONS DURING ALD OF Ru STUDIED BY 18O2...

142

Page 150: Atomic layer deposition of ruthenium films : properties and surface

143

Chapter 8 “October knew, of course, that the action of turning a page, of ending a chapter or of

shutting a book, did not end a tale.” ‒ Neil Gaiman

Conclusions and recommendations

As mentioned in Chapter 1, ultrathin layers of ruthenium are becoming of high interest for various applications, especially in the interconnect technology and in metal-insulator-metal (MIM) capacitors. For these types of applications, the films are typically required to be nanometer-thin, smooth, and conformal and they should also have a good adhesion to different types of substrates. Additionally, it is preferable for the films to have bulk-like properties in terms of mass density and conductivity. Over the last 10 years, ALD has demonstrated its potential as a well-suited deposition method for the preparation of such Ru films.

The work described in this thesis has focused on the deposition of Ru thin films using ALD, on the characterization of the films, and most of all, on the in-vestigation of the underlying reaction mechanisms leading to Ru films for the particular ALD process used. The essential new insights gained in the thesis project are briefly summarized below:

In Chapter 4, it was shown that thermal and plasma-assisted ALD of Ru can be achieved using the heteroleptic cyclopentadienyl (Cp)-based metal pre-cursor CpRu(CO)2Et and O2 as a co-reactant. The Ru films obtained were dense, polycrystalline with a hexagonal closed packed crystal structure and had a resistivity close to bulk resistivity. These properties make the ALD-prepared Ru films suitable to be used as electrode layers in MIM capacitors. The merits of these new ALD processes lie in the relatively high growth rate (1 Å/cycle) obtained and, specifically for the O2 plasma-based process, in the reduction of the nucleation delay down to ~ 45 ALD cycles. However, despite the improved nucleation, the plasma-assisted ALD Ru films had a higher surface roughness than the thermal ALD films which appears to be

Page 151: Atomic layer deposition of ruthenium films : properties and surface

8. CONCLUSIONS AND RECOMMENDATIONS

144

correlated to the nucleation of the films. Under our operating conditions, no major benefits have emerged for the plasma-assisted ALD process of Ru over the thermal process. Therefore, the remainder of the work focused on Ru ALD using O2 gas rather than O2 plasma as a co-reactant.

As ALD of Pt-group metals has two main growth regimes, i.e. the nuclea-tion phase and the steady-state layer-by-layer growth of the metal, it was crucial to develop a tool monitoring the Ru film thickness in situ, and measure the thickness real time. In Chapter 5, we discussed the optical method of spectroscopic ellipsometry to study the film thicknesses and film growth, and we monitored the dielectric function of the films in the range of 0.75-5 eV. Using a Kramers-Kronig consistent variational parametrization to account for the thickness-dependent dielectric func-tions, we were able to obtain accurate film thickness values at almost any stage of the ALD process. Based on this model, we found the nucleation delay and the growth per cycle values to be 40, 20 and 100 cycles and 0.55 Å/cycle, 1.6 Å/cycle, 0.17 Å/cycle for Pt, Ru, and Pd, respectively.

To extract the electrical properties of the metal films, such as their resistivi-ty, the spectral range was expanded to 0.04-5 eV by combining reflectance data with ellipsometry data (both obtained ex situ). The variational parametrization of the metal films served as a basis for the more tradition-ally used Drude-Lorentz parametrization. Both parametrizations exhibited very similar shapes, showing that the variational parametrization already delivered a physical dielectric function for the Pt, Ru and Pd films. Further-more, due to the additional reflectance data, it was possible to obtain a unique and accurate Drude-Lorentz model. The optical resistivity associat-ed with the Drude term of the parametrization was in relatively good agreement with that measured by the four point probe technique, espe-cially in the case of the Pt films. The parameters of the Lorentz oscillators were in line with previous ellipsometry reports on Pt, Ru and Pd films. Therefore, the Drude-Lorentz parameters of this study can serve as refer-ence data for future Pt, Ru and Pd films deposited by ALD.

Page 152: Atomic layer deposition of ruthenium films : properties and surface

8. CONCLUSIONS AND RECOMMENDATIONS

145

In order to gain insights into the reaction mechanism, we studied the thermal Ru ALD process with quadrupole mass spectrometry (QMS) in Chapter 7. By using 18O2 isotope labeling, it was possible to distinguish be-tween C16O+ and C2H4

+, which were both originating from potential reac-tion by-products, especially during the CpRu(CO)2Et pulse. The gas-phase reaction products were CO2, CO, H2, and H2O during the Ru precursor pulse, and mainly CO2 and CO during the O2 pulse. Most of the reaction products were formed during the Ru pulse of the ALD process: approximately 70% of the carbon contained in the precursor was released during this pulse, of which ~90% was in the form of CO2 and the rest as CO. During the same metal precursor pulse, ~70% of the carbonyl ligands reacting with the sur-face were oxidized, and the remaining ~30% were liberated during the O2 pulse in the form of C16O and C16O18O. Furthermore, during the Ru pulse, ~100% of the hydrogen contained in the precursor was released, in the form of H2 (85%) and H2O (15%).

To interpret these experimental results, we applied the information availa-ble from the surface science and catalysis literature (surveyed in Chapter 2). Before this study, the predominant reaction pathway of Ru ALD using a metalorganic precursor and O2 gas was assumed to rely on combustion-like reactions. The interpretation of the QMS data permitted to refine this mechanism and to propose that catalytic combustion reactions govern the entire thermal Ru ALD process. These reactions lead to the removal of the precursor ligands and, ultimately, allow the metal film to grow. The reac-tions consist of the dissociative chemisorption of specific species on a cata-lytic surface (organic ligands during the Ru pulse, and oxygen in the O2 pulse) accompanied with the oxidation of the C- and H-rich species with the adsorbed oxygen.

The dissociative chemisorption reactions of organic ligands occurring on the Ru surface are called dehydrogenation reactions. In Chapter 6, dehy-drogenation reactions (reported extensively in the catalysis literature, see Chapter 2) were presented and it was shown that they play a significant role in the overall reaction mechanism during Ru ALD with CpRu(CO)2Et as a precursor, especially during the CpRu(CO)2Et pulsing.

Page 153: Atomic layer deposition of ruthenium films : properties and surface

8. CONCLUSIONS AND RECOMMENDATIONS

146

From the results and insights gained from this study, a number of recom-mendations can be given concerning the deposition and characterization of Ru films deposited with ALD as well as the study of the reaction mechanism of the film growth: As mentioned in Chapter 4, from the perspective of process integration,

the thermal ALD process with CpRu(CO)2Et precursor and O2 gas delivers Ru films with high film roughness values. So far, two main routes have been proposed to obtain smoother films. The first being the use of zero-valent Ru precursors in order to obtain a higher density of Ru nuclei on the sub-strate surface in the very initial cycles of ALD.1,2 The lower nucleation delay of these processes resulted in smoother films than those reported in Chap-ter 4. The second route consists of using O2-free plasmas, such as N2/H2 or NH3/H2-plasmas.3 This approach has been successfully implemented in the fabrication of the Ru layers of Ru‒STO3‒Ru MIM capacitors by ALD.4 In order to deposit smooth Ru films by ALD, it is recommended to use either one, or a combination of these two options.

The use of the plasma-assisted ALD process for the deposition of Ru films was not further pursued in this dissertation, but it might show clear bene-fits over thermal ALD when temperature sensitive substrates are used, for example. It was demonstrated that high-quality Pt films can be deposited on a diversity of materials (polymers, textile and paper substrates) using plasma-assisted ALD.5 This could potentially also hold for the plasma-assisted ALD process for Ru.

From the optical characterization of Ru films presented in Chapter 5, it be-came apparent that the lack of thick spacer layer of a transparent material (e.g., Al2O3, SiO2) between the substrate and the Ru film impeded an accu-rate optical modeling of Ru films thicker than ~25 nm. It is therefore rec-ommended to use such a spacer layer, as it was done for the Pt and Pd films, especially when the Ru films to study exceed a thickness of ~25 nm.

The variational parametrization of the optical SE data on Ru films delivered the most accurate thickness values for the steady-state growth mode of

Page 154: Atomic layer deposition of ruthenium films : properties and surface

8. CONCLUSIONS AND RECOMMENDATIONS

147

the metal. The thickness values obtained during the initial film nucleation phase, however, were rather indicative than absolute. As a future investiga-tion, the use of a Kramers-Kronig variational parametrization could also be applied to SE and reflectance data during the initial island-like growth of Ru ALD. Recent work on noble metal islands suggests that the electrical perco-lation threshold, amongst others, can be extracted from such optical measurements.6,7 The information on the percolation threshold might be very useful for the deposition of Pt-group metal particles for catalytic appli-cations.

In the Chapters 6 and 7, a reaction mechanism of Ru ALD was proposed for CpRu(CO)2Et as the precursor and O2 gas as the co-reactant. This mecha-nism was based on the on the quadrupole mass spectrometry measure-ments in combination with surface science and catalysis literature on Ru. Mass spectrometry, however, is an analytical method probing the ambient gas rather than the Ru surface directly. To confirm, for example, the pres-ence of a carbon-rich layer on top of the Ru surface at the end of the CpRu(CO)2Et pulse, additional experiments are recommended using in situ diagnostics sensitive to surface species. For example, in situ Fourier trans-form infrared spectroscopy or in situ X-ray fluorescence of the surface spe-cies at the end of each half-cycle could corroborate the mechanism pro-posed. In situ thermal programmed desorption performed between each half-cycle may also provide more direct evidence for the mechanism pro-posed in Chapters 6 and 7.

On the basis of the surface science and catalysis literature on Ru surfaces, we speculated on specific properties of the Ru ALD process in Chapter 7: the role of RuO2 and Ru due to the different catalytic activity of both mate-rials, the deposition temperature dependence, and on the reactions re-sponsible for the ALD saturation behavior. These speculations might also be substantiated or discredited during such surface-sensitive experiments, allowing for the mechanism to be refined. Ultimately, the Ru ALD processes, optimized precursor usage and film properties could be tailored further on the basis of a better understanding of the surface reactions.

Page 155: Atomic layer deposition of ruthenium films : properties and surface

8. CONCLUSIONS AND RECOMMENDATIONS

148

Provided that the reaction mechanism proposed in this dissertation is based on the right assumptions, this mechanism should also hold in a more general way for other metalorganic precursors than CpRuEt(CO)2. Therefore, similar mechanistic studies as the one presented in Chapter 6 and 7 could be carried out for other Ru precursors and compared to this work.

References

(1) Hong, T. E.; Choi, S.-H.; Yeo, S.; Park, J.-Y.; Kim, S.-H.; Cheon, T.; Kim, H.; Kim, M.-K.; Kim, H. ECS J. Solid State Sci. Technol. 2013, 2 , P47‒P53.

(2) Eom, T. K.; Sari, W.; Choi, K. J.; Shin, W. C.; Kim, J. H.; Lee, D. J.; Kim, K. B.; Sohn, H.; Kim, S. H. Electrochem. Solid-State Lett. 2009, 12, D85‒D88.

(3) Swerts, J.; Delabie, A.; Salimullah, M. M.; Popovici, M.; Kim, M.-S.; Schaekers, M.; Van Elshocht, S. Electrochem. Solid-State Lett. 2012, 1 , P19‒P21.

(4) Popovici, M.; Swerts, J.; Redolfi, A.; Kaczer, B.; Aoulaiche, M.; Radu, I.; Clima, S.; Everaert, J.-L.; Van Elshocht, S.; Jurczak, M. Appl. Phys. Lett. 2014, 104, 082908.

(5) Mackus, A. J. M.; Garcia-Alonso, D.; Knoops, H. C. M.; Bol, A. A.; Kessels, W. M. M. Chem. Mater. 2013, 25, 1769‒1774.

(6) Beyene, H. T.; Weber, J. W.; Jongen, J. C. B.; Verheijen, M. A.; Van de Sanden, M. C. M.; Creatore, M. Nano Res. 2012, 5, 513‒520.

(7) Hövel, M.; Alws, M.; Gompf, B.; Dressel, M. Phys. Rev. B 2010, 81, 035402.

Page 156: Atomic layer deposition of ruthenium films : properties and surface

8. CONCLUSIONS AND RECOMMENDATIONS

149

Page 157: Atomic layer deposition of ruthenium films : properties and surface

8. CONCLUSIONS AND RECOMMENDATIONS

150

Page 158: Atomic layer deposition of ruthenium films : properties and surface

151

Summary Atomic Layer Deposition of Ruthenium Films Properties and Surface Reactions

Over the last decade, the downscaling of electronic devices by the semi-conductor industry has been the driving force to explore the deposition of na-nometer-sized thin films and 3D-structures. For long, the microelectronics in-dustrial roadmap has been dominated by Si-based materials, such as SiO2 and SiNx dielectrics, and conductors such as doped polycrystalline silicon and TiN, which are both highly Si-compatible. However, on the nanometer scale, these materials start to reach their physical limitations in terms of next-generation device requirements, especially in terms of leakage current, dielectric constant, conductivity and 3D-conformaility. Therefore, great efforts have been and are being undertaken in the investigation of new materials. Here, atomic layer deposition (ALD) has emerged as a versatile deposition technique able to cope with the demand for atomically thin films and nanostructures.

Pt-group metals, such as platinum (Pt), iridium (Ir), palladium (Pd) and ru-

thenium (Ru), are catalytic in nature and are widely used in heterogeneous ca-

talysis, but they can also be applied as electrodes in microelectronic devices.

ALD of these metals yields high-quality films and benefits from excellent thick-

ness control at the atomic level. However, the ALD processes still call for further

optimization and tailoring of the film properties. To this end, Ru ALD using a

metalorganic precursor and O2 as a co-reactant was selected for investigation

in this dissertation.

The work of this dissertation can be divided into three distinct parts. The

first part includes the deposition of Ru-films with O2 gas and O2 plasma, and a

comparison of both Ru ALD processes and the respective material properties.

The second part presents the use of in situ spectroscopic ellipsometry as a

means to determine the Pt, Pd and Ru film thicknesses that evolve during ALD

growth. The third part deals with the investigation of the surface chemistry

Page 159: Atomic layer deposition of ruthenium films : properties and surface

SUMMARY

152

leading to Ru film growth, which could serve as a representative benchmark for

ALD growth mechanisms of other Pt-group metals.

This dissertation has focused on Ru ALD using (C5H5)Ru(CH2CH3)(CO)2, also written as CpRuEt(CO)2, as a precursor. In most experiments O2 was used as a co-reactant and the process can serve as a model system for other Ru ALD pro-cesses using metalorganic precursors and O2. The O2-gas based process was compared with the O2-plasma based ALD process in terms of thickness in-crease, nucleation delay and the material properties investigated included the resistivity, the mass density, the crystal orientation and the film roughness. The thermal and plasma-assisted ALD processes delivered high growth per cycle values (~1 Å/cycle) which were found to be a clear benefit of these ALD pro-cesses compared to other existing processes using different precursors. With the plasma-assisted process the nucleation delay could be reduced considera-bly, however, at the expense of a high surface roughness of the films. Therefore, the thermal ALD process with O2 gas was chosen as the main research topic in this thesis.

In order to monitor the film growth in situ, the growth of Ru, Pt and Pd films was studied by spectroscopic ellipsometry (SE). To take into account the change in dielectric function of metal films below 10 nm in thickness, the data were parametrized using polynomial base functions. This variational parametrization of the SE data allowed the observation of the initial incubation time characteristic for ALD of metals. Not only could accurate thickness values be extracted but also the dielectric functions of the films could be retrieved from the parametrization at any stage of the metal deposition. Furthermore, by adding Fourier-transform infrared reflectance data to the SE data, it was possi-ble to probe the optical conductivity of the film and to obtain a unique oscilla-tor-based parametrization of the three types of metal films. Most of these val-ues, conductivities and oscillator parameters, were in agreement with previous reports.

The surface science and catalysis community has provided a vast amount of information on the surface reactions occurring on (O-covered) Ru surfaces. One approach adopted in this dissertation was to apply insights on the Ru sur-face chemistry gained from the available literature to the experimental data obtained during the Ru ALD process. Basic properties of Ru surfaces were in-troduced and specific reactions relevant to the reaction mechanism were re-

Page 160: Atomic layer deposition of ruthenium films : properties and surface

SUMMARY

153

viewed. Surface reactions on Pt-group metal surfaces can be quite complex due to their unique electronic d-band structure which can lead to the dissocia-tive chemisorption of adsorbing gaseous species, such as O2 and CxHy. An ex-ample of the latter is the liberation of H2 as a result of surface dehydrogenation reactions and the carbon layer remaining on the Ru surface, once the dehydro-genation reactions are terminated.

Quadrupole mass spectrometry (QMS), including experiments with the ox-ygen isotope 18O2, was used to probe the gas-phase products originating from the surface reactions during Ru thermal ALD. In agreement with other Ru ALD studies, it was found that the process relies on the dissociative chemisorption of O2 on the surface during the oxygen exposure. The resulting surface O-atoms are reactive enough to combust the hydrocarbon ligands to CO2 and H2O. These combustion products were also observed during the precursor pulse confirming the presence of O on the metal surface at the start of the ALD cycle. In addition to breaking O-O bonds, the QMS data substantiated that the metal’s catalytic activity was also responsible for the scission of C-H bonds of the hydrocarbon ligands adsorbed at the surface. These C and H atoms were formed to be either oxidized to CO2 and H2O or the H atoms desorbed from the surface as H2. This finding implies that the overall reaction mechanism of Ru ALD can be summarized by catalytic combustion reactions, occurring in each half-reaction. These reactions consist of the dissociative chemisorption of spe-cific species on a catalytic surface (organic ligands during the Ru pulse, and ox-ygen in the O2 pulse) accompanied with the combustion of C- and H-rich spe-cies by the oxygen atoms at the surface.

This dissertation has highlighted various facets of Ru ALD which most likely hold for ALD of Pt-group metals in general. The results and conclusions pre-sented and derived in this thesis work are important to probe and control film growth of Pt-group metals by ALD.

Page 161: Atomic layer deposition of ruthenium films : properties and surface

154

List of publications related to this work “Good writing is rewriting.” ‒ Truman Capote Atomic Layer Deposition of Platinum Nanoparticles on Titanium Oxide and Tungsten Oxide Using Platinum(II) Hexafluoroacetylacetonate and Formalin as the Reactants

V. R. Anderson, N. Leick, J. W. Clancey, K. E. Hurst, K. M. Jones, A. C. Dillon, and S. M. George, J. Phys. Chem. C, 17 8960 (2014)

Catalytic Combustion Reactions During Atomic Layer Deposition of Ru Studied Using 18O2 Isotope Labeling

N. Leick, S. Agarwal, A.J.M. Mackus, S. E. Potts, and W. M. M. Kessels, J. Phys. Chem. C, 41, 21320 (2013)

Plasma-Assisted Atomic Layer Deposition of SrTiO3: Stoichiometry and Crystallinity Studied by Spectroscopic Ellipsometry

V. Longo, N. Leick, F. Roozeboom, and W. M. M. Kessels, ECS J. Solid State Sci. Technol., 2, N15 (2013)

Influence of Oxygen Exposure on the Nucleation of Platinum Atomic Layer Deposition: Consequences for Film Growth, Nanopatterning, and Nanoparticle Synthesis

A. J. M. Mackus, M. A. Verheijen, N. Leick, A. A. Bol, and W. M. M. Kessels, Chem. Mater. 9, 1905 (2013)

Page 162: Atomic layer deposition of ruthenium films : properties and surface

155

Dehydrogenation Reactions during Atomic Layer Deposition of Ru Using O2

N. Leick, S. Agarwal, A. J. M. Mackus, and W.M.M. Kessels, Chem. Mater. 24, 3696 (2012)

Catalytic Combustion and Dehydrogenation Reactions during Atomic Layer Deposition of Platinum

A.J.M. Mackus, N. Leick, L. Baker, and W.M.M. Kessels, Chem. Mater. 24, 1752 (2012)

Atomic layer deposition of Ru from CpRu(CO)2Et using O2 gas and O2 plasma

N. Leick, R.O.F. Verkuijlen, L. Lamagna, E. Langereis, S. Rushworth, F. Roozeboom, M. C. M. van de Sanden, and W. M. M. Kessels, J. Vac. Sci. Technol. A 29, 021016-1 (2011)

Bis(cyclopentadienyl) zirconium(IV) amides as possible precursors for low pressure CVD and plasma-enhanced ALD

S. E. Potts, C. J. Carmalt, C. S. Blackman, F. Abou-Chahine, N. Leick, W. M. M. Kessels, H.O. Davies, and P.N. Heys, Inorg. Chim. Acta 363, 1077 (2010)

Page 163: Atomic layer deposition of ruthenium films : properties and surface

156

Page 164: Atomic layer deposition of ruthenium films : properties and surface

157

Acknowledgements “Some people grumble that roses have thorns; I am grateful that thorns have roses.” ‒ Alphonse Karr

Because you hold this dissertation in your hands, I can say: “All is well that ends well.” Had anyone asked about my thesis at any moment within the last two years, its completion would have been questionable. This entire enterprise seemed like a big black shadow haunting me every single day. And out of the blue, this book came to life. Its content, its shape and its flavor were greatly influenced by a multitude of people. This thesis work has definitely been the fruit of interactions with many different people; and in the following lines I would like to express my gratitude to a selected number of these people.

I would like to start with my daily supervisor, Erwin Kessels, who may have been very frustrated with many quirks of mine. Erwin, without you, this work would not have been possible, on all levels possible. You’ve continuously set an example of perseverance and of hard work that was impossible for me to meet. You taught me the importance of details, the way to interpret data and present a scientific story. I would also like to thank you for your constant availability via emails. I always knew that I could count on an answer within 48 hours at the latest, regardless your GPS coordinate.

From my second promoter, Fred Roozeboom, I learned the importance of enthusiasm and ... ofvanadium oxide. Fred, the discussions with you were always full of meanders, and each of them was very useful and helpful, both scientifically and socially. I must thank you for your critical chemical point of view, especially for teaching me about the Mars-Van Krevelen-mechanism which I would probably have missed without you.

Adriana Creatore and Richard Engeln, the staff-members who gave me various opportunities to revise my perspective on life and on science, be it by making me laugh or telling me anecdotes of other people. Adriana, thank you for being such a great room-mate at conferences! Richard, thank you for being there... physically or virtually, I know you’ve always had an eye on me.

Sumit Agarwal, who I had the pleasure to work with over a period of 6 months, has great inertia. Sumit, once started, nothing was stopping you from

Page 165: Atomic layer deposition of ruthenium films : properties and surface

158

getting “things” done... from going over the experimental procedure to data interpretation, everything was checked thrice or more, if necessary. The same holds for your writing habits. On the day you decided to write up the joint paper, you not only made sure that it got done, but you also made justice to the adjective “joint”. Doing “live” corrections, directly on the Word-file, with you was extremely valuable to my learning process and was also extremely time efficient. Thank you for allowing me to witness your working habits!

During my 3rd year as a PhD-student, I was lucky enough to spend three months as a visiting student at Steven George’s ALD research group. Steve, I would like to thank you and your students for welcoming me in your group with open arms, and for your composure, your trust and optimism in your students.

Janneke, Ries, Joris, Wytze and Cristian, thank you for standing by my side in the course of my experimental work. As you know, it would have been impossible for me to open a precursor bubbler, hadn’t it been for you! Wytze, thank you for getting me started at the FlexALTM system and its in situ diagnostics. Cristian, I am so grateful that you went along with any special request I came up with; be it by disconnecting the gate valve of the plasma chamber for the QMS studies or hooking up and adapting the manometer for the oxygen isotope, to name a few. Thank you.

Jeanne and Lianne, you never got annoyed when explaining once again how to fill in a certain form. You never got tired of the inquiry about Erwin’s whereabouts! Thank you both for your patience with me and for your good mood.

When I started to work at PMP, I was assigned a room with great room mates. Gianfranco, thank you for hiding your Italian cursing habits for 3 weeks only; I would have missed them. Thanks to you, I started swearing in Italian at home, too. Nick, thank you for starting the habit of “the quote of the week” in our room, and thank you for sharing music with me. I loved it, along with our discussions which were not only interesting but also very helpful! Jan-Willem, thank you for your “gezelligheid” and putting people at ease. We really got to know each other over the SE course; of course driving for 4 hours in car makes one talk. Thank you for your moral support but also for your sense of humor; you’ve become a close friend over the years. You’ve lent a shoulder for me to cry on, but you also made sure that we could crack up about anything ranging

Page 166: Atomic layer deposition of ruthenium films : properties and surface

159

from fonts to green grass! On a more professional note, thank you also for your helpful input on the ellispometry work of this thesis.

Apart from my room-mates, so many other colleagues made this PhD-journey a real treat. Thank you Harm and Stephen for your patience in answering all the questions I came to you with. You both also had the right daily ratio of laughing and working. This is one reason why our trip to Korea remains a joyful memory. Adrie, thank you for your curiosity which made our discussions both, helpful and stimulating. I’ve enjoyed working with you very much, and without your input the interpretation of the reaction mechanism of Ru growth during ALD wouldn’t have come as far as it actually has. Jan-Pieter, I don’t even remember how we started to get along; probably first by laughing together at something random in the corridors. When I try to think back, it just seemed like you’ve always been there with your good mood, high spirits and your graphics-design skills.

Mikahil, thank you for opening up in the course of time and letting me in your discussions. Thank you for encouraging me to join the swimming lessons at the sports center; it was a lot of fun!

Stefan, Gott sei Dank kamst du noch in die Gruppe hinzu. So konnte ich nicht nur mein Deutsch aufrechterhalten sondern auch über Atmosphären-druckplasmas reden. Es hat gut getan, sich mit dir auszutauschen, egal worüber!

Aafke, where to begin...probably at the microwave in the “koopzaal” in N-laag. This is where we first met and where you explained to me what your project encompassed. Who would have guessed that we would laugh and cry together through the next four years? Who would have guessed that I would spend ~1,5 years living at your house during the week and have so much fun? At the time, I didn’t. But I am glad to have gained a true friend who shares my taste in eating chocolate (in its various forms), reading fiction, drinking tea, gossiping and discussing science. Thank you!

Also, my big thanks to Ivo, Matthieu, Valentino, Chai, Srinath, Erik, Terje, Anu, Kashish for having made this PhD period so much fun! Virginia, thank you so much for your warm welcome during my stay in Boulder!

Ich möchte mich ebenfalls bei den Schülern des Königlichen Athenäums Sankt Vith bedanken, die ich von 2013 bis 2014 unterrichtet habe. Tag ein, Tag aus habt ihr mir gezeigt, wie einzigartig und wie wertvoll Lernerfahrungen sind.

Page 167: Atomic layer deposition of ruthenium films : properties and surface

160

Finalement, merci Maman pour m’avoir encouragé à commencer une thèse.

Merci Marraine pour m’avoir encouragé à la finir, par tes paroles et tes actes. Merci Lena de m’avoir montré la vie à ta manière, pour m’avoir encouragé,

sans le savoir, à donner du meilleur de moi-même, à la maison comme au travail. Depuis ta naissance, ta gaité me fait sourire et me donne de la force. Merci mon ange !

Page 168: Atomic layer deposition of ruthenium films : properties and surface

161

Page 169: Atomic layer deposition of ruthenium films : properties and surface

162

Page 170: Atomic layer deposition of ruthenium films : properties and surface

163

Curriculum Vitæ

04 October 1982 Born in Thionville, France

2002 Allgemeine Hochschulreife (Abitur), Erzbischöfliches Ursulinengymnasium, Köln, Germany

2002 ‒ 2007 Master of Science degree Fundamental and Applied Physics, University of Parix XI, Paris, France

Traineeship (June 2003) in the group of Prof. Schneider, Max Planck Institute for Radioastronomy, Bonn, Germany

Traineeship (June 2004) in the group of Dr. Konczewicz, Laboratory for Semiconductors, University of Montpellier, Montpellier, France

Three months (Spring 2006) traineeship in the group of Dr. Wenger at the ATLAS division of the European Organization for Nuclear Research (CERN), Geneva, Switzerland

Traineeship (June 2006) in the group of Prof. Pop, Laboratory for Crystallography, Babes-Bolyai University, Cluj-Napoca, Romania

Master thesis project in the group of Prof. Rousseau at the Low Temperature Plasma Laboratory, École Polytechnique, Paris, France

2008 ‒ 2012 PhD project in the group Plasma and Materials processing, Department of Applied Physics, Eindhoven University of Technology, Eindhoven, The Netherlands

Three months (spring 2011) working visit to the group of Prof. S. George, Department of Chemistry, University of Colorado, Boulder, CO, U.S.A.

Second place of “Spectroscopic ellipsometry ‒ focus session” award at the 58th International Symposium American Vacuum Society (November 2011)

Page 171: Atomic layer deposition of ruthenium films : properties and surface

164